Satellite Cells and Skeletal Muscle Regeneration Nicolas A. Dumont,1 C. Florian Bentzinger,1,2 Marie-Claude Sincennes,1 and Michael A. Rudnicki*1,3 ABSTRACT Skeletal muscles are essential for vital functions such as movement, postural support, breathing, and thermogenesis. Muscle tissue is largely composed of long, postmitotic multinucleated fibers. The life-long maintenance of muscle tissue is mediated by satellite cells, lying in close proximity to the muscle fibers. Muscle satellite cells are a heterogeneous population with a small subset of muscle stem cells, termed satellite stem cells. Under homeostatic conditions all satellite cells are poised for activation by stimuli such as physical trauma or growth signals. After activation, satellite stem cells undergo symmetric divisions to expand their number or asymmetric divisions to give rise to cohorts of committed satellite cells and thus progenitors. Myogenic progenitors proliferate, and eventually differentiate through fusion with each other or to damaged fibers to reconstitute fiber integrity and function. In the recent years, research has begun to unravel the intrinsic and extrinsic mechanisms controlling satellite cell behavior. Nonetheless, an understanding of the complex cellular and molecular interactions of satellite cells with their dynamic microenvironment remains a major challenge, especially in pathological conditions. The goal of this review is to comprehensively summarize the current knowledge on satellite cell characteristics, functions, and behavior in muscle regeneration and in pathological conditions. © 2015 American Physiological Society. Compr Physiol 5:1027-1059, 2015.

Introduction Striated skeletal muscle is the most abundant tissue of the human body, representing approximately 35% to 45% of total body mass (238). In contrast to heart muscle, skeletal muscle is controlled by the somatic nervous system that enables voluntary contraction. There are more than 600 skeletal muscles of different sizes and contractile properties that allow for various functions such as locomotion, postural support, powerful or precise movements, and breathing. One of the most remarkable capacities of skeletal muscle is its very high adaptive potential. For example, resistance training is well known to induce muscle hypertrophy and increase muscle strength. In contrast, many different diseases or conditions such as cancer, chronic obstructive pulmonary diseases, heart failure, aging, denervation, or disuse can lead to muscle atrophy and loss of function (10, 98). Next to their high plasticity, skeletal muscles also possess exceptional regenerative capacity. Only few weeks after a major injury destroying fiber integrity, skeletal muscle structure and function can be completely restored (24). The mediators of muscle regeneration are a population of small adult stem cells, termed satellite cells, located in close proximity to muscle fibers. Under resting conditions, adult satellite cells are quiescent, but they can quickly re-enter the cell cycle following injuries or growth signals. Activated satellite cells will migrate extensively, proliferate, differentiate, and fuse to form regenerating myofibers. Only a subset of self-renewing cells is able to withstand differentiation during muscle injury and will return to quiescence after completion of muscle regeneration to replenish the satellite cell pool and to prepare the tissue for a subsequent injury. Importantly, the fate of satellite

Volume 5, July 2015

cells is greatly influenced by intrinsic and extrinsic factors. Dynamic interactions with inflammatory cells, stromal cells, trophic signals, and extracellular matrix (ECM) components guide satellite cells through the regeneration process. However, many pathological conditions, such as muscular dystrophies (MDs) or muscle wasting, provide inadequate cues to the satellite cells and thereby impair their regenerative potential. This review will first describe satellite cell characteristics and their developmental origin. Subsequently, satellite cell behavior during the different steps of muscle regeneration will be discussed and will be followed by a review of the intrinsic and extrinsic factors regulating satellite cell fate during adult myogenesis. Finally, the effects of diverse degenerative conditions on satellite cells, as well as the different therapeutic avenues will be summarized.

Historical Perspective It has been known since the 19th century that muscle tissue possesses great regenerative potential even following serious traumas such as ischemia, burns, incisions, or crush injuries * Correspondence

to [email protected] Centre for Stem Cell Research, Ottawa Hospital Research Institute, Ottawa, Ontario, Canada 2 Nestl´ e Institute of Health Sciences, EPFL Campus, Lausanne, Switzerland 3 Faculty of Medicine, Department of Cellular and Molecular Medicine, University of Ottawa, Ottawa, Ontario, Canada Published online, July 2015 (comprehensivephysiology.com) DOI: 10.1002/cphy.c140068 Copyright © American Physiological Society. 1 Sprott

1027

Satellite Cells and Skeletal Muscle Regeneration

(264). Initially, the basic characteristics of muscle regeneration such as leukocyte infiltration, fibrotic tissue formation, revascularization, and repair of damaged fibers were characterized (64). However, the origin of de novo myofiber formation was unclear and various hypotheses were proposed, for instance, the splitting of intact myofibers, the elongation of damaged myofibers, the formation of new myofibers from circulating leukocytes, or the proliferation of myonuclei. The first evidence for the presence of muscle resident cells with myogenic potential was revealed by an in vitro study showing that culture of small pieces of chick embryonic muscles results in expansion of a myoblast population and formation of multinucleated myotubes (175). Alexander Mauro was the first to identify in vivo a population of small mononuclear cells that he named satellite cells based on their location between the plasma membrane of the myofiber and the basement membrane (192). First described in frogs, these satellite cells were also observed in skeletal muscles from various mammalian species, including humans (140). Mauro suggested that satellite cells are a dormant cell type that could be activated to promote regeneration upon fiber damage. The development of radioactive nucleoside labeling techniques using tritiated thymidine demonstrated that only mononuclear cells, and not myonuclei, were actively synthesizing new DNA strands (29). However, a short radioactive pulse followed by muscle injury showed that a radioactive signal could be observed in the nuclei of newly regenerated myotubes, which suggested that activated satellite cells are able to fuse to damaged myofibers. The fusion ability of myoblasts was confirmed using allophenic mouse models, where hybrid fibers formed from the fusion of cells with different genetic backgrounds were observed (45, 202). Subsequently, single fiber isolation techniques were useful to

Osteotendinous junction

Comprehensive Physiology

confirm that satellite cells lying underneath the basal lamina are the cell type responsible for myoblast proliferation and myotube formation (31). The discovery of myogenic regulatory factors (MRFs) and paired box proteins then allowed to further characterize myogenic cell behavior during the different steps of myogenesis (102, 120, 198, 269, 316). More recently, another important hallmark of satellite cell behavior was discovered. Using transplantation experiments in radiation-ablated muscles, it was shown that freshly isolated satellite cells possess a high self-renewal capacity (67, 163, 255). These results were also accompanied by the discovery of considerable heterogeneity within the satellite cell population. Satellite cells can, for instance, be subdivided into a subpopulation of satellite cells that are already committed to become myogenic progenitors and a subpopulation of satellite stem cells with higher self-renewal capacity (163). The latter population was also shown to be able to perform asymmetric division and give rise to committed satellite cells (163). Thus, the ability of a subset of satellite cells to self-renew and to perform asymmetric division confirmed these cells as adult stem cells. The following sections of this review will describe our current understanding of satellite cell biology.

Anatomy Skeletal muscles are mostly composed of long parallel cylindrical fibers attached together by a complex and structured ECM (Fig. 1). Each fiber is enveloped by layers of ECM, namely, the endomysium and the basement membrane (188). The endomysium is the most exterior membrane and is mainly composed of collagen fibrils interlaced with a vast network of Myotendinous junction

Skeletal muscle

Bone Basal lamina

Joint

Satellite Blood cell vessel

Multinucleated muscle fiber

Figure 1 Skeletal muscle anatomy. Schematic illustrating the structure of skeletal muscle. Skeletal muscles are divided into bundles containing numerous myofibers. Myofibers are multinucleated cylindrical cells that are surrounded by a vast network of blood vessels and nerves. Satellite cells are small mononuclear cells located between the plasmalemma of the myofibers and the basal membrane.

1028

Volume 5, July 2015

Comprehensive Physiology

Satellite Cells and Skeletal Muscle Regeneration

Figure 2 Satellite cell markers. Images of single extensor digitorium longus (EDL) myofiber labeled for different markers of satellite cells. The left image shows a multinucleated myofiber with two satellite cells stained by Pax7 antibody (red) and a nuclear dye (DAPI, blue). The right picture displays a satellite cell stained with M-cadherin (red) on the apical side and α7-integrin (white) on the basal side.

blood vessels and nerves. Lying between the endomysium and the fiber membrane (plasmalemma), the basement membrane can be further divided into the reticular lamina (collagen fibrils) and the basal lamina (collagen IV, fibronectin, laminin, and proteoglycans) (167, 262). The plasmalemma contains numerous proteins such as the dystroglycan-sarcoglycandystrophin complex that are involved in linking the fiber to the ECM. Skeletal muscle fibers are distinct from every other cell type of the body. Their interior is mainly constituted of myofibrils, which contain actin and myosin filaments responsible for muscle contraction. Muscle fibers are very long cells that can measure up to 20 to 30 cm in humans (131). To fulfill the high-protein synthesis demand of these long contractile cells, muscle fibers possess numerous myonuclei. Myonuclei cover a relatively constant territory ranging from 2 × 104 to 5 × 104 μm3 , although the existence of a defined nuclear domain is controversial (121, 178, 193). Importantly, myonuclei are postmitotic and are not able to divide. The satellite cells that populate these fibers are wedged between the basal lamina and the plasmalemma and more than 80% to 90% of them are located only a few micrometers away from a blood vessel (Fig. 1) (61). Satellite cells constitute 2% to 10% of total myonuclei, which represents approximately 2 × 105 to 1 × 106 satellite cells per gram of muscle (25,128). However, the proportion of satellite cell depends on many factors such as muscle type, age, and species. For example, slow oxidative muscles contain more satellite cells than fast glycolytic muscles, probably to satisfy the repair demand of these highly active postural muscles (109).

Satellite Cell Characteristics Apart from their anatomic location, satellite cells can also be distinguished based on their cellular characteristics, for instance, their high nucleus-to-cytoplasm ratio. In addition,

Volume 5, July 2015

many different proteins were discovered as markers of satellite cells. Some markers are located in the nucleus such as the paired box proteins PAX3 and PAX7 (Fig. 2). In adult muscles, Pax7 is highly and constantly expressed in satellite cells, while Pax3 is usually expressed at very low levels except in a subset of muscles such as the diaphragm (152,206). Other transcription factors such as myogenic factor 5 (Myf5) and myogenic differentiation 1 (MyoD) can be specifically expressed by myogenic cells, but are not expressed in quiescent satellite cells. Finally, several cell surface membrane proteins were identified as markers of satellite cells such as, M-cadherin, α7- and β1-integrins, c-Met, C-X-C chemokine receptor type 4 (CXCR4), syndecan-3 and -4, calcitonin receptor, calveolin-1, CD34, vascular cell adhesion molecule-1 (VCAM1), and neural cell adhesion molecule-1 (NCAM1) (Fig. 2) (104, 113). Histologically, anti-laminin combined with anti-Mcadherin immunostaining allows for the identification of satellite cells in their niche (Fig. 3). However, Pax7 is the most widely recognized intracellular marker because it is ubiquitously expressed in all satellite cells in a wide variety of species. It is important to keep in mind that Pax7 is also expressed in activated satellite cells and is thus not a specific marker for quiescent satellite cells. Transgenic mice carrying Pax7 promoter-driven fluorochrome expression, such as Pax7zsgreen mice or tamoxifen-inducible fluorochrome expression in Pax7creER -YFP (yellow fluorescent protein) mice are very convenient research tools to isolate and/or analyze satellite cell behavior (23, 37). However, without these reporter mouse models Pax7 is an unpractical marker for viable satellite cell isolation by FACS due to its nuclear location. For cell sorting, positive selection for the satellite cell surface markers α7-integrin and CD34 combined with negative selection for markers of other cell types CD45 (hematopoietic lineage), CD11b (leukocytes), CD31 (endothelial cells and leukocytes), and Sca-1 (mesenchymal and hematopoietic stem cells) has been shown to isolate a pure population of satellite

1029

Satellite Cells and Skeletal Muscle Regeneration

Comprehensive Physiology

Figure 3

The satellite cell niche. Cross-section of a tibialis anterior (TA) muscle immunostained with laminin (white), M-cadherin (green), Pax7 (red), and a nuclear dye (DAPI, blue). All muscle fibers are surrounded by laminin. The insert in the top right corner shows a magnification of a satellite cell (Pax7-positive) in its niche.

cells from skeletal muscles (229). A monoclonal antibody, SM/C-2.6, was also developed to bind specifically to satellite cells although its antigen is still unknown (103).

Developmental Origin of Satellite Cells Three germ layers are formed during mammal embryogenesis: the ectoderm, mesoderm, and endoderm. The mesoderm can be further subdivided into lateral, intermediate, and paraxial mesoderm. Bilateral paired blocks, named somites, form from the paraxial mesoderm and develop along the anteriorposterior axis (head to tail) of the embryo (Fig. 4) (25). Somites give rise to a variety of tissues such as vertebral bones, tendons, cartilage, dermis, and skeletal muscles. All skeletal muscles, except for some head muscles, are derived from the dorsal portion of the somites, named the dermomyotome. Dorsal muscles arise from the epaxial part of the dermomyotome (adjacent to the neural tube), whereas limb muscles originate from the hypaxial part of the dermomyotome. A vast network of transcription factors controls myogenic cell specification, determination and differentiation (Fig. 5). Paired box domain transcription factors PAX3 and/or PAX7 are expressed by dermomyotome cells and are important upstream regulators of muscle tissue formation. Pax3 is critical during muscle development especially for the formation of hypaxial domain. Pax3 loss-of-function leads to absence of diaphragm and limb muscles, but this effect is less severe on epaxial-derived muscles (35, 301). Pax3 has been demonstrated to target the hepatocyte growth factor (HGF) receptor c-met, which is essential for the delamination and migration of the myogenic progenitors (96). Moreover,

1030

Figure 4

Embryonic muscle development. The image shows a mouse embryo transgenic for a fluorescent expression reporter (TdTomato) of the myogenic transcription factor Myf5 at 10.5 days after fertilization. At this developmental stage Myf5 is highly expressed in the somites. Somites develop antero-posteriorly, with the anterior somites being the most mature ones and the posterior somites the least developed.

Pax3 has been shown to regulate the expression of Myf5 and MyoD in the embryos, which triggers myogenic determination (239). Despite its essential role in muscle formation, absence of Pax3 in postnatal muscles does not impair adult muscle regeneration (245, 290). On the other hand, Pax7 deletion is much less severe for muscle development in the embryo, but is absolutely essential for satellite cell formation (162, 221, 244, 269). In adult myogenesis, the transcriptional

Pax3/7 + MRF–

Myf5 Pax3/7

Myogenin MRF4

MyoD Myogenic specification

Myogenic determination

Myogenic differentiation

Figure 5

Genetic programme regulating skeletal muscle stem cell fate. Hierarchical network of the transcription factors controlling embryonic progenitor cell fate during limb and trunk muscle development. Pax3 and Pax7 control the myogenic specification of embryonic progenitors. Pax3/7 can directly target Myf5 and MyoD expression and thus myogenic determination. Myf5 also acts partially upstream of MyoD and can trigger its expression. Upregulation of myogenin and MRF4 induces myogenic differentiation. A population of Pax3/7+ progenitor cells does not express MRFs and gives rise to satellite stem cells during late fetal myogenesis.

Volume 5, July 2015

Comprehensive Physiology

dominance of Pax7 over Pax3 is explained by its higher binding affinity to the homeodomain binding motif (283). Pax7-null mice show gradual loss of satellite cells, reduced fiber size, and impaired muscle regeneration resulting in early postnatal death (122, 309). Lineage tracing experiments confirmed that Pax3-positive cells are essential for embryonic myogenesis, while Pax7-positive cells are more important for late fetal and postnatal myogenesis (137). Thus, it is hypothesized that Pax3-positive founder cells form a template of initial fibers, while Pax7-positive cells form secondary fibers and generate the satellite cell population (25, 30). Progression of embryonic progenitors (Pax3/7+ cells) through the myogenic program is regulated by different MRFs. MRFs are a family of four basic helix-loop-helix transcription factors: Myf5, MyoD, myogenin, and myogenic regulatory factor 4 (MRF4). Myf5 is the first MRF to be expressed, initially in the epaxial domain and subsequently in the hypaxial domain of the dermomyotome (42). MyoD is expressed after Myf5 in the hypaxial and epaxial domains. Myf5 and MyoD are critical factors for myogenic cell determination. Knockout of Myf5 or MyoD alone have a relatively mild effect on muscle development due to the functional overlap between these two factors (41, 251, 252). However, Myf5:MyoD double knockout mice show complete lack of skeletal muscle formation (252). Moreover, Myf5:MyoD double knockout mice do not express myogenin, which demonstrates the hierarchal relationship among the MRFs. Myogenin is essential for the terminal differentiation of myogenic cells and myogenin-null mice quickly die after birth from severe and global muscle deficiency (127). Similarly, MRF4null mice display myogenin overexpression, which suggests a role for this factor in late differentiation (242, 331). It has been suggested that MRF4 could also play a partial role as a determination factor in absence of Myf5 and MyoD; however, it has not been established whether it does the same in presence of Myf5 and MyoD (151). Globally, Myf5 and MyoD are determination factors required to establish myogenic identity, and act upstream of the differentiation factors myogenin and MRF4, but the exact hierarchical organization of the different MRFs is still a matter of debate (Fig. 5). While most Pax3/7+ cells express MRFs and commit to myogenic progenitors during development, there is a subset of Pax3/7+ cells that proliferates without upregulating the different MRFs (152, 246). While this might be true at least for the expression of Myf5, different reporter mice demonstrated that virtually all satellite cells transiently express MyoD prenatally (149). Nonetheless, it has been suggest that late in fetal development Pax3/7+MRF− cells align to the nascent myotubes and eventually adopt the satellite cell position under the basal lamina in postnatal muscles (Fig. 5) (246). Upon arrival to nascent muscles, the majority of Pax3/7+MRF− cells will downregulate Pax3 and quickly upregulate Myf5 (152). The presence of Pax7+Myf5- satellite stem cells in adult muscle might correspond to a lineage continuum of the Pax3/7+MRF− progenitors.

Volume 5, July 2015

Satellite Cells and Skeletal Muscle Regeneration

Overall, there are many similarities between embryonic and adult myogenesis and many common transcription factors and signaling pathways are involved. Nevertheless, there are also many differences, especially in the microenvironment of the cells and in their level of activity.

Satellite Cell Heterogeneity Originally considered as a homogeneous population, different satellite cell subpopulations have recently been characterized. Some evidence suggests that there is a heterogeneity in satellite cells residing in slow versus fast muscles, for example, that satellite cells from slow muscles give rise to myofibers expressing slow myosin isoforms while satellite cells from fast muscles tend to form fast myosin (17, 97, 135, 148, 250). However, there is no clear data demonstrating intrinsic differences of satellite cells based on fiber type, particularly not in humans (36, 95). Expression of different markers also varies considerably between satellite cells. Paired box protein homologs Pax3 and Pax7 expression show a discrepancy depending on the muscle. For instance, more than 80% of satellite cells in diaphragm coexpress Pax3 and Pax7, while Pax3-expressing satellite cells are not found in hindlimb muscles (162). Expression of other satellite cell markers such as CD34, Myf5, and Mcadherin also vary among the satellite cell population (19). These markers are almost always coexpressed together in quiescent satellite cells, but they account only to approximately 80% of the total satellite cell population. Subpopulations of CD34neg , Myf5neg , and/or M-cadherinneg satellite cells were discovered and it was speculated that these cells could represent uncommitted progenitors responsible for the maintenance of the satellite cell population. Notably, the proportion of CD34neg and M-cadherinneg satellite cells decreases during aging (237). The presence of a satellite stem cell population within the satellite cell pool has been confirmed using Myf5-cre/R26R-YFP mice in which Myf5cre expression leads to permanent labeling with YFP. Using this mouse line it has been shown that approximately 10% of total satellite cells have never expressed Myf5 during development or postnatally (163). Transplantation experiments showed that Myf5neg satellite cells possess higher selfrenewal and engraftment capacity and are less prone to precocious differentiation than Myf5-positive (Myf5pos ) satellite cells. Myf5neg satellite cells are also able to perform asymmetric divisions and give rise to Myf5pos satellite cells (163). Myf5neg satellite cells are also relatively slow dividing relative to Myf5pos cells. These results confirm that there is heterogeneity within the satellite cell population, with a subpopulation of committed satellite cells (Myf5pos ) and a subpopulation of satellite stem cells (Myf5neg ). Another transgenic mouse model showed that satellite cells could be subdivided based on their level of Pax7 expression. Pax7Hi cells possess lower metabolic rate and slower

1031

Satellite Cells and Skeletal Muscle Regeneration

Comprehensive Physiology

division ability than Pax7Lo satellite cells that rather tend to express higher levels of myogenin (247). Consistently, labeling of satellite cells with the proliferation markers PKH26 or BrDU showed that most activated satellite cells undergo fast division, but a minority are slow-dividing (219,266,277). This observation is supported by a study using label retention as marker of satellite cell proliferative history throughout animal lifespan (53). In this study, it was shown that a subpopulation of satellite cells divides less frequently and retains the labeling (label-retaining cells) while another subpopulation loses the labeling over time (nonlabel retaining cells). Label retention experiments also demonstrated that slow-dividing cells can generate distinct daughter cells that are more prone to differentiation (52). However, the immortal strand hypothesis suggesting that one daughter stem cell retains the template DNA while the more committed cell is composed of the newly synthesized DNA strand, is still debated and no clear mechanism describing this phenomenon has yet been demonstrated (166). Taken together, the idea that satellite cells are heterogeneous and can be divided at least into a committed progenitor and a muscle stem cell subpopulation is experimentally well supported.

Satellite Cells in Muscle Regeneration Skeletal muscles are able to quickly regenerate mature myofibers following massive injuries that completely destroy muscle structure and integrity (Fig. 6). Deletion models confirmed that this remarkable regenerative capacity relies on the myogenic potential of satellite cells. Pax7-null mice have small myofibers at birth and are not able to form a functional satellite cell pool, which leads to rapid postnatal death (within few weeks of life) (162, 269). Conditional knockout models were more helpful to delineate the role of satellite cells in adult muscle regeneration. Pax7 depletion in adult mice using inducible Pax7flox/CreERT2 tamoxifen-treated mice strongly impaired the regenerative potential of skeletal muscle, indicating that Pax7 is also essential for the normal function of postnatal satellite cells (122, 309). In agreement with these findings, conditional depletion of satellite cells using mice Uninjured

2d

carrying the diphtheria toxin A gene under the control of the Pax7 locus showed that absence of satellite cells in injured muscles results in strong decrease in myofiber number and size, accompanied by massive fibrosis and intracellular fat accumulation (172,209,259). The deleterious effects of satellite cell ablation could be rescued by satellite cell transplantation into the depleted muscles (259). Intriguingly, satellite cell depletion completely blunts the regeneration of damaged muscles but it does not significantly impair the short-term hypertrophic capacity of muscle fibers in the absence of injury (194). Mechanical overload induced similar increases in muscle fiber cross sectional area and muscle force in presence or absence of satellite cells over 2 weeks. These results suggest that muscle fibers possess adaptive potential, that is, in part independent of satellite cells. However, the increase in myonuclei number usually associated with fiber hypertrophy is not observed in satellite celldeficient muscles, suggesting that the hypertrophic process is perturbed and that the myofibers are probably less functional. Moreover, the model of satellite cell depletion used by the authors (conditional depletion using diphtheria toxin A gene under the control of the Pax7 locus) leads to 90% depletion, leaving the possibility that the remaining satellite cells could contribute to the hypertrophic process (194). Muscle regeneration can be divided in several stages that are characterized by the expression of different MRFs (Fig. 7). In the quiescent stage satellite cells are inactive but are poised to activation. Quiescent satellite cells generally express markers such as Pax7 and Myf5 (except for the Myf5neg satellite stem cell population). After muscle injury, satellite cells are stimulated by the various signals arising from the damaged environment. Satellite cells then migrate toward the injury site and re-enter the cell cycle to proliferate. At this stage, committed satellite cells are termed myoblasts and express the myogenic markers Pax7, and/or Myf5, and/or MyoD (Fig. 8). After the proliferative phase, myoblasts exit the cell cycle and differentiate into mature myocytes. This differentiation process is accompanied by a decreased expression of Pax7 and Myf5, and by an increase in myogenin (MyoG) and MRF4 levels (Fig. 9). Finally, myocytes fuse together to form multinucleated myotubes and/or fuse to damaged myofibers. 5d

2mo

Figure 6 Skeletal muscle regeneration. H&E staining (nuclei stained in dark blue and cytosolic proteins stained in red) illustrating the different stages of muscle regeneration following cardiotoxin injury. Two days following injury, myofiber structure is severely impaired and a multitude of mononuclear cells (myoblasts, inflammatory cells, fibroblasts, etc.) are present in the damaged area. At five days postinjury, there is the formation of small new myofibers with centralized nuclei and many cells remain in the exudate. Two months after the injury, muscle architecture is similar to uninjured muscle.

1032

Volume 5, July 2015

Comprehensive Physiology

Embryonic progenitors

Satellite Cells and Skeletal Muscle Regeneration

Satellite stem cells Pax7

Satellite cells Pax7 Myf5

Myoblasts Pax7 Myf5 MyoD

Myocytes MyoD Myogenin MRF4

Myotubes Myogenin MRF4 MyHC

Selfrenewal

Figure 7 Myogenic lineage progression. Following muscle injury quiescent satellite cells (Pax7+ and Myf5+/−) are activated and differentiate to myoblasts (Pax7+, Myf5+, and MyoD+). After several rounds of proliferation, myoblasts exit the cell cycle and become myocytes (Pax7−, MyoD+, myogenin+, and MRF4+). Myocytes can undergo a fusion process to form multinucleated myotubes (myogenin+, MRF4+, and MyHC+) that eventually mature into myofibers. The satellite stem cell subpopulation (Pax7+ and Myf5−) can also self-renew to replenish the satellite cell pool.

Figure 8

Myogenic regulatory factors in proliferating myoblasts. Images from single EDL myofibers cultured ex vivo for 60 h (dashed line delineates the myofiber membrane). Staining for Pax7 (green) and MyoD (red) shows that at this stage cells express either only Pax7 (arrowhead), only MyoD, or both (arrow).

At this stage, MyoD levels become reduced while myosin heavy chain (MyHC) and other contractile proteins start to be expressed. The balance between the different MRFs is critical to control cell fate. It was suggested that a high Pax7:MyoD ratio favors quiescence, an intermediate ratio enables proliferation but not differentiation, while a low Pax7:MyoD ratio promotes differentiation (218). During the myogenic lineage progression, satellite cell fate is influenced by a complex balance between intrinsic factors and external cues (24).

Satellite Cell Quiescence In adult resting muscles, satellite cells are in a quiescent state characterized by the absence of cell cycling (G0 phase), a low rate of metabolism and low RNA content (60). Genome-wide gene expression analysis comparing quiescent to activated satellite cells revealed more than 500 genes upregulated in quiescent satellite cells, which suggests that maintaining quiescence is an active process (104). These genes are mainly

Figure 9

Myogenic regulatory factors in differentiating myoblasts. Pictures from EDL isolated myofibers cultured ex vivo for 72 h (dashed line delineates the myofiber membrane). Staining for Pax7 (green) and myogenin (red) shows that cells that acquire the differentiation marker myogenin lose expression of Pax7.

Volume 5, July 2015

1033

Satellite Cells and Skeletal Muscle Regeneration

related to cell-cell interactions, formation of ECM, and regulation of cell growth. Some proteins such as Odz4 and the calcitonin receptor are almost exclusively expressed in dormant satellite cells and not expressed in activated satellite cells (325). Cell-cycle regulators, such as the retinoblastoma tumor suppressor protein Rb, are also highly enriched in quiescent satellite cells (104). In proliferating myoblasts, Rb protein is phosphorylated leading to its inactivation, while its dephosphorylation during terminal myogenic differentiation induces withdrawal from the cell cycle (71). Tamoxifen inducible ablation of Rb1 under the control of a Pax7 promoter leads to activation of quiescent satellite cells (134). Conditional Rb1 knockout triggers an expansion of the satellite cell population but also decreases the regenerative capacity of muscle due to impaired differentiation. Consistently, regulators of Rb protein, the cyclin-dependent kinase inhibitors (CKIs), are upregulated in quiescent satellite cells compared to activated satellite cells (104). Furthermore, the cell-cycle inhibitor p27kip1 is highly expressed in quiescent satellite cell (especially in the stem cell subpopulation) and knockout of this factor promotes proliferation and differentiation of satellite cells (52). Overall, there are many negative regulators that prevent cell cycle entry in quiescent satellite cells. The position of the satellite cells between the basal lamina and the fiber membrane suggests that the microenvironment of the cells could be involved in their regulation (205). For example, M-cadherin proteins present on both the satellite cell and the myofiber membrane interact together and “immobilize” the satellite cell (139). Adhesion molecules such as α7- and β1-integrins also interact with the basal lamina to stabilize the satellite cell in its niche. Moreover, quiescent satellite cells express high levels of tissue inhibitor of metalloproteinases (TIMPs) to block ECM degradation by metalloproteinases (MMPs) and therefore favor the integrity of the niche (223). A key signaling system involved in the control of satellite cell quiescence is the Notch pathway. The Notch ligand Delta1 is expressed at the fiber membrane and it appears likely that it interacts with Notch receptors present on the satellite cell membrane (70). In agreement with this idea, the Notch pathway is active in quiescent satellite cells. The cooperation of Notch receptor with Syndecan-3 is important to facilitate Notch signal transduction (233). Notch signaling activates the transcription factor Recombining binding protein suppressor of hairless (RBPJ) that triggers the activation of multiple genes such as Hes1, Hes5, Hey, and HeyL (319). Hey1 and HeyL double-knockout results in cell cycle re-entry of satellite cells and stimulates MyoD and myogenin expression (105). Similarly, Syndecan-3 knockout satellite cells display impaired Notch signaling leading to satellite cell exit from quiescence and lack of self-renewal (233). Thus, inhibition of the Notch pathway leads to activation and differentiation of satellite cells that ultimately results in the depletion of the satellite cell pool (33, 163, 208). On the other hand, overexpression of the constitutively active Notch intracellular domain (NICD)

1034

Comprehensive Physiology

Table 1

Pathways Involved in the Different Myogenic Stages

Myogenic stage

Pathways

References

Quiescent satellite cells

Notch

Conboy 2002, Fukada 2011, Bjornson 2012, Mourikis 2012, Wen 2012 Gopinath 2014 Shea 2010, Abou-Khalil 2009

FOXO3 ERK1/2 (Sprouty1) Satellite cell activation

P38α/β MAPK Calcium (NFATc) Akt/mTOR NF-κB

Jones 2006 Liu 2014 Machida 2003 Acharyya 2010

Myogenic cell proliferation

JAK/STAT

Serrano 2008, Sun 2007 Otis 1994, Guttridge 1999 Perdigureo 2007, Alter 2008 Gillepsie 2009

NF-κB JNK P38γMAPK Myogenic differentiation/ fusion

Canonical Wnt p38αMAPK Calcium (calcineurin/ CaMK)

Myofiber hypertrophy

Akt/mTOR Myostatin

Self-renewal

Noncanonical Wnt (planar cell polarity) Notch

Brack 2008 Wu 2000, Perdiguero 2007, Lluis 2005 Friday 2003, Friday 2000, Lu 2000 Park 2005, Von Maltzahn 2012 McPherron 1997, Bogdanovich 2002 Legrand 2009, Bentzinger 2013 Gopinath 2014, Wen 2012, Kuang 2007

This table summarizes the major pathways involved in the different myogenic stages; satellite cell quiescence, activation, proliferation, differentiation/fusion, fiber hypertrophy, and self-renewal.

downregulates MyoD expression, reduces entry in S phase and decreases proliferation (319). In contrast to signals that actively maintain the dormant state, pathways involved in proliferation must be repressed in quiescent satellite cells (Table 1). Fibroblast growth factor-2 (FGF2) is an important signaling molecule involved in satellite cell activation (327). FGF2 signaling can be repressed by Sprouty-1, a receptor tyrosine kinase inhibitor. Sprouty-1 is highly enriched in quiescent satellite cells, is down-regulated in activated satellite cells, but is re-expressed when satellite cells return to quiescence. Sprouty-1-null mice display overactivated ERK1/2 pathway that lead to the failure of satellite cells to return to quiescence (274). Other signaling transduction inhibitors are enriched in quiescent satellite cells. For instance, Insulin growth factor binding protein-6 (Igfbp-6) is highly expressed by quiescent satellite cells and potentially antagonizes IGF signaling (223). It has been shown that quiescent satellite cells express Tie-2, a receptor that can interact with angiopoietin-1 (Ang1) secreted in part by vascular smooth muscle cells. Blocking Tie-2 results in increased numbers of cycling satellite cells,

Volume 5, July 2015

Comprehensive Physiology

while overexpression of Ang-1 increases the number of quiescent satellite cells (2). Ang-1/Tie-2 signals through ERK1/2 to promote the G0 state and decrease proliferation and differentiation. These results suggest that neighboring vascular cells can help maintain satellite cell quiescence. MicroRNAs have been suggested to play a key role for the regulation of satellite cell quiescence. These small noncoding RNA molecules generally bind mRNA strands and posttranscriptionally regulate their translation. Hundreds of microRNAs were discovered so far, but a role for only a few of them have been described in muscle. Importantly, ablation of Dicer, the processing enzyme responsible for pre-microRNA hairpin cleavage and activation, leads to satellite cell exit of quiescence and entry into the cell cycle (59). By comparing microRNA expression of quiescent versus activated satellite cells, it was shown that 22 microRNAs are highly expressed in quiescent satellite cells and strongly downregulated following activation (59). The effects of some of these microRNAs were partially elucidated. For instance, miR-489 targets the oncogenic protein Dek that is involved in myogenic commitment and expansion of the myogenic progenitors (59). Another microRNA, miR-31, was shown to directly target Myf5 mRNA (75). Importantly, the myogenic determination factor Myf5 is expressed at the mRNA level in quiescent satellite cells and thus must be regulated to avoid premature satellite cell activation. MiR-31 is sequestered with Myf5 in mRNP granules where it ensures silencing and prevents Myf5 mRNA accumulation in resting satellite cells. Inhibition of miR-31 leads to Myf5 protein accumulation and cell cycle re-entry (75). Overall, quiescence is maintained in satellite cells through different mechanisms such as repression of cell cycle, regulation of intracellular pathways, and interactions with molecular and cellular components of the satellite cell niche. However, these quiescent satellite cells are primed or “poised” to be activated within a short period of time following muscle injury. Thus, quiescence is an active process and complex regulatory pathways are needed to guarantee efficient re-entry into the cell cycle upon muscle injury.

Satellite Cell Activation Quiescent satellite cells can quickly respond to changes in their niche or to specific signals sent by their microenvironment. For instance, the basal lamina contains molecules that can entrap growth factors. Following muscle injury, damage to the ECM releases these growth factors and some of them can interact with the satellite cells to stimulate their activation. FGF2 is one of these factors (88). FGF receptors are upregulated in activated satellite cells compared to quiescent satellite cells (153). The presence of Syndecan-4 on satellite cells is required to mediate FGF2 signaling (73). FGF2 induces a quick increase in intracellular calcium concentration through TRPC channels leading to NFATc translocation into the nucleus and satellite cell activation (181). FGF2 also

Volume 5, July 2015

Satellite Cells and Skeletal Muscle Regeneration

activates the p38 MAPK pathway and inhibition of this pathway prevents satellite cell activation (144, 189). Other isoforms of FGF such as FGF6 have been shown to play a role in muscle regeneration through the stimulation of satellite cell activation and/or proliferation (99). Another critical growth factor for satellite cell activation is the HGF. Similar to FGF2, HGF is located in the ECM of uninjured muscle and is released following muscle injury (295). HGF binds to c-Met, a receptor expressed on quiescent satellite cells. Colocalization of HGF and c-Met was shown on satellite cells in regenerating muscle. In vitro HGF promotes satellite cell entry into the cell cycle (6). Similar to FGF, the presence of Syndecan-4 is necessary to facilitate HGF signaling (73). Local expression of IGF-1 using a muscle-restricted transgene showed that overexpression of IGF-1 can also activate satellite cells (210). In vivo, IGF-1 is produced by different local cell types such as fibroblasts and myofibers (231). IGF-1 is well known to stimulate the Akt/mTOR pathway and consequently downregulate the activity of the transcription factor FOXO. Downregulation of FOXO1 by IGF-1 was shown to inactivate the cell cycle repressor p27kip and induce cycling of satellite cells (187). Nitric oxide (NO) is a molecule almost instantly generated following damage. Inhibition of NO synthase, the enzyme responsible for NO production, was shown to decrease the immediate response of satellite cells to injury (8). It has been demonstrated that NO stimulates MMP expression and increases the release of growth factors from the ECM (294). A cytokine, TNF-α, is also quickly produced and released following muscle injury. Injection of TNF-α in healthy uninjured muscles results in satellite cell activation and entry in the cell cycle (176). Activation of the NF-kB pathway by TNF-α silences gene expression of Notch1, which results in satellite cell activation (3). Lipid mediator pathways are also involved in satellite cell activation, particularly Sphingosine-1-phosphate (S1P). S1P is released through the metabolism of sphingomyelin located on the inner surface of the plasma membrane. Sphingomyelin is abundant in the membrane of quiescent, but not proliferating, satellite cells (213). Consistently, S1P stimulation promotes satellite cell entry in the cell cycle while inhibition of this pathway decreases the response of satellite cells to mitogen stimulation (214). Altogether, these studies indicate that the exit from quiescence into a proliferative state is mediated by various extracellular signals released from the damaged muscle. A cocktail of cytokines and growth factors stimulates receptors present on the quiescent satellite cell membrane and activates different signaling pathways that promote cell cycle re-entry (Table 1).

Asymmetric Division of Satellite Cells Upon activation, Myf5pos committed satellite cells can proliferate and differentiate. In contrast, Myf5neg satellite stem

1035

Satellite Cells and Skeletal Muscle Regeneration

Comprehensive Physiology

Asymmetric self-renewal and differentiation

Stochastic self-renewal and differentiation

Satellite stem cell

Extracellular matrix

Asymmetric division

Muscle fiber

Symmetric division

Committed satellite cell

Figure 10 Asymmetric and stochastic self-renewal. Satellite cells can be divided in two subpopulations: committed satellite cells and satellite stem cells. Satellite stem cells can perform asymmetric division in an apical-basal orientation or symmetric division in a planar orientation (relative to the myofiber). Asymmetric division allows for the formation of one committed daughter cell that will participate in myogenesis, and for the maintenance of the original satellite stem cell. Symmetric division of satellite stem cells produces two daughter stem cells and favors satellite stem cell expansion. Committed satellite cells can divide to form two committed daughter cells that participate in the myogenesis process.

cells are able to self-renew through symmetric or asymmetric division upon entry into the cell cycle. Asymmetric divisions produce one daughter stem cell and one daughter committed cell (Fig. 10). Alternatively, symmetric divisions give rise to two identical daughter stem cells. Although the elements controlling satellite stem cell fate are not well understood, different factors and mechanisms of action begin to be unraveled. A long-standing question was how Myf5 expression is regulated in Myf5neg satellite stem cell during asymmetric division. Indeed, Pax7 is known to induce Myf5 expression and it is present in both Myf5pos and Myf5neg cells after

(A)

(B)

asymmetric division. Interestingly, it was demonstrated that Pax7 needs to be methylated by the arginine methyltransferase Carm1 to induce Myf5 transcription (155,195). During asymmetric division, in the Myf5neg satellite stem cells Pax7 does not interact with Carm1 and therefore Myf5 expression is prevented. The satellite stem cell niche has a critical influence on satellite stem cell fate. It has been shown that muscle stem cells dividing in an apical-basal position (perpendicular to the myofiber) perform mostly asymmetric divisions, while planar divisions (parallel to the myofiber) are symmetric (Fig. 11)

(C)

Figure 11

Symmetric versus asymmetric divisions. Images show myofibers isolated from EDL muscles of Myf5-cre R26R-YFP stained with Pax7 (red), YFP (for Myf5, green), and a nuclear dye (DAPI, blue). The myofiber membrane is outlined by a dashed line. In these mice, expression of Myf5 leads to permanent yellow-fluorescent protein (YFP) staining. Most of the satellite cells express Myf5 (YFP+) during development, but a subpopulation of satellite stem cells never expressed Myf5 (YFP−). YFP− satellite stem cells can perform asymmetric division (A) and give rise to a committed progenitor (YFP+), or perform symmetric division (B) and produce two satellite stem cells (YFP−). Asymmetric divisions occur mostly in an apical-basal orientation (A), while symmetric divisions occur with planar orientation (B and C).

1036

Volume 5, July 2015

Comprehensive Physiology

(163). Generally, during apical-basal asymmetric division, the cell that remains attached to the basal lamina stays Myf5neg while the cell that is pushed toward the myofiber becomes a Myf5pos committed progenitor. In vitro culture of myogenic progenitors in ECM-coated micropatterns showed that when a myogenic progenitor cell has symmetrical adhesion to the ECM it is most likely to perform random DNA segregation. On the other hand, a myogenic progenitor cell that has asymmetric adherence to the ECM will often perform nonrandom DNA segregation (326). On asymmetric ECM micropatterns the cell that retains the old template DNA is preferentially located in the low adhesion side while the committed cell localizes to the high adherence side. Interestingly, Pax7hi cells that intrinsically tend to perform asymmetric division are not affected by symmetric or asymmetric micropatterns, which suggests that intrinsic factors also have an influence on stem cell fate. However, it may be difficult to extrapolate these in vitro findings to in situ satellite cells. Nevertheless, these findings suggest that ECM adherence in the satellite cell niche could influence muscle stem cell fate decision. In asymmetric divisions, many different proteins have been shown to preferentially locate to either the mother or the daughter cell. The partitioning defective (PAR) complex comprising PAR-3, PAR-6, and atypical protein kinase C (aPKC) is polarized in the committed daughter cell and activates p38α/β MAPK asymmetrically (302). Activation of p38α/β MAPK leads to induction of MyoD and consequently induces commitment and proliferation of the daughter cell. The original template DNA that is asymmetrically located in the mother cell has been shown to colocalize with the Notch signaling inhibitor Numb (277). However, another study suggests that Numb is asymmetrically located in the committed daughter cell (70). Consistent with these latter results it has been shown that the mother stem cell expresses high levels of Notch-3 while the committed daughter cell highly expresses Notch ligand Delta-1 (163). Interactions between the mother and the daughter cell could then strengthen the cell fate of each cell. The stimulation of the Notch pathway in the mother cell is important to maintain its stemness as demonstrated by the increased commitment in the presence of Notch inhibitors (163). Overall, during asymmetric division, a variety of proteins are differentially distributed between the cells. Some proteins such as Notch-3 stay in the mother stem cell, while other proteins like PAR3-PAR6-aPCK complex and phosphop38MAPK are located in the committed daughter cell and trigger the myogenic program through Myf5 and/or MyoD activation.

Satellite Stem Cell Self-Renewal During muscle regeneration, the vast majority of satellite cells becomes activated, proliferates and differentiates to repair myofibers. However, to be prepared for subsequent injuries,

Volume 5, July 2015

Satellite Cells and Skeletal Muscle Regeneration

skeletal muscle needs to replenish its satellite cell population. Importantly, even following multiple injuries, the satellite cell population remains constant (276). This capacity to restore the overall population is due to the high self-renewal ability of satellite cells. Transplantation experiments nicely illustrate this enormous self-renewing ability. Grafting of a single intact myofiber containing few satellite cells into a radiated muscle can regenerate up to a hundred myofibers (67). Similarly, a single transplanted satellite cell is able to undergo massive waves of proliferation leading to satellite cell repopulation (255). Importantly, the engraftment potential of the Myf5neg satellite stem cell subpopulation is much higher than the one of committed satellite cells (163). Symmetric self-renewal of Myf5neg satellite stem cells involves the polarized distribution of proteins during division. The planar cell polarity (PCP) pathway has been shown to regulate oriented divisions during development (118). The PCP pathway is a noncanonical Wnt pathway activated by the interaction of wingless (Wnt) ligands with Frizzled receptors (Fzd). Binding of Wnt to Fzd induces the cytoplasmic scaffolding protein Disheveled to activate Rac/JNK and Rho/ROCK pathways (158). These downstream pathways regulate cytoskeleton reorganization and gene expression. Analysis of mRNA expression from FACS sorted Myf5neg stem cells versus committed satellite cells revealed that Fzd7 is highly enriched in the stem cell subpopulation (169). Moreover, Wnt7a is upregulated during muscle regeneration and interacts with Fzd7 to mediate symmetric divisions of Myf5neg satellite stem cells (169) (Fig. 12). Fzd7 induces polarized expression of Vangl2, a well-known effector of the PCP pathway. Knockdown of Vangl2 reduces symmetric division and favors apical-basal asymmetric divisions. Planar symmetric expansion allows each daughter cell to receive equal feedback from their niche (e.g., from the basal lamina and the plasmalemma). Physical interaction with the microenvironment is critical to support self-renewal of satellite stem cells (163). For instance, the binding of Fzd7 to the ECM protein fibronectin through syndecan-4 coreceptor provides feedback and enhances Wnt7a/Fzd7-mediated symmetric proliferation of satellite stem cells (26). Collagen VI is another ECM component that is found in the satellite cell niche and that affects satellite cell self-renewal (304). Lack of collagen VI leads to impaired muscle regeneration and exhaustion of the satellite cell pool following multiple injuries. In addition to the symmetric expansion of satellite stem cells, it was also hypothesized that committed satellite cells can withdraw from terminal differentiation to replenish the satellite cell population. These observations were first shown in vitro using C2C12 myoblasts cultured in low-serum differentiation medium. While most of the myoblasts differentiate, a small subpopulation of myoblasts called “reserve cells” is unable to differentiate and shows greatly reduced MyoD and Myf5 expression (328). One must keep in mind that these cells are immortalized myoblasts that have different behavior than satellite cells in vivo. Nonetheless, culture of isolated

1037

Satellite Cells and Skeletal Muscle Regeneration

Comprehensive Physiology

Satellite stem cells FN

Symmetric cell division

Myogenic progenitors

Cell migration

Cytoskeleton

Muscle fibers

Muscle hypertrophy

Figure 12

Wnt7a-activated pathways in myogenic cells. In myogenic cells, Wnt7a binds to Fzd7 and induces various responses depending on the myogenic stage. In satellite stem cells, Wnt7a forms a complex with the Fzd7 receptor, Scd4 coreceptor and fibronectin to activate the planar cell polarity pathway. This noncanonical pathway leads to the symmetric stem cell division. In myogenic progenitors, Wnt7a leads to rearrangements of the cytoskeleton and increases directional cell migration. Lastly, in muscle fibers, Wnt7a activates the Akt/mTOR pathway, in an IGF-1-independent manner, and promotes myofiber hypertrophy.

myofibers ex vivo also showed that Pax7-positive activated satellite cells can lose MyoD expression and exit the cell cycle to maintain the satellite cell pool (329). These cells can potentially be reactivated in a future muscle injury. However, the mechanism of myoblast dedifferentiation remains to be clarified.

1038

Self-renewal of the satellite cell pool also depends on the ability of the cells to withdraw from cell cycle and return to quiescence. Indeed, it was observed that following muscle regeneration in the presence of the proliferation label BrdU, quiescent satellite cells were marked, which indicates that these cells had previously divided (266). Cell-cycle

Volume 5, July 2015

Comprehensive Physiology

withdrawal is necessary for both differentiation or return to quiescence. Thus, the molecular mechanisms mediating cell fate decision need to be extraordinarily well defined to regulate both processes simultaneously. Microarray analysis of human satellite cells returning to quiescence versus proliferating satellite cells revealed substantial differences in gene expression (160). The authors of this study observed that there is an overall decrease in total mRNAs in quiescent satellite cells compared to proliferating cells. Consistently, the activity of RNAseL, an enzyme involved in RNA destruction, increases when satellite cells return to quiescence. MicroRNAs are also globally downregulated in satellite cells that become dormant (160). Regulation of the ERK1/2 pathway is important to mediate the return to quiescence. As discussed previously, fibroblasts and smooth muscle cells located in the microenvironment of activated satellite cells during muscle regeneration can secrete Ang-1 that binds to Tie-2 receptor on satellite cells. This interaction promotes cell cycle exit, represses differentiation and supports return to quiescence, through the ERK1/2 pathway (2). Moreover, the transcription factor Six1 has been shown to repress the ERK pathway by regulating Dusp6, a phosphatase that inhibits MAPK phosphorylation. Downregulation of Six1 or Dusp6 supports the return to quiescence and satellite cell niche occupancy (168). Moreover, the receptor tyrosine kinase inhibitor, Sprouty1, is upregulated when Pax7-positive cells re-enter quiescence to inhibit FGF2 signaling and induces cell cycle exit (274). Constitutively active Notch signaling has been shown to inhibit myoblast differentiation and promote satellite cell self-renewal (Table 1) (319). As previously discussed, Syndecan-3, a protein highly expressed in quiescent satellite cells, was shown to directly bind and activate Notch to promote satellite cell return to quiescence (233). Reactivation of Notch signaling is also driven by the FOXO3 transcription factor, and these two proteins are involved into a regulatory feedback loop (115). FOXO3 stimulates Notch1 and Notch3 receptor expression, while activation of Notch results in increased FOXO3 expression. FOXO is directly downregulated by the Akt/mTOR pathway. Therefore, during regeneration, the release of growth factors such as IGF-1 stimulates the Akt/mTOR pathway that inhibits FOXO and thus downregulates Notch. This cascade enables satellite cell proliferation and differentiation. The decrease in growth factors present in the microenvironment at the end of the regeneration process then appears to induce an increase in FOXO and Notch signaling that favors self-renewal and the return to quiescence. In summary, signaling pathways that promote the return to quiescence only begin to be unraveled and the significance of this process is likely underestimated (Table 1). Emerging evidence in the literature suggests that this important process is affected in aging and is gradually lost during various degenerative conditions (see section on pathologies and degenerative conditions).

Volume 5, July 2015

Satellite Cells and Skeletal Muscle Regeneration

Myogenic Cell Migration Satellite cells are distributed all along the length of myofibers. Following muscle injury, activated satellite cells migrate from their original location to the site of the lesion (267). Real time video microscopy of satellite cells on isolated fibers ex vivo clearly illustrates the high migratory capacity of activated satellite cells (278). Graft experiments of labeled cells revealed that satellite cells can migrate from one fiber to another over many millimeters (142, 313). In vitro migration assays and in vivo transplantation experiments indicate that satellite cell migration capacity is higher before they start to proliferate (255, 278). Analysis of satellite cell activity following muscle injury demonstrated that activated satellite cells first migrate into the damaged area and then increase their mitotic activity at the site of injury (267). Adhesion molecules, soluble molecules, and guidance cues are important to direct migration of satellite cells (Fig. 13). Adhesion of myogenic cells to the ECM is important to mediate migration. Fibronectin and collagen are able to increase migration in vitro, but generally, laminin is considered the most potent inducer of satellite cell migration (278). Expression of integrin-α7 and -β1 on the satellite cell membrane allows them to bind to laminin. CD34, a membrane protein highly expressed in satellite cells also mediates cell motility (5). In CD34-deficient mice, the migration of satellite cells along the myofiber is decreased and muscle regeneration is impaired. Deficiency for CD44, another cell surface protein, leads to decreased motility of myoblasts and to a reduced response to soluble chemotactic molecules (212). CD44 can interact with various ECM proteins such as hyaluronan, collagens, fibronectin, or laminin (174). Although the ECM provides support for cell migration, it can also represent a physical barrier for migrating satellite cells especially in presence of fibrosis. The secretion of different Matrix Metalloproteinases (MMPs) by myogenic cells is important to facilitate their migration. MMPs are a family of enzymes that can specifically digest precise components of the ECM. For example, MMP-1 can specifically degrade collagen type I, II, and III to facilitate myoblast migration (147). MMP2 and MMP-9 are also upregulated during the early regeneration process and could degrade type-IV collagen that is enriched in the basal lamina (156). Overexpression of MMP9 was shown to increase migration of transplanted myoblasts (207). MMP activity is regulated by different enzymes, named Tissue Inhibitors of MMP (TIMPs). TIMP expression is usually increased during muscle regeneration and can positively or negatively affect MMP activity depending on the TIMP isoform and concentration (150). In vitro experiments demonstrated that MMP inhibition appears to decrease myoblast migration (217). Soluble molecules liberated after muscle injury are important to drive satellite cell migration toward the site of injury. Crushed muscle extract is able to mediate myoblast chemotaxis, by promoting directional migration along a gradient of attracting molecules (32). Among the different factors present

1039

Satellite Cells and Skeletal Muscle Regeneration

Comprehensive Physiology

Figure 13

Satellite cell migration. Schematic representing a migrating satellite cell. Satellite cells receive many guidance cues from their microenvironment that promote chemotaxis toward the damaged area (ex: HGF) or repulse the cell away from uninjured region (ex: Ephrins).

in the crushed muscle extract, TGF-β and HGF have been shown to directly increase the motility of myogenic cells. FGF-2 and FGF-6 can also stimulate satellite cell motility (215, 278). Wnt7a expressed during regeneration can guide myogenic cells by activating the PCP pathway through the Fzd7 receptor (Fig. 12) (23). Similarly, vascular endothelial growth factor (VEGF) affects chemotaxis of myogenic cells in regenerating muscle. Myoblasts express both VEGF and its receptors, a characteristic that could potentially allow for autoregulatory interactions (108). Interleukine-4 (IL-4) also stimulates myoblast migration by increasing the expression of specific integrins, such as β1-integrin (165). In addition to the various chemoattractant molecules that induce migration toward the site of injury, there are also molecules that have chemorepulsive effects. Activated satellite cells and muscle fiber both express Ephrin ligand and Eph receptor. Multiple Ephrins were shown to induce repulsion of satellite cells in vitro (286). Ephrins are expressed in healthy fibers and could thus favor migration of satellite cells away from undamaged fibers. Overall, satellite cells are highly motile during muscle regeneration. This migration capacity is mediated by a

1040

multitude of physical and chemical guidance cues. Migration of satellite cells is essential to direct myogenic precursors to the site of injury, where they can then proliferate and fuse to regenerate the damaged muscle.

Myoblast Proliferation Activated myogenic progenitors undergo several rounds of proliferation to increase the myogenic pool needed for tissue repair. Inhibition of cell proliferation makes muscle regeneration impossible (232). The majority of the growth factors known to activate satellite cells also facilitates cell division. Addition of FGF2 to the medium of primary myoblasts or in fiber cultures in vitro is commonly used to stimulate myogenic cell proliferation. However, FGF2 is not a mitogen and its effect on proliferation is probably due to increased satellite cell activation (323). Several studies have also revealed the role of IGF-1 for the stimulation of cell-cycle progression of satellite cells. Satellite cells from mice overexpressing IGF-1 have enhanced proliferative lifespan and can generate a much bigger myogenic cell population (54). In vitro

Volume 5, July 2015

Comprehensive Physiology

Satellite Cells and Skeletal Muscle Regeneration

experiments have shown that addition of low concentration of HGF promotes myoblast proliferation while higher concentrations lead to decreased proliferation (324). Following muscle damage, proinflammatory cytokines are enriched in the injured tissue. This boost in proinflammatory cytokines, such as IL-6, IL-1β, and TNF-α, coincides with the initiation of the proliferative phase of myoblasts (298). IL-6 can be produced by myoblasts or myofibers and acts in an autocrine or paracrine manner to stimulate myoblast proliferation (44,312). IL-6 deficiency reduces myoblast proliferation by impairing STAT3 signaling and its downstream target, the cell cycle protein cyclin-D1 (272). IL-1β and TNF-α during the early regenerative phase are also known to stimulate IL-6 expression (185, 236). Moreover, these factors can directly promote myoblast proliferation, possibly through the NF-kB pathway (220). Similar to NF-kB, the JNK pathway favors myoblast proliferation by promoting the transcription of cellcycle proteins like Cyclin-D1 (Table 1) (7, 124, 230). On the other hand, certain signaling pathways encourage proliferation by repressing premature differentiation. p38γ can directly phosphorylate MyoD leading to the formation of a repressive transcriptional complex that will occupy the myogenin promoter and repress premature activation (112). Similarly, JAK1/STAT1 promotes myoblast proliferation and inhibits premature differentiation by repressing differentiation genes such as MEF2 (289). Several microRNAs were shown to control the switch from proliferation to differentiation. For instance, miR-221 and miR-222 are more expressed in proliferating myoblasts than in differentiating myotubes (46). These two microRNAs repress the cell cycle inhibitor p27 and the differentiation factor myogenin, which results in delayed withdrawal from the cell cycle. MiR-133 also promotes myoblast proliferation by repressing the serum response factor (SRF), which is involved in myogenic differentiation and expression of myofiber-specific genes (56, 315). Overall, there is a multitude of different factors and signaling pathways involved in myoblast proliferation. These extrinsic factors promote myoblast proliferation either directly by increasing cell cycle progression and/or by repressing precocious differentiation. 0d

1d

Myogenic Differentiation Following several rounds of proliferation, myogenic cells exit the cell cycle and start to differentiate. Thereafter, myogenic cells fuse together to form multinuclear myotubes or fuse to existing damaged fibers. Different cues control this switch from the proliferation to the differentiation state. In vitro, myoblasts can be expanded in the presence of high concentrations of serum and switching to low-serum medium induces their differentiation and fusion to the neighboring cells (Fig. 14). MyoD is a key factor for myogenic progression and subsequent differentiation. MyoD−/− myoblasts fail to upregulate myogenin and MRF4, display differentiation defects and impaired muscle regeneration (72,198,254). MyoD facilitates the transition from myoblast proliferation to differentiation through the induction of cell cycle inhibitors such as p21 and p57, and by inducing myogenin expression (125, 132). Downstream of MyoD, myogenin triggers the expression of genes involved in muscle contractility such as myosin light chain, myosin heavy chain, muscle creatine kinase, α-actinin, troponin, and voltage-dependent calcium channel (80). Similarly, MRF4, an MRF with functions related to myogenin, is involved in late differentiation (242). MRF activity during differentiation is also controlled by the MEF2 family of transcription factors. For instance, MEF2C can physically interact with MyoD or myogenin and synergistically activate different muscle gene promoters (204). Moreover, there is a positive regulatory feedback loop between MEF2 and MRF expression. Therefore, forced expression of MyoD or myogenin can induce MEF2 expression, while overexpression of MEF2, in turn, increases MyoD expression and myoblast differentiation (76, 154). Conditional deletion of different MEF2 genes (MEF2A, MEF2C, and MEF2D) in satellite cells revealed that absence of only one isoform can be compensated by the others. However, combined deletion of the different MEF2 genes leads to impaired myogenic differentiation and muscle regeneration (180). Certain microRNAs are involved in the transition from myogenic cell proliferation to differentiation. For instance, miR-1, miR-133, miR-206, and miR-486, are upregulated 3d

5d

Figure 14

Myotube formation. Pictures from primary myoblasts cultured in vitro in differentiation medium (low-serum medium). Cells are stained with a nuclear dye (DAPI, blue) and myosin heavy chain (MyHC, in green). Almost no myoblasts express MyHC before the addition of differentiation medium. After 1 day in differentiation medium, the cells start to express MyHC. At three days, myoblasts fuse to form small multinucleated myotubes (few myonuclei per myotube). After 5 days of differentiation, myotubes become larger and contain more nuclei.

Volume 5, July 2015

1041

Satellite Cells and Skeletal Muscle Regeneration

during myoblast differentiation and myotube formation (159, 160). These microRNAs enhance differentiation by directly inhibiting Pax7 transcription, which consequently results in increased MyoD activity (57, 87). Distinct signaling pathways regulate MRF expression and thus affect the switch from myoblast proliferation to differentiation (Table 1). A switch from Notch to the canonical Wnt signaling pathway has been shown to be critical for the transition to myoblast differentiation (39). The cross-talk between these two pathways is mediated by the protein kinase Glycogen synthase kinase (GSK3). The MAPK signaling protein p38α is also involved in myoblast differentiation (321). p38α decreases myoblast proliferation by antagonizing the effect of the proliferation-promoting pathway JNK (230). Moreover, p38α stimulates MyoD transcriptional activity and consequently the transcription of muscle-specific genes (182). p38α also directly phosphorylates MEF2A and MEF2C, the latter being a coactivator of MyoD (321). Calcium-activated pathways have been shown to be involved in myogenic differentiation. Calcium is a major regulator of muscle cell function and is thus tightly controlled. Increased intracellular calcium concentrations lead to the rapid activation of calcineurin and calcium/calmodulindependent protein kinase (CaMK) pathways. Calcineurin activates MyoD through MEF2C (101). Inhibition of calcineurin leads to decreased myogenin expression and impaired differentiation (100). CaMK also increases MEF2 activity by removing the inhibitory action of HDAC proteins on MEF2 target gene expression (184). Thus, p38α and the calcium-activated pathways both promote in a nonredundant manner MyoD and MEF2 activities resulting in increased myogenin expression and myogenic differentiation (322).

Myogenic Cell Fusion and Myotube Formation Following the induction of the differentiation program, myogenic cells undergo a cell-cell fusion process that will strongly modify cell shape and function (Fig. 14). To fuse, myocytes must first recognize and adhere to each other. Myocyte fusion has mostly been studied in Drosophila because in this experimental system muscle formation is easily visualized. In Drosophila, interaction between Dumbfounded (Duf), Sticks and stones (Sns), Roughest (Rst), and Hibris (Hbs) proteins mediate cell recognition and adhesion (248). Duf and Rst both interact with Sns and have redundant function so that depletion of both proteins is necessary to cause a fusion deficit (288). Notch signaling downregulates the expression of adhesion proteins such as Sns, that are important for fusion (111). In mice, related adhesion molecules such as Nephrin (Sns homolog), β1-integrin, Focal adhesion kinase, M-cadherin, and N-cadherin have been described (1, 282). Upon myoblast-myoblast contact, Duf and Sns proteins form a ring-like structure called fusion-restricted

1042

Comprehensive Physiology

myogenic-adhesive structure (FuRMAS) in Drosophila. The formation of this structure is hypothesized to be a key event leading to the recruitment of the molecular fusion machinery. Following FuRMAS establishment, actin filaments accumulate in the center of the fusion site. Rho GTPase family members such as Rac1 play an essential role to drive actin polymerization through activation of Actin related protein (Arp2/3) (287). During the next steps of the fusion process, intracellular perfusion vesicles accumulate at the contact site of the two adherent cells. Actin was hypothesized to partially mediate the recruitment of these vesicles but its exact role remains unclear (157, 248). Thereafter, one or many small pores in the membranes are created. These pores will then enlarge to eventually dissolve completely to give rise to a multinucleated cell. Importantly, re-expression of adhesion molecules by the newly formed multinucleated cell is necessary to allow further myocyte recognition and to continue the fusion process. Until recently, the proteins regulating myoblast fusion in mammals were elusive. However, a muscle-specific membrane protein named myomaker was shown to control myoblast fusion (200). Absence of myomaker in mice leads to perinatal death due to lack of multinucleated myofibers. Accordingly, inducible deletion of myomaker specifically in satellite cells (using pax7-creERT2 mice) strongly perturbs muscle regeneration (201). Interestingly, myomaker-deficient mice express normal level of MyoD and myogenin indicating that muscle precursor cells are able to commit into the myogenic lineage up to a certain point. Moreover, MyoD and myogenin were shown to induce myomaker transcription. Mechanistically, myomaker requires actin cytoskeletal rearrangement to induce fusion (200). The fusion process can be divided into early and late fusion. Early fusion involves myocyte-myocyte contact, while late fusion implies the fusion of myocytes to nascent multinucleated myotubes. Although they share many similarities, these two modes of fusion involve different adhesion molecules and signaling pathways. For example, β1-integrin mediates myocyte-myocyte fusion. Consequently, β1-integrin-deficient myoblasts are able to adhere to each other but cannot fuse (268). On the other hand, proteins such as the actin binding protein FilaminC are involved in myocyte-myotube fusion. In vitro, knockdown of FilaminC in differentiating myoblasts results in formation of small and round myotubes with limited numbers of nuclei (78). During late fusion, the calcium activated transcription factor NFATc2 is upregulated. NFATc2 controls the expression of IL-4. Myoblasts from IL-4-deficient mice are able to fuse but form smaller myotubes with less myonuclei (133). Thus, after initial myoblast-myoblast fusion, the nascent myotubes secrete IL-4 to recruit more myoblasts and perform late fusion. At this stage, interactions of myotubes with the ECM are particularly important to stimulate myotube formation. In vitro studies showed that laminin promotes myotube alignment and survival (63). Generally, the proteins and signaling pathways involved in myocyte fusion show some discrepancies between

Volume 5, July 2015

Comprehensive Physiology

the different species but the major events during the fusion process are conserved.

Myotube Maturation and Myofiber Hypertrophy Nascent myotubes need to undergo a maturation process to become fully functional myofibers. Mature muscle fibers are highly specialized and need to acquire a competent excitationcontraction coupling, contractile, and metabolic machineries. The Akt/mTOR pathway has a pivotal role in myotube maturation and hypertrophy (Table 1) (225, 226). Interaction of IGF-1 with its receptor activates Akt and subsequently its downstream effector mTOR (249). In absence of endogenous mTOR, C2C12 myoblasts are able to fuse together, but they cannot form mature myotubes and express contractile proteins. The Akt-1/mTOR pathway induces various trophic effects via several pathways (18). mTOR activation stimulates protein synthesis by phosphorylating p70S6k that activates the ribosomal protein S6. mTOR also exerts anabolic effects by inhibiting the translation repressor 4E-protein binding 1 (4EBP1). Akt, in turn, has been shown to support hypertrophy by inhibiting the activity of FOXO, a factor involved in protein catabolism. FOXO can induce muscle ring finger-1 (MuRF1) and atrogin-1 expression, two E3-ubiquitin ligases strongly associated with muscle wasting (261). Although IGF-1 is the most well described factor activating the Akt/mTOR signaling, other proteins and growth factors were shown to induce this pathway. Interestingly, Wnt7a, which stimulates satellite stem cell expansion (see section on self-renewal), can also act directly on myotubes and induce a hypertrophic response through the activation of the Akt/mTOR pathway (Fig. 12) (308, 311). Surprisingly, long-term overstimulation of mTOR induces a strong feedback mechanism characterized by a powerful induction of E3-ubiquitin ligases and muscle atrophy (21). Thus, Akt/mTOR pathway needs to be tightly regulated to stimulate muscle hypertrophy. It has been suggested that increases in calcium signaling can stimulate muscle fiber hypertrophy. However, these results are still controversial (141, 227, 249). The main function of calcium-dependent signaling pathways, such as the Calcineurin and CaMK cascades, in differentiated fibers appears to be the determination of muscle fiber type (e.g., slow vs. fast). Briefly, Calcineurin activation leads to NFAT translocation in the nucleus, which stimulates slow fiber gene transcription. Similarly, CaMK liberates and activates MEF2 transcription factors that promote the slow fiber gene program. Muscle fiber phenotype and plasticity have been extensively reviewed by Blaauw B and colleagues (34).

Epigenetics The previous sections illustrate that myogenesis is a complex and coordinated process relying on various extracellular cues and intracellular pathways. In addition to these diverse signals

Volume 5, July 2015

Satellite Cells and Skeletal Muscle Regeneration

there is also other intrinsic factors that will determine satellite cell fate during myogenesis. During the last few decades, epigenetics emerged as a new field of research and many different studies indicated its crucial effect on myogenesis. Epigenetics involve modifications in gene expression that occur in absence of any modification of the primary DNA sequence. These alterations in transcriptional potential can be mediated by a multitude of external cues. Epigenetics include dynamic and reversible changes in the chromatin structure due to histone modifications or nucleosome positioning, as well as DNA methylation. Epigenetic changes can turn on or turn off genes and thus influence which protein will be expressed. Histone modifications can be associated with active gene expression, such as histone acetylation (H3Ac and H4Ac) or trimethylation of histone H3 lysine 4 (H3K4me3). Alternatively, other modifications are associated to gene repression, including trimethylation of histone H3 lysine 27 (H3K27me3) and methylation of lysine 9 of histone H3 (H3K9me) (Table 2) (335). Combinations of these various modifications have different consequences on gene activity, particularly modifications taking place on histone tails, known as the “histone code.” The study of stem cell epigenetics is technically challenging, because adult stem cells are very limited in numbers and do not represent a uniform population in terms of cell-cycle Table 2 Histone Marks Present on Different Groups of Genes During Myogenic Progression

Cell type

Family of genes (example)

Histone mark

Embryonic stem cells

Myogenic developmental regulators (Pax7)

H3K27me3 (repressive), H3K4me3 (active)

Quiescent satellite cells

Nonmyogenic developmental regulators (Neurog1) Satellite cell-specific (Pax7) Muscle-specific (Myog)

H3K27me3 (repressive), H3K4me3 (active)

PAX7 target genes (satellite cell-specific) (Myf5) Muscle-specific (Myog)

H3K4me3 (active)

Satellite cell-specific (Pax7) Muscle-specific (Myog)

H3K27me3 (repressive)

Proliferating myoblasts

Myotubes

H3K4me3 (active) unmarked (not H3K4me3, not H3K27me3)

H3deAc, H4deAc, H3K9me2, H3K27me3 (all repressive)

H3Ac, H4Ac, H3K4me3 (all active)

In embryonic stem cells and in quiescent satellite cells, genes encoding developmental regulators possess bivalent (repressive and active) histone marks. In quiescent satellite cells, stemness genes like Pax7 are marked by active chromatin modifications, whereas muscle-specific genes are mostly unmarked. In proliferating myoblasts, satellite cellspecific genes like Myf5 are active, while muscle-specific genes possess repressive histone marks. When cells differentiate into myotubes, satellite cell-specific genes are repressed, whereas muscle-specific genes like myogenin (Myog) present active chromatin marks.

1043

Satellite Cells and Skeletal Muscle Regeneration

status and in their ability to regenerate a given tissue. Therefore, technical issues limit the usage of classical epigenetic approaches like chromatin immunoprecipitation (ChIP) and ChIP-seq (sequencing) for studying satellite cell epigenetics. As a consequence, most studies on satellite cell epigenetics have been performed in myoblasts. Thus, because myoblasts are proliferating whereas satellite cells are mostly quiescent, these experiments do not perfectly reflect the biology of satellite cells, particularly for the regulation of quiescence and selfrenewal. Recently, ChIP-seq and RNA-seq techniques were adapted for their use with small numbers of cells (4, 292). These techniques will open new perspectives in the study of stem cell epigenetic regulation.

Epigenetics of Satellite Cells and Myoblasts: Repression of a Muscle Genetic Program Satellite stem cells present a particular epigenetic state known as “poised” state. Many genes encoding developmental regulators possess at the same time some repressive (H3K27me3) and some active (H3K4me3) histone marks. This coexpression of active and repressive marks is observed in embryonic stem (ES) cells for the Pax3, Pax7, Myod1, and Mrf4 loci (171, 334). The coexistence of negative and positive (“bivalent”) histone modifications is thought to keep developmental regulators in a silenced state, keeping them poised for rapid activation upon reception of appropriate cues to differentiate along a specific lineage. In quiescent satellite cells, many developmental regulators important for nonmuscle lineages, including a large number of transcription factors, possess bivalent chromatin marks, consistent with the stemness characteristics of satellite cells (179). In fact, quiescent satellite cell chromatin is mostly in a permissive conformation, with about half of the annotated genes being marked with the active H3K4me3 mark. Satellite cell activation is characterized by the retention of the H3K4me3 mark and the acquisition of the repressive H3K27me3 mark on specific loci (Table 2) (179). In agreement with the previous findings, ChIP-seq analysis in myoblasts demonstrated that many genes that are almost not expressed in myoblasts, but are upregulated in myotubes, display RNA polymerase II (PolII) binding. The presence of an enzyme involved in RNA transcription on inactive genes reflects that these genes are in a poised state prior to their maximal expression (13). To prevent premature differentiation, satellite cells and myoblasts must possess a regulatory system that represses muscle gene expression. DNA methylation is one important mechanism that represses differentiation. Indeed, treatment with 5-azacytidine, a DNA methyltransferase inhibitor, has been shown to induce fibroblast transdifferentiation into myoblasts (296). In undifferentiated myoblasts, genes that are not expressed mainly possess repressive histone marks in the vicinity of their transcriptional start site (TSS). Core histones H3 and H4 are hypoacetylated and methylated

1044

Comprehensive Physiology

at specific lysine residues (H3K9me2, H3K27me3) (13, 47, 106, 190, 330). Repressive histone marks are also found in quiescent satellite cells from old mice, whereas in young mice the frequency of H3K27me3 mark is relatively low (179). These repressive histone marks are deposited by histone deacetylases (HDACs) and histone methyltransferases (HMTs). HDACs are often found in large protein complexes together with HMTs and are recruited at chromatin regulatory regions of inactive muscle genes (47, 305, 330). In proliferating myoblasts, class II HDACs are found, together with the HMT SUV39H1, at the myogenin gene promoter (Fig. 15). Their presence is associated with hypoacetylation of histone H3 and methylation of H3K9 (190, 330). Reducing Suv39h1 levels using siRNA leads to muscle gene expression (190). HDACs also suppress myogenic differentiation by being recruited to muscle gene promoters via interaction with MEF2 and MyoD (184, 190, 240). Moreover, the histone deacetylase SIRT1 is expressed in proliferating myoblasts and its expression decreases as the cells differentiate into myotubes (Fig. 15) (106). SIRT1 is present on muscle regulatory elements and its downregulation using shRNAs results in increased muscle gene expression and myogenic differentiation (106). HDACs and HMTs also have the property to modify nonhistone proteins, including transcription factors. For example, MyoD is acetylated by PCAF, and deacetylated by HDAC1 and SIRT1 (106,190,263). Similarly, MEF2 is acetylated by p300 and deacetylated by SIRT1 (186, 333). One of the most studied methyltransferase complexes is the Polycomb Repressive Complex 2 (PRC2). Its catalytic subunit, the HMT EZH2, is highly expressed during development and in proliferating satellite cells, and is downregulated when myogenic differentiation is induced (47). In undifferentiated myoblasts, EZH2 is found at the regulatory regions of muscle specific genes together with HDAC1 and the transcription factor YY1 (Fig. 15). These genes are marked by H3K27me3 (13, 47). Conditional deletion of Ezh2 in satellite cells compromises satellite cell proliferation and differentiation, with an impact on muscle regeneration (145). Mice with a conditional deletion of Ezh2 have reduced muscle mass and present smaller myofibers. In addition, deletion of Ezh2 in satellite cells leads to derepression of genes that are normally expressed in nonmuscle lineages, underscoring the importance of EZH2 in establishing satellite cell transcriptional program (145). An alternative epigenetic modification is to replace canonical histones with specific histone isoforms that are known to induce repressed or activated chromatin. For example, histone H1b and its interacting homeodomain partner MSX1 are present in a Myod1 regulatory enhancer, where they induce a repressive chromatin state and inhibit muscle differentiation (170). In contrast, the presence of histone H3.3 in the Myod1 promoter is associated with MyoD expression and epigenetic memory (216). The role of specific histone isoforms in satellite cell epigenetics is still unexplored and it would be of great interest to expand our knowledge in this field.

Volume 5, July 2015

Comprehensive Physiology

Satellite Cells and Skeletal Muscle Regeneration

Lineage determination genes PRC2

Suv39H1 HDAC

Undifferentiated satellite cells

Stemness genes

SIRT1

SWI/SNF

MLL

HAT

HDAC

RNA pol II

YY1

YY1 HDAC MEF2

MyoD

For example, Pax7

For example, Myogenin Relocalization of chromatin modifiers

UTX MLL

SWI/SNF HAT

PRC2 RNA pol ll

Differentiating satellite cells MEF2

MyoD

For example, Myogenin

HDAC

For example, Pax7

Figure 15 Epigenetic control of stemness and lineage determination genes during myogenic differentiation. In undifferentiated satellite cells and myoblasts, lineage determination gene regulatory regions such as the myogenin promoter are repressed through a combination of epigenetic mechanisms. MyoD is repressed by its association with SIRT1 and HDACs. MEF2 also recruits HDACs as well as SUV39H1, the histone methyltransferase responsible for H3K9 methylation. The Polycomb Repressive complex 2 (PRC2) containing the histone methyltransferase EZH2 is also recruited on the myogenin promoter by the transcription factor YY1. Together, these protein complexes are responsible to establish heterochromatin formation that is repressive for transcription. In undifferentiated satellite cells and myoblasts, stemness genes like Pax7 are active and are thus occupied by histone acetyltransferases (HAT) as well as the SWI/SNF chromatin remodeling complex and the MLL/Trithorax complex, responsible for H3K4me3 deposition. These proteins promote the formation of euchromatin, which is permissive for transcription. When cells receive differentiation cues, these epigenetic complexes are relocalized to different regulatory regions. Repressive complexes leave the promoters of lineage determination genes, which become active, and could be recruited to the promoters of stemness genes that need to be silenced. In contrast, proteins involved in active chromatin configuration leave the promoters of stemness genes and are relocalized at muscle specific promoters (myogenin). The myogenin promoter is thus occupied by the MLL/Trithorax complex, by HATs, by the SWI/SNF complex as well as by the demethylase UTX.

While muscle specific genes need to be repressed in satellite cells, other genes important for satellite cell function need to be expressed. Pax7 is an important transcription factor that controls genetic programs taking place in satellite cells. Recent ChIP-seq and gene expression data analysis determined that Pax7 activates genes involved in myoblast growth and proliferation whereas it represses genes involved in myogenic differentiation (283). Pax7 cooperates with the Trithorax methyltransferase complex through direct interaction with the MLL2 subunit (155). Trithorax complex activates gene expression by methylating the lysine 4 of histone H3. In myoblasts, Pax7 is responsible for the recruitment of the Trithorax complex to the Myf5 promoter, a known Pax7 target gene, to mark the gene with H3K4me3 (195). It is likely that PAX7 activates the expression of its other target genes involved in the promotion of cell proliferation by a similar mechanism (283).

Muscle Differentiation: Reconfiguration of the Chromatin When satellite cells receive prodifferentiation cues, the chromatin structure close to muscle-specific genes needs to be remodeled to be permissive for transcription (Fig. 15). The repressive marks on muscle regulatory regions need to

Volume 5, July 2015

be removed and replaced by active histone modifications (Table 2). For example, H3K27 becomes hypomethylated at specific muscle regulatory regions after induction of myogenesis (47). EZH2, the protein responsible for methylation of H3K27, is downregulated at the transcriptional level as well as by the action of specific microRNAs (47, 146, 320). The demethylase UTX mediates the removal of the repressive H3K27me3 marks upon activation of the myogenic program (270). Following induction of myogenic differentiation, the HMT SUV39H1 dissociates from MyoD at muscle gene promoters that show reduced methylation at histone H3 lysine 9 (H3K9) (190), which is indicative of gene activation. The histone demethylase JMJD2A is recruited to the myogenin promoter following induction of differentiation, where it catalyzes the removal of the H3K9 methylation mark (306). Accordingly, following promyogenic signals, HDACs leave the promoters of muscle-specific genes and the histone tails surrounding these genes become acetylated (330). The expression of the deacetylase SIRT1 is also decreased as cells differentiate, consistent with an increased acetylation at muscle regulatory regions (106). Another elegant mechanism for the downregulation of HDAC activity at myogenic promoters is the titration of HDACs to other protein complexes. For example, HDAC1 interacts with MyoD and is recruited on muscle gene promoters in proliferating myoblasts. Upon induction of myogenesis, the transcription factor pRb is

1045

Satellite Cells and Skeletal Muscle Regeneration

dephosphorylated, which promotes the formation of a complex between pRb and HDAC1 that coincides with the disassembly of the HDAC1-MyoD complex (240). It is interesting to note that pharmacological inhibition of HDACs can enhance muscle differentiation (138). Histone acetyltransferases (HATs) are recruited at specific muscle gene loci by different transcription factors to promote histone acetylation of their regulatory elements (197). Nucleosomes need to be distributed in a specific manner to provide access of DNA regulatory elements to the transcription factors and machinery. Nucleosome positioning is regulated by chromatin remodeling complexes that mobilize nucleosomes following ATP hydrolysis, but can also be determined by transcription factors binding to specific loci. Some “pioneer” transcription factors can penetrate repressive chromatin to bind their regulatory elements and initiate chromatin opening (62). One example is PBX1/MEIS which is constitutively bound to muscle regulatory regions even in the absence of differentiation cues (27, 82). PBX1/MEIS interacts with MyoD and could recruit MyoD to specific genes from the muscle lineage, even in the context of repressive chromatin, thus establishing myogenic potential (27). This may be a mechanism through which MyoD has the potential to mediate “myogenic conversion,” for example, by reprogramming the genome of a nonmuscle cell toward the expression of musclespecific genes (317). Transcription factor binding to a specific DNA element facilitates access to secondary transcription factors on adjacent sites. Alternatively, transcription factors can recruit chromatin remodeling enzymes that displace nucleosomes and facilitate transcription factor recruitment. MyoD interacts with the SWI/SNF chromatin remodeling complex and could recruit this complex on muscle regulatory elements (Fig. 15) (81, 82). The arginine methyltransferase PRMT5, which dimethylates histone H3 arginine 8 (H3R8me2), as well as the Scaffold Attachment Factor B1 (SAFB1), are also involved in the binding of the SWI/SNF complex at the myogenin promoter (77, 130). Recruitment of the SWI/SNF complex, in turn, is required for the binding of other myogenic transcription factors including MyoD itself, myogenin and MEF2 (82).

Signaling to chromatin during muscle differentiation The p38 MAPK is an important mediator of muscle cell differentiation. Signaling by p38 regulates muscle gene expression by different epigenetic mechanisms. p38 is detected by chromatin immunoprecipitation on the regulatory elements of muscle genes (279), where it phosphorylates multiple targets. One direct target of p38-mediated phosphorylation is the ubiquitous bHLH protein E47. Phosphorylation of E47 promotes its heterodimerization with MyoD and increases the transcriptional activity of E47/MyoD that is essential for myogenesis (182). MEF2 proteins are also directly phosphorylated by p38, with an increased transcriptional activity as a result (126, 332). In addition, phosphorylation of MEF2D by

1046

Comprehensive Physiology

p38 mediates the recruitment of Trithorax methyltransferase complexes to MEF2D target genes, which methylates histone H3 lysine 4 (H3K4me3), a chromatin mark associated with active gene expression (241). The SWI/SNF chromatin remodeling complex is also directly targeted by p38, since the BAF60 subunit of the complex is phosphorylated by p38, promoting the recruitment of the SWI/SNF complex on myogenic regulatory regions and favoring gene expression (279). On the other hand, genes important for satellite cell proliferation and expansion need to be repressed upon p38-induced muscle differentiation. One example is the Pax7 gene that is silenced by the action of the PRC2 complex. EZH2, the enzymatic subunit of the complex, is phosphorylated by p38, which promotes the formation of repressive chromatin on the Pax7 promoter (222). It appears that during myogenic differentiation, EZH2 (and possibly other proteins involved in chromatin repression) leaves the promoters of muscle-specific genes to associate with and repress genes important for satellite cell proliferation and expansion. How these changes are regulated by p38 and other signaling pathways remains to be determined. Blockade of p38 can efficiently repress muscle gene transcription, however the recruitment of muscle-regulatory transcription factors or acetyltransferases to muscle specific loci is not compromised (279). These observations suggest that other signaling pathways converge to chromatin to activate muscle gene expression. One example is the IGF1/PI3K/AKT pathway that promotes the association between MyoD and the acetyltransferases p300 and PCAF (271). p300 is a direct target of phosphorylation by AKT1/2 (271). The calciumdependent pathway CaMK is another important player in muscle differentiation. CaMK signaling disrupts the interaction between HDAC4/5 and MEF2, allowing MEF2 to recruit coactivators and chromatin modifiers to activate gene transcription at MEF2 target genes (184, 330). In addition, HDAC4/5 are directly phosphorylated by CaMK, and this phosphorylation drives the nuclear export of HDAC4/5 to the cytoplasm, preventing the binding of HDAC to myogenic gene promoters (196). In addition to the different pathways described here, there are many more signaling events that must converge to the nucleus where the ultimate effect is the remodeling of chromatin to modify gene expression. The specific effect of these pathways on epigenetic regulation remains to be further investigated.

Inflammatory processes Satellite cells are essential for muscle regeneration, however efficient muscle recovery relies on many other cell types (Fig. 16). Muscle damage initially triggers a complex inflammatory process. This inflammatory reaction involves the interaction of resident and infiltrating inflammatory cell types, for example, leukocytes. Inflammation following muscle injury has long been known for its role in the clearance of cell debris and/or pathogens. However, leukocytes have also been shown

Volume 5, July 2015

Comprehensive Physiology

Satellite Cells and Skeletal Muscle Regeneration

Committed satellite cells

Blood vessel Vessel-associated cell types

Basal lamina Muscle fiber

Muscle fiber nucleus

Specialized ECM

Growth factors

Directly interacting cell types

Satellite cell

Cell types secreting ECM and growth factors

Fascia

Figure 16 The satellite cell microenvironment. The satellite cell niche is composed of various ECM proteins and cell types. Cell types such as inflammatory cells and stromal cells can physically interact with satellite cells or release various cytokines, growth factors and ECM components that will influence satellite cell behavior.

to play significant roles in regulating satellite cell behavior (Fig. 17). The inflammatory process following muscle damage can be divided in three major steps: initiation, development, and dampening. In the earliest stage after injury the different leukocytes residing in the tissue, for example, mast cells, macrophages and a subset of “patrolling” circulating monocytes, sense the perturbation of tissue homeostasis (14, 107). Damaged myofibers release molecules called damaged-associated molecular pattern (DAMPs) such as DNA, RNA, metabolites, and others that will be recognized by resident leukocytes. Mast cells are the first cell type to be activated in the course of this process. These cells can release preformed and newly synthesized cytokines such as TNF-α, tryptase, and IL-6 within minutes. At physiological concentration, these cytokines are known to stimulate satellite cell activation and proliferation (58, 91, 272). In vitro coculture experiments confirmed that

Volume 5, July 2015

factors secreted by mast cells enhance myoblast proliferation (90). In vivo, the inhibition of mast cell secretion delays muscle regeneration (93). Thus, the initiation of the inflammatory process is closely linked to the activation of satellite cells. Next to the activation of satellite cells, the burst of cytokines released by the resident leukocytes also leads to the recruitment of blood-circulating leukocytes. Granulocytes, such as neutrophils and eosinophils, are the first nonresident cell types to invade the injured muscle. Granulocytes possess a high phagocytosis potential and have a critical role in clearing muscle debris. The proinflammatory environment that granulocytes generate influences myogenesis. Indeed, an in vitro study suggested that oxidative stress induced by neutrophils can exacerbate membrane damage on myotubes (234). However, it appears that the effect of neutrophils on myogenesis is generally dependent on the injury type (92,235). At the opposite, eosinophils have recently been demonstrated to promote

1047

Comprehensive Physiology

Myofiber growth Differentiation/fusion Proliferation Activation

Inflammatory cell activation

Myogenic stages

Satellite Cells and Skeletal Muscle Regeneration

MC

Injury

PMN

1d

M1

3d

M2

5-7d

Figure 17

Inflammation and myogenesis. Inflammation is characterized by the activation of mast cells (MC), followed by the early recruitment of granulocytes (polymorphonuclear cells, PMN), and the accumulation of monocytes/macrophages. Macrophage accumulation is classically composed of an initial proinflammatory phase (M1 macrophages) followed by an anti-inflammatory phase (M2 macrophages). These different leukocytes secrete factors that directly influence the function of myogenic cells. Initial burst of leukocytes has been shown to stimulate satellite cell activation and proliferation. M1 macrophages promote myoblast proliferation and repress early differentiation, while M2 macrophages stimulate differentiation and myofiber growth.

granulocytes and favor nonphlogistic phagocytosis (e.g., that does not stimulate inflammation) of apoptotic neutrophils (281). In addition to dampening of the inflammatory reaction, cytokines and growth factors such as IL-4, IL-10, and IGF-1 released by M2 macrophages also influence myogenesis (11, 183). IL-4 promotes myotube formation during the late fusion of myoblasts into myotubes (133). IGF-1 stimulates myotube hypertrophy (94). Consistently, the suppression of M2 macrophages in vivo leads to reduced myogenin expression and impaired fiber growth (86, 253). In regenerating human muscle, it has been shown that M2 macrophages are located in close proximity to myocytes expressing myogenin (257). Physical contact appears to increase macrophage-induced myogenic effects through a specific set of adhesion molecules (284). This proximity might also be important to allow macrophages to reconstitute the satellite cell niche through the secretion of different ECM proteins. M2 macrophages were shown to secrete a high amount of fibronectin and collagen VI, two proteins involved in satellite cell self-renewal (26, 117, 265, 304). Taken together, dampening of the inflammatory reaction in later stages after injury promotes myogenesis by facilitating differentiation, fusion and self-renewal.

Nonsatellite cell types with myogenic potential muscle regeneration (129). IL-4 secreted by eosinophils indirectly promotes myogenesis by stimulating fibro-adipogenic progenitors. The proinflammatory environment that neutrophils generate also encourages the recruitment of monocytes. Two subsets of monocytes accumulate in muscles following injury, the classical monocytes (Ly6C+) and the nonclassical monocytes (Ly6C-) (11). Classical monocytes are predominant during the first few days after an injury and are known to have proinflammatory phenotype. Classical monocytes secrete higher level of inflammatory factors such as TNF-α and IL-1β. This cocktail of proinflammatory molecules promotes myoblast proliferation and delays differentiation (11, 298). Thereafter, nonclassical monocytes become the main monocyte population present in the damaged tissue. These cells rather release anti-inflammatory molecules such as IL-10 and TGF-β (11). Nonclassical monocytes decrease myoblast proliferation and stimulate differentiation and fusion (298). Thus, the inflammatory reaction directed by leukocytes regulates satellite cell activation, proliferation, and differentiation (Fig. 17). In the course of the inflammatory process, infiltrated monocytes will differentiate and mature into macrophages. The local microenvironment polarizes macrophages into two different subpopulations, proinflammatory M1 macrophages and anti-inflammatory M2 macrophages (116). M2 macrophages are critical for mediating the resolution of inflammation. Anti-inflammatory molecules and proresolving mediators released by these cells will stop the further recruitment of

1048

Distinct ECM proteins influence the function of satellite cells during muscle regeneration. Connective tissue fibroblasts residing in the interstitium between myofibers are major producers and organizers of ECM in skeletal muscle (Fig. 16). Fibroblasts strongly proliferate following muscle injury and their expansion closely correlates with the increase in satellite cell numbers (209). Partial depletion of fibroblasts during muscle regeneration leads to decreased numbers of satellite cells and premature differentiation of myoblasts. Another cell type that participates in ECM formation is the fibroadipogenic progenitors (FAPs) that can give rise to fibrocytes or adipocytes. FAPs were shown to proliferate during muscle injury, but contrary to fibroblasts they promote myogenic differentiation instead of proliferation (143). This myogenic effect of FAPs could be mediated by the release of different myogenic factors, such as IGF-1, or by increased phagocytic debris removal (129, 143). Certain cell types that participate in myogenesis can also acquire Pax7 expression and become satellite cells (224). Heterogeneous population of cells with hematopoietic potential resident in the bone marrow as well as in muscle tissue called side population (SP) cells can give rise to satellite cells and contribute to muscle fibers upon transplantation (12,123,211,291). Lineage tracing of SP cells has shown that they display a strong proliferative response following muscle injury but only marginally contribute to the formation of myofibers (89). Germline Abgc2 knockout mice do not have SP cells in muscle and display impaired adult myogenesis, reduced satellite cell numbers, and fewer immune cells during

Volume 5, July 2015

Comprehensive Physiology

regeneration (89). This suggests that under physiological conditions SP cells could have an accessory function for adult myogenesis by regulating the immune response. Mature muscle tissue also contains a population of Pax7 negative and PW1 (PEG3) positive interstitial cells termed PICs (PW1+ interstitial cells) (203). Upon transplantation, PICs engraft into muscle, contribute to muscle fibers and can generate Pax7-positive satellite cells. The population of PICs with myogenic potential in adult muscle has been shown to be Sca1 and CD34-positive but Pax7-negative. Importantly, freshly sorted PICs can acquire Pax7 and MyoD in vitro when cultured with myoblasts (203). Acquisition of a myogenic fate by PICs appears to be dependent on the induction of Pax7 expression. These results suggest that PICs represent a bipotent resident stem cell in skeletal muscle. Similar to other tissues, the microvasculature in adult skeletal muscle contains a population of contractile cells called pericytes that are wrapped around endothelial cells (114). Pericytes isolated from muscle do not express myogenic markers or endothelial markers but are positive for neuro-glial 2 proteoglycan (NG2), platelet derived growth factor receptor-beta (PDGFR-β), alpha-smooth muscle actin (αSMA) and alkaline phosphatase (AP) (85). In vitro, pericytes can differentiate into muscle fibers and coculture with myoblasts enhances this process. Pericytes can be expanded with high efficiency ex vivo and contribute to muscle fibers and satellite cells after infusion. Lineage tracing showed that pericytes, but not endothelial cells, are able to fuse to developing myofibers and contribute to myogenesis (85). This study also revealed that pericytes expand after muscle injury and that a relatively low number of pericytes can engraft in the satellite cell compartment. In addition to the above cell types that are present in the stem cell micro- or macroniche, other local or exogenous cell types such as CD133 and Integrin-β4 positive cells have been demonstrated to have myogenic potential (177, 260, 300). However, a role of these cell types during muscle regeneration remains to be determined. Moreover, as mentioned earlier, several ablation studies indicated that satellite cell depletion completely impairs muscle regeneration (see section on satellite cell and regeneration) (172,209,259,309). Consequently, satellite cells remain the principal myogenic precursor cells and no other cell type can compensate for their loss. On the other hand, in vitro experiments showed that nonsatellite cell types support adult myogenesis through indirect mechanisms that require coculture with myoblasts to participate to myotube formation. Thus, the absence of satellite cells in ablation experiments could also have indirectly affected the myogenic potential of nonsatellite cells. Overall, many cell types can contribute to myogenesis either indirectly by interacting with satellite cells or potentially by directly fusing to myofibers or engrafting into the satellite cell niche. Such cell types possess an important therapeutic potential and are currently under investigation by many groups.

Volume 5, July 2015

Satellite Cells and Skeletal Muscle Regeneration

Pathologies and Degenerative Conditions Aging The regenerative capacity of skeletal muscle decreases slowly with aging. Regeneration in aged injured muscle results in more fibrotic tissue, increased intramuscular fat accumulation, and smaller myofibers (119, 258). Aging is also accompanied by a decrease in satellite cell numbers (275). This effect varies depending on the species, the age, the type of muscle, and the markers used (40). In addition to the exhaustion of the satellite cell pool, the myogenic capacity of individual satellite cells also becomes impaired during aging. For instance, in vitro culture revealed that aged myoblasts are able to proliferate and differentiate albeit to a slower rate than younger myoblasts (16,55). Whether this decrease in satellite cell function is fundamentally intrinsic, or extrinsically induced by the changing environment, remains a matter of debate. Increasing evidence suggests that both processes are involved (40). Extrinsic influences affecting satellite cells during aging can be systemic (ex: changes in hormonal levels), or local (ex: muscle fibrosis). In vitro primary myoblasts exposed to the serum of young mice have a higher myogenic potential than cells treated with old serum (49). Moreover, it has been shown by transplantation experiments that the age of the recipient defines the regenerative capacity of satellite cells (48). Therefore, old muscles that were transplanted into a young recipient have a higher regenerative capacity than young muscles transplanted into an old recipient. Importantly, it has been shown that aged muscles fail to increase levels of the Notch ligand Delta during muscle injury (68). As previously discussed, Notch has multiple essential roles in satellite cell activation, proliferation, and self-renewal (Table 1). Parabiotic pairings (e.g., a shared circulatory system between two mice) has revealed that Notch signaling can be restored in aged satellite cells by exposure to young serum (69). These results demonstrate that under certain conditions the satellite cell environment can partially overcome possible intrinsic defects in aged satellite cells. Old muscles were also shown to secrete excessive amounts of TGF-β, which overactivates pSmad3 and impairs satellite cell function (50). Interestingly, the pSmad3 and Notch pathways appear to antagonize each other. Activated aged satellite cells also exhibit a higher canonical Wnt signaling activity, which is potentially mediated by increased Wnt protein levels in the serum (40). Canonical Wnt signaling in old activated satellite cells results in the conversion of myogenic cells to fibrogenic cells. These results clearly illustrate the importance of the micro- and macroenvironment in the regulation of satellite cell myogenic potential. It has been suggested that a loss of satellite cell quiescence is responsible for the exhaustion of satellite cell pool and the reduced regenerative potential of aged muscles. Aged muscle fibers were shown to express more FGF2, which activates satellite cells and decreases their self-renewal potential (53). Moreover, loss of self-renewal in old satellite cells has

1049

Satellite Cells and Skeletal Muscle Regeneration

Comprehensive Physiology

been demonstrated to arise from an impaired FGF receptor response and increased p38α/β signaling (28). Increased p38α/β signaling in aged satellite cells has also been shown to stimulate the expression of senescence markers (74). Loss of reversible quiescence in aged satellite cells is accompanied by derepression of the CKI Cdkn2a (285). The above studies provided evidence that loss of quiescence and self-renewal is irreversible and cannot be restored completely by a youthful environment. Recently, it has been shown that the JAK/STAT signaling pathway is upregulated in aged satellite cells (237, 299). Increased Stat3 signaling is at least partially mediated by elevated level of IL-6 inflammatory cytokine (299). Stat3 overexpression leads to increased asymmetric division of satellite stem cells and promotes myogenic lineage progression through MyoD1 expression. Increased asymmetric division is correlated with an exhaustion of satellite stem cell subpopulation (237). Knockdown of Jak2 and/or Stat3 improves satellite stem cell symmetric division and increases the selfrenewal potential upon transplantation. Moreover, pharmacological inhibitors of Jak/Stat improve muscle regeneration and muscle force. Telomere shortening associated with aging has been hypothesized to be responsible for cellular senescence, e.g. the cessation of cell division caused by aging (43). A correlation between telomere length and satellite cell proliferative potential has been described (83). However, human satellite cells show only marginal signs of telomere shortening during aging (84). Thus, the exact role of telomere shortening on aged satellite cell behavior remains to be determined. Taken together, satellite cell functions rely on a fine-tuned balance of both intrinsic and extrinsic factors that can be disturbed by aging.

Muscular Dystrophy MDs are characterized by muscle fragility and weakness. Duchenne muscular dystrophy (DMD) is the most common form of MD. DMD patients show early signs of muscle H&E

weakness during childhood that progressively worsens and leads to an inability to walk and premature death around 20 to 30 years of age (9). DMD is caused by a genetic mutation that results in the absence of dystrophin, a structural protein involved in the interaction of the muscle fiber with the ECM. In absence of dystrophin, muscle fibers are fragile and prone to injury leading to continuous cycles of degeneration and regeneration (273). The chronically degenerative environment in MDs has negative effects on satellite cell function. In dystrophic muscles, satellite cells are constantly exposed to an activating environment. It has been speculated that this continuous activation results in premature exhaustion of the satellite cell pool. Primary cultures of myoblasts from dystrophic muscles show a much lower proliferative potential in vitro than healthy myoblasts, a phenomenon that is accentuated by aging (314). Moreover, it has been demonstrated that telomeres are significantly shorter in dystrophic satellite cells and it has been speculated that this phenomenon could be responsible for the decreased regenerative ability of muscle satellite cell observed in MDs (83, 256). Contradicting these ideas, biopsies from humans MD patients or dystrophic mice are characterized by elevated numbers of satellite cells even in advanced stages of dystrophy (52, 161, 243). In dystrophic muscles, satellite cells are surrounded by an altered microenvironment composed of fibrotic tissue, fat, and inflammatory cells (Fig. 18). Under dystrophic conditions injured myofibers appear to send cues to FAPs that favor adipogenesis and fibrosis (143, 303). A reduction of fibrosis in dystrophin-null mice (mdx) mediated by the application of TGF-β inhibitors decreases myofiber necrosis and improves muscle regeneration (66, 293). Similarly, inflammatory cells are deregulated in dystrophic muscles and impair muscle regeneration and satellite cell function. The chronic degenerative environment encourages the polarization of leukocytes, such as macrophages, toward the proinflammatory phenotype (M1) (307). Thus, anti-inflammatory corticosteroids remain one of the few therapeutic options that can temporarily ameliorate the symptoms of MDs (191).

Sirius red

Trichrome masson

Figure 18

Degeneration of dystrophic muscles. Images from dystrophic tibialis anterior muscles. H&E staining showing centronucleated fibers that indicate muscle regeneration. Sirius red (fibrotic tissue in red) and trichrome masson (fibrotic tissue in blue) illustrate the massive fibrosis response in the dystrophic muscles.

1050

Volume 5, July 2015

Comprehensive Physiology

Next to the altered extracellular environment, certain evidence suggests that intrinsic changes can affect satellite cell function in muscle dystrophy. Dystrophin stabilizes cells by binding to the dystroglycan (DAG) complex on the cell membrane surface. The DAG also serves as anchoring protein for ECM proteins such as laminin. In absence of dystrophin the DAG complex is strongly perturbed. Interestingly, it was found that disruption of DAG1 specifically in the myofibers (e.g., not in the satellite cells) results in a remarkably mild dystrophic phenotype (65). These results suggest that myofiber fragility is not the only problem underlying MD and that satellite cell dysfunction could play a role in the pathogenesis of this group of diseases. Overall, there are many conflicting results in the literature with regard to the role of satellite cells in MDs. One reason underlying this lack of consensus is the absence of an experimental model that appropriately mimics human DMD. Deletion of dystrophin gene in mdx mice does not result in a phenotype that is similar to the one observed in humans. Mdx mice appear to be able to compensate much more efficiently for the absence of dystrophin than humans and other higher organisms. It was hypothesized that this higher resistance to muscle degeneration could be mediated by compensation by a related structural protein such as utrophin or by a higher regenerative potential of mouse satellite cells (318).

Muscle Stem Cell Regenerative Therapies The myogenic potential of satellite cells can be perturbed by different pathologies and conditions. Research on regenerative therapies for skeletal muscle often focuses on delivering healthy myogenic cells into diseased muscles or investigates the possibility to improve or restore the endogenous myogenic potential of satellite cells. Muscle satellite cell transplantation has been studied extensively in the last decades and holds great therapeutic potential for diseases such as MD. In addition to restoring the satellite cell pool and to repairing the host myofibers, healthy myogenic donor cells can serve as vectors to mediate expression of normal alleles in muscle fibers they fuse to (Fig. 19) (228). Thus, fusion of healthy transplanted donor cells to dystrophic myofibers will allow the donor nuclei to express the normal dystrophin allele in the host muscle fibers. Unfortunately, such therapeutic approaches have many technical issues. There is for instance a need for immunosuppression, donor cells often show poor survival after injection and a high number of injections are required to get an effect on a given muscle. Moreover, the engraftment potential of in vitro cultured myoblasts is very low and the surviving myoblasts rapidly fuse to the fibers and show no long-term engraftment into the stem cell compartment (136). In contrast to in vitro cultured myoblasts it has been shown that freshly isolated satellite cells (especially the satellite stem cell subpopulation) are capable of long-term engraftment and can sustain multiple round of injuries (51, 67, 163, 255). However, obtaining

Volume 5, July 2015

Satellite Cells and Skeletal Muscle Regeneration

sufficient numbers of freshly isolated satellite cells to treat multiple muscles is impractical. Thus, so far the success of myogenic cell therapy in humans is very limited (297). Niche components such as ECM proteins involved in selfrenewal (e.g., collagen VI or fibronectin) have been used to mimic the satellite cell microenvironment in vitro and can lead to improved transplantation efficiency of cultured cells (26, 110, 304). It was also shown that it is feasible to isolate viable satellite cells from postmortem tissue, which may allow for obtaining sufficient numbers of cells for therapy. Another source for myogenic cells could be the conditional expression of Pax7 in embryonic stem cells or inducible pluripotent stem cells. Myogenic cells generated with this technique can engraft in the satellite cell niche and contribute to muscle fibers (79). Myogenic cells have also been coinjected with growth factors or with other cell types that improve cell survival, migration, and engraftment (20, 38, 164, 173). For instance, short ex vivo treatment of satellite cells with Wnt7a considerably increases dispersion and engraftment that ultimately results in improved function of dystrophic muscles (23). However, despite these various improvements, myogenic cell transplantation is still far from being an efficient clinical therapy. An option that allows for circumventing the huge amount of intramuscular injections needed for transplantation (25-100 injections/cm3 ) is to use cell types that have the potential to be delivered systemically and migrate through the vasculature to engraft as satellite cells (280). Mesoangioblasts, CD133+, SP cells and others have shown potential for systemic delivery. However, the engraftment potential of these cell types into the satellite cell compartment appears to be limited (15, 85, 123, 199, 260). A promising avenue to improve muscle regeneration in disease is to boost the endogenous myogenic potential of satellite cells. Intramuscular injection of Wnt7a can increase satellite cell numbers, myofiber size and muscle force in mdx muscles (310). Pharmacological inhibition of p38 MAPK could be another strategy to rejuvenate old satellite cells (28, 74). Inhibitors of p38 are already investigated in different clinical trials for various inflammatory conditions (22). Similarly, JAK/STAT inhibitors were also demonstrated to restore old muscle satellite cell regenerative potential (237). However, administration of systemic inhibitors strongly increases the chances of adverse side effects.

Conclusion In this review, we summarized the characteristics of satellite cells and their functions in healthy and regenerating skeletal muscle. Especially, we showed that during muscle regeneration satellite cell fate is tightly regulated by intrinsic and extrinsic factors, a fragile balance that can be perturbed by diseases or degenerative conditions. Moreover, the relatively recent discovery of satellite cell subpopulations with selfrenewal potential greatly improved our understanding of the

1051

Satellite Cells and Skeletal Muscle Regeneration

Satellite cellderived myogenic cells in culture

Comprehensive Physiology

Freshly isolated satellite cells Donor muscle tissue

Injection of donor cells

Host skeletal muscle

Mosaic muscle

Host muscle fiber

Donor cells fusing to host fibers Genetically corrected host fiber

Donor nucleus

Figure 19 Satellite cell therapy. Schematic illustrating the different steps of cell therapy. Healthy satellite cells are isolated from donor muscle tissue and are cultured in vitro or directly transplanted into diseased host muscle. Healthy satellite cells will then fuse to host myofibers to form hybrid fibers. Addition of new healthy nuclei to host myofibers can partially correct genetic disorders such as muscular dystrophies. However, this therapeutic avenue has major technical limitations such as short-term engraftment, limited availability of donor cells, and poor cell migration and survival.

mechanisms that control the overall satellite cell pool. Future advances in molecular and genetic technologies will help unravel the mechanisms controlling these stem cells and will hopefully allow for the development of efficient therapies for muscle diseases.

Acknowledgements The studies from the laboratory of M. A. Rudnicki were carried out with support of grants from the Canadian

1052

Institutes for Health Research (MOP-81288, MOP-12080, and ERA-132935), the US National Institutes for Health (R01AR044031), the Muscular Dystrophy Association, Muscular Dystrophy Canada, the Stem Cell Network, and the Ontario Ministry of Economic Development and Innovation, and the Canada Research Chair Program. M. A. Rudnicki holds a Canada Research Chair in Molecular genetics. N. A. Dumont is supported by a Postdoctoral Fellowship from the Canadian Institutes for Health Research. We thank Caroline Brun for reviewing the manuscript.

Volume 5, July 2015

Comprehensive Physiology

References 1. 2.

3.

4. 5.

6. 7. 8. 9. 10. 11.

12. 13.

14.

15.

16. 17.

18. 19.

20.

21.

22. 23.

Abmayr SM, Pavlath GK. Myoblast fusion: Lessons from flies and mice. Development 139: 641-656, 2012. Abou-Khalil R, Le Grand F, Pallafacchina G, Valable S, Authier FJ, Rudnicki MA, Gherardi RK, Germain S, Chretien F, Sotiropoulos A, Lafuste P, Montarras D, Chazaud B. Autocrine and paracrine angiopoietin 1/Tie-2 signaling promotes muscle satellite cell self-renewal. Cell Stem Cell 5: 298-309, 2009. Acharyya S, Sharma SM, Cheng AS, Ladner KJ, He W, Kline W, Wang H, Ostrowski MC, Huang TH, Guttridge DC. TNF inhibits Notch1 in skeletal muscle cells by Ezh2 and DNA methylation mediated repression: Implications in duchenne muscular dystrophy. PloS One 5: e12479, 2010. Adli M, Bernstein BE. Whole-genome chromatin profiling from limited numbers of cells using nano-ChIP-seq. Nat Protoc 6: 1656-1668, 2011. Alfaro LA, Dick SA, Siegel AL, Anonuevo AS, McNagny KM, Megeney LA, Cornelison DD, Rossi FM. CD34 promotes satellite cell motility and entry into proliferation to facilitate efficient skeletal muscle regeneration. Stem Cells 29: 2030-2041, 2011. Allen RE, Sheehan SM, Taylor RG, Kendall TL, Rice GM. Hepatocyte growth factor activates quiescent skeletal muscle satellite cells in vitro. J Cell Physiol 165: 307-312, 1995. Alter J, Rozentzweig D, Bengal E. Inhibition of myoblast differentiation by tumor necrosis factor alpha is mediated by c-Jun N-terminal kinase 1 and leukemia inhibitory factor. J Biol Chem 283: 23224-23234, 2008. Anderson JE. A role for nitric oxide in muscle repair: Nitric oxidemediated activation of muscle satellite cells. Mol Biol Cell 11: 18591874, 2000. Anderson MS, Kunkel LM. The molecular and biochemical basis of Duchenne muscular dystrophy. Trends Biochem Sci 17: 289-292, 1992. Argiles JM, Busquets S, Felipe A, Lopez-Soriano FJ. Muscle wasting in cancer and ageing: Cachexia versus sarcopenia. Adv Gerontol 18: 39-54, 2006. Arnold L, Henry A, Poron F, Baba-Amer Y, van Rooijen N, Plonquet A, Gherardi RK, Chazaud B. Inflammatory monocytes recruited after skeletal muscle injury switch into antiinflammatory macrophages to support myogenesis. J Exp Med 204: 1057-1069, 2007. Asakura A, Seale P, Girgis-Gabardo A, Rudnicki MA. Myogenic specification of side population cells in skeletal muscle. J Cell Biol 159: 123-134, 2002. Asp P, Blum R, Vethantham V, Parisi F, Micsinai M, Cheng J, Bowman C, Kluger Y, Dynlacht BD. Genome-wide remodeling of the epigenetic landscape during myogenic differentiation. Proc Natl Acad Sci U S A 108: E149-E158, 2011. Auffray C, Fogg D, Garfa M, Elain G, Join-Lambert O, Kayal S, Sarnacki S, Cumano A, Lauvau G, Geissmann F. Monitoring of blood vessels and tissues by a population of monocytes with patrolling behavior. Science 317: 666-670, 2007. Bachrach E, Li S, Perez AL, Schienda J, Liadaki K, Volinski J, Flint A, Chamberlain J, Kunkel LM. Systemic delivery of human microdystrophin to regenerating mouse dystrophic muscle by muscle progenitor cells. Proc Natl Acad Sci U S A 101: 3581-3586, 2004. Barani AE, Durieux AC, Sabido O, Freyssenet D. Age-related changes in the mitotic and metabolic characteristics of muscle-derived cells. J Appl Physiol (1985) 95: 2089-2098, 2003. Barjot C, Cotten ML, Goblet C, Whalen RG, Bacou F. Expression of myosin heavy chain and of myogenic regulatory factor genes in fast or slow rabbit muscle satellite cell cultures. J Muscle Res Cell Motil 16: 619-628, 1995. Bassel-Duby R, Olson EN. Signaling pathways in skeletal muscle remodeling. Annu Rev Biochem 75: 19-37, 2006. Beauchamp JR, Heslop L, Yu DS, Tajbakhsh S, Kelly RG, Wernig A, Buckingham ME, Partridge TA, Zammit PS. Expression of CD34 and Myf5 defines the majority of quiescent adult skeletal muscle satellite cells. J Cell Biol 151: 1221-1234, 2000. Bencze M, Negroni E, Vallese D, Yacoub-Youssef H, Chaouch S, Wolff A, Aamiri A, Di Santo JP, Chazaud B, Butler-Browne G, Savino W, Mouly V, Riederer I. Proinflammatory macrophages enhance the regenerative capacity of human myoblasts by modifying their kinetics of proliferation and differentiation. Mol Ther 20: 2168-2179, 2012. Bentzinger CF, Lin S, Romanino K, Castets P, Guridi M, Summermatter S, Handschin C, Tintignac LA, Hall MN, Ruegg MA. Differential response of skeletal muscles to mTORC1 signaling during atrophy and hypertrophy. Skelet Muscle 3: 6, 2013. Bentzinger CF, Rudnicki MA. Rejuvenating aged muscle stem cells. Nat Med 20: 234-235, 2014. Bentzinger CF, von Maltzahn J, Dumont NA, Stark DA, Wang YX, Nhan K, Frenette J, Cornelison DD, Rudnicki MA. Wnt7a stimulates myogenic stem cell motility and engraftment resulting in improved muscle strength. J Cell Biol 205: 97-111, 2014.

Volume 5, July 2015

Satellite Cells and Skeletal Muscle Regeneration

24. Bentzinger CF, Wang YX, Dumont NA, Rudnicki MA. Cellular dynamics in the muscle satellite cell niche. EMBO Rep 14: 1062-1072, 2013. 25. Bentzinger CF, Wang YX, Rudnicki MA. Building muscle: Molecular regulation of myogenesis. Cold Spring Harb Perspect Biol 4: pii: a008342, 2012. 26. Bentzinger CF, Wang YX, von Maltzahn J, Soleimani VD, Yin H, Rudnicki MA. Fibronectin regulates Wnt7a signaling and satellite cell expansion. Cell Stem Cell 12: 75-87, 2013. 27. Berkes CA, Bergstrom DA, Penn BH, Seaver KJ, Knoepfler PS, Tapscott SJ. Pbx marks genes for activation by MyoD indicating a role for a homeodomain protein in establishing myogenic potential. Mol Cell 14: 465-477, 2004. 28. Bernet JD, Doles JD, Hall JK, Kelly Tanaka K, Carter TA, Olwin BB. 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. 29. Bintliff S, Walker BE. Radioautographic study of skeletal muscle regeneration. Am J Anat 106: 233-245, 1960. 30. Biressi S, Molinaro M, Cossu G. Cellular heterogeneity during vertebrate skeletal muscle development. Dev Biol 308: 281-293, 2007. 31. Bischoff R. Regeneration of single skeletal muscle fibers in vitro. Anat Rec 182: 215-235, 1975. 32. Bischoff R. Chemotaxis of skeletal muscle satellite cells. Dev Dyn 208: 505-515, 1997. 33. Bjornson CR, Cheung TH, Liu L, Tripathi PV, Steeper KM, Rando TA. Notch signaling is necessary to maintain quiescence in adult muscle stem cells. Stem Cells 30: 232-242, 2012. 34. Blaauw B, Schiaffino S, Reggiani C. Mechanisms modulating skeletal muscle phenotype. Compr Physiol 3: 1645-1687, 2013. 35. Bober E, Franz T, Arnold HH, Gruss P, Tremblay P. Pax-3 is required for the development of limb muscles: A possible role for the migration of dermomyotomal muscle progenitor cells. Development 120: 603612, 1994. 36. Bonavaud S, Agbulut O, Nizard R, D’Honneur G, Mouly V, ButlerBrowne G. A discrepancy resolved: Human satellite cells are not preprogrammed to fast and slow lineages. Neuromuscul Disord 11: 747-752, 2001. 37. Bosnakovski D, Xu Z, Li W, Thet S, Cleaver O, Perlingeiro RC, Kyba M. Prospective isolation of skeletal muscle stem cells with a Pax7 reporter. Stem Cells 26: 3194-3204, 2008. 38. Bouchentouf M, Benabdallah BF, Bigey P, Yau TM, Scherman D, Tremblay JP. Vascular endothelial growth factor reduced hypoxiainduced death of human myoblasts and improved their engraftment in mouse muscles. Gene Ther 15: 404-414, 2008. 39. Brack AS, Conboy IM, Conboy MJ, Shen J, Rando TA. A temporal switch from notch to Wnt signaling in muscle stem cells is necessary for normal adult myogenesis. Cell Stem Cell 2: 50-59, 2008. 40. Brack AS, Rando TA. Intrinsic changes and extrinsic influences of myogenic stem cell function during aging. Stem Cell Rev 3: 226-237, 2007. 41. Braun T, Rudnicki MA, Arnold HH, Jaenisch R. Targeted inactivation of the muscle regulatory gene Myf-5 results in abnormal rib development and perinatal death. Cell 71: 369-382, 1992. 42. Buckingham M, Rigby PW. Gene regulatory networks and transcriptional mechanisms that control myogenesis. Dev Cell 28: 225-238, 2014. 43. Campisi J, Kim SH, Lim CS, Rubio M. Cellular senescence, cancer and aging: The telomere connection. Exp Gerontol 36: 1619-1637, 2001. 44. Cantini M, Massimino ML, Rapizzi E, Rossini K, Catani C, Dalla Libera L, Carraro U. Human satellite cell proliferation in vitro is regulated by autocrine secretion of IL-6 stimulated by a soluble factor(s) released by activated monocytes. Biochem Biophys Res Commun 216: 49-53, 1995. 45. Capers CR. Multinucleation of skeletal muscle in vitro. J Biophy Biochem Cytol 7: 559-565, 1960. 46. Cardinali B, Castellani L, Fasanaro P, Basso A, Alema S, Martelli F, Falcone G. Microrna-221 and microrna-222 modulate differentiation and maturation of skeletal muscle cells. PloS One 4: e7607, 2009. 47. Caretti G, Di Padova M, Micales B, Lyons GE, Sartorelli V. The Polycomb Ezh2 methyltransferase regulates muscle gene expression and skeletal muscle differentiation. Genes Dev 18: 2627-2638, 2004. 48. Carlson BM, Faulkner JA. Muscle transplantation between young and old rats: Age of host determines recovery. Am J Physiol 256: C1262C1266, 1989. 49. Carlson ME, Conboy IM. Loss of stem cell regenerative capacity within aged niches. Aging Cell 6: 371-382, 2007. 50. Carlson ME, Hsu M, Conboy IM. Imbalance between pSmad3 and Notch induces CDK inhibitors in old muscle stem cells. Nature 454: 528-532, 2008. 51. Cerletti M, Jurga S, Witczak CA, Hirshman MF, Shadrach JL, Goodyear LJ, Wagers AJ. Highly efficient, functional engraftment of skeletal muscle stem cells in dystrophic muscles. Cell 134: 37-47, 2008.

1053

Satellite Cells and Skeletal Muscle Regeneration

52.

53. 54.

55. 56.

57. 58. 59. 60. 61.

62. 63. 64. 65.

66.

67.

68. 69. 70. 71. 72. 73.

74.

75.

Chakkalakal JV, Christensen J, Xiang W, Tierney MT, Boscolo FS, Sacco A, Brack AS. Early forming label-retaining muscle stem cells require p27kip1 for maintenance of the primitive state. Development 141: 1649-1659, 2014. Chakkalakal JV, Jones KM, Basson MA, Brack AS. The aged niche disrupts muscle stem cell quiescence. Nature 490: 355-360, 2012. Chakravarthy MV, Abraha TW, Schwartz RJ, Fiorotto ML, Booth FW. Insulin-like growth factor-I extends in vitro replicative life span of skeletal muscle satellite cells by enhancing G1/S cell cycle progression via the activation of phosphatidylinositol 3’-kinase/Akt signaling pathway. J Biol Chem 275: 35942-35952, 2000. Charge SB, Brack AS, Hughes SM. Aging-related satellite cell differentiation defect occurs prematurely after Ski-induced muscle hypertrophy. Am J Physiol Cell Physiol 283: C1228-C1241, 2002. Chen JF, Mandel EM, Thomson JM, Wu Q, Callis TE, Hammond SM, Conlon FL, Wang DZ. The role of microRNA-1 and microRNA-133 in skeletal muscle proliferation and differentiation. Nat Genet 38: 228233, 2006. Chen JF, Tao Y, Li J, Deng Z, Yan Z, Xiao X, Wang DZ. microRNA-1 and microRNA-206 regulate skeletal muscle satellite cell proliferation and differentiation by repressing Pax7. J Cell Biol 190: 867-879, 2010. Chen SE, Jin B, Li YP. TNF-alpha regulates myogenesis and muscle regeneration by activating p38 MAPK. Am J Physiol Cell Physiol 292: C1660-C1671, 2007. Cheung TH, Quach NL, Charville GW, Liu L, Park L, Edalati A, Yoo B, Hoang P, Rando TA. Maintenance of muscle stem-cell quiescence by microRNA-489. Nature 482: 524-528, 2012. Cheung TH, Rando TA. Molecular regulation of stem cell quiescence. Nat Rev Mol Cell Biol 14: 329-340, 2013. Christov C, Chretien F, Abou-Khalil R, Bassez G, Vallet G, Authier FJ, Bassaglia Y, Shinin V, Tajbakhsh S, Chazaud B, Gherardi RK. Muscle satellite cells and endothelial cells: Close neighbors and privileged partners. Mol Biol Cell 18: 1397-1409, 2007. Cirillo LA, Lin FR, Cuesta I, Friedman D, Jarnik M, Zaret KS. Opening of compacted chromatin by early developmental transcription factors HNF3 (FoxA) and GATA-4. Mol Cell 9: 279-289, 2002. Clark P, Coles D, Peckham M. Preferential adhesion to and survival on patterned laminin organizes myogenesis in vitro. Exp Cell Res 230: 275-283, 1997. Clark WE, Blomfield LB. The efficiency of intramuscular anastomoses, with observations on the regeneration of devascularized muscle. J Anat 79: 15-32.4, 1945. Cohn RD, Henry MD, Michele DE, Barresi R, Saito F, Moore SA, Flanagan JD, Skwarchuk MW, Robbins ME, Mendell JR, Williamson RA, Campbell KP. Disruption of DAG1 in differentiated skeletal muscle reveals a role for dystroglycan in muscle regeneration. Cell 110: 639648, 2002. Cohn RD, van Erp C, Habashi JP, Soleimani AA, Klein EC, Lisi MT, Gamradt M, ap Rhys CM, Holm TM, Loeys BL, Ramirez F, Judge DP, Ward CW, Dietz HC. Angiotensin II type 1 receptor blockade attenuates TGF-beta-induced failure of muscle regeneration in multiple myopathic states. Nat Med 13: 204-210, 2007. Collins CA, Olsen I, Zammit PS, Heslop L, Petrie A, Partridge TA, Morgan JE. Stem cell function, self-renewal, and behavioral heterogeneity of cells from the adult muscle satellite cell niche. Cell 122: 289-301, 2005. Conboy IM, Conboy MJ, Smythe GM, Rando TA. Notch-mediated restoration of regenerative potential to aged muscle. Science 302: 15751577, 2003. Conboy IM, Conboy MJ, Wagers AJ, Girma ER, Weissman IL, Rando TA. Rejuvenation of aged progenitor cells by exposure to a young systemic environment. Nature 433: 760-764, 2005. Conboy IM, Rando TA. The regulation of Notch signaling controls satellite cell activation and cell fate determination in postnatal myogenesis. Dev Cell 3: 397-409, 2002. Corbeil HB, Whyte P, Branton PE. Characterization of transcription factor E2F complexes during muscle and neuronal differentiation. Oncogene 11: 909-920, 1995. Cornelison DD, Olwin BB, Rudnicki MA, Wold BJ. MyoD(-/-) satellite cells in single-fiber culture are differentiation defective and MRF4 deficient. Dev Biol 224: 122-137, 2000. Cornelison DD, Wilcox-Adelman SA, Goetinck PF, Rauvala H, Rapraeger AC, Olwin BB. Essential and separable roles for Syndecan3 and Syndecan-4 in skeletal muscle development and regeneration. Genes Dev 18: 2231-2236, 2004. Cosgrove BD, Gilbert PM, Porpiglia E, Mourkioti F, Lee SP, Corbel SY, Llewellyn ME, Delp SL, Blau HM. Rejuvenation of the muscle stem cell population restores strength to injured aged muscles. Nat Med 20: 255-264, 2014. Crist CG, 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.

1054

Comprehensive Physiology

76. Cserjesi P, Olson EN. Myogenin induces the myocyte-specific enhancer binding factor MEF-2 independently of other muscle-specific gene products. Mol Cell Biol 11: 4854-4862, 1991. 77. Dacwag CS, Ohkawa Y, Pal S, Sif S, Imbalzano AN. The protein arginine methyltransferase Prmt5 is required for myogenesis because it facilitates ATP-dependent chromatin remodeling. Mol Cell Biol 27: 384-394, 2007. 78. Dalkilic I, Schienda J, Thompson TG, Kunkel LM. Loss of FilaminC (FLNc) results in severe defects in myogenesis and myotube structure. Mol Cell Biol 26: 6522-6534, 2006. 79. Darabi R, Arpke RW, Irion S, Dimos JT, Grskovic M, Kyba M, Perlingeiro RC. Human ES- and iPS-derived myogenic progenitors restore DYSTROPHIN and improve contractility upon transplantation in dystrophic mice. Cell Stem Cell 10: 610-619, 2012. 80. Davie JK, Cho JH, Meadows E, Flynn JM, Knapp JR, Klein WH. Target gene selectivity of the myogenic basic helix-loop-helix transcription factor myogenin in embryonic muscle. Dev Biol 311: 650-664, 2007. 81. de la Serna IL, Carlson KA, Imbalzano AN. Mammalian SWI/SNF complexes promote MyoD-mediated muscle differentiation. Nat Genet 27: 187-190, 2001. 82. de la Serna IL, Ohkawa Y, Berkes CA, Bergstrom DA, Dacwag CS, Tapscott SJ, Imbalzano AN. MyoD targets chromatin remodeling complexes to the myogenin locus prior to forming a stable DNA-bound complex. Mol Cell Biol 25: 3997-4009, 2005. 83. Decary S, Hamida CB, Mouly V, Barbet JP, Hentati F, Butler-Browne GS. Shorter telomeres in dystrophic muscle consistent with extensive regeneration in young children. Neuromuscul Disord 10: 113-120, 2000. 84. Decary S, Mouly V, Hamida CB, Sautet A, Barbet JP, Butler-Browne GS. Replicative potential and telomere length in human skeletal muscle: Implications for satellite cell-mediated gene therapy. Hum Gene Ther 8: 1429-1438, 1997. 85. Dellavalle A, Sampaolesi M, Tonlorenzi R, Tagliafico E, Sacchetti B, Perani L, Innocenzi A, Galvez BG, Messina G, Morosetti R, Li S, Belicchi M, Peretti G, Chamberlain JS, Wright WE, Torrente Y, Ferrari S, Bianco P, Cossu G. Pericytes of human skeletal muscle are myogenic precursors distinct from satellite cells. Nat Cell Biol 9: 255-267, 2007. 86. Deng B, Wehling-Henricks M, Villalta SA, Wang Y, Tidball JG. IL-10 triggers changes in macrophage phenotype that promote muscle growth and regeneration. J Immunol 189: 3669-3680, 2012. 87. Dey BK, Gagan J, Dutta A. miR-206 and -486 induce myoblast differentiation by downregulating Pax7. Mol Cell Biol 31: 203-214, 2011. 88. DiMario J, Buffinger N, Yamada S, Strohman RC. Fibroblast growth factor in the extracellular matrix of dystrophic (mdx) mouse muscle. Science 244: 688-690, 1989. 89. Doyle MJ, Zhou S, Tanaka KK, Pisconti A, Farina NH, Sorrentino BP, Olwin BB. Abcg2 labels multiple cell types in skeletal muscle and participates in muscle regeneration. J Cell Biol 195: 147-163, 2011. 90. Duchesne E, Bouchard P, Roussel MP, Cote CH. Mast cells can regulate skeletal muscle cell proliferation by multiple mechanisms. Muscle Nerve 48: 403-414, 2013. 91. Duchesne E, Tremblay MH, Cote CH. Mast cell tryptase stimulates myoblast proliferation; A mechanism relying on protease-activated receptor-2 and cyclooxygenase-2. BMC Musculoskelet Disord 12: 235, 2011. 92. Dumont N, Bouchard P, Frenette J. Neutrophil-induced skeletal muscle damage: A calculated and controlled response following hindlimb unloading and reloading. Am J Physiol Regul Integr Comp Physiol 295: R1831-R1838, 2008. 93. Dumont N, Lepage K, Cote CH, Frenette J. Mast cells can modulate leukocyte accumulation and skeletal muscle function following hindlimb unloading. J Appl Physiol (1985) 103: 97-104, 2007. 94. Dumont NA, Frenette J. Macrophage colony-stimulating factor-induced macrophage differentiation promotes regrowth in atrophied skeletal muscles and C2C12 myotubes. Am J Pathol 182: 505-515, 2013. 95. Edom F, Mouly V, Barbet JP, Fiszman MY, Butler-Browne GS. Clones of human satellite cells can express in vitro both fast and slow myosin heavy chains. Dev Biol 164: 219-229, 1994. 96. Epstein JA, Shapiro DN, Cheng J, Lam PY, Maas RL. Pax3 modulates expression of the c-Met receptor during limb muscle development. Proc Natl Acad Sci U S A 93: 4213-4218, 1996. 97. Feldman JL, Stockdale FE. Skeletal muscle satellite cell diversity: Satellite cells form fibers of different types in cell culture. Dev Biol 143: 320-334, 1991. 98. Fitts RH, Riley DR, Widrick JJ. Physiology of a microgravity environment invited review: Microgravity and skeletal muscle. J Appl Physiol (1985) 89: 823-839, 2000. 99. Floss T, Arnold HH, Braun T. A role for FGF-6 in skeletal muscle regeneration. Genes Dev 11: 2040-2051, 1997. 100. Friday BB, Horsley V, Pavlath GK. Calcineurin activity is required for the initiation of skeletal muscle differentiation. J Cell Biol 149: 657-666, 2000.

Volume 5, July 2015

Comprehensive Physiology

101. Friday BB, Mitchell PO, Kegley KM, Pavlath GK. Calcineurin initiates skeletal muscle differentiation by activating MEF2 and MyoD. Differentiation 71: 217-227, 2003. 102. Fuchtbauer EM, Westphal H. MyoD and myogenin are coexpressed in regenerating skeletal muscle of the mouse. Dev Dyn 193: 34-39, 1992. 103. Fukada S, Higuchi S, Segawa M, Koda K, Yamamoto Y, Tsujikawa K, Kohama Y, Uezumi A, Imamura M, Miyagoe-Suzuki Y, Takeda S, Yamamoto H. Purification and cell-surface marker characterization of quiescent satellite cells from murine skeletal muscle by a novel monoclonal antibody. Exp Cell Res 296: 245-255, 2004. 104. Fukada S, Uezumi A, Ikemoto M, Masuda S, Segawa M, Tanimura N, Yamamoto H, Miyagoe-Suzuki Y, Takeda S. Molecular signature of quiescent satellite cells in adult skeletal muscle. Stem Cells 25: 24482459, 2007. 105. Fukada S, Yamaguchi M, Kokubo H, Ogawa R, Uezumi A, Yoneda T, Matev MM, Motohashi N, Ito T, Zolkiewska A, Johnson RL, Saga Y, Miyagoe-Suzuki Y, Tsujikawa K, Takeda S, Yamamoto H. Hesr1 and Hesr3 are essential to generate undifferentiated quiescent satellite cells and to maintain satellite cell numbers. Development 138: 4609-4619, 2011. 106. Fulco M, Schiltz RL, Iezzi S, King MT, Zhao P, Kashiwaya Y, Hoffman E, Veech RL, Sartorelli V. Sir2 regulates skeletal muscle differentiation as a potential sensor of the redox state. Mol Cell 12: 51-62, 2003. 107. Galli SJ, Borregaard N, Wynn TA. Phenotypic and functional plasticity of cells of innate immunity: Macrophages, mast cells and neutrophils. Nat Immunol 12: 1035-1044, 2011. 108. Germani A, Di Carlo A, Mangoni A, Straino S, Giacinti C, Turrini P, Biglioli P, Capogrossi MC. Vascular endothelial growth factor modulates skeletal myoblast function. Am J Pathol 163: 1417-1428, 2003. 109. Gibson MC, Schultz E. Age-related differences in absolute numbers of skeletal muscle satellite cells. Muscle Nerve 6: 574-580, 1983. 110. Gilbert PM, Havenstrite KL, Magnusson KE, Sacco A, Leonardi NA, Kraft P, Nguyen NK, Thrun S, Lutolf MP, Blau HM. Substrate elasticity regulates skeletal muscle stem cell self-renewal in culture. Science 329: 1078-1081, 2010. 111. Gildor B, Schejter ED, Shilo BZ. Bidirectional Notch activation represses fusion competence in swarming adult Drosophila myoblasts. Development 139: 4040-4050, 2012. 112. Gillespie MA, Le Grand F, Scime A, Kuang S, von Maltzahn J, Seale V, Cuenda A, Ranish JA, Rudnicki MA. p38-{gamma}-dependent gene silencing restricts entry into the myogenic differentiation program. J Cell Biol 187: 991-1005, 2009. 113. Gnocchi VF, White RB, Ono Y, Ellis JA, Zammit PS. Further characterisation of the molecular signature of quiescent and activated mouse muscle satellite cells. PloS One 4: e5205, 2009. 114. Gomez-Gaviro MV, Lovell-Badge R, Fernandez-Aviles F, Lara-Pezzi E. The vascular stem cell niche. J Cardiovasc Transl Res 5: 618-630, 2012. 115. Gopinath SD, Webb AE, Brunet A, Rando TA. FOXO3 promotes quiescence in adult muscle stem cells during the process of self-renewal. Stem Cell Reports 2: 414-426, 2014. 116. Gordon S. Alternative activation of macrophages. Nat Rev Immunol 3: 23-35, 2003. 117. Gratchev A, Guillot P, Hakiy N, Politz O, Orfanos CE, Schledzewski K, Goerdt S. Alternatively activated macrophages differentially express fibronectin and its splice variants and the extracellular matrix protein betaIG-H3. Scand J Immunol 53: 386-392, 2001. 118. Gros J, Serralbo O, Marcelle C. WNT11 acts as a directional cue to organize the elongation of early muscle fibres. Nature 457: 589-593, 2009. 119. Grounds MD. Age-associated changes in the response of skeletal muscle cells to exercise and regeneration. Ann N Y Acad Sci 854: 78-91, 1998. 120. Grounds MD, Garrett KL, Lai MC, Wright WE, Beilharz MW. Identification of skeletal muscle precursor cells in vivo by use of MyoD1 and myogenin probes. Cell Tissue Res 267: 99-104, 1992. 121. Gundersen K, Bruusgaard JC. Nuclear domains during muscle atrophy: Nuclei lost or paradigm lost? J Physiol 586: 2675-2681, 2008. 122. Gunther S, Kim J, Kostin S, Lepper C, Fan CM, Braun T. Myf5-positive satellite cells contribute to Pax7-dependent long-term maintenance of adult muscle stem cells. Cell Stem Cell 13: 590-601, 2013. 123. Gussoni E, Soneoka Y, Strickland CD, Buzney EA, Khan MK, Flint AF, Kunkel LM, Mulligan RC. Dystrophin expression in the mdx mouse restored by stem cell transplantation. Nature 401: 390-394, 1999. 124. Guttridge DC, Albanese C, Reuther JY, Pestell RG, Baldwin AS, Jr. NFkappaB controls cell growth and differentiation through transcriptional regulation of cyclin D1. Mol Cell Biol 19: 5785-5799, 1999. 125. Halevy O, Novitch BG, Spicer DB, Skapek SX, Rhee J, Hannon GJ, Beach D, Lassar AB. Correlation of terminal cell cycle arrest of skeletal muscle with induction of p21 by MyoD. Science 267: 1018-1021, 1995. 126. Han J, Jiang Y, Li Z, Kravchenko VV, Ulevitch RJ. Activation of the transcription factor MEF2C by the MAP kinase p38 in inflammation. Nature 386: 296-299, 1997.

Volume 5, July 2015

Satellite Cells and Skeletal Muscle Regeneration

127. Hasty P, Bradley A, Morris JH, Edmondson DG, Venuti JM, Olson EN, Klein WH. Muscle deficiency and neonatal death in mice with a targeted mutation in the myogenin gene. Nature 364: 501-506, 1993. 128. Hawke TJ, Garry DJ. Myogenic satellite cells: Physiology to molecular biology. J Appl Physiol (1985) 91: 534-551, 2001. 129. Heredia JE, Mukundan L, Chen FM, Mueller AA, Deo RC, Locksley RM, Rando TA, Chawla A. Type 2 innate signals stimulate fibro/adipogenic progenitors to facilitate muscle regeneration. Cell 153: 376-388, 2013. 130. Hernandez-Hernandez JM, Mallappa C, Nasipak BT, Oesterreich S, Imbalzano AN. The Scaffold attachment factor b1 (Safb1) regulates myogenic differentiation by facilitating the transition of myogenic gene chromatin from a repressed to an activated state. Nucleic Acids Res 41: 5704-5716, 2013. 131. Heron MI, Richmond FJ. In-series fiber architecture in long human muscles. J Morphol 216: 35-45, 1993. 132. Hollenberg SM, Cheng PF, Weintraub H. Use of a conditional MyoD transcription factor in studies of MyoD trans-activation and muscle determination. Proc Natl Acad Sci U S A 90: 8028-8032, 1993. 133. Horsley V, Jansen KM, Mills ST, Pavlath GK. IL-4 acts as a myoblast recruitment factor during mammalian muscle growth. Cell 113: 483494, 2003. 134. Hosoyama T, Nishijo K, Prajapati SI, Li G, Keller C. Rb1 gene inactivation expands satellite cell and postnatal myoblast pools. J Biol Chem 286: 19556-19564, 2011. 135. Huang YC, Dennis RG, Baar K. Cultured slow vs. fast skeletal muscle cells differ in physiology and responsiveness to stimulation. Am J Physiol Cell Physiol 291: C11-C17, 2006. 136. Huard J, Bouchard JP, Roy R, Malouin F, Dansereau G, Labrecque C, Albert N, Richards CL, Lemieux B, Tremblay JP. Human myoblast transplantation: Preliminary results of 4 cases. Muscle Nerve 15: 550560, 1992. 137. Hutcheson DA, Zhao J, Merrell A, Haldar M, Kardon G. Embryonic and fetal limb myogenic cells are derived from developmentally distinct progenitors and have different requirements for beta-catenin. Genes Dev 23: 997-1013, 2009. 138. Iezzi S, Cossu G, Nervi C, Sartorelli V, Puri PL. Stage-specific modulation of skeletal myogenesis by inhibitors of nuclear deacetylases. Proc Natl Acad Sci U S A 99: 7757-7762, 2002. 139. Irintchev A, Zeschnigk M, Starzinski-Powitz A, Wernig A. Expression pattern of M-cadherin in normal, denervated, and regenerating mouse muscles. Dev Dyn 199: 326-337, 1994. 140. Ishikawa H. The fine structure of myo-tendon junction in some mammalian skeletal muscles. Arch Histol Jpn 25: 275-296, 1965. 141. Ito N, Ruegg UT, Kudo A, Miyagoe-Suzuki Y, Takeda S. Activation of calcium signaling through Trpv1 by nNOS and peroxynitrite as a key trigger of skeletal muscle hypertrophy. Nat Med 19: 101-106, 2013. 142. Jockusch H, Voigt S. Migration of adult myogenic precursor cells as revealed by GFP/nLacZ labelling of mouse transplantation chimeras. J Cell Sci 116: 1611-1616, 2003. 143. Joe AW, Yi L, Natarajan A, Le Grand F, So L, Wang J, Rudnicki MA, Rossi FM. Muscle injury activates resident fibro/adipogenic progenitors that facilitate myogenesis. Nat Cell Biol 12: 153-163, 2010. 144. Jones NC, Tyner KJ, Nibarger L, Stanley HM, Cornelison DD, Fedorov YV, Olwin BB. The p38alpha/beta MAPK functions as a molecular switch to activate the quiescent satellite cell. J Cell Biol 169: 105-116, 2005. 145. Juan AH, Derfoul A, Feng X, Ryall JG, Dell’Orso S, Pasut A, Zare H, Simone JM, Rudnicki MA, Sartorelli V. Polycomb EZH2 controls selfrenewal and safeguards the transcriptional identity of skeletal muscle stem cells. Genes Dev 25: 789-794, 2011. 146. Juan AH, Kumar RM, Marx JG, Young RA, Sartorelli V. Mir-214dependent regulation of the polycomb protein Ezh2 in skeletal muscle and embryonic stem cells. Mol Cell 36: 61-74, 2009. 147. Kaar JL, Li Y, Blair HC, Asche G, Koepsel RR, Huard J, Russell AJ. Matrix metalloproteinase-1 treatment of muscle fibrosis. Acta Biomater 4: 1411-1420, 2008. 148. Kalhovde JM, Jerkovic R, Sefland I, Cordonnier C, Calabria E, Schiaffino S, Lomo T. “Fast” and “slow” muscle fibres in hindlimb muscles of adult rats regenerate from intrinsically different satellite cells. J Physiol 562: 847-857, 2005. 149. Kanisicak O, Mendez JJ, Yamamoto S, Yamamoto M, Goldhamer DJ. Progenitors of skeletal muscle satellite cells express the muscle determination gene, MyoD. Dev Biol 332: 131-141, 2009. 150. Karalaki M, Fili S, Philippou A, Koutsilieris M. Muscle regeneration: Cellular and molecular events. In Vivo 23: 779-796, 2009. 151. Kassar-Duchossoy L, Gayraud-Morel B, Gomes D, Rocancourt D, Buckingham M, Shinin V, Tajbakhsh S. Mrf4 determines skeletal muscle identity in Myf5:Myod double-mutant mice. Nature 431: 466-471, 2004. 152. Kassar-Duchossoy L, Giacone E, Gayraud-Morel B, Jory A, Gomes D, Tajbakhsh S. Pax3/Pax7 mark a novel population of primitive myogenic cells during development. Genes Dev 19: 1426-1431, 2005.

1055

Satellite Cells and Skeletal Muscle Regeneration

153. Kastner S, Elias MC, Rivera AJ, Yablonka-Reuveni Z. Gene expression patterns of the fibroblast growth factors and their receptors during myogenesis of rat satellite cells. J Histochem Cytochem 48: 1079-1096, 2000. 154. Kaushal S, Schneider JW, Nadal-Ginard B, Mahdavi V. Activation of the myogenic lineage by MEF2A, a factor that induces and cooperates with MyoD. Science 266: 1236-1240, 1994. 155. Kawabe Y, Wang YX, McKinnell IW, Bedford MT, Rudnicki MA. Carm1 regulates Pax7 transcriptional activity through MLL1/2 recruitment during asymmetric satellite stem cell divisions. Cell Stem Cell 11: 333-345, 2012. 156. Kherif S, Lafuma C, Dehaupas M, Lachkar S, Fournier JG, VerdiereSahuque M, Fardeau M, Alameddine HS. Expression of matrix metalloproteinases 2 and 9 in regenerating skeletal muscle: A study in experimentally injured and mdx muscles. Dev Biol 205: 158-170, 1999. 157. Kim S, Shilagardi K, Zhang S, Hong SN, Sens KL, Bo J, Gonzalez GA, Chen EH. A critical function for the actin cytoskeleton in targeted exocytosis of prefusion vesicles during myoblast fusion. Dev Cell 12: 571-586, 2007. 158. Komiya Y, Habas R. Wnt signal transduction pathways. Organogenesis 4: 68-75, 2008. 159. Koning M, Werker PM, van der Schaft DW, Bank RA, Harmsen MC. MicroRNA-1 and microRNA-206 improve differentiation potential of human satellite cells: A novel approach for tissue engineering of skeletal muscle. Tissue Eng Part A 18: 889-898, 2012. 160. Koning M, Werker PM, van Luyn MJ, Krenning G, Harmsen MC. A global downregulation of microRNAs occurs in human quiescent satellite cells during myogenesis. Differentiation 84: 314-321, 2012. 161. Kottlors M, Kirschner J. Elevated satellite cell number in Duchenne muscular dystrophy. Cell Tissue Res 340: 541-548, 2010. 162. Kuang S, Charge SB, Seale P, Huh M, Rudnicki MA. Distinct roles for Pax7 and Pax3 in adult regenerative myogenesis. J Cell Biol 172: 103-113, 2006. 163. Kuang S, Kuroda K, Le Grand F, Rudnicki MA. Asymmetric selfrenewal and commitment of satellite stem cells in muscle. Cell 129: 999-1010, 2007. 164. Lafreniere JF, Caron MC, Skuk D, Goulet M, Cheikh AR, Tremblay JP. Growth factor coinjection improves the migration potential of monkey myogenic precursors without affecting cell transplantation success. Cell Transplant 18: 719-730, 2009. 165. Lafreniere JF, Mills P, Bouchentouf M, Tremblay JP. Interleukin-4 improves the migration of human myogenic precursor cells in vitro and in vivo. Exp Cell Res 312: 1127-1141, 2006. 166. Lansdorp PM. Immortal strands? Give me a break. Cell 129: 1244-1247, 2007. 167. Laurie GW, Leblond CP, Martin GR. Localization of type IV collagen, laminin, heparan sulfate proteoglycan, and fibronectin to the basal lamina of basement membranes. J Cell Biol 95: 340-344, 1982. 168. Le Grand F, Grifone R, Mourikis P, Houbron C, Gigaud C, Pujol J, Maillet M, Pages G, Rudnicki M, Tajbakhsh S, Maire P. Six1 regulates stem cell repair potential and self-renewal during skeletal muscle regeneration. J Cell Biol 198: 815-832, 2012. 169. Le Grand F, Jones AE, Seale V, Scime A, Rudnicki MA. Wnt7a activates the planar cell polarity pathway to drive the symmetric expansion of satellite stem cells. Cell Stem Cell 4: 535-547, 2009. 170. Lee H, Habas R, Abate-Shen C. MSX1 cooperates with histone H1b for inhibition of transcription and myogenesis. Science 304: 1675-1678, 2004. 171. Lee TI, Jenner RG, Boyer LA, Guenther MG, Levine SS, Kumar RM, Chevalier B, Johnstone SE, Cole MF, Isono K, Koseki H, Fuchikami T, Abe K, Murray HL, Zucker JP, Yuan B, Bell GW, Herbolsheimer E, Hannett NM, Sun K, Odom DT, Otte AP, Volkert TL, Bartel DP, Melton DA, Gifford DK, Jaenisch R, Young RA. Control of developmental regulators by Polycomb in human embryonic stem cells. Cell 125: 301-313, 2006. 172. Lepper C, Partridge TA, Fan CM. An absolute requirement for Pax7positive satellite cells in acute injury-induced skeletal muscle regeneration. Development 138: 3639-3646, 2011. 173. Lesault PF, Theret M, Magnan M, Cuvellier S, Niu Y, Gherardi RK, Tremblay JP, Hittinger L, Chazaud B. Macrophages improve survival, proliferation and migration of engrafted myogenic precursor cells into MDX skeletal muscle. PloS One 7: e46698, 2012. 174. Lesley J, Hyman R, Kincade PW. CD44 and its interaction with extracellular matrix. Adv Immunol 54: 271-335, 1993. 175. Lewis WH, Lewis MR. Behavior of cross striated muscle in tissue cultures. Am J Anat 22: 169-194, 1917. 176. Li YP. TNF-alpha is a mitogen in skeletal muscle. Am J Physiol Cell Physiol 285: C370-C376, 2003. 177. Liadaki K, Casar JC, Wessen M, Luth ES, Jun S, Gussoni E, Kunkel LM. beta4 integrin marks interstitial myogenic progenitor cells in adult murine skeletal muscle. J Histochem Cytochem 60: 31-44, 2012.

1056

Comprehensive Physiology

178. Liu JX, Hoglund AS, Karlsson P, Lindblad J, Qaisar R, Aare S, Bengtsson E, Larsson L. Myonuclear domain size and myosin isoform expression in muscle fibres from mammals representing a 100,000-fold difference in body size. Exp Physiol 94: 117-129, 2009. 179. Liu L, Cheung TH, Charville GW, Hurgo BM, Leavitt T, Shih J, Brunet A, Rando TA. Chromatin modifications as determinants of muscle stem cell quiescence and chronological aging. Cell Reports 4: 189-204, 2013. 180. Liu N, Nelson BR, Bezprozvannaya S, Shelton JM, Richardson JA, Bassel-Duby R, Olson EN. Requirement of MEF2A, C, and D for skeletal muscle regeneration. Proc Natl Acad Sci U S A 111: 41094114, 2014. 181. Liu Y, Schneider MF. FGF2 activates TRPC and Ca(2+) signaling leading to satellite cell activation. Front Physiol 5: 38, 2014. 182. Lluis F, Ballestar E, Suelves M, Esteller M, Munoz-Canoves P. E47 phosphorylation by p38 MAPK promotes MyoD/E47 association and muscle-specific gene transcription. EMBO J 24: 974-984, 2005. 183. Lu H, Huang D, Saederup N, Charo IF, Ransohoff RM, Zhou L. Macrophages recruited via CCR2 produce insulin-like growth factor-1 to repair acute skeletal muscle injury. FASEB J 25: 358-369, 2011. 184. Lu J, McKinsey TA, Nicol RL, Olson EN. Signal-dependent activation of the MEF2 transcription factor by dissociation from histone deacetylases. Proc Natl Acad Sci U S A 97: 4070-4075, 2000. 185. Luo G, Hershko DD, Robb BW, Wray CJ, Hasselgren PO. IL-1beta stimulates IL-6 production in cultured skeletal muscle cells through activation of MAP kinase signaling pathway and NF-kappa B. Am J Physiol Regul Integr Comp Physiol 284: R1249-R1254, 2003. 186. Ma K, Chan JK, Zhu G, Wu Z. Myocyte enhancer factor 2 acetylation by p300 enhances its DNA binding activity, transcriptional activity, and myogenic differentiation. Mol Cell Biol 25: 3575-3582, 2005. 187. Machida S, Spangenburg EE, Booth FW. Forkhead transcription factor FoxO1 transduces insulin-like growth factor’s signal to p27Kip1 in primary skeletal muscle satellite cells. J Cell Physiol 196: 523-531, 2003. 188. MacIntosh BR, Gardiner PF, McComas AJ. Skeletal Mmuscle: Form and Function, second edition. Champaign, Ill.; Leeds: Human Kinetics, 2006. 189. Maher P. p38 mitogen-activated protein kinase activation is required for fibroblast growth factor-2-stimulated cell proliferation but not differentiation. J Biol Chem 274: 17491-17498, 1999. 190. Mal AK. Histone methyltransferase Suv39h1 represses MyoDstimulated myogenic differentiation. EMBO J 25: 3323-3334, 2006. 191. Manzur AY, Kuntzer T, Pike M, Swan A. Glucocorticoid corticosteroids for Duchenne muscular dystrophy. Cochrane Database Syst Rev 1: CD003725, 2008. 192. Mauro A. Satellite cell of skeletal muscle fibers. J Biophy Biochem Cytol 9: 493-495, 1961. 193. McCall GE, Allen DL, Linderman JK, Grindeland RE, Roy RR, Mukku VR, Edgerton VR. Maintenance of myonuclear domain size in rat soleus after overload and growth hormone/IGF-I treatment. J Appl Physiol (1985) 84: 1407-1412, 1998. 194. McCarthy JJ, Mula J, Miyazaki M, Erfani R, Garrison K, Farooqui AB, Srikuea R, Lawson BA, Grimes B, Keller C, Van Zant G, Campbell KS, Esser KA, Dupont-Versteegden EE, Peterson CA. Effective fiber hypertrophy in satellite cell-depleted skeletal muscle. Development 138: 3657-3666, 2011. 195. McKinnell IW, Ishibashi J, Le Grand F, Punch VG, Addicks GC, Greenblatt JF, Dilworth FJ, Rudnicki MA. Pax7 activates myogenic genes by recruitment of a histone methyltransferase complex. Nat Cell Biol 10: 77-84, 2008. 196. McKinsey TA, Zhang CL, Lu J, Olson EN. Signal-dependent nuclear export of a histone deacetylase regulates muscle differentiation. Nature 408: 106-111, 2000. 197. McKinsey TA, Zhang CL, Olson EN. Control of muscle development by dueling HATs and HDACs. Curr Opin Genet Dev 11: 497-504, 2001. 198. Megeney LA, Kablar B, Garrett K, Anderson JE, Rudnicki MA. MyoD is required for myogenic stem cell function in adult skeletal muscle. Genes Dev 10: 1173-1183, 1996. 199. Meng J, Chun S, Asfahani R, Lochmuller H, Muntoni F, Morgan J. Human skeletal muscle-derived CD133(+) cells form functional satellite cells after intramuscular transplantation in immunodeficient host mice. Mol Ther 22: 1008-1017, 2014. 200. Millay DP, O’Rourke JR, Sutherland LB, Bezprozvannaya S, Shelton JM, Bassel-Duby R, Olson EN. Myomaker is a membrane activator of myoblast fusion and muscle formation. Nature 499: 301-305, 2013. 201. Millay DP, Sutherland LB, Bassel-Duby R, Olson EN. Myomaker is essential for muscle regeneration. Genes Dev 28: 1641-1646, 2014. 202. Mintz B, Baker WW. Normal mammalian muscle differentiation and gene control of isocitrate dehydrogenase synthesis. Proc Natl Acad Sci U S A 58: 592, 1967. 203. Mitchell KJ, Pannerec A, Cadot B, Parlakian A, Besson V, Gomes ER, Marazzi G, Sassoon DA. Identification and characterization of a nonsatellite cell muscle resident progenitor during postnatal development. Nat Cell Biol 12: 257-266, 2010.

Volume 5, July 2015

Comprehensive Physiology

204. Molkentin JD, Black BL, Martin JF, Olson EN. Cooperative activation of muscle gene expression by MEF2 and myogenic bHLH proteins. Cell 83: 1125-1136, 1995. 205. Montarras D, L’Honore A, Buckingham M. Lying low but ready for action: The quiescent muscle satellite cell. FEBS J 280: 4036-4050, 2013. 206. Montarras D, Morgan J, Collins C, Relaix F, Zaffran S, Cumano A, Partridge T, Buckingham M. Direct isolation of satellite cells for skeletal muscle regeneration. Science 309: 2064-2067, 2005. 207. Morgan J, Rouche A, Bausero P, Houssaini A, Gross J, Fiszman MY, Alameddine HS. MMP-9 overexpression improves myogenic cell migration and engraftment. Muscle Nerve 42: 584-595, 2010. 208. Mourikis P, Sambasivan R, Castel D, Rocheteau P, Bizzarro V, Tajbakhsh S. A critical requirement for notch signaling in maintenance of the quiescent skeletal muscle stem cell state. Stem Cells 30: 243-252, 2012. 209. Murphy MM, Lawson JA, Mathew SJ, Hutcheson DA, Kardon G. Satellite cells, connective tissue fibroblasts and their interactions are crucial for muscle regeneration. Development 138: 3625-3637, 2011. 210. Musaro A, McCullagh K, Paul A, Houghton L, Dobrowolny G, Molinaro M, Barton ER, Sweeney HL, Rosenthal N. Localized Igf-1 transgene expression sustains hypertrophy and regeneration in senescent skeletal muscle. Nat Genet 27: 195-200, 2001. 211. Muskiewicz KR, Frank NY, Flint AF, Gussoni E. Myogenic potential of muscle side and main population cells after intravenous injection into sub-lethally irradiated mdx mice. J Histochem Cytochem 53: 861-873, 2005. 212. Mylona E, Jones KA, Mills ST, Pavlath GK. CD44 regulates myoblast migration and differentiation. J Cell Physiol 209: 314-321, 2006. 213. Nagata Y, Kobayashi H, Umeda M, Ohta N, Kawashima S, Zammit PS, Matsuda R. Sphingomyelin levels in the plasma membrane correlate with the activation state of muscle satellite cells. J Histochem Cytochem 54: 375-384, 2006. 214. Nagata Y, Partridge TA, Matsuda R, Zammit PS. Entry of muscle satellite cells into the cell cycle requires sphingolipid signaling. J Cell Biol 174: 245-253, 2006. 215. Neuhaus P, Oustanina S, Loch T, Kruger M, Bober E, Dono R, Zeller R, Braun T. Reduced mobility of fibroblast growth factor (FGF)-deficient myoblasts might contribute to dystrophic changes in the musculature of FGF2/FGF6/mdx triple-mutant mice. Mol Cell Biol 23: 6037-6048, 2003. 216. Ng RK, Gurdon JB. Epigenetic memory of an active gene state depends on histone H3.3 incorporation into chromatin in the absence of transcription. Nat Cell Biol 10: 102-109, 2008. 217. Nishimura T, Nakamura K, Kishioka Y, Kato-Mori Y, Wakamatsu J, Hattori A. Inhibition of matrix metalloproteinases suppresses the migration of skeletal muscle cells. J Muscle Res Cell Motil 29: 37-44, 2008. 218. Olguin HC, Yang Z, Tapscott SJ, Olwin BB. Reciprocal inhibition between Pax7 and muscle regulatory factors modulates myogenic cell fate determination. J Cell Biol 177: 769-779, 2007. 219. Ono Y, Masuda S, Nam HS, Benezra R, Miyagoe-Suzuki Y, Takeda S. Slow-dividing satellite cells retain long-term self-renewal ability in adult muscle. J Cell Sci 125: 1309-1317, 2012. 220. Otis JS, Niccoli S, Hawdon N, Sarvas JL, Frye MA, Chicco AJ, Lees SJ. Pro-inflammatory mediation of myoblast proliferation. PloS One 9: e92363, 2014. 221. Oustanina S, Hause G, Braun T. Pax7 directs postnatal renewal and propagation of myogenic satellite cells but not their specification. EMBO J 23: 3430-3439, 2004. 222. Palacios D, Mozzetta C, Consalvi S, Caretti G, Saccone V, Proserpio V, Marquez VE, Valente S, Mai A, Forcales SV, Sartorelli V, Puri PL. TNF/p38alpha/polycomb signaling to Pax7 locus in satellite cells links inflammation to the epigenetic control of muscle regeneration. Cell Stem Cell 7: 455-469, 2010. 223. Pallafacchina G, Francois S, Regnault B, Czarny B, Dive V, Cumano A, Montarras D, Buckingham M. 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. 224. Pannerec A, Marazzi G, Sassoon D. Stem cells in the hood: The skeletal muscle niche. Trends Mol Med 18: 599-606, 2012. 225. Park IH, Chen J. Mammalian target of rapamycin (mTOR) signaling is required for a late-stage fusion process during skeletal myotube maturation. J Biol Chem 280: 32009-32017, 2005. 226. Park IH, Erbay E, Nuzzi P, Chen J. Skeletal myocyte hypertrophy requires mTOR kinase activity and S6K1. Exp Cell Res 309: 211-219, 2005. 227. Parsons SA, Millay DP, Wilkins BJ, Bueno OF, Tsika GL, Neilson JR, Liberatore CM, Yutzey KE, Crabtree GR, Tsika RW, Molkentin JD. Genetic loss of calcineurin blocks mechanical overload-induced skeletal muscle fiber type switching but not hypertrophy. J Biol Chem 279: 26192-26200, 2004.

Volume 5, July 2015

Satellite Cells and Skeletal Muscle Regeneration

228. Partridge TA, Morgan JE, Coulton GR, Hoffman EP, Kunkel LM. Conversion of mdx myofibres from dystrophin-negative to -positive by injection of normal myoblasts. Nature 337: 176-179, 1989. 229. Pasut A, Oleynik P, Rudnicki MA. Isolation of muscle stem cells by fluorescence activated cell sorting cytometry. Methods Mol Biol 798: 53-64, 2012. 230. Perdiguero E, Ruiz-Bonilla V, Serrano AL, Munoz-Canoves P. Genetic deficiency of p38alpha reveals its critical role in myoblast cell cycle exit: The p38alpha-JNK connection. Cell Cycle 6: 1298-1303, 2007. 231. Perrone CE, Fenwick-Smith D, Vandenburgh HH. Collagen and stretch modulate autocrine secretion of insulin-like growth factor-1 and insulinlike growth factor binding proteins from differentiated skeletal muscle cells. J Biol Chem 270: 2099-2106, 1995. 232. Pietsch P. Differentiation in regeneration. I. The development of muscle and cartilage following deplantation of of regenerating limb blastemata of Amblystoma larvae. Dev Biol 3: 255-264, 1961. 233. Pisconti A, Cornelison DD, Olguin HC, Antwine TL, Olwin BB. Syndecan-3 and Notch cooperate in regulating adult myogenesis. J Cell Biol 190: 427-441, 2010. 234. Pizza FX, McLoughlin TJ, McGregor SJ, Calomeni EP, Gunning WT. Neutrophils injure cultured skeletal myotubes. Am J Physiol Cell Physiol 281: C335-C341, 2001. 235. Pizza FX, Peterson JM, Baas JH, Koh TJ. Neutrophils contribute to muscle injury and impair its resolution after lengthening contractions in mice. J Physiol 562: 899-913, 2005. 236. Prelovsek O, Mars T, Jevsek M, Podbregar M, Grubic Z. High dexamethasone concentration prevents stimulatory effects of TNF-alpha and LPS on IL-6 secretion from the precursors of human muscle regeneration. Am J Physiol Regul Integr Comp Physiol 291: R1651-R1656, 2006. 237. Price FD, von Maltzahn J, Bentzinger CF, Dumont NA, Yin H, Chang NC, Wilson DH, Frenette J, Rudnicki MA. Inhibition of JAK-STAT signaling stimulates adult satellite cell function. Nat Med 20: 11741181, 2014. 238. Proctor DN, O’Brien PC, Atkinson EJ, Nair KS. Comparison of techniques to estimate total body skeletal muscle mass in people of different age groups. Am J Physiol 277: E489-E495, 1999. 239. Punch VG, Jones AE, Rudnicki MA. Transcriptional networks that regulate muscle stem cell function. Wiley Interdiscip Rev Syst Biol Med 1: 128-140, 2009. 240. Puri PL, Iezzi S, Stiegler P, Chen TT, Schiltz RL, Muscat GE, Giordano A, Kedes L, Wang JY, Sartorelli V. Class I histone deacetylases sequentially interact with MyoD and pRb during skeletal myogenesis. Mol Cell 8: 885-897, 2001. 241. Rampalli S, Li L, Mak E, Ge K, Brand M, Tapscott SJ, Dilworth FJ. p38 MAPK signaling regulates recruitment of Ash2L-containing methyltransferase complexes to specific genes during differentiation. Nat Struct Mol Biol 14: 1150-1156, 2007. 242. Rawls A, Valdez MR, Zhang W, Richardson J, Klein WH, Olson EN. Overlapping functions of the myogenic bHLH genes MRF4 and MyoD revealed in double mutant mice. Development 125: 2349-2358, 1998. 243. Reimann J, Irintchev A, Wernig A. Regenerative capacity and the number of satellite cells in soleus muscles of normal and mdx mice. Neuromuscul Disord 10: 276-282, 2000. 244. Relaix F, Montarras D, Zaffran S, Gayraud-Morel B, Rocancourt D, Tajbakhsh S, Mansouri A, Cumano A, Buckingham M. Pax3 and Pax7 have distinct and overlapping functions in adult muscle progenitor cells. J Cell Biol 172: 91-102, 2006. 245. Relaix F, Rocancourt D, Mansouri A, Buckingham M. Divergent functions of murine Pax3 and Pax7 in limb muscle development. Genes Dev 18: 1088-1105, 2004. 246. Relaix F, Rocancourt D, Mansouri A, Buckingham M. A Pax3/Pax7dependent population of skeletal muscle progenitor cells. Nature 435: 948-953, 2005. 247. Rocheteau P, Gayraud-Morel B, Siegl-Cachedenier I, Blasco MA, Tajbakhsh S. A subpopulation of adult skeletal muscle stem cells retains all template DNA strands after cell division. Cell 148: 112-125, 2012. 248. Rochlin K, Yu S, Roy S, Baylies MK. Myoblast fusion: When it takes more to make one. Dev Biol 341: 66-83, 2010. 249. Rommel C, Bodine SC, Clarke BA, Rossman R, Nunez L, Stitt TN, Yancopoulos GD, Glass DJ. Mediation of IGF-1-induced skeletal myotube hypertrophy by PI(3)K/Akt/mTOR and PI(3)K/Akt/GSK3 pathways. Nat Cell Biol 3: 1009-1013, 2001. 250. Rosenblatt JD, Parry DJ, Partridge TA. Phenotype of adult mouse muscle myoblasts reflects their fiber type of origin. Differentiation 60: 39-45, 1996. 251. Rudnicki MA, Braun T, Hinuma S, Jaenisch R. Inactivation of MyoD in mice leads to up-regulation of the myogenic HLH gene Myf-5 and results in apparently normal muscle development. Cell 71: 383-390, 1992. 252. Rudnicki MA, Schnegelsberg PN, Stead RH, Braun T, Arnold HH, Jaenisch R. MyoD or Myf-5 is required for the formation of skeletal muscle. Cell 75: 1351-1359, 1993.

1057

Satellite Cells and Skeletal Muscle Regeneration

253. Ruffell D, Mourkioti F, Gambardella A, Kirstetter P, Lopez RG, Rosenthal N, Nerlov C. A CREB-C/EBPbeta cascade induces M2 macrophage-specific gene expression and promotes muscle injury repair. Proc Natl Acad Sci U S A 106: 17475-17480, 2009. 254. Sabourin LA, Girgis-Gabardo A, Seale P, Asakura A, Rudnicki MA. Reduced differentiation potential of primary MyoD-/- myogenic cells derived from adult skeletal muscle. J Cell Biol 144: 631-643, 1999. 255. Sacco A, Doyonnas R, Kraft P, Vitorovic S, Blau HM. Self-renewal and expansion of single transplanted muscle stem cells. Nature 456: 502-506, 2008. 256. Sacco A, Mourkioti F, Tran R, Choi J, Llewellyn M, Kraft P, Shkreli M, Delp S, Pomerantz JH, Artandi SE, Blau HM. Short telomeres and stem cell exhaustion model Duchenne muscular dystrophy in mdx/mTR mice. Cell 143: 1059-1071, 2010. 257. Saclier M, Yacoub-Youssef H, Mackey AL, Arnold L, Ardjoune H, Magnan M, Sailhan F, Chelly J, Pavlath GK, Mounier R, Kjaer M, Chazaud B. Differentially activated macrophages orchestrate myogenic precursor cell fate during human skeletal muscle regeneration. Stem Cells 31: 384-396, 2013. 258. Sadeh M. Effects of aging on skeletal muscle regeneration. J Neurol Sci 87: 67-74, 1988. 259. Sambasivan R, Yao R, Kissenpfennig A, Van Wittenberghe L, Paldi A, Gayraud-Morel B, Guenou H, Malissen B, Tajbakhsh S, Galy A. Pax7expressing satellite cells are indispensable for adult skeletal muscle regeneration. Development 138: 3647-3656, 2011. 260. Sampaolesi M, Torrente Y, Innocenzi A, Tonlorenzi R, D’Antona G, Pellegrino MA, Barresi R, Bresolin N, De Angelis MG, Campbell KP, Bottinelli R, Cossu G. Cell therapy of alpha-sarcoglycan null dystrophic mice through intra-arterial delivery of mesoangioblasts. Science 301: 487-492, 2003. 261. Sandri M, Sandri C, Gilbert A, Skurk C, Calabria E, Picard A, Walsh K, Schiaffino S, Lecker SH, Goldberg AL. Foxo transcription factors induce the atrophy-related ubiquitin ligase atrogin-1 and cause skeletal muscle atrophy. Cell 117: 399-412, 2004. 262. Sanes JR. The basement membrane/basal lamina of skeletal muscle. J Biol Chem 278: 12601-12604, 2003. 263. Sartorelli V, Puri PL, Hamamori Y, Ogryzko V, Chung G, Nakatani Y, Wang JY, Kedes L. Acetylation of MyoD directed by PCAF is necessary for the execution of the muscle program. Mol Cell 4: 725-734, 1999. 264. Scharner J, Zammit PS. The muscle satellite cell at 50: The formative years. Skelet Muscle 1: 28, 2011. 265. Schnoor M, Cullen P, Lorkowski J, Stolle K, Robenek H, Troyer D, Rauterberg J, Lorkowski S. Production of type VI collagen by human macrophages: A new dimension in macrophage functional heterogeneity. J Immunol 180: 5707-5719, 2008. 266. Schultz E. Satellite cell proliferative compartments in growing skeletal muscles. Dev Biol 175: 84-94, 1996. 267. Schultz E, Jaryszak DL, Valliere CR. Response of satellite cells to focal skeletal muscle injury. Muscle Nerve 8: 217-222, 1985. 268. Schwander M, Leu M, Stumm M, Dorchies OM, Ruegg UT, Schittny J, Muller U. Beta1 integrins regulate myoblast fusion and sarcomere assembly. Dev Cell 4: 673-685, 2003. 269. Seale P, Sabourin LA, Girgis-Gabardo A, Mansouri A, Gruss P, Rudnicki MA. Pax7 is required for the specification of myogenic satellite cells. Cell 102: 777-786, 2000. 270. Seenundun S, Rampalli S, Liu QC, Aziz A, Palii C, Hong S, Blais A, Brand M, Ge K, Dilworth FJ. UTX mediates demethylation of H3K27me3 at muscle-specific genes during myogenesis. EMBO J 29: 1401-1411, 2010. 271. Serra C, Palacios D, Mozzetta C, Forcales SV, Morantte I, Ripani M, Jones DR, Du K, Jhala US, Simone C, Puri PL. Functional interdependence at the chromatin level between the MKK6/p38 and IGF1/PI3K/AKT pathways during muscle differentiation. Mol Cell 28: 200-213, 2007. 272. Serrano AL, Baeza-Raja B, Perdiguero E, Jardi M, Munoz-Canoves P. Interleukin-6 is an essential regulator of satellite cell-mediated skeletal muscle hypertrophy. Cell Metab 7: 33-44, 2008. 273. Serrano AL, Mann CJ, Vidal B, Ardite E, Perdiguero E, MunozCanoves P. Cellular and molecular mechanisms regulating fibrosis in skeletal muscle repair and disease. Curr Top Dev Biol 96: 167-201, 2011. 274. Shea KL, Xiang W, LaPorta VS, Licht JD, Keller C, Basson MA, Brack AS. Sprouty1 regulates reversible quiescence of a self-renewing adult muscle stem cell pool during regeneration. Cell Stem Cell 6: 117-129, 2010. 275. Shefer G, Van de Mark DP, Richardson JB, Yablonka-Reuveni Z. Satellite-cell pool size does matter: Defining the myogenic potency of aging skeletal muscle. Dev Biol 294: 50-66, 2006. 276. Shi X, Garry DJ. Muscle stem cells in development, regeneration, and disease. Genes Dev 20: 1692-1708, 2006. 277. Shinin V, Gayraud-Morel B, Gomes D, Tajbakhsh S. Asymmetric division and cosegregation of template DNA strands in adult muscle satellite cells. Nat Cell Biol 8: 677-687, 2006.

1058

Comprehensive Physiology

278. Siegel AL, Atchison K, Fisher KE, Davis GE, Cornelison DD. 3D timelapse analysis of muscle satellite cell motility. Stem Cells 27: 25272538, 2009. 279. Simone C, Forcales SV, Hill DA, Imbalzano AN, Latella L, Puri PL. p38 pathway targets SWI-SNF chromatin-remodeling complex to musclespecific loci. Nat Genet 36: 738-743, 2004. 280. Skuk D, Goulet M, Roy B, Tremblay JP. Efficacy of myoblast transplantation in nonhuman primates following simple intramuscular cell injections: Toward defining strategies applicable to humans. Exp Neurol 175: 112-126, 2002. 281. Soehnlein O, Lindbom L. Phagocyte partnership during the onset and resolution of inflammation. Nat Rev Immunol 10: 427-439, 2010. 282. Sohn RL, Huang P, Kawahara G, Mitchell M, Guyon J, Kalluri R, Kunkel LM, Gussoni E. A role for nephrin, a renal protein, in vertebrate skeletal muscle cell fusion. Proc Natl Acad Sci U S A 106: 9274-9279, 2009. 283. Soleimani VD, Punch VG, Kawabe Y, Jones AE, Palidwor GA, Porter CJ, Cross JW, Carvajal JJ, Kockx CE, van IWF, Perkins TJ, Rigby PW, Grosveld F, Rudnicki MA. Transcriptional dominance of Pax7 in adult myogenesis is due to high-affinity recognition of homeodomain motifs. Dev Cell 22: 1208-1220, 2012. 284. Sonnet C, Lafuste P, Arnold L, Brigitte M, Poron F, Authier FJ, Chretien F, Gherardi RK, Chazaud B. Human macrophages rescue myoblasts and myotubes from apoptosis through a set of adhesion molecular systems. J Cell Sci 119: 2497-2507, 2006. 285. Sousa-Victor P, Gutarra S, Garcia-Prat L, Rodriguez-Ubreva J, Ortet L, Ruiz-Bonilla V, Jardi M, Ballestar E, Gonzalez S, Serrano AL, Perdiguero E, Munoz-Canoves P. Geriatric muscle stem cells switch reversible quiescence into senescence. Nature 506: 316-321, 2014. 286. Stark DA, Karvas RM, Siegel AL, Cornelison DD. Eph/ephrin interactions modulate muscle satellite cell motility and patterning. Development 138: 5279-5289, 2011. 287. Stradal TE, Rottner K, Disanza A, Confalonieri S, Innocenti M, Scita G. Regulation of actin dynamics by WASP and WAVE family proteins. Trends Cell Biol 14: 303-311, 2004. 288. Strunkelnberg M, Bonengel B, Moda LM, Hertenstein A, de Couet HG, Ramos RG, Fischbach KF. rst and its paralogue kirre act redundantly during embryonic muscle development in Drosophila. Development 128: 4229-4239, 2001. 289. Sun L, Ma K, Wang H, Xiao F, Gao Y, Zhang W, Wang K, Gao X, Ip N, Wu Z. JAK1-STAT1-STAT3, a key pathway promoting proliferation and preventing premature differentiation of myoblasts. J Cell Biol 179: 129-138, 2007. 290. Tajbakhsh S. Skeletal muscle stem cells in developmental versus regenerative myogenesis. J Intern Med 266: 372-389, 2009. 291. Tanaka KK, Hall JK, Troy AA, Cornelison DD, Majka SM, Olwin BB. Syndecan-4-expressing muscle progenitor cells in the SP engraft as satellite cells during muscle regeneration. Cell Stem Cell 4: 217-225, 2009. 292. Tang F, Barbacioru C, Nordman E, Li B, Xu N, Bashkirov VI, Lao K, Surani MA. RNA-Seq analysis to capture the transcriptome landscape of a single cell. Nat Protoc 5: 516-535, 2010. 293. Taniguti AP, Pertille A, Matsumura CY, Santo Neto H, Marques MJ. Prevention of muscle fibrosis and myonecrosis in mdx mice by suramin, a TGF-beta1 blocker. Muscle Nerve 43: 82-87, 2011. 294. Tatsumi R. Mechano-biology of skeletal muscle hypertrophy and regeneration: Possible mechanism of stretch-induced activation of resident myogenic stem cells. Anim Sci J 81: 11-20, 2010. 295. Tatsumi R, Anderson JE, Nevoret CJ, Halevy O, Allen RE. HGF/SF is present in normal adult skeletal muscle and is capable of activating satellite cells. Dev Biol 194: 114-128, 1998. 296. Taylor SM, Jones PA. Multiple new phenotypes induced in 10T1/2 and 3T3 cells treated with 5-azacytidine. Cell 17: 771-779, 1979. 297. Tedesco FS, Dellavalle A, Diaz-Manera J, Messina G, Cossu G. Repairing skeletal muscle: Regenerative potential of skeletal muscle stem cells. J Clin Invest 120: 11-19, 2010. 298. Tidball JG, Villalta SA. Regulatory interactions between muscle and the immune system during muscle regeneration. Am J Physiol Regul Integr Comp Physiol 298: R1173-R1187, 2010. 299. Tierney MT, Aydogdu T, Sala D, Malecova B, Gatto S, Puri PL, Latella L, Sacco A. STAT3 signaling controls satellite cell expansion and skeletal muscle repair. Nat Med 20: 1182-1186, 2014. 300. Torrente Y, Belicchi M, Sampaolesi M, Pisati F, Meregalli M, D’Antona G, Tonlorenzi R, Porretti L, Gavina M, Mamchaoui K, Pellegrino MA, Furling D, Mouly V, Butler-Browne GS, Bottinelli R, Cossu G, Bresolin N. Human circulating AC133(+) stem cells restore dystrophin expression and ameliorate function in dystrophic skeletal muscle. J Clin Invest 114: 182-195, 2004. 301. Tremblay P, Dietrich S, Mericskay M, Schubert FR, Li Z, Paulin D. A crucial role for Pax3 in the development of the hypaxial musculature and the long-range migration of muscle precursors. Dev Biol 203: 49-61, 1998.

Volume 5, July 2015

Comprehensive Physiology

302. Troy A, Cadwallader AB, Fedorov Y, Tyner K, Tanaka KK, Olwin BB. Coordination of satellite cell activation and self-renewal by Parcomplex-dependent asymmetric activation of p38alpha/beta MAPK. Cell Stem Cell 11: 541-553, 2012. 303. 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. 304. Urciuolo A, Quarta M, Morbidoni V, Gattazzo F, Molon S, Grumati P, Montemurro F, Tedesco FS, Blaauw B, Cossu G, Vozzi G, Rando TA, Bonaldo P. Collagen VI regulates satellite cell self-renewal and muscle regeneration. Nat Commun 4: 1964, 2013. 305. van der Vlag J, Otte AP. Transcriptional repression mediated by the human polycomb-group protein EED involves histone deacetylation. Nat Genet 23: 474-478, 1999. 306. Verrier L, Escaffit F, Chailleux C, Trouche D, Vandromme M. A new isoform of the histone demethylase JMJD2A/KDM4A is required for skeletal muscle differentiation. PLoS Genet 7: e1001390, 2011. 307. Villalta SA, Nguyen HX, Deng B, Gotoh T, Tidball JG. Shifts in macrophage phenotypes and macrophage competition for arginine metabolism affect the severity of muscle pathology in muscular dystrophy. Hum Mol Genet 18: 482-496, 2009. 308. von Maltzahn J, Bentzinger CF, Rudnicki MA. Wnt7a-Fzd7 signalling directly activates the Akt/mTOR anabolic growth pathway in skeletal muscle. Nat Cell Biol 14: 186-191, 2012. 309. von Maltzahn J, Jones AE, Parks RJ, Rudnicki MA. Pax7 is critical for the normal function of satellite cells in adult skeletal muscle. Proc Natl Acad Sci U S A 110: 16474-16479, 2013. 310. von Maltzahn J, Renaud JM, Parise G, Rudnicki MA. Wnt7a treatment ameliorates muscular dystrophy. Proc Natl Acad Sci U S A 109: 2061420619, 2012. 311. von Maltzahn J, Zinoviev R, Chang NC, Bentzinger CF, Rudnicki MA. A truncated Wnt7a retains full biological activity in skeletal muscle. Nat Commun 4: 2869, 2013. 312. Wang X, Wu H, Zhang Z, Liu S, Yang J, Chen X, Fan M, Wang X. Effects of interleukin-6, leukemia inhibitory factor, and ciliary neurotrophic factor on the proliferation and differentiation of adult human myoblasts. Cell Mol Neurobiol 28: 113-124, 2008. 313. Watt DJ, Morgan JE, Clifford MA, Partridge TA. The movement of muscle precursor cells between adjacent regenerating muscles in the mouse. Anat Embryol (Berl) 175: 527-536, 1987. 314. Webster C, Blau HM. Accelerated age-related decline in replicative life-span of Duchenne muscular dystrophy myoblasts: Implications for cell and gene therapy. Somat Cell Mol Genet 16: 557-565, 1990. 315. Wei L, Zhou W, Croissant JD, Johansen FE, Prywes R, Balasubramanyam A, Schwartz RJ. RhoA signaling via serum response factor plays an obligatory role in myogenic differentiation. J Biol Chem 273: 30287-30294, 1998. 316. Weintraub H, Davis R, Tapscott S, Thayer M, Krause M, Benezra R, Blackwell TK, Turner D, Rupp R, Hollenberg S, et al. The myoD gene family: Nodal point during specification of the muscle cell lineage. Science 251: 761-766, 1991. 317. Weintraub H, Tapscott SJ, Davis RL, Thayer MJ, Adam MA, Lassar AB, Miller AD. Activation of muscle-specific genes in pigment, nerve, fat, liver, and fibroblast cell lines by forced expression of MyoD. Proc Natl Acad Sci U S A 86: 5434-5438, 1989. 318. Weir AP, Morgan JE, Davies KE. A-utrophin up-regulation in mdx skeletal muscle is independent of regeneration. Neuromuscul Disord 14: 19-23, 2004.

Volume 5, July 2015

Satellite Cells and Skeletal Muscle Regeneration

319. Wen Y, Bi P, Liu W, Asakura A, Keller C, Kuang S. Constitutive Notch activation upregulates Pax7 and promotes the self-renewal of skeletal muscle satellite cells. Mol Cell Biol 32: 2300-2311, 2012. 320. Wong CF, Tellam RL. MicroRNA-26a targets the histone methyltransferase Enhancer of Zeste homolog 2 during myogenesis. J Biol Chem 283: 9836-9843, 2008. 321. Wu Z, Woodring PJ, Bhakta KS, Tamura K, Wen F, Feramisco JR, Karin M, Wang JY, Puri PL. p38 and extracellular signal-regulated kinases regulate the myogenic program at multiple steps. Mol Cell Biol 20: 3951-3964, 2000. 322. Xu Q, Yu L, Liu L, Cheung CF, Li X, Yee SP, Yang XJ, Wu Z. p38 Mitogen-activated protein kinase-, calcium-calmodulin-dependent protein kinase-, and calcineurin-mediated signaling pathways transcriptionally regulate myogenin expression. Mol Biol Cell 13: 1940-1952, 2002. 323. Yablonka-Reuveni Z, Rivera AJ. Influence of PDGF-BB on proliferation and transition through the MyoD-myogenin-MEF2A expression program during myogenesis in mouse C2 myoblasts. Growth Factors 15: 1-27, 1997. 324. Yamada M, Tatsumi R, Yamanouchi K, Hosoyama T, Shiratsuchi S, Sato A, Mizunoya W, Ikeuchi Y, Furuse M, Allen RE. High concentrations of HGF inhibit skeletal muscle satellite cell proliferation in vitro by inducing expression of myostatin: A possible mechanism for reestablishing satellite cell quiescence in vivo. Am J Physiol Cell Physiol 298: C465-C476, 2010. 325. Yamaguchi M, Ogawa R, Watanabe Y, Uezumi A, Miyagoe-Suzuki Y, Tsujikawa K, Yamamoto H, Takeda S, Fukada S. Calcitonin receptor and Odz4 are differently expressed in Pax7-positive cells during skeletal muscle regeneration. J Mol Histol 43: 581-587, 2012. 326. 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 Reports 7: 961-970, 2014. 327. Yin H, Price F, Rudnicki MA. Satellite cells and the muscle stem cell niche. Physiol Rev 93: 23-67, 2013. 328. Yoshida N, Yoshida S, Koishi K, Masuda K, Nabeshima Y. Cell heterogeneity upon myogenic differentiation: Down-regulation of MyoD and Myf-5 generates ‘reserve cells’. J Cell Sci 111(Pt 6): 769-779, 1998. 329. Zammit PS, Golding JP, Nagata Y, Hudon V, Partridge TA, Beauchamp JR. Muscle satellite cells adopt divergent fates: A mechanism for selfrenewal? J Cell Biol 166: 347-357, 2004. 330. Zhang CL, McKinsey TA, Olson EN. Association of class II histone deacetylases with heterochromatin protein 1: Potential role for histone methylation in control of muscle differentiation. Mol Cell Biol 22: 7302-7312, 2002. 331. Zhang W, Behringer RR, Olson EN. Inactivation of the myogenic bHLH gene MRF4 results in up-regulation of myogenin and rib anomalies. Genes Dev 9: 1388-1399, 1995. 332. Zhao M, New L, Kravchenko VV, Kato Y, Gram H, di Padova F, Olson EN, Ulevitch RJ, Han J. Regulation of the MEF2 family of transcription factors by p38. Mol Cell Biol 19: 21-30, 1999. 333. Zhao X, Sternsdorf T, Bolger TA, Evans RM, Yao TP. Regulation of MEF2 by histone deacetylase 4- and SIRT1 deacetylase-mediated lysine modifications. Mol Cell Biol 25: 8456-8464, 2005. 334. Zhao XD, Han X, Chew JL, Liu J, Chiu KP, Choo A, Orlov YL, Sung WK, Shahab A, Kuznetsov VA, Bourque G, Oh S, Ruan Y, Ng HH, Wei CL. Whole-genome mapping of histone H3 Lys4 and 27 trimethylations reveals distinct genomic compartments in human embryonic stem cells. Cell Stem Cell 1: 286-298, 2007. 335. Zhou VW, Goren A, Bernstein BE. Charting histone modifications and the functional organization of mammalian genomes. Nat Rev Genet 12: 7-18, 2011.

1059

Satellite Cells and Skeletal Muscle Regeneration.

Skeletal muscles are essential for vital functions such as movement, postural support, breathing, and thermogenesis. Muscle tissue is largely composed...
15MB Sizes 4 Downloads 11 Views