Review

Myogenesis in the Genomics Era

Alexandre Blais Ottawa Institute of Systems Biology, Department of Biochemistry, Microbiology and Immunology, Faculty of Medicine, University of Ottawa, 451 Smyth Road, Ottawa, ON, K1H 8M5, Canada

Correspondence to Alexandre Blais: [email protected] http://dx.doi.org/10.1016/j.jmb.2015.02.009 Edited by M. Yaniv

Abstract Skeletal myogenesis is the process of formation of the muscles that enable movement and breathing. Muscles form after the fate determination and differentiation of precursor cells. Being an extraordinarily complex process, myogenesis is regulated at multiple levels, and transcriptional regulation naturally plays a big part in the making of muscle. A significant part of what we know today of the transcriptional regulatory networks overseeing myogenesis comes from large-scale functional genomics studies. The objective of this review is to provide an overview of the various genomics techniques that have been employed over the years to understand myogenic regulation, to give a sense of the degree of understanding they have provided us up to now, and to highlight the next challenges to be overcome. © 2015 Elsevier Ltd. All rights reserved.

Introduction At its simplest level, myogenesis is the formation of muscle precursor cells, often referred to as myoblasts, and their subsequent differentiation into contractile cells termed myofibers or myotubes. Two hallmarks of muscle differentiation are that it coincides with the fusion of several myoblasts together to form syncytial myofibers, as well as with an irreversible mitotic exit. In vertebrates, skeletal muscles of the trunk and limbs originate from mesodermal precursors arising in the dermomyotome, the dorsal part of the somites (reviewed in Refs. [1–3]). Under the influence of morphogens secreted by the surrounding tissues, dermomyotomal cells commit to the myogenic lineage and start their differentiation to acquire a muscle identity. Dermomyotomal precursors eventually migrate ventrally to populate the myotome, the site of primary myogenesis. Later, the dermomyotome undergoes de-epithelialization to allow a second wave of muscle fiber formation. Muscles of the limbs originate from cells at the ventrolateral region of the dermomyotome that delaminate and migrate to the limb buds, after which they undergo terminal differentiation. Myogenesis also occurs after birth, in postnatal growth, and during regeneration after damage. Muscle 0022-2836/© 2015 Elsevier Ltd. All rights reserved.

stem cells called satellite cells are responsible for both (recently reviewed in Refs. [4] and [5]). Satellite cells are undifferentiated precursors committed to the muscle lineage. They are wedged between the basal lamina and the sarcolemma of the myofibers [6]. During the first 3 weeks of a mouse postnatal life, muscle tissue growth occurs predominantly through the fusion of satellite-cell-derived myoblasts with existing fibers [7]. In the adult, after an injury involving myotrauma, satellite cells become activated and leave their state of profound quiescence. After extensive proliferation, the myoblasts produced engage in the differentiation route and fuse to damaged fibers for repair, or they fuse together to form new myofibers. The maintenance of the pool of satellite cells is thought to be enabled by the asymmetric divisions that satellite cells engage in, early in the activation process, or by a return to quiescence of actively proliferating cells. As in any other cellular system, the profound cell identity changes that occur during the prenatal or postnatal myogenesis programs are put in place by remodeling of the cellular transcriptome, such that the genes required by a given cell type to enable them to carry out their intended functions (e.g., alertness to muscle damage, proliferation, and contraction, to name just a few) are expressed at the appropriate levels. The J Mol Biol (2015) 427, 2023–2038

2024 changes to the complement of genes expressed by muscle cells are quite formidable, both in terms of the amplitude of the expression change that is often witnessed and in terms of the sheer number of genes being regulated concurrently. For example, hundreds of cell cycle effector genes simultaneously go from highly expressed to permanently silenced, when myoblasts differentiate into myotubes [8]. Gene expression changes require to be made in an orderly and concerted fashion and are thus orchestrated by an elaborate transcriptional regulatory network [9]. At its core are sequence-specific transcription factors that control which genes will be expressed and which ones should become or remain silent as cells progress through myogenesis. Because transcription factors perform their duties in the context of a chromatinized template, rather than naked DNA, their activity is tightly connected to the structure of that template: it is thought that transcription factor function can be both a cause and a consequence of specific chromatin states, often referred to as the epigenome [10]. Although a very large number of transcription factors are expressed in muscle cells at some point during development and/or in adulthood, some of them are in fact uniquely expressed, or strongly enriched, in skeletal muscle. The transcription factors most intimately connected to myogenesis are the myogenic regulatory factors (MRFs), a group of four related basic helix–loop–helix factors that are at the center of skeletal muscle formation (reviewed in Ref. [11]). The first member identified, MyoD, was identified by virtue of its ability to convert fibroblasts to the muscle lineage [12,13]. The MRFs do not act alone but cooperate with many other muscle-enriched factors, notably the homeodomain factors Six1, Six4, and Pbx1 [14,15]; the MADS box factors of the Mef2 family [16,17]; and the Rel-homology domain transcription factors of the NFATc family [18,19]. Other transcription factors are also involved in myogenesis, either by controlling the expression of the MRFs or by other means (reviewed in Ref. [20]). Notable cases include the paired-box transcription factors Pax3 and Pax7 [21,22], as well as the homeodomain factor Pitx2 [23] and Nfix [24]. Pax7 is especially important in the case of adult muscle myogenesis, as its expression is an essential and defining factor of satellite cells [25] and its absence, just like the absence of satellite cells themselves, abrogates regeneration after injury [26–29]. A detailed understanding of how myogenesis is regulated by these and other transcription factors, as well as by the epigenome, would be beneficial for multiple reasons, beyond simply improving our general knowledge: understanding of regulatory principles that may also prevail in other biological systems, elucidating the etiology of muscular diseases, and obtaining important clues on how we may be able to treat the afflicted patients. For instance, reprogramming of pluripotent stem cells (embryonic or induced) in view of cell therapy requires an in-depth knowledge of the

Review: Myogenesis in the Genomics Era

myogenic regulatory network [30–35]. Beyond the obvious involvement of skeletal muscle in breathing and movement, this tissue also plays an important role in energy metabolism due to its high total mass and it being responsible of about 20% of the body's resting energy expenditure. Skeletal muscle takes care of the bulk of glucose uptake in response to insulin, and deregulation in muscle composition or homeostasis is intimately connected to energy metabolism. For instance, obese insulin-resistant (type 2) diabetic patients tend to have a lower proportion of type I oxidative muscle fibers [36] and have malfunctioning skeletal muscle mitochondria [37]. Likewise, diet-resistant and diet-sensitive obese patients tend to have different muscle characteristics, with diet-sensitive patients having more type I oxidative muscle fibers compared to diet-resistant individuals [38]. Thyroid hormone signals to skeletal muscle to help establish the basal metabolic rate and to elicit adaptive thermogenesis (reviewed in Ref. [39]). Understanding the mechanisms controlling muscle formation, adaptation, and homeostasis can therefore inform us on the regulation of energy balance and on how metabolic changes are brought about by variations in physical activity, endocrine status, mode of life, or disease states. Thus, we can identify three broad goals to be achieved for an understanding of the myogenesis transcriptional regulatory network: (1) to obtain an accurate representation of transcriptomes of cells at the successive stages of myogenesis, (2) to identify the transcription factors and their target loci involved in establishing these changes, and (3) to determine the state of the epigenome at those stages and identify its relationship to transcription factor function. Over the last 10–15 years, one arguably successful way of achieving these goals has been the use of a systems biology approach, where genomics tools are employed to provide us with a global picture of the myogenic regulatory network. The next sections summarize the progress made and highlight areas for future research.

Genomics Phase I: Gene Expression Profiling The invention of gene expression profiling by DNA microarrays [40] revolutionized the way we study genome regulation, and in fact, it was instrumental to the emergence of systems biology as an experimental science [41]. The first microarray studies of gene expression in skeletal muscle were performed using the murine C2C12 myoblast cell line, an in vitro model of myogenesis. Derived from a hindlimb muscle, the cells proliferate rapidly under high-growth-factor conditions and differentiate when they reach confluence and are switched to low serum medium [42,43]. They are often used as a model of activated satellite cells, as they share many properties with the adult

Review: Myogenesis in the Genomics Era

muscle stem cells: the ability to proliferate, to fuse together and differentiate to give rise to multinucleated contractile myotubes, and to do so concomitantly with an irreversible mitotic exit. The adoption of C2C12 as a favorite model to study myogenesis has been facilitated by the facts that they can be cultured in large amounts indefinitely (being immortal at least due to their loss of tumor suppressor Arf [44]), that they are clonal and thus more homogeneous than primary myoblast cultures, and that they recapitulate many fundamental steps of myogenic differentiation. Two early microarray studies reported gene expression profiles of C2C12 myoblasts on a time course of differentiation [8,45]. The importance of this early work was in providing an idea of the scale of transcriptome changes that occur during myogenesis and revealing the identity of hundreds of differentially expressed genes in myoblasts and myotubes. The existence of large clusters of co-expressed genes was also put forth, confirming and reinforcing the notion that myogenic differentiation is a highly coordinated, and hence organized, process. A later study looked at gene expression changes that occur in the entire tibialis anterior muscle during regeneration following a massive necrotizing injury induced by the injection of the snake venom toxin cardiotoxin [46]. The results reaffirmed the notions that massive expression changes occur during myogenesis and that these transcriptome alterations are highly coordinated. This study also revealed the limits and challenges inherent to studying complex tissues as opposed to pure cell cultures. For instance, the massive infiltration of macrophages in the injured muscle was witnessed by the appearance of clusters of inflammatory function genes. However, this was resolved years later by the expression profiling of quiescent and activated satellite cells purified by virtue of their expression of specific cell surface markers [47]. Interestingly, clusters of genes associated to wound healing and chemoattraction were found to be specifically expressed in activated satellite cells, providing strong evidence that the muscle stem cells play an active role in the essential inflammatory process that follows injury. The advent of RNA sequencing (RNA-seq) to perform gene expression profiling has been a huge leap forward in terms of accuracy, sensitivity, and comprehensiveness of transcriptome surveys [48]. RNA-seq allows to discover unannotated transcripts and to look at alternative transcription start site usage and alternative splicing in an unbiased manner. Profiling of C2C12 myoblasts on a differentiation time course has revealed the extent by which alternative splicing occurs and changes with differentiation [49] and should help identifying the mechanisms behind these fluctuations. In particular, alternative splicing of transcripts from the Mef2D gene gives rise to two proteins with radically different transcriptional regulatory functions in myogenesis [50], and it is probably not exaggerated to believe that similar mechanisms control

2025 the function of other key players in the myogenic regulatory network. The increased sensitivity of RNA-seq has allowed the identification of important non-coding, enhancer-derived RNAs (eRNAs) implicated in myogenesis regulation [51,52]. They tend to be transcribed from loci neighboring highly expressed protein-coding genes that are associated to musclerelated functions. Two of these are transcribed from well-characterized enhancers of Myod, the core enhancer region and distal regulatory region. They are exclusively nuclear, unlike protein-coding transcripts, and knockdown experiments suggest that the eRNA derived from the core enhancer region is necessary for normal Myod transcription and might act in cis. The eRNA arising from MyoD's distal regulatory region has been termed MUNC, for “MyoD upstream non-coding RNA” [52]. There is conflicting evidence on whether MUNC is involved in the control of Myod expression, but it is clear that it is required for MyoD target gene activation (action in trans). Interestingly, MUNC also appears to play a transcriptional activation role independently of MyoD induction [52]. The two eRNAs would act by altering the opening of the structure of chromatin and influencing RNA pol II recruitment at proximal promoters. The discovery of these eRNAs adds to the growing list of non-coding RNAs that play a role in controlling myogenesis [53,54]. More recently, expression profiling has been applied to satellite cells purified by virtue of their expression of a fluorescent marker gene [55] or cell surface proteins [47,56]. One goal common to all studies was to identify genes that would be characteristic of the quiescent state and was achieved by comparing the transcriptome of quiescent satellite cells with that of activated satellite cells (isolated from actively regenerating muscle or cultured in vitro). This is something that could not be achieved using C2C12 cells, which mimic satellite cells in the active state and are grown outside of the muscle niche. In particular, these studies allowed us to identify negative regulators of the cell cycle and proteins involved in the interaction between these stem cells and their niche. Gene expression profiling can reveal the state of a cellular transcriptome. Bioinformatic algorithms to analyze transcriptome data on a time course of differentiation or on cells with gain or loss of function of a transcription factor can reveal the regulatory logic that exists in the system being studied. Segal et al. showed that, from expression profiling data, one can identify clusters of co-expressed genes that are likely under the control of the same, co-expressed transcription factors [57]. A similar strategy has been employed to various large expression datasets, incorporating or not knowledge of transcription factor DNA binding preferences [58–62]. However, these methods have important limits. For example, steady-state mRNA levels are not very accurate proxies to estimate transcription factor activity, since this does

2026 not take into account post-transcriptional levels of regulation (e.g., by alternative splicing, post-translational modifications, or interactions with co-regulators). They also cannot always resolve the contribution of highly related transcription factors, for example, those that would be co-expressed and expected to bind related or identical DNA sequences, as would be the case of MyoD and Myogenin. A second gene-expression-based approach used to identify regulatory relationships in the myogenesis control network is expression profiling after gene gain or loss of function. The assumption is simple: genes whose mRNA is deregulated by altered transcription factor activity might be under the control of that factor. One of the first studies was performed to identify MyoD-regulated genes in fibroblasts undergoing myogenic conversion [63], and the approach has been extended to many other transcriptional regulators. For example, profiling of wild-type and Six1/Six4 knockout embryos has revealed a basis for muscle-fiber-type development program [64], comparison of wild-type and Mef2c knockout mouse embryonic limbs has shown the importance of this factor in sarcomeric organization [65], the important role played by Runx1 in preventing muscle wasting after denervation was revealed by microarray profiling of muscle tissue of adult Runx1 knockout mice [66], putative Pax3 target genes in the developing somites were identified by their up-regulation in embryos expressing the transcriptionally potent Pax3-FKHR fusion protein [67], and possible genes mediating the proliferative effect of the Hippo-Yap pathway in satellite cells were uncovered by profiling the transcriptome of activated satellite cells with constitutively active Yap [68]. The approach has the power to reveal the identity of genes located downstream of the factors of interest, in the regulatory network. One drawback of the method is the difficulty to distinguish direct from indirect effects, which can occur any time a transcription factor regulates the expression of another protein acting in gene regulation; the regulation of Myog by MyoD is a prime example. While this problem can be circumvented by performing the expression profiling in the presence of the protein synthesis inhibitor cycloheximide, as was performed to identify direct targets of Myod [63], this proves impossible when work is being conducted in developing embryos.

Genomics Phase II: Genome-Wide Location Analysis of Transcription Factors The development of methods that allow us to identify the genomic sites of binding of transcription factors has been of tremendous accelerator of our understanding of myogenesis. Experiments using chromatin immunoprecipitation of transcription factors

Review: Myogenesis in the Genomics Era

followed by identification of the associated genomic DNA fragments by microarrays (ChIP-on-chip [69]) or by high-throughput sequencing (ChIP-seq [70,71]) tell investigators where their factor of interest locates in the genome, essentially revealing the identity of direct target genes. The first ChIP-on-chip studies performed in muscle looked at the binding sites of Myod, Myogenin, and Mef2c in C2C12 myoblasts and myotubes and in MRF-expressing fibroblasts [72,73]. The microarrays used for target DNA identification were limited to proximal regulatory regions of a subset of all known genes and therefore provided only a limited, biased view of the myogenic regulatory network. However, they revealed, for example, that Myod and Myogenin, despite having very similar DNA binding domains, do not have identical target gene sets. These studies also highlighted the existence of condition-specific targets of these transcription factors bound preferentially or more strongly in myoblasts or in myotubes. These are important considerations because they hint to the fact that DNA sequence is not the sole determinant of transcription factor binding. This has been echoed by the limited success of bioinformatic methods attempting to predict physiologically relevant transcription factor binding sites, even when parameters such as clustering into cis-regulatory modules and phylogenetic conservation are taken into account [74]. In these studies, the transcription factor binding profiles were combined to gene expression profiles of myoblasts prior to and after differentiation or MRF expression. What emerged is that the correlation between target gene binding by a transcription factor and transcriptional modulation of the genes is far from perfect: the expression of several direct targets of the MRFs is stable during myoblast differentiation. Beyond the trivial possibilities (e.g., false-positive binding sites and wrong gene annotations), it might be because the target genes are expressed equally in myoblasts and myotubes or because additional stimuli (or lifting of repressive activities) are necessary for transcriptional activation to occur. ChIP-seq mapping of MyoD targets in myoblasts and fibroblasts [75] generated quite a surprise because of the sheer number of binding sites discovered: the unbiased screened revealed that MyoD binds more than 20,000 genomic locations. Similar experiments performed by the laboratory of B. Wold (Caltech) under the ENCODE project or the laboratory of V. Sartorelli [51] came to comparable numbers. Again, most genes with neighboring MyoD binding sites turned out to have expression profiles that are unchanged during myogenic differentiation. However, analysis of covalent histone modification marks in fibroblasts expressing MyoD or not revealed that the majority of these binding events are at least partially functional: although most MyoD-bound regions do not display enhancer activity in transcriptional assays, the global trend is that MyoD binding correlates with acquisition of histone H4

2027

Review: Myogenesis in the Genomics Era

acetylation, a mark associated to transcribed loci. In contrast, there is no clear association between MyoD binding and trimethylation of H3K4, a mark associated mostly to the transcription start site of expressed genes. What then determines if MyoD will give rise to productive transcriptional activation? One possibility is that it must cooperate with other transcription factors that stimulate complementary essential steps of the transcription process. For instance, we found that binding sites for the homeodomain transcription factors of the Six family have a striking overlap with those of MyoD [14], and it was shown that Six4 helps recruit the H3K27me3 demethylase Utx to the Myog and Ckm loci for derepression in differentiating myoblasts [76]. Furthermore, Six1 is necessary for MyoD expression and for MyoD binding to DNA at its own core enhancer, in myoblasts [77]. Genome-wide binding profiles of Mef2c and Mef2D are also suggestive of a significant overlap with MyoD target sites [50,72], and Mef2D controls MyoD target gene transcription in at least two ways. First, Mef2D can repress MyoD target gene expression in proliferating myoblasts, through the recruitment of histone deacetylase Hdac9 [50]. Second, it participates in the recruitment of the Trithorax group MLL H3K4 methyltransferase, which contributes to the transcriptional activation of MyoD targets [78]. There is also a partial overlap between binding sites for the Tead4 transcription factor and those of MyoD [79], although the molecular basis for a possible cooperation between Tead and MyoD is not well understood, beyond the fact that it involves the vestigial-like 2 co-activator [80]. Other myogenesis regulators may function by antagonizing MyoD function. Such is the case for Snail1, which was shown to recruit the histone deacetylase Hdac1 to MyoD target sites in myoblasts and block MyoD binding to its targets before the onset of differentiation [81]. Down-regulation of Snail by microRNAs whose expression is induced by MyoD at the onset of differentiation would represent a mechanism that might render the differentiation process irreversible. Pax3 and Pax7 are two transcription factors essential for myogenesis. Although they are expressed in overlapping groups of cells and share significant structural homology, they do not carry out exactly the same function. In particular, Pax7 is essential in satellite cells while Pax3 is dispensable, and Pax3 cannot prevent satellite cell apoptosis triggered by the loss of Pax7 [22,82]. ChIP-seq analysis of TAP-tagged Pax3 and Pax7 in primary adult mouse myoblasts has shed light on the molecular basis of these functional differences by showing that Pax7 binds to a much larger number of genomic loci than Pax3 (more than 52,000 versus 4600) and the two overlap at only approximately 1200 loci [83]. This is reflected by apparent preferences of the two factors for binding DNA through the paired-box, in the case of Pax3, or the

homeodomain, for Pax7, and it is hypothesized that the dominance of Pax7 is due to its higher affinity for the homeodomain binding sequence motif “TAAT”.

Genomics Phase III: Contribution of the Epigenome Transcription factors function on a chromatin template. Histone proteins are subject to a myriad of covalent modifications that correlate with, and sometimes regulate, the transcriptional state of genes. The Dynlacht laboratory has used ChIP-seq to map in C2C12 myoblasts and myotubes an impressive number of histone marks associated to gene activity or silencing and to identify the binding sites of the protein complexes that write these marks to control myogenesis [84–87]. In particular, the profiling of multiple marks including H3K4me1 and H3K27ac and the recruitment of MyoD and the histone acetyltransferase p300 and RNA pol II allowed the identification of several thousands of active transcriptional enhancers, a large fraction of which are MyoD dependent in that the histone marks are greatly diminished in MyoD knockout myoblasts [86]. MyoD plays an active role in establishing the histone mark landscape of enhancers, in part by recruiting the H3K4 methyltransferase Setd7 and enzyme required for efficient myogenic differentiation [88]. It is interesting that although the H3K4me1 mark is found at enhancers independently of their activity state, the mark is also found at proximal promoters of silent but inducible genes [87]. The MLL3 and MLL4 methyltransferases are responsible for establishing this repressive chromatin environment at the proximal promoter of certain genes in myoblasts, and knockdown of MLL3/MLL4 in myoblasts leads to the up-regulation of a group of muscle genes. At the onset of differentiation, recruitment of the H3K4 demethylase Lsd1 is associated with the removal of the repressive mark and with the MLL1-dependent trimethylation of H3K4 (a mark that cannot be erased by Lsd1), allowing activation of the genes. H3K4me3 constitutes a recognition platform for ING1, which can recruit a Sin3a/Lsd1 complex that will further antagonize H3K4me1 once differentiation is initiated. Interestingly, MLL4 has been shown to be essential for skeletal myogenesis and adipogenesis, as mice without MLL4 in Myf5-expressing cells are characterized by deficient back muscle formation [89]. In the context of myogenic conversion of preadipocytes by MyoD overexpression, loss of MLL4 was associated to attenuated muscle gene induction, suggesting that the protein is involved in activating their expression during myogenesis. In concordance with that, ChIP-seq for MLL4 in MyoD-overexpressing preadipocytes revealed that the protein is strongly enriched at MyoD-bound enhancers categorized as “active” for bearing the H3K27ac mark. In fact, MLL4 was shown

2028 to be necessary not only for the monomethylation of H3K4 at muscle gene enhancers but also for their acetylation on H3K27. These two studies [87,89] on MLL4 are therefore in apparent contradiction since they propose that MLL4 functions at promoters or enhancers as inhibitor or activator of muscle gene expression, respectively. However, very different systems were used to look at MLL4 in myogenesis; thus, it is not inconceivable that the protein is involved in both aspects depending on the exact myogenesis context. One could imagine that MLL4 is required for myogenic determination, in the developing embryo or in heterologous cells with ectopic expression of MyoD, and that it is also needed to prevent the premature differentiation of committed myogenic progenitors that are already expressing MyoD. Another transcriptionally repressive histone mark that must be removed for myogenesis to occur is H3K27me3, a mark written by the polycomb repressive complex 2 (PRC2). The association of H3K27me3 with muscle gene repression was first revealed by Caretti et al., who showed that the activity of Ezh2, the catalytic subunit of PRC2, antagonizes muscle gene induction in order to prevent premature myogenic differentiation [90]. They showed that the Yy1 transcription factor is involved in recruiting Ezh2 to muscle genes in order to repress their expression by trimethylation of H3K27. The onset of differentiation coincides with dissociation of Yy1 and Ezh2 from these genes, lowered H3K27me3 levels, and gene activation. While Ezh2 and the rest of the PRC2 complex are present on the promoter of many muscle genes in myoblasts, more recent evidence suggests that, in most instances, Yy1 is not part of the recruitment mechanism since the genomic binding profiles of both proteins are very different [91]. The decrease in H3K27me3 levels when myoblasts engage in the differentiation path could occur through “passive dilution”, whereby H3K27me3-containing nucleosomes are replaced during cell division by nucleosomes devoid of this repressive mark. However, more recent work demonstrated the importance of an “active removal” mechanism. Demethylation of H3K27me3 by the Trithorax group protein Utx at muscle gene promoters is an important event for myogenesis [76]. Utx associates with the transcriptional elongation factor Spt6, a histone chaperone known to help the passage of RNA pol II through chromatin by destabilizing nucleosomes [92]. Spt6 knockdown is associated to a failure to undergo myogenesis and lower transcription of muscle genes. This is tied to higher H3K27me3 levels. The authors found that the two proteins are part of the same complex. In fact, using ChIP-seq allowed them to demonstrate a genome-wide correspondence between recruitment of Spt6 and the demethylase Utx and conversely absence of H3K27me3 signal. The proof that the role of Spt6-Utx in myogenesis is to erase the repressive H3K27me3 mark came from rescue experiments: knocking down the PRC2 component

Review: Myogenesis in the Genomics Era

Ezh2 rescues the effect of Spt6 knockdown on muscle gene expression. Genome-wide histone mark analyses are not limited to C2C12 cells. The Rando laboratory used ChIP-seq to look at the relationship between aging and epigenetic marks in satellite cells [47]. By profiling quiescent and activated young and old satellite cells for the presence of H3K4me3, H3K27me3, and H3K36me3, they found that quiescent and activated satellite cells have globally comparable profiles of the H3K4me3 mark, which suggests that genes induced during the transition from quiescence to activation are already marked for expression. Still, a small number of important myogenic regulators, such as Myogenin, acquire H3K4me3 only once cells become activated. The situation with H3K4me3 is in contrast to what happens to the repressive H3K27me3: the number of loci repressed by the H3K27me3 mark that greatly increase during activation. This brings up the question of how the repressive mark is removed from satellite cells that return to quiescence instead of differentiating, following activation. Presumably, a histone demethylase acting on H3K27me3, such as Utx or Jmjd3 [93], must come into play. In satellite cells from old mice, H3K4me3 is unaffected but there is an increase in the number of genomic loci marked by H3K27me3 in quiescent cells compared to young mice, suggesting that perhaps the demethylation mechanisms involved have lower activity with age or that PRC2 activity is increased. Access to DNA is controlled by the remodeling of chromatin, the addition or displacement of nucleosomes, or the replacement of canonical histones with variants. There are several protein complexes involved in chromatin remodeling and controlling access of transcription factors to DNA. One of the complexes shown to be important for myogenesis is mSWI/SNF, also called the BAF complex (reviewed in Refs. [94–96]). MyoD has been shown to participate in the recruitment of mSWI/SNF during myogenesis [97]. Interactions between MyoD and one of the subunits named Baf60c control recruitment of the rest of mSWI/ SNF in a manner that is dependent on phosphorylation of Baf60c by p38 [98]. It will be interesting to see how widespread this mechanism is and, more generally, to understand how mSWI/SNF and the other nucleosome-displacing complexes contribute to myogenesis. Several genome-scale methods exist to explore chromatin structure, such as DNAse-seq [99], FAIRE-seq [100], and ATAC-seq [101], and applying these techniques to the myogenesis system should reveal a great deal about its regulatory network. The ENCODE project includes FAIRE-seq and DNAseseq datasets from a large number of tissues or cell lines, including skeletal muscle, but an in-depth analysis of the observed patterns of DNA accessibility still remains to be conducted. There is ample evidence demonstrating that histone variants, loaded into chromatin independently of DNA

Review: Myogenesis in the Genomics Era

replication, contribute to regulating DNA accessibility. Genome-wide in vivo studies suggest that the presence of histone variants H3.3 and H2A.Z contributes to rendering nucleosomes more labile, which is postulated to facilitate access to DNA by transcription factors [102]. Loading of H3.3 is also associated with preventing the formation of higher-ordered chromatin, which could also help transcription factor activity [103]. Genome-wide maps of H3.3-containing nucleosomes in C2C12 myoblasts have been reported [104]. They reveal that this histone variant is loaded preferentially at genes to be induced during myogenesis and that this activity depends on MyoD's ability to recruit the SNF2-related enzyme Chd2. It has been shown that nucleosomes located immediately downstream of the transcription start site constitute a barrier to the progression of RNA polymerase II but that this block is easier to overcome when the nucleosomes contain H2A.Z in lieu of H2A [105]. Although both p400/NuA4 [106] and the SNF2-related complex SRCAP [107] can replace H2A with H2A.Z near transcription start sites, only SRCAP is necessary for the induction of Myog at the onset of C2C12 myoblast differentiation [108]. A genome-wide view of H2a.z location in the muscle system remains to be obtained and will be especially interesting to analyze in the context of the maps of H3.3 and histone marks mentioned above.

Challenges and Opportunities Ahead Despite the terrific progress that functional genomics approaches have collectively generated to understand the myogenic regulatory network, there is obviously a lot that remains to be discovered. Some areas of research, despite being challenging, have enormous potential. The vast amount of genomic data that have been garnered so far on the myogenic regulatory network is truly impressive; each study puts a new piece of the puzzle into place, such that we understand a little bit better every time how our regulatory network functions. One approach that holds great promise is the use of computational algorithms that integrate various data elements into descriptive and predictive models. Among others, it has been applied successfully to explore development regulatory pathways in the fruit fly [109–114], in Caenorhabditis elegans [115,116], in mouse heart formation [117], and in embryonic stem cells [118,119]. One benefit of these approaches is to process a collection of correlations (e.g., factor A binds gene set B, which carries histone mark C) into cause–effect relationships. The predictive nature of these models is also extremely useful because it helps to identify new regulatory nodes in the transcription network, which can then be studied re-iteratively using the array of techniques described above. Considering how much information has already been obtained

2029 from C2C12 myoblasts (see Table 1), there is little doubt that computational biologists in the muscle field have in hand a good starting collection of datasets. A second challenge resides in the fact that muscle tissue is very complex, being composed of a multitude of cell types, and that there is even heterogeneity within “defined” populations. Even within a single myofiber, not all nuclei transcribe the same genes at the same levels due to their varying distance from the neuromuscular junction (discussed in Ref. [120]). A lot of what we currently know about the myogenic regulatory network has been discovered using C2C12 or primary myoblasts in culture. C2C12 cells, being clonal, arose from a single, unidentified myogenic cell type out of the many known (reviewed in Ref. [121]). They spontaneously became immortal with serial passaging and have undergone poorly characterized chromosomal rearrangements. On the other hand, primary myoblast cultures used in genomics experiments can be considered genetically normal, but the issues reside elsewhere. Primary myoblasts are typically isolated in bulk from a mixture of limb muscle groups (e.g., see Ref. [122]) and used directly in genomic experiments (e.g., see Ref. [83]) or subjected first to enrichment based on cell surface (e.g., see Refs. [47] and [56]) or genetic markers (e.g., see Ref. [55]). Even with significant marker enrichment, the populations used can still be quite heterogeneous. Considering that satellite cells are profoundly affected by their niche [123] and that the niche may vary significantly according to anatomical location, slow- or fast-twitch myofiber association, proximity to blood vessels, or neuromuscular junctions, it seems logical to propose that not all satellite cells are equivalent. There is ample experimental evidence supporting the notion of satellite cell heterogeneity (reviewed in Refs. [124] and [125]). From there, it would make sense to assume that the myogenic regulatory network is not in quite the same state in all satellite cells either. Seemingly pure cell populations exhibit significant variability in the genes they express once individual cells are considered independently. This has been highlighted very clearly in a recent study where single-cell RNA-seq of primary human myoblasts was conducted [126]. Even in clonal cell lines, variability in gene expression is seen between individual cells [127]. Analysis of large cell pools provides a population average that can mask minor yet important trends in gene expression by groups of cells, for instance, bimodality of gene expression (i.e., the existence of subpopulations) and heterogeneity in gene expression levels and alternative splicing isoform selection [128,129]. Whole-transcript single-cell RNA-seq is now feasible [127] and could be applicable to isolated satellite cells as well; this should reveal the extent of transcriptome variability within this cell population. Another promising technique that may soon help deconvolute satellite cell subpopulations is fluorescent in situ RNA-seq (FISSEQ), which has the added benefit of functioning in the intact tissue [130].

2030

Table 1. A non-exhaustive list of genomics datasets relevant to myogenesis Description Expression profiling C2C12 differentiation time course

Cell type

Technology

References

Mapping of TF binding sites MyoD in myoblasts and myotubes

C2C12, growing and differentiated

ChIP-on-chip, proximal promoter array, ChIP-seq

Myogenin in myotubes

C2C12 myotubes

ChIP-on-chip, proximal promoter array, ChIP-seq

Mef2c in myotubes MyoD in myoblasts, fibroblasts

C2C12 myotubes C2C12, primary myoblasts, embryonic fibroblasts P19 murine embryonal carcinoma C2C12 myotubes Primary myoblasts, growing and differentiating C3H10T1/2 fibroblast Primary myoblasts Primary myoblasts

ChIP-on-chip, proximal promoter array, ChIP-seq

ENCODE (Caltech), [51,72,73,132] ENCODE (Caltech), [51,72] [72] [75,133]

ChIP-seq ChIP-seq ChIP-seq

[134] [132] [81]

ChIP-seq ChIP-seq ChIP-seq

[81] [81] [81]

Satellite cells, quiescent and in vitro activated

MyoD in pluripotent cells Smad3 in myotubes MyoD in myoblasts, TAP-tagged MyoD in fibroblasts, TAP-tagged Myf5 in myoblasts, TAP-tagged E47, TAP-tagged

Review: Myogenesis in the Genomics Era

Satellite cells from resting and regenerating tibialis anterior of young and aged mice Mouse embryos, wild-type and Six1/Six4 knockouts

Affymetrix microarrays; RNA-seq; Agilent microarrays, [8,14,45,49] Illumina Mouse WG-6 Expression Bead Arrays Mouse tissue cDNA microarray [46] C3H10T1/2 fibroblasts [63] Primary myoblasts Affymetrix microarrays [72] C2C12 myotubes Affymetrix microarrays [131] C2C12, growing and differentiating Agilent microarrays [14] C2C12, differentiating Agilent microarrays [14] C2C12 myotubes RNA-seq [79] C2C12, differentiating RNA-seq [50] Primary myoblasts RNA-seq [81] Primary myoblasts Affymetrix microarrays [83] Adult mouse satellite cells purified by FACS Affymetrix microarrays [56] (SM/C-2.6) Adult mouse satellite cells purified by FACS Affymetrix microarrays [55] (Pax3-GFP) [47] Adult mouse satellite cells purified by Affymetrix microarrays FACS (VCAM1+/CD31−/CD45−/Sca1−) E10.5 somites Affymetrix microarrays [64]

Tibialis anterior regeneration time course Fibroblast myogenic conversion Primary myoblasts, wild-type and Myod knockouts Myotubes after pRb knockdown Myoblasts after Six1 and Six4 knockdown Myoblasts after Myog knockdown Myoblasts after Tead4 knockdown Myoblasts after Mef2D knockdown Myoblasts after Snai1 or Snai2 knockdown Myoblasts with Pax3 or Pax7 overexpression Satellite cells, quiescent and in vitro activated

C2C12 mouse myoblasts

Mapping of co-regulators Utx in myoblasts and myotubes Spt6 in myoblasts and myotubes Ezh1 in myoblasts and myotubes Ezh2 in myoblasts and myotubes Suz12 in myoblasts and myotubes RNA pol II in differentiating myoblasts, all forms (8WG16) RNA pol II in differentiating myoblasts, CTD S5P (paused) MLL1 in myoblasts and myotubes MLL2 in myoblasts and myotubes MLL4 before and after myogenic conversion Ing1 in myoblasts and myotubes Lsd1 in myoblasts and myotubes Sin3a and Sin3b in myotubes p300 in myoblasts and myotubes Hdac1 in myoblasts, TAP-tagged Hdac2 in myoblasts, TAP-tagged Chromatin structure and modifications H3K4me3 and H4ac in fibroblasts, with and without MyoD H3K4me1, H3K4me2, H3K4me3, H3K9me3, H3K36me3, H3K27me3, H3K9ac, H4K12ac, H3K18ac, in myoblasts and myotubes H2bK120Ub, in myoblasts and myotubes H3K27me3 Histone variant H3.3 in myoblasts H3K4me3 H3K36me3 in myoblasts H3K4me3, H3K27me3, H3K36me3, in quiescent and activated satellite cells

Primary myoblasts Primary myoblasts Primary myoblasts C2C12 myotubes C2C12, growing and differentiating C2C12 myotubes C2C12, growing and differentiating C2C12, growing and differentiated C2C12, differentiating RD cells and normal human muscle tissue RD cells

ChIP-seq ChIP-seq ChIP-seq ChIP-seq ChIP-seq ChIP-on-chip, Agilent promoter ChIP-on-chip, Agilent extended promoters ChIP-on-chip, Agilent promoter microarrays ChIP-seq ChIP-seq ChIP-seq

[81] [83] [83] [135] [91] [79] [14] [136] [50] [140] [139]

C2C12, growing and differentiated ChIP-seq C2C12, growing and differentiated ChIP-seq C2C12, growing and differentiated ChIP-seq C2C12, growing ChIP-seq C2C12, growing and differentiated ChIP-seq C2C12, differentiating ChIP-seq C2C12, differentiating ChIP-seq C2C12, growing and differentiated ChIP-seq C2C12, growing and differentiated ChIP-seq Preadipocytes, preadipocytes with MyoD ChIP-seq overexpression C2C12, growing and differentiated ChIP-seq C2C12, growing and differentiated ChIP-seq C2C12 myotubes ChIP-seq C2C12, growing and differentiated ChIP-seq Primary myoblasts ChIP-seq Primary myoblasts ChIP-seq

[92] [92] [160] [91,139] [139] [139] [139] [87] [87] [89]

Embryonic fibroblasts C2C12, growing and differentiated

ChIP-seq ChIP-seq

[75] [84]

C2C12, growing and differentiated C2C12, growing C2C12, growing C2C12, growing C2C12, differentiating FACS-purified mouse satellite cells

ChIP-seq ChIP-seq ChIP-seq ChIP-seq ChIP-seq ChIP-seq

[85] [91,139] [104] [139] [139] [47]

Review: Myogenesis in the Genomics Era

Snai1 in myoblasts, TAP-tagged Pax3 in myoblasts, TAP-tagged Pax7 in myoblasts, TAP-tagged Msx1 in myotubes, Flag-tagged YY1 in myoblasts Tead4 in myotubes, Flag-HA-tagged Six1 in myoblasts and myotubes E2f3a and E2f3b in myoblasts Mef2D in myotubes MyoD in rhabdomyosarcoma cells Musculin in rhabdomyosarcoma cells

[87] [87] [140] [86,87] [81] [81]

2031

2032 Both approaches would lead to a better understanding of satellite cell gene and protein expression, which is essential if we want to devise strategies to separate them to a greater degree of purity for other genomic analyses. Genetic isolation of cells based on their expression of a specific gene is possible, and for instance, it is possible to isolate Pax3- or Pax7-expressing cells by virtue of their expression of fluorescent protein markers whose expression is controlled by the regulatory elements of the cognate genes [141–144]. However, pure populations cannot be isolated by virtue of their expression of a single gene. Strategies employing a combination of multiple genetic markers may be required but are currently limited by the number of fluorescent protein-coding genes that can be used concurrently in flow cytometry applications. Synthetic transcription-based circuits incorporating multi-input AND logic gates [145], possibly combined with split fluorescent proteins or enzymes as reporters [146], may one day be amenable to transgenic animal applications and would greatly increase the specificity of cell selection for genomic analyses. It should be noted that, for ChIP-seq studies, the BiTS-ChIP method makes it possible to isolate, from complex tissues that have been cross-linked with formaldehyde, only the nuclei of interest on the basis of reporter gene expression [147]. Single-cell ChIP-seq is not currently feasible, but for histone mark mapping, as low as 1000 cells can be used [148]. It should be possible to acquire this much satellite cells that are pure enough for such studies. Another avenue of great interest will be to explore, with a genomics perspective, the function of the various myogenic and non-myogenic cells that support muscle development and regeneration (e.g., infiltrating macrophages, PW1-expressing interstitial cells, fibroadipogenic progenitors, and fibroblasts). Some of these populations are becoming very well defined in terms of their surface markers, which should enable transcriptomic and various ChIP-seq analyses. Such work will allow us to compare the regulatory networks controlling myogenesis in these various cell types with myogenic potential and would help reveal how non-myogenic cells exert their influence on satellite cells and vice versa.

Conclusion From the discovery of MyoD to mapping of its transcriptional targets and to elucidation of epigenetic marks associated to muscle transcription factor function, the field of skeletal myogenesis research has made tremendous progress. Every novel finding sheds some new light on the transcriptional regulatory network that controls this process. The advent of new technologies, with increased analytical resolution, sensitivity, and/or precision has heightened the pace of discoveries, and there is little doubt that progress will

Review: Myogenesis in the Genomics Era

continue. Genomics studies in muscular disease models or clinical samples are also providing important insight about what is the diseased state precisely is or how it came to be (e.g., for rhabdomyosarcoma [137– 138], [149–151], muscular dystrophy [152–157], and muscle wasting [158–161]). Constantly improving our understanding of what keeps our muscles in good function and health and of what is deregulated during illnesses is the key to finding treatments to muscular diseases.

Acknowledgements The author would like to acknowledge the help of laboratory members for insightful discussions. Research in the author's laboratory is supported by an operating grant from the Canadian Institutes of Health Research (MOP 119458). Received 9 December 2014; Received in revised form 4 February 2015; Accepted 5 February 2015 Available online 14 February 2015 Keywords: myogenesis; transcriptional regulatory network; transcription factors; epigenetics; chromatin structure

Abbreviations used: MRF, myogenic regulatory factor; RNA-seq, RNA sequencing; PRC2, polycomb repressive complex 2.

References [1] Buckingham M. Skeletal muscle formation in vertebrates. Curr Opin Genet Dev 2001;11:440–8. [2] Bryson-Richardson RJ, Currie PD. The genetics of vertebrate myogenesis. Nat Rev Genet 2008;9:632–46. http://dx. doi.org/10.1038/nrg2369. [3] Buckingham M, Vincent SD. Distinct and dynamic myogenic populations in the vertebrate embryo. Curr Opin Genet Dev 2009;19:444–53. http://dx.doi.org/10.1016/j.gde.2009.08.001. [4] Montarras D, L'honore A, Buckingham M. Lying low but ready for action: the quiescent muscle satellite cell. FEBS J 2013;280:4036–50. http://dx.doi.org/10.1111/febs.12372. [5] Chang NC, Rudnicki MA. Satellite cells: the architects of skeletal muscle. Curr Top Dev Biol 2014;107:161–81. http://dx. doi.org/10.1016/B978-0-12-416022-4.00006-8.

Review: Myogenesis in the Genomics Era

[6] Mauro A. Satellite cell of skeletal muscle fibers. J Biophys Biochem Cytol 1961;9:493–5. [7] White RB, Bierinx AS, Gnocchi VF, Zammit PS. Dynamics of muscle fibre growth during postnatal mouse development. BMC Dev Biol 2010;10. http://dx.doi.org/10.1186/1471-213X10-21. [8] Shen X, Collier JM, Hlaing M, Zhang L, Delshad EH, Bristow J, et al. Genome-wide examination of myoblast cell cycle withdrawal during differentiation. Dev Dyn 2003;226:128–38. http://dx.doi.org/10.1002/dvdy.10200. [9] Dilworth FJ, Blais A. Epigenetic regulation of satellite cell activation during muscle regeneration. Stem Cell Res Ther 2011;2:18. http://dx.doi.org/10.1186/scrt59. [10] Berger SL. The complex language of chromatin regulation during transcription. Nature 2007;447:407–12. http://dx.doi. org/10.1038/nature05915. [11] Berkes CA, Tapscott SJ. MyoD and the transcriptional control of myogenesis. Semin Cell Dev Biol 2005;16:585–95. [12] Davis RL, Weintraub H, Lassar AB. Expression of a single transfected cDNA converts fibroblasts to myoblasts. Cell 1987;51:987–1000. [13] Tapscott SJ, Davis RL, Thayer MJ, Cheng PF, Weintraub H, Lassar AB. MyoD1: a nuclear phosphoprotein requiring a Myc homology region to convert fibroblasts to myoblasts. Science 1988;242:405–11. [14] Liu Y, Chu A, Chakroun I, Islam U, Blais A. Cooperation between myogenic regulatory factors and SIX family transcription factors is important for myoblast differentiation. Nucleic Acids Res 2010;38:6857–71. http://dx.doi.org/10. 1093/nar/gkq585. [15] 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 2004; 14:465–77. [16] Edmondson DG, Cheng TC, Cserjesi P, Chakraborty T, Olson EN. Analysis of the myogenin promoter reveals an indirect pathway for positive autoregulation mediated by the muscle-specific enhancer factor MEF-2. Mol Cell Biol 1992; 12:3665–77. [17] Yu YT, Breitbart RE, Smoot LB, Lee Y, Mahdavi V, NadalGinard B. Human myocyte-specific enhancer factor 2 comprises a group of tissue-restricted MADS box transcription factors. Genes Dev 1992;6:1783–98. [18] Daou N, Lecolle S, Lefebvre S, della Gaspera B, Charbonnier F, Chanoine C, et al. A new role for the calcineurin/NFAT pathway in neonatal myosin heavy chain expression via the NFATc2/MyoD complex during mouse myogenesis. Development 2013;140:4914–25. http://dx.doi.org/10.1242/dev. 097428. [19] Armand AS, Bourajjaj M, Martinez-Martinez S, el Azzouzi H, da Costa Martins PA, Hatzis P, et al. Cooperative synergy between NFAT and MyoD regulates myogenin expression and myogenesis. J Biol Chem 2008;283:29004–10. http://dx. doi.org/10.1074/jbc.M801297200. [20] Buckingham M, Rigby PW. Gene regulatory networks and transcriptional mechanisms that control myogenesis. Dev Cell 2014;28:225–38. http://dx.doi.org/10.1016/j.devcel.2013.12. 020. [21] Tajbakhsh S, Rocancourt D, Cossu G, Buckingham M. Redefining the genetic hierarchies controlling skeletal myogenesis: Pax-3 and Myf-5 act upstream of MyoD. Cell 1997;89:127–38.

2033 [22] Relaix F, Montarras D, Zaffran S, Gayraud-Morel B, Rocancourt D, Tajbakhsh S, et al. Pax3 and Pax7 have distinct and overlapping functions in adult muscle progenitor cells. J Cell Biol 2006;172:91–102. http://dx.doi.org/10.1083/jcb. 200508044. [23] L'honore A, Ouimette JF, Lavertu-Jolin M, Drouin J. Pitx2 defines alternate pathways acting through MyoD during limb and somitic myogenesis. Development 2010;137:3847–56. http://dx.doi.org/10.1242/dev.053421. [24] Messina G, Biressi S, Monteverde S, Magli A, Cassano M, Perani L, et al. Nfix regulates fetal-specific transcription in developing skeletal muscle. Cell 2010;140:554–66. http://dx. doi.org/10.1016/j.cell.2010.01.027. [25] Seale P, Sabourin LA, Girgis-Gabardo A, Mansouri A, Gruss P, Rudnicki MA. Pax7 is required for the specification of myogenic satellite cells. Cell 2000;102:777–86. [26] Sambasivan R, Yao R, Kissenpfennig A, Van Wittenberghe L, Paldi A, Gayraud-Morel B, et al. Pax7-expressing satellite cells are indispensable for adult skeletal muscle regeneration. Development 2011;138:3647–56. http://dx.doi.org/10.1242/ dev.067587. [27] Lepper C, Partridge TA, Fan CM. An absolute requirement for Pax7-positive satellite cells in acute injury-induced skeletal muscle regeneration. Development 2011;138: 3639–46. http://dx.doi.org/10.1242/dev.067595. [28] 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 2013;110:16474–9. http://dx. doi.org/10.1073/pnas.1307680110. [29] 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 2013;13:590–601. http://dx.doi.org/10.1016/j.stem.2013. 07.016. [30] Ryan T, Liu J, Chu A, Wang L, Blais A, Skerjanc IS. Retinoic acid enhances skeletal myogenesis in human embryonic stem cells by expanding the premyogenic progenitor population. Stem Cell Rev 2012;8:482–93. http://dx.doi.org/10.1007/ s12015-011-9284-0. [31] Shelton M, Metz J, Liu J, Carpenedo RL, Demers SP, Stanford WL, et al. Derivation and expansion of PAX7positive muscle progenitors from human and mouse embryonic stem cells. Stem Cell Rep 2014;3:516–29. http://dx.doi. org/10.1016/j.stemcr.2014.07.001. [32] Voronova A, Coyne E, Al Madhoun A, Fair JV, Bosiljcic N, St-Louis C, et al. Hedgehog signaling regulates MyoD expression and activity. J Biol Chem 2013;288:4389–404. http://dx.doi.org/10.1074/jbc.M112.400184. [33] Wong J, Mehta V, Voronova A, Coutu J, Ryan T, Shelton M, et al. beta-Catenin is essential for efficient in vitro premyogenic mesoderm formation but can be partially compensated by retinoic acid signalling. PLoS One 2013; 8:e57501. http://dx.doi.org/10.1371/journal.pone. 0057501. [34] Darabi R, Perlingeiro RC. Derivation of skeletal myogenic precursors from human pluripotent stem cells using conditional expression of PAX7. Methods Mol Biol 2014. http://dx. doi.org/10.1007/7651_2014_134. [35] Magli A, Schnettler E, Swanson SA, Borges L, Hoffman K, Stewart R, et al. Pax3 and Tbx5 specify whether PDGFRalpha + cells assume skeletal or cardiac muscle fate in differentiating embryonic stem cells. Stem Cells 2014;32: 2072–83. http://dx.doi.org/10.1002/stem.1713.

2034 [36] Hickey MS, Carey JO, Azevedo JL, Houmard JA, Pories WJ, Israel RG, et al. Skeletal muscle fiber composition is related to adiposity and in vitro glucose transport rate in humans. Am J Physiol 1995;268:E453–7. [37] Kelley DE, He J, Menshikova EV, Ritov VB. Dysfunction of mitochondria in human skeletal muscle in type 2 diabetes. Diabetes 2002;51:2944–50. [38] Gerrits MF, Ghosh S, Kavaslar N, Hill B, Tour A, Seifert EL, et al. Distinct skeletal muscle fiber characteristics and gene expression in diet-sensitive versus diet-resistant obesity. J Lipid Res 2010;51:2394–404. http://dx.doi.org/10.1194/jlr. P005298. [39] Silva JE. Thyroid hormone control of thermogenesis and energy balance. Thyroid 1995;5:481–92. [40] Schena M, Shalon D, Davis RW, Brown PO. Quantitative monitoring of gene expression patterns with a complementary DNA microarray. Science 1995;270:467–70. [41] Ideker T, Galitski T, Hood L. A new approach to decoding life: systems biology. Annu Rev Genomics Hum Genet 2001;2:343–72. http://dx.doi.org/10.1146/annurev.genom. 2.1.343. [42] Yaffe D, Saxel O. Serial passaging and differentiation of myogenic cells isolated from dystrophic mouse muscle. Nature 1977;270:725–7. [43] Blau HM, Chiu CP, Webster C. Cytoplasmic activation of human nuclear genes in stable heterocaryons. Cell 1983; 32:1171–80. [44] Pajcini KV, Corbel SY, Sage J, Pomerantz JH, Blau HM. Transient inactivation of Rb and ARF yields regenerative cells from postmitotic mammalian muscle. Cell Stem Cell 2010;7: 198–213. http://dx.doi.org/10.1016/j.stem.2010.05.022. [45] Moran JL, Li Y, Hill AA, Mounts WM, Miller CP. Gene expression changes during mouse skeletal myoblast differentiation revealed by transcriptional profiling. Physiol Genomics 2002;10:103–11. [46] Yan Z, Choi S, Liu X, Zhang M, Schageman JJ, Lee SY, et al. Highly coordinated gene regulation in mouse skeletal muscle regeneration. J Biol Chem 2003;278:8826–36. http://dx.doi.org/10.1074/jbc.M209879200. [47] Liu L, Cheung TH, Charville GW, Hurgo BM, Leavitt T, Shih J, et al. Chromatin modifications as determinants of muscle stem cell quiescence and chronological aging. Cell Rep 2013;4:189–204. http://dx.doi.org/10.1016/j.celrep.2013. 05.043. [48] Mortazavi A, Williams BA, McCue K, Schaeffer L, Wold B. Mapping and quantifying mammalian transcriptomes by RNA-Seq. Nat Methods 2008;5:621–8. http://dx.doi.org/10. 1038/nmeth.1226. [49] Trapnell C, Williams BA, Pertea G, Mortazavi A, Kwan G, van Baren MJ, et al. Transcript assembly and quantification by RNA-Seq reveals unannotated transcripts and isoform switching during cell differentiation. Nat Biotechnol 2010;28: 511–5. http://dx.doi.org/10.1038/nbt.1621. [50] Sebastian S, Faralli H, Yao Z, Rakopoulos P, Palii C, Cao Y, et al. Tissue-specific splicing of a ubiquitously expressed transcription factor is essential for muscle differentiation. Genes Dev 2013;27:1247–59. http://dx.doi.org/10.1101/ gad.215400.113. [51] Mousavi K, Zare H, Dell'orso S, Grontved L, Gutierrez-Cruz G, Derfoul A, et al. eRNAs promote transcription by establishing chromatin accessibility at defined genomic loci. Mol Cell 2013;51:606–17. http://dx.doi.org/10.1016/j.molcel. 2013.07.022.

Review: Myogenesis in the Genomics Era

[52] Mueller AC, Cichewicz MA, Dey BK, Layer R, Reon BJ, Gagan JR, et al. MUNC, a long noncoding RNA that facilitates the function of MyoD in skeletal myogenesis. Mol Cell Biol 2015;35:498–513. http://dx.doi.org/10.1128/MCB. 01079-14. [53] Sohi G, Dilworth FJ. Non-coding RNAs as epigenetic mediators of skeletal muscle regeneration. FEBS J 2014. http://dx.doi.org/10.1111/febs.13170. [54] Ballarino M, Cazzella V, D'Andrea D, Grassi L, Bisceglie L, Cipriano A, et al. Novel long noncoding RNAs (lncRNAs) in myogenesis: a miR-31 overlapping lncRNA transcript controls myoblast differentiation. Mol Cell Biol 2015;35: 728–36. http://dx.doi.org/10.1128/MCB.01394-14. [55] Pallafacchina G, Francois S, Regnault B, Czarny B, Dive V, Cumano A, et al. An adult tissue-specific stem cell in its niche: a gene profiling analysis of in vivo quiescent and activated muscle satellite cells. Stem Cell Res 2009;4:77–91. [56] Fukada S, Uezumi A, Ikemoto M, Masuda S, Segawa M, Tanimura N, et al. Molecular signature of quiescent satellite cells in adult skeletal muscle. Stem Cells 2007;25:2448–59. http://dx.doi.org/10.1634/stemcells.2007-0019. [57] Segal E, Shapira M, Regev A, Pe'er D, Botstein D, Koller D, et al. Module networks: identifying regulatory modules and their condition-specific regulators from gene expression data. Nat Genet 2003;34:166–76. [58] Wang C, Xuan J, Chen L, Zhao P, Wang Y, Clarke R, et al. Motif-directed network component analysis for regulatory network inference. BMC Bioinforma 2008;9. http://dx.doi. org/10.1186/1471-2105-9-S1-S21. [59] Gong T, Xuan J, Wang C, Li H, Hoffman E, Clarke R, et al. Gene module identification from microarray data using nonnegative independent component analysis. Gene Regul Syst Biol 2008;1:349–63. [60] Kim TM, Chung YJ, Rhyu MG, Jung MH. Inferring biological functions and associated transcriptional regulators using gene set expression coherence analysis. BMC Bioinforma 2007;8:453. http://dx.doi.org/10.1186/14712105-8-453. [61] Perez-Montarelo D, Hudson NJ, Fernandez AI, RamayoCaldas Y, Dalrymple BP, Reverter A. Porcine tissue-specific regulatory networks derived from meta-analysis of the transcriptome. PLoS One 2012;7:e46159. http://dx.doi.org/10. 1371/journal.pone.0046159. [62] Hudson NJ, Reverter A, Wang Y, Greenwood PL, Dalrymple BP. Inferring the transcriptional landscape of bovine skeletal muscle by integrating co-expression networks. PLoS One 2009;4:e7249. http://dx.doi.org/10.1371/journal.pone. 0007249. [63] Bergstrom DA, Penn BH, Strand A, Perry RL, Rudnicki MA, Tapscott SJ. Promoter-specific regulation of MyoD binding and signal transduction cooperate to pattern gene expression. Mol Cell 2002;9:587–600. [64] Niro C, Demignon J, Vincent S, Liu Y, Giordani J, Sgarioto N, et al. Six1 and Six4 gene expression is necessary to activate the fast-type muscle gene program in the mouse primary myotome. Dev Biol 2009;338:168–82. [65] Potthoff MJ, Arnold MA, McAnally J, Richardson JA, BasselDuby R, Olson EN. Regulation of skeletal muscle sarcomere integrity and postnatal muscle function by Mef2c. Mol Cell Biol 2007;27:8143–51. http://dx.doi.org/10.1128/MCB. 01187-07. [66] Wang X, Blagden C, Fan J, Nowak SJ, Taniuchi I, Littman DR, et al. Runx1 prevents wasting, myofibrillar

Review: Myogenesis in the Genomics Era

[67]

[68]

[69]

[70]

[71]

[72]

[73]

[74]

[75]

[76]

[77]

[78]

[79]

[80]

[81]

disorganization, and autophagy of skeletal muscle. Genes Dev 2005;19:1715–22. http://dx.doi.org/10.1101/gad.1318305. Lagha M, Sato T, Regnault B, Cumano A, Zuniga A, Licht J, et al. Transcriptome analyses based on genetic screens for Pax3 myogenic targets in the mouse embryo. BMC Genomics 2010;11. http://dx.doi.org/10.1186/1471-2164-11-696. Judson RN, Tremblay AM, Knopp P, White RB, Urcia R, De Bari C, et al. The Hippo pathway member Yap plays a key role in influencing fate decisions in muscle satellite cells. J Cell Sci 2012;125:6009–19. http://dx.doi.org/10.1242/jcs. 109546. Ren B, Robert F, Wyrick JJ, Aparicio O, Jennings EG, Simon I, et al. Genome-wide location and function of DNA binding proteins. Science 2000;290:2306–9. Johnson DS, Mortazavi A, Myers RM, Wold B. Genomewide mapping of in vivo protein–DNA interactions. Science 2007;316:1497–502. http://dx.doi.org/10.1126/science. 1141319. Impey S, McCorkle SR, Cha-Molstad H, Dwyer JM, Yochum GS, Boss JM, et al. Defining the CREB regulon: a genome-wide analysis of transcription factor regulatory regions. Cell 2004; 119:1041–54. http://dx.doi.org/10.1016/j.cell.2004.10.032. Blais A, Tsikitis M, Acosta-Alvear D, Sharan R, Kluger Y, Dynlacht BD. An initial blueprint for myogenic differentiation. Genes Dev 2005;19:553–69. Cao Y, Kumar RM, Penn BH, Berkes CA, Kooperberg C, Boyer LA, et al. Global and gene-specific analyses show distinct roles for Myod and Myog at a common set of promoters. EMBO J 2006;25:502–11. Wasserman WW, Sandelin A. Applied bioinformatics for the identification of regulatory elements. Nat Rev Genet 2004;5: 276–87. http://dx.doi.org/10.1038/nrg1315. Cao Y, Yao Z, Sarkar D, Lawrence M, Sanchez GJ, Parker MH, et al. Genome-wide MyoD binding in skeletal muscle cells: a potential for broad cellular reprogramming. Dev Cell 2010;18:662–74. Seenundun S, Rampalli S, Liu QC, Aziz A, Palii C, Hong S, et al. UTX mediates demethylation of H3K27me3 at musclespecific genes during myogenesis. EMBO J 2010;29: 1401–11. Liu Y, Chakroun I, Yang D, Horner E, Liang J, Aziz A, et al. Six1 regulates MyoD expression in adult muscle progenitor cells. PLoS One 2013;8:e67762. http://dx.doi.org/10.1371/ journal.pone.0067762. Rampalli S, Li L, Mak E, Ge K, Brand M, Tapscott SJ, et al. p38 MAPK signaling regulates recruitment of Ash2L-containing methyltransferase complexes to specific genes during differentiation. Nat Struct Mol Biol 2007;14:1150–6. http://dx.doi. org/10.1038/nsmb1316. Benhaddou A, Keime C, Ye T, Morlon A, Michel I, Jost B, et al. Transcription factor TEAD4 regulates expression of myogenin and the unfolded protein response genes during C2C12 cell differentiation. Cell Death Differ 2012;19: 220–31. http://dx.doi.org/10.1038/cdd.2011.87. Maeda T, Chapman DL, Stewart AF. Mammalian vestigiallike 2, a cofactor of TEF-1 and MEF2 transcription factors that promotes skeletal muscle differentiation. J Biol Chem 2002;277:48889–98. http://dx.doi.org/10.1074/jbc. M206858200. Soleimani VD, Yin H, Jahani-Asl A, Ming H, Kockx CE, van Ijcken WF, et al. Snail regulates MyoD binding-site occupancy to direct enhancer switching and differentiation-specific transcription in myogenesis. Mol Cell 2012;47:457–68. http://dx.doi.org/10.1016/j.molcel.2012.05.046.

2035 [82] Kuang S, Charge SB, Seale P, Huh M, Rudnicki MA. Distinct roles for Pax7 and Pax3 in adult regenerative myogenesis. J Cell Biol 2006;172:103–13. [83] Soleimani VD, Punch VG, Kawabe Y, Jones AE, Palidwor GA, Porter CJ, et al. Transcriptional dominance of Pax7 in adult myogenesis is due to high-affinity recognition of homeodomain motifs. Dev Cell 2012;22:1208–20. http://dx.doi.org/ 10.1016/j.devcel.2012.03.014. [84] Asp P, Blum R, Vethantham V, Parisi F, Micsinai M, Cheng J, et al. Genome-wide remodeling of the epigenetic landscape during myogenic differentiation. Proc Natl Acad Sci U S A 2011;108:E149–58. http://dx.doi.org/10.1073/pnas. 1102223108. [85] Vethantham V, Yang Y, Bowman C, Asp P, Lee JH, Skalnik DG, et al. Dynamic loss of H2B ubiquitylation without corresponding changes in H3K4 trimethylation during myogenic differentiation. Mol Cell Biol 2012;32:1044–55. http://dx.doi. org/10.1128/MCB.06026-11. [86] Blum R, Vethantham V, Bowman C, Rudnicki M, Dynlacht BD. Genome-wide identification of enhancers in skeletal muscle: the role of MyoD1. Genes Dev 2012;26:2763–79. http://dx.doi.org/10.1101/gad.200113.112. [87] Cheng J, Blum R, Bowman C, Hu D, Shilatifard A, Shen S, et al. A role for H3K4 monomethylation in gene repression and partitioning of chromatin readers. Mol Cell 2014;53: 979–92. http://dx.doi.org/10.1016/j.molcel.2014.02.032. [88] Tao Y, Neppl RL, Huang ZP, Chen J, Tang RH, Cao R, et al. The histone methyltransferase Set7/9 promotes myoblast differentiation and myofibril assembly. J Cell Biol 2011;194: 551–65. http://dx.doi.org/10.1083/jcb.201010090. [89] Lee JE, Wang C, Xu S, Cho YW, Wang L, Feng X, et al. H3K4 mono- and di-methyltransferase MLL4 is required for enhancer activation during cell differentiation. Elife 2013;2: e01503. http://dx.doi.org/10.7554/eLife.01503. [90] 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 2004;18:2627–38. [91] Lu L, Sun K, Chen X, Zhao Y, Wang L, Zhou L, et al. Genome-wide survey by ChIP-seq reveals YY1 regulation of lincRNAs in skeletal myogenesis. EMBO J 2013;32: 2575–88. http://dx.doi.org/10.1038/emboj.2013.182. [92] Wang AH, Zare H, Mousavi K, Wang C, Moravec CE, Sirotkin HI, et al. The histone chaperone Spt6 coordinates histone H3K27 demethylation and myogenesis. EMBO J 2013;32:1075–86. http://dx.doi.org/10.1038/emboj. 2013.54. [93] Lan F, Bayliss PE, Rinn JL, Whetstine JR, Wang JK, Chen S, et al. A histone H3 lysine 27 demethylase regulates animal posterior development. Nature 2007;449:689–94. [94] Puri PL, Mercola M. BAF60 A, B, and Cs of muscle determination and renewal. Genes Dev 2012;26:2673–83. http://dx.doi.org/10.1101/gad.207415.112. [95] Hargreaves DC, Crabtree GR. ATP-dependent chromatin remodeling: genetics, genomics and mechanisms. Cell Res 2011;21:396–420. http://dx.doi.org/10.1038/cr.2011.32. [96] Lessard JA, Crabtree GR. Chromatin regulatory mechanisms in pluripotency. Annu Rev Cell Dev Biol 2010;26: 503–32. http://dx.doi.org/10.1146/annurev-cellbio-051809102012. [97] de la Serna IL, Ohkawa Y, Berkes CA, Bergstrom DA, Dacwag CS, Tapscott SJ, et al. MyoD targets chromatin remodeling complexes to the myogenin locus prior to forming a stable DNA-bound complex. Mol Cell Biol 2005;

2036

[98]

[99]

[100]

[101]

[102]

[103]

[104]

[105]

[106]

[107]

[108]

[109]

[110]

[111]

[112]

25:3997–4009. http://dx.doi.org/10.1128/MCB.25.10.39974009.2005. Forcales SV, Albini S, Giordani L, Malecova B, Cignolo L, Chernov A, et al. Signal-dependent incorporation of MyoDBAF60c into Brg1-based SWI/SNF chromatin-remodelling complex. EMBO J 2012;31:301–16. http://dx.doi.org/10. 1038/emboj.2011.391. Boyle AP, Davis S, Shulha HP, Meltzer P, Margulies EH, Weng Z, et al. High-resolution mapping and characterization of open chromatin across the genome. Cell 2008;132: 311–22. http://dx.doi.org/10.1016/j.cell.2007.12.014. Gaulton KJ, Nammo T, Pasquali L, Simon JM, Giresi PG, Fogarty MP, et al. A map of open chromatin in human pancreatic islets. Nat Genet 2010;42:255–9. http://dx.doi.org/ 10.1038/ng.530. Buenrostro JD, Giresi PG, Zaba LC, Chang HY, Greenleaf WJ. Transposition of native chromatin for fast and sensitive epigenomic profiling of open chromatin, DNA-binding proteins and nucleosome position. Nat Methods 2013;10: 1213–8. http://dx.doi.org/10.1038/nmeth.2688. Jin C, Zang C, Wei G, Cui K, Peng W, Zhao K, et al. H3.3/ H2A.Z double variant-containing nucleosomes mark “nucleosome-free regions” of active promoters and other regulatory regions. Nat Genet 2009;41:941–5. http://dx.doi. org/10.1038/ng.409. Chen P, Zhao J, Wang Y, Wang M, Long H, Liang D, et al. H3.3 actively marks enhancers and primes gene transcription via opening higher-ordered chromatin. Genes Dev 2013;27: 2109–24. http://dx.doi.org/10.1101/gad.222174.113. Harada A, Okada S, Konno D, Odawara J, Yoshimi T, Yoshimura S, et al. Chd2 interacts with H3.3 to determine myogenic cell fate. EMBO J 2012;31:2994–3007. http://dx. doi.org/10.1038/emboj.2012.136. Weber CM, Ramachandran S, Henikoff S. Nucleosomes are context-specific, H2A.Z-modulated barriers to RNA polymerase. Mol Cell 2014;53:819–30. http://dx.doi.org/10. 1016/j.molcel.2014.02.014. Gevry N, Chan HM, Laflamme L, Livingston DM, Gaudreau L. p21 transcription is regulated by differential localization of histone H2A.Z. Genes Dev 2007;21:1869–81. http://dx.doi. org/10.1101/gad.1545707. Wong MM, Cox LK, Chrivia JC. The chromatin remodeling protein, SRCAP, is critical for deposition of the histone variant H2A.Z at promoters. J Biol Chem 2007;282: 26132–9. http://dx.doi.org/10.1074/jbc.M703418200. Cuadrado A, Corrado N, Perdiguero E, Lafarga V, MunozCanoves P, Nebreda AR. Essential role of p18Hamlet/ SRCAP-mediated histone H2A.Z chromatin incorporation in muscle differentiation. EMBO J 2010;29:2014–25. http://dx. doi.org/10.1038/emboj.2010.85. Zinzen RP, Girardot C, Gagneur J, Braun M, Furlong EE. Combinatorial binding predicts spatio-temporal cis-regulatory activity. Nature 2009;462:65–70. Rhee DY, Cho DY, Zhai B, Slattery M, Ma L, Mintseris J, et al. Transcription factor networks in Drosophila melanogaster. Cell Rep 2014;8:2031–43. http://dx.doi.org/10.1016/ j.celrep.2014.08.038. Junion G, Spivakov M, Girardot C, Braun M, Gustafson EH, Birney E, et al. A transcription factor collective defines cardiac cell fate and reflects lineage history. Cell 2012;148: 473–86. http://dx.doi.org/10.1016/j.cell.2012.01.030. Wilczynski B, Liu YH, Yeo ZX, Furlong EE. Predicting spatial and temporal gene expression using an integrative model of transcription factor occupancy and chromatin

Review: Myogenesis in the Genomics Era

[113]

[114]

[115]

[116]

[117]

[118]

[119]

[120] [121]

[122]

[123]

[124]

[125]

[126]

[127]

[128]

state. PLoS Comput Biol 2012;8:e1002798. http://dx.doi. org/10.1371/journal.pcbi.1002798. Marbach D, Roy S, Ay F, Meyer PE, Candeias R, Kahveci T, et al. Predictive regulatory models in Drosophila melanogaster by integrative inference of transcriptional networks. Genome Res 2012;22:1334–49. http://dx.doi.org/10.1101/ gr.127191.111. modENCODE ConsortiumRoy S, Ernst J, Kharchenko PV, Kheradpour P, Negre N, et al. Identification of functional elements and regulatory circuits by Drosophila modENCODE. Science 2010;330:1787–97. http://dx.doi.org/10. 1126/science.1198374. Gerstein MB, Lu ZJ, Van Nostrand EL, Cheng C, Arshinoff BI, Liu T, et al. Integrative analysis of the Caenorhabditis elegans genome by the modENCODE project. Science 2010;330:1775–87. http://dx.doi.org/10.1126/science. 1196914. Van Nostrand EL, Kim SK. Integrative analysis of C. elegans modENCODE ChIP-seq data sets to infer gene regulatory interactions. Genome Res 2013;23:941–53. http://dx.doi.org/10.1101/gr.152876.112. Chen H, VanBuren V. A provisional gene regulatory atlas for mouse heart development. PLoS One 2014;9:e83364. http://dx.doi.org/10.1371/journal.pone.0083364. Sasai M, Kawabata Y, Makishi K, Itoh K, Terada TP. Time scales in epigenetic dynamics and phenotypic heterogeneity of embryonic stem cells. PLoS Comput Biol 2013;9: e1003380. http://dx.doi.org/10.1371/journal.pcbi.1003380. Walker E, Ohishi M, Davey RE, Zhang W, Cassar PA, Tanaka TS, et al. Prediction and testing of novel transcriptional networks regulating embryonic stem cell self-renewal and commitment. Cell Stem Cell 2007;1:71–86. http://dx.doi.org/ 10.1016/j.stem.2007.04.002. Burden SJ. Building the vertebrate neuromuscular synapse. J Neurobiol 2002;53:501–11. Bentzinger CF, Wang YX, von Maltzahn J, Rudnicki MA. The emerging biology of muscle stem cells: implications for cell-based therapies. Bioessays 2013;35:231–41. http://dx. doi.org/10.1002/bies.201200063. Rando TA, Blau HM. Primary mouse myoblast purification, characterization, and transplantation for cell-mediated gene therapy. J Cell Biol 1994;125:1275–87. Bentzinger CF, Wang YX, Dumont NA, Rudnicki MA. Cellular dynamics in the muscle satellite cell niche. EMBO Rep 2013;14:1062–72. http://dx.doi.org/10.1038/embor. 2013.182. Biressi S, Rando TA. Heterogeneity in the muscle satellite cell population. Semin Cell Dev Biol 2010;21:845–54. http:// dx.doi.org/10.1016/j.semcdb.2010.09.003. Motohashi N, Asakura A. Muscle satellite cell heterogeneity and self-renewal. Front Cell Dev Biol 2014;2:1. http://dx.doi. org/10.3389/fcell.2014.00001. Trapnell C, Cacchiarelli D, Grimsby J, Pokharel P, Li S, Morse M, et al. The dynamics and regulators of cell fate decisions are revealed by pseudotemporal ordering of single cells. Nat Biotechnol 2014;32:381–6. http://dx.doi.org/10.1038/nbt. 2859. Picelli S, Bjorklund AK, Faridani OR, Sagasser S, Winberg G, Sandberg R. Smart-seq2 for sensitive full-length transcriptome profiling in single cells. Nat Methods 2013; 10:1096–8. http://dx.doi.org/10.1038/nmeth.2639. Shalek AK, Satija R, Adiconis X, Gertner RS, Gaublomme JT, Raychowdhury R, et al. Single-cell transcriptomics reveals bimodality in expression and splicing in immune

Review: Myogenesis in the Genomics Era

[129]

[130]

[131]

[132]

[133]

[134]

[135]

[136]

[137]

[138]

[139]

[140]

[141]

[142]

cells. Nature 2013;498:236–40. http://dx.doi.org/10.1038/ nature12172. Hough SR, Thornton M, Mason E, Mar JC, Wells CA, Pera MF. Single-cell gene expression profiles define self-renewing, pluripotent, and lineage primed states of human pluripotent stem cells. Stem Cell Rep 2014;2:881–95. http://dx.doi.org/ 10.1016/j.stemcr.2014.04.014. Lee JH, Daugharthy ER, Scheiman J, Kalhor R, Yang JL, Ferrante TC, et al. Highly multiplexed subcellular RNA sequencing in situ. Science 2014;343:1360–3. http://dx.doi. org/10.1126/science.1250212. Blais A, van Oevelen CJ, Margueron R, Acosta-Alvear D, Dynlacht BD. Retinoblastoma tumor suppressor proteindependent methylation of histone H3 lysine 27 is associated with irreversible cell cycle exit. J Cell Biol 2007;179: 1399–412. Mullen AC, Orlando DA, Newman JJ, Loven J, Kumar RM, Bilodeau S, et al. Master transcription factors determine celltype-specific responses to TGF-beta signaling. Cell 2011; 147:565–76. http://dx.doi.org/10.1016/j.cell.2011.08.050. Yao Z, Fong AP, Cao Y, Ruzzo WL, Gentleman RC, Tapscott SJ. Comparison of endogenous and overexpressed MyoD shows enhanced binding of physiologically bound sites. Skelet Muscle 2013;3. http://dx.doi.org/10.1186/2044-50403-8. Fong AP, Yao Z, Zhong JW, Cao Y, Ruzzo WL, Gentleman RC, et al. Genetic and epigenetic determinants of neurogenesis and myogenesis. Dev Cell 2012;22:721–35. http://dx.doi. org/10.1016/j.devcel.2012.01.015. Wang J, Kumar RM, Biggs VJ, Lee H, Chen Y, Kagey MH, et al. The Msx1 homeoprotein recruits polycomb to the nuclear periphery during development. Dev Cell 2011;21: 575–88. http://dx.doi.org/10.1016/j.devcel.2011.07.003. Asp P, Acosta-Alvear D, Tsikitis M, van Oevelen C, Dynlacht BD. E2f3b plays an essential role in myogenic differentiation through isoform-specific gene regulation. Genes Dev 2009; 23:37–53. http://dx.doi.org/10.1101/gad.1727309. MacQuarrie KL, Yao Z, Fong AP, Diede SJ, Rudzinski ER, Hawkins DS, et al. Comparison of genome-wide binding of MyoD in normal human myogenic cells and rhabdomyosarcomas identifies regional and local suppression of promyogenic transcription factors. Mol Cell Biol 2013;33:773–84. http://dx. doi.org/10.1128/MCB.00916-12. Macquarrie KL, Yao Z, Fong AP, Tapscott SJ. Genome-wide binding of the basic helix–loop–helix myogenic inhibitor musculin has substantial overlap with MyoD: implications for buffering activity. Skelet Muscle 2013;3. http://dx.doi.org/10. 1186/2044-5040-3-26. Mousavi K, Zare H, Wang AH, Sartorelli V. Polycomb protein Ezh1 promotes RNA polymerase II elongation. Mol Cell 2012;45:255–62. http://dx.doi.org/10.1016/j.molcel. 2011.11.019. van Oevelen C, Bowman C, Pellegrino J, Asp P, Cheng J, Parisi F, et al. The mammalian Sin3 proteins are required for muscle development and sarcomere specification. Mol Cell Biol 2010. http://dx.doi.org/10.1128/MCB.00975-10. Gayraud-Morel B, Chretien F, Jory A, Sambasivan R, Negroni E, Flamant P, et al. Myf5 haploinsufficiency reveals distinct cell fate potentials for adult skeletal muscle stem cells. J Cell Sci 2012;125:1738–49. http://dx.doi.org/10.1242/jcs.097006. Montarras D, Morgan J, Collins C, Relaix F, Zaffran S, Cumano A, et al. Direct isolation of satellite cells for skeletal muscle regeneration. Science 2005;309:2064–7.

2037 [143] Sambasivan R, Gayraud-Morel B, Dumas G, Cimper C, Paisant S, Kelly RG, et al. Distinct regulatory cascades govern extraocular and pharyngeal arch muscle progenitor cell fates. Dev Cell 2009;16:810–21. http://dx.doi.org/10. 1016/j.devcel.2009.05.008. [144] Bosnakovski D, Xu Z, Li W, Thet S, Cleaver O, Perlingeiro RC, et al. Prospective isolation of skeletal muscle stem cells with a Pax7 reporter. Stem Cells 2008;26:3194–204. http://dx.doi. org/10.1634/stemcells.2007-1017. [145] Moon TS, Lou C, Tamsir A, Stanton BC, Voigt CA. Genetic programs constructed from layered logic gates in single cells. Nature 2012;491:249–53. http://dx.doi.org/10.1038/ nature11516. [146] Cabantous S, Terwilliger TC, Waldo GS. Protein tagging and detection with engineered self-assembling fragments of green fluorescent protein. Nat Biotechnol 2005;23:102–7. http://dx.doi.org/10.1038/nbt1044. [147] Bonn S, Zinzen RP, Perez-Gonzalez A, Riddell A, Gavin AC, Furlong EE. Cell type-specific chromatin immunoprecipitation from multicellular complex samples using BiTS-ChIP. Nat Protoc 2012;7:978–94. http://dx.doi.org/10.1038/nprot.2012. 049. [148] Liu S, Brind'Amour J, Karimi MM, Shirane K, Bogutz A, Lefebvre L, et al. Setdb1 is required for germline development and silencing of H3K9me3-marked endogenous retroviruses in primordial germ cells. Genes Dev 2014;28:2041–55. http:// dx.doi.org/10.1101/gad.244848.114. [149] Lae M, Ahn EH, Mercado GE, Chuai S, Edgar M, Pawel BR, et al. Global gene expression profiling of PAX-FKHR fusionpositive alveolar and PAX-FKHR fusion-negative embryonal rhabdomyosarcomas. J Pathol 2007;212:143–51. http://dx. doi.org/10.1002/path.2170. [150] De Pitta C, Tombolan L, Albiero G, Sartori F, Romualdi C, Jurman G, et al. Gene expression profiling identifies potential relevant genes in alveolar rhabdomyosarcoma pathogenesis and discriminates PAX3-FKHR positive and negative tumors. Int J Cancer 2006;118:2772–81. http://dx. doi.org/10.1002/ijc.21698. [151] Khan J, Simon R, Bittner M, Chen Y, Leighton SB, Pohida T, et al. Gene expression profiling of alveolar rhabdomyosarcoma with cDNA microarrays. Cancer Res 1998;58:5009–13. [152] Krom YD, Thijssen PE, Young JM, den Hamer B, Balog J, Yao Z, et al. Intrinsic epigenetic regulation of the D4Z4 macrosatellite repeat in a transgenic mouse model for FSHD. PLoS Genet 2013;9:e1003415. http://dx.doi.org/10.1371/journal. pgen.1003415. [153] Young JM, Whiddon JL, Yao Z, Kasinathan B, Snider L, Geng LN, et al. DUX4 binding to retroelements creates promoters that are active in FSHD muscle and testis. PLoS Genet 2013;9:e1003947. http://dx.doi.org/10.1371/journal. pgen.1003947. [154] Xu X, Tsumagari K, Sowden J, Tawil R, Boyle AP, Song L, et al. DNaseI hypersensitivity at gene-poor, FSH dystrophy-linked 4q35.2. Nucleic Acids Res 2009;37:7381–93. http://dx.doi.org/ 10.1093/nar/gkp833. [155] Haslett JN, Sanoudou D, Kho AT, Bennett RR, Greenberg SA, Kohane IS, et al. Gene expression comparison of biopsies from Duchenne muscular dystrophy (DMD) and normal skeletal muscle. Proc Natl Acad Sci U S A 2002;99: 15000–5. http://dx.doi.org/10.1073/pnas.192571199. [156] Bakay M, Zhao P, Chen J, Hoffman EP. A web-accessible complete transcriptome of normal human and DMD muscle. Neuromuscul Disord 2002;12:S125–41.

2038 [157] Saccone V, Consalvi S, Giordani L, Mozzetta C, Barozzi I, Sandona M, et al. HDAC-regulated myomiRs control BAF60 variant exchange and direct the functional phenotype of fibro-adipogenic progenitors in dystrophic muscles. Genes Dev 2014;28:841–57. http://dx.doi.org/10.1101/gad. 234468.113. [158] Giresi PG, Stevenson EJ, Theilhaber J, Koncarevic A, Parkington J, Fielding RA, et al. Identification of a molecular signature of sarcopenia. Physiol Genomics 2005;21:253–63. http://dx.doi.org/10.1152/physiolgenomics.00249.2004. [159] Drummond MJ, McCarthy JJ, Sinha M, Spratt HM, Volpi E, Esser KA, et al. Aging and microRNA expression in human skeletal muscle: a microarray and bioinformatics analysis.

Review: Myogenesis in the Genomics Era

Physiol Genomics 2011;43:595–603. http://dx.doi.org/10. 1152/physiolgenomics.00148.2010. [160] Barns M, Gondro C, Tellam RL, Radley-Crabb HG, Grounds MD, Shavlakadze T. Molecular analyses provide insight into mechanisms underlying sarcopenia and myofibre denervation in old skeletal muscles of mice. Int J Biochem Cell Biol 2014;53: 174–85. http://dx.doi.org/10.1016/j.biocel.2014.04.025. [161] Jackman RW, Wu CL, Kandarian SC. The ChIP-seqdefined networks of Bcl-3 gene binding support its required role in skeletal muscle atrophy. PLoS One 2012;7:e51478. http://dx.doi.org/10.1371/journal.pone.0051478.

Myogenesis in the genomics era.

Skeletal myogenesis is the process of formation of the muscles that enable movement and breathing. Muscles form after the fate determination and diffe...
385KB Sizes 0 Downloads 12 Views