REVIEW ARTICLE

Noncoding RNAs as epigenetic mediators of skeletal muscle regeneration Gurjeev Sohi1 and Francis Jeffrey Dilworth1,2 1 Sprott Center for Stem Cell Research, Regenerative Medicine Program, Ottawa Hospital Research Institute, Canada 2 Department of Cellular and Molecular Medicine, University of Ottawa, Canada

Keywords differentiation, epigenetics, gene expression, lncRNAs, miRNAs, MyoD, myogenesis, myoMiRs, satellite cells, skeletal muscle Correspondence F. J. Dilworth, Sprott Center for Stem Cell Research, Ottawa Hospital Research Institute, 501 Smyth Rd, Mailbox 511, Ottawa, ON, Canada K1H 8L6 Fax: +1 613 739 6294 Tel.: +1 613 737 8899 ext 70339 E-mail: [email protected]

Skeletal muscle regeneration is a well-characterized biological process in which resident adult stem cells must undertake a series of cell-fate decisions to ensure efficient repair of the damaged muscle fibers while also maintaining the stem cell niche. Satellite cells, the main stem cell contributing to the repaired muscle fiber, are maintained in a quiescent state in healthy muscle. Upon injury, the satellite cells become activated, and proliferate to expand the muscle progenitor cell population before returning to the quiescent state or differentiating to become myofibers. Importantly, the determination of cell fate is controlled at the epigenetic level in response to environmental cues. In this review, we discuss our current understanding of the role played by noncoding RNAs (both miRNAs and long-noncoding RNAs) in the epigenetic control of muscle regeneration.

(Received 21 October 2014, revised 1 December 2014, accepted 2 December 2014) doi:10.1111/febs.13170

Introduction Over the last half century, RNA has emerged as a key player in regulating cellular function. Extending from its role as an essential intermediate in the conversion of genomic information to protein products, RNA is now regarded as an important mediator of gene expression and protein function [1,2]. With the recent advancement of unbiased RNA detection methods for mammalian transcriptome analysis, several studies have now established that the vast majority of transcription results in noncoding RNA sequences, with

only 2% of the genome encoding for proteins [3–8]. This has coincided with an increase in the number of scientific publications dedicated to identifying and understanding the function of nonprotein-coding RNAs (ncRNAs). The ncRNA family has emerged as a functionally diverse group of nucleotide entities that possess intrinsic abilities to serve as potential protein ligands, nucleotide base-pairing partners, molecular scaffolds and enzymes. These functions have added to the complexity with which gene expression and protein

Abbreviations Ago, argonaute; CE, core enhancer; DRR, distal regulatory region; eRNA, enhancer RNA; GW182, glycine–tryptophan repeat-containing protein of 182 kDa; IGF, insulin-like growth factor; lncRNA, long nonprotein coding RNA; mRNP, mRNA-containing ribonucleoprotein; Myog, myogenin; MRFs, myogenic regulatory factors; myomiR, muscle-specific miRNA; ncRNA, nonprotein-coding RNA; PRC2, polycomb repressive complex 2; RISC, RNA-induced silencing complex; SC, satellite cell; SR, serine/arginine-rich; Yam, YY1-associated muscle lncRNA.

FEBS Journal (2014) ª 2014 FEBS

1

The role of ncRNAs in muscle regeneration

function are regulated and cellular processes influenced [9]. Specifically, two of the ncRNA classes that have received extensive attention are the miRNAs and long ncRNAs (lncRNAs). Their functional versatility has conferred on them the ability to control different aspects of cellular development, from progenitor cell maintenance to commitment and subsequently differentiation.

miRNAs The miRNA class of small ncRNAs is a group of RNA polymerase II (RNA Pol II)-derived transcripts that undergo extensive processing to achieve their characteristic length of ~ 22 nucleotides [10]. Initially, miRNAs are transcribed as primary transcripts (primiRNAs) that can be several kilobases long [11]. Within the nucleus, the RNase III enzyme Drosha cleaves the pri-mRNA transcripts to produce ~ 70 nucleotide precursor miRNAs [12]. The precursor miRNA is then exported out to the cytoplasm via Exportin-5 [13]. Subsequently, the RNase III enzyme Dicer cleaves precursor miRNAs to give rise to mature miRNAs in the cytoplasm [12]. This processing mechanism is widely used in cells because miRBase (www.mirbase.org – release 21), an online repository for published miRNA sequences, now lists ~ 1800 human miRNAs. However, we note that a subgroup of miRNAs can bypass Drosha processing during their maturation [14]. miRNAs are proposed to regulate the expression of > 60% of human genes through a mechanism in which they guide a diverse set of multi-protein RNA-induced silencing complexes (RISC) to specific target mRNAs [15]. The miRNA-associated RISC complexes consist of a combination of the argonaute (Ago) protein family, the glycine–tryptophan (GW) repeat-containing protein of 182 kDa (GW182) protein family, and other

G. Sohi and F. J. Dilworth

accessory proteins. The combinatory nature of the RISC complex components and the degree of miRNA–mRNA complementarity have been demonstrated to dictate the distinct pathway with which post-translational repression of mRNA targets can occur. Although RISC complexes containing Ago1–4 can repress translation and induce mRNA decay of mRNAs targeted through miRNA-mediated base pairing, only Ago2-containing complexes possess the cleavage activity required for inducing strand breaks in mRNAs that show perfect complementarity of base pairing with miRNAs [16,17]. Both these mRNA translational silencing mechanisms are based on targetsite complementarity localized to the 30 -UTR of the transcript. However, miRNAs also display complementarity for gene promoters. In this context, miRNAs have also been shown to mediate transcriptional activation [18] or silencing [19] through targeting of Ago/GW182-containing complexes to promoter regions. Thus, miRNAs can influence the contribution of specific genes to the proteome through multiple different mechanisms. The overall impact of mature miRNAs on muscle development can be appreciated from the study by O’Rourke et al., in which skeletal muscle-specific inactivation of Dicer during embryonic myogenesis led to decreases in muscle fiber development, reduction in muscle mass and perinatal lethality in mice [20]. A large number of miRNAs have been characterized for their ability to functionally silence transcripts in muscle (see Table 1). However, the impact of distinct RISC complexes to modulate muscle regeneration has yet to be characterized.

lncRNAs The lncRNA class of ncRNAs regroups all transcripts of > 200 nucleotides. They represent the most numer-

Table 1. MyomiRs and their validated targets in muscle MyomiR

Validated targets in muscle

miR-1

DLL1 [116], HDAC4 [76], RASGAP [117], CDK9 [117], FN [117], RHEB [117], PKCe [118], HSP60 [118], HAND2 [119], PAX3 [93], PAX7 [120], FZD7 [121], FRS2 [121], NCX1 [122], ANXA5 [122], HSP70 [123], MYLK3 [124], CALM1 [124], CALM2 [124], SRI [125], CDC42 [126], KLF4 [127], TWF1 [128], IGF1 [129], IGF1R [129], c-MET [130], BCL2 [131], MEF2A [132], PP2A Regulatory Subunit B56a [133], NOTCH3 [134] CCND2 [135], SRF [76], RhoA [136], CDC42 [136], NELFA [136], IGF1R [79], FGFR1 [137], PP2AC [137], PRDM16 [77,78,138], NFATc4 [139], KLF15 [140], CTGF [141], nPTB [142] GJA1 [143], PAX3 [93], PAX7 [92,120], c-MET [130], KSRP [144], HDAC4 [145], NOTCH3 [134], HMGB3 [146], TIMP3 [147], HSP60 [148] p21 [149] PAX7 [92], PTEN [150], FOXO1 [150] PDCD4 [151], PACS2 [151], DYRK2 [151], SOX6 [152], SK3 [153]

miR-133 miR-206 miR-208b miR-486 miR-499

2

FEBS Journal (2014) ª 2014 FEBS

G. Sohi and F. J. Dilworth

ous and functionally diverse class of ncRNAs. Many lncRNAs share features common to mRNAs, such as 50 -capping, poly-adenylation and exon/intron splicing, yet they are distinct in that they are devoid of ORFs, which precludes their translation to proteins [21,22]. These lncRNAs can be transcribed in both the sense and antisense directions by RNA Pol II from genomic sequences found within promoters, exonic and intronic sequences of protein-coding genes [23]. The lncRNAs that lack poly(A) tails are generally transcribed by RNA Pol III [24,25]. In most cases, lncRNA transcripts are less abundant than protein-coding transcripts. Furthermore, the expression of lncRNAs is generally more tightly regulated, both spatially and temporally, compared with the expression of proteincoding genes [8,26]. The primary nucleotide sequence encoded by the lncRNA may allow for the targeting of specific genomic loci, whereas their complex secondary structure allows for the interaction with DNA, RNA and proteins to confer functionally diverse roles in the cell. Importantly, given that lncRNAs can localize to both the nucleus and the cytoplasm, lncRNAs have the potential to act on every aspect of gene expression [27,28]. This functional diversity of lncRNAs, alongside their cell-type- or development stagespecific expressional pattern, positions them as important molecular factors controlling cell identity and lineage commitment [29,30]. In the nucleus, lncRNAs can act either in cis (acting on a gene within the same locus) or trans (acting on a gene at a different locus) to alter transcription via recruitment of a variety of chromatin-modifying enzymes. As a consequence, they have an ability to regulate the chromatin environment at a particular locus or an entire chromosome [31–35]. Interestingly, genome-wide RNA immunoprecipitation sequencing revealed that thousands of lncRNAs associate with the Ezh2-containing polycomb repressive complex 2 (PRC2) [36]. The mechanism whereby lncRNAs interact with chromatin-remodeling complexes, such as PRC2, and help guide them to a particular locus in cis or trans is the subject of ongoing investigations. For instance, evidence exists of specific lncRNAs that serve as guiding scaffolds via their ability to directly bind DNA elements via RNA : DNA complementarity [33]. Moreover, the same lncRNA can often scaffold with different epigenetic complexes. Nuclear lncRNAs have also been shown to interact with the splicing factors and impact splicing events, by influencing their activity and their distribution within nuclear (speckle) domains [37,38]. Moreover, some nuclear lncRNAs interact with RNA sequences to influence their splicing [39]. The ability of nuclear lncRNAs to favor or disrupt FEBS Journal (2014) ª 2014 FEBS

The role of ncRNAs in muscle regeneration

chromosome interactions also enables them to influence chromosome looping [40–42]. Moreover, not only do nuclear lncRNAs act as protein scaffolds, but they can also serve as structural scaffolds helping regulate the formation of nuclear compartments such as speckles, paraspeckels and polycomb bodies [37,43,44]. In the cytoplasm, lncRNAs can hybridize to transcripts and modulate their translational control. They do so by regulating polysome loading to the mRNA transcript [45–47], Staufen1-mediated mRNA decay [48,49] and miRNA-mediated mRNA degradation [29,30,50–52]. miRNA-mediated mRNA degradation is primarily inhibited by cytoplasmic lncRNAs, referred to as competing endogenous RNAs, which serve as molecular sponges for miRNAs, thereby leading to a reduction in levels of unbound miRNAs available to target their respective mRNA targets. Although competing endogenous RNAs were first described in plants [53], subsequent studies in mammals have implicated these cytoplasmic lncRNAs in a number of cellular processes, including cell differentiation and pluripotency [29,30,50]. More recently, circular RNAs have also been demonstrated to serve as miRNA sponges. Because circular RNAs have greater stability and a lower turnover, they have the potential to offer prolonged control of miRNA sponge activity [51,52,54]. Although we continue to uncover the different functions of lncRNA, we are only beginning to elucidate their roles in regulating muscle regeneration.

Muscle regeneration In postnatal life, primary muscle stem cells involved in injury repair are satellite cells (SCs) that reside along the muscle fiber between the sarcolemma and the basal lamina [55]. When associated with a healthy muscle fiber, a SC is maintained in a quiescent state [56]. However, upon injury, the SC becomes activated and begins to proliferate to expand the muscle progenitor pool. The SC must then make a cell-fate decision to either self-renew for maintenance of the stem cell population or undergo muscle differentiation to establish new muscle fibers [57,58]. Over the course of 28 days, SCs that survive injury-induced damage give rise to a complete regeneration of the muscle through a multistep process. These steps include: (a) maintenance of the SC quiescent state; (b) activation of SC to either self-renew or commit to the muscle-progenitor lineage; (c) expansion of the muscle progenitor population; and (d) exit of the cell cycle and terminal differentiation to form myofibers (see Fig. 1). Thus muscle regeneration represents an excellent model for studying the control of cell-fate decisions in adult tissue. The role of epigenetic enzymes in 3

G. Sohi and F. J. Dilworth

ous and functionally diverse class of ncRNAs. Many lncRNAs share features common to mRNAs, such as 50 -capping, poly-adenylation and exon/intron splicing, yet they are distinct in that they are devoid of ORFs, which precludes their translation to proteins [21,22]. These lncRNAs can be transcribed in both the sense and antisense directions by RNA Pol II from genomic sequences found within promoters, exonic and intronic sequences of protein-coding genes [23]. The lncRNAs that lack poly(A) tails are generally transcribed by RNA Pol III [24,25]. In most cases, lncRNA transcripts are less abundant than protein-coding transcripts. Furthermore, the expression of lncRNAs is generally more tightly regulated, both spatially and temporally, compared with the expression of proteincoding genes [8,26]. The primary nucleotide sequence encoded by the lncRNA may allow for the targeting of specific genomic loci, whereas their complex secondary structure allows for the interaction with DNA, RNA and proteins to confer functionally diverse roles in the cell. Importantly, given that lncRNAs can localize to both the nucleus and the cytoplasm, lncRNAs have the potential to act on every aspect of gene expression [27,28]. This functional diversity of lncRNAs, alongside their cell-type- or development stagespecific expressional pattern, positions them as important molecular factors controlling cell identity and lineage commitment [29,30]. In the nucleus, lncRNAs can act either in cis (acting on a gene within the same locus) or trans (acting on a gene at a different locus) to alter transcription via recruitment of a variety of chromatin-modifying enzymes. As a consequence, they have an ability to regulate the chromatin environment at a particular locus or an entire chromosome [31–35]. Interestingly, genome-wide RNA immunoprecipitation sequencing revealed that thousands of lncRNAs associate with the Ezh2-containing polycomb repressive complex 2 (PRC2) [36]. The mechanism whereby lncRNAs interact with chromatin-remodeling complexes, such as PRC2, and help guide them to a particular locus in cis or trans is the subject of ongoing investigations. For instance, evidence exists of specific lncRNAs that serve as guiding scaffolds via their ability to directly bind DNA elements via RNA : DNA complementarity [33]. Moreover, the same lncRNA can often scaffold with different epigenetic complexes. Nuclear lncRNAs have also been shown to interact with the splicing factors and impact splicing events, by influencing their activity and their distribution within nuclear (speckle) domains [37,38]. Moreover, some nuclear lncRNAs interact with RNA sequences to influence their splicing [39]. The ability of nuclear lncRNAs to favor or disrupt FEBS Journal (2014) ª 2014 FEBS

The role of ncRNAs in muscle regeneration

chromosome interactions also enables them to influence chromosome looping [40–42]. Moreover, not only do nuclear lncRNAs act as protein scaffolds, but they can also serve as structural scaffolds helping regulate the formation of nuclear compartments such as speckles, paraspeckels and polycomb bodies [37,43,44]. In the cytoplasm, lncRNAs can hybridize to transcripts and modulate their translational control. They do so by regulating polysome loading to the mRNA transcript [45–47], Staufen1-mediated mRNA decay [48,49] and miRNA-mediated mRNA degradation [29,30,50–52]. miRNA-mediated mRNA degradation is primarily inhibited by cytoplasmic lncRNAs, referred to as competing endogenous RNAs, which serve as molecular sponges for miRNAs, thereby leading to a reduction in levels of unbound miRNAs available to target their respective mRNA targets. Although competing endogenous RNAs were first described in plants [53], subsequent studies in mammals have implicated these cytoplasmic lncRNAs in a number of cellular processes, including cell differentiation and pluripotency [29,30,50]. More recently, circular RNAs have also been demonstrated to serve as miRNA sponges. Because circular RNAs have greater stability and a lower turnover, they have the potential to offer prolonged control of miRNA sponge activity [51,52,54]. Although we continue to uncover the different functions of lncRNA, we are only beginning to elucidate their roles in regulating muscle regeneration.

Muscle regeneration In postnatal life, primary muscle stem cells involved in injury repair are satellite cells (SCs) that reside along the muscle fiber between the sarcolemma and the basal lamina [55]. When associated with a healthy muscle fiber, a SC is maintained in a quiescent state [56]. However, upon injury, the SC becomes activated and begins to proliferate to expand the muscle progenitor pool. The SC must then make a cell-fate decision to either self-renew for maintenance of the stem cell population or undergo muscle differentiation to establish new muscle fibers [57,58]. Over the course of 28 days, SCs that survive injury-induced damage give rise to a complete regeneration of the muscle through a multistep process. These steps include: (a) maintenance of the SC quiescent state; (b) activation of SC to either self-renew or commit to the muscle-progenitor lineage; (c) expansion of the muscle progenitor population; and (d) exit of the cell cycle and terminal differentiation to form myofibers (see Fig. 1). Thus muscle regeneration represents an excellent model for studying the control of cell-fate decisions in adult tissue. The role of epigenetic enzymes in 3

G. Sohi and F. J. Dilworth

activation. Interestingly, miR-489 was found to posttranscriptionally suppress the oncogene Dek, a protein that promotes the proliferative expansion of quiescent SCs [65]. More recently, another study identified miR195 and miR-497 as novel regulators of SC selfrenewal through the attenuation of cell-cycle progression in muscle progenitor cells. In this case, it was shown that miR-195 and miR-497 cause cell-cycle exit by targeting Cdc25a, Cdc25b and Ccnd2 to establish the quiescent state [66]. The lncRNA Uc.283 + A has been demonstrated to complement with the immature pri-miR-195 transcript and prevent its cleavage by Drosha, which is required for the formation of a functional miR-195 molecule [39]. Consistent with mechanisms existing to maintain quiescent cells that are poised for activation, lncRNA Uc.283 + A may turn out to be a key regulator that can quickly disrupt the availability of functional miR-195 required for maintaining the quiescence of SCs. Under such conditions, rapid degradation of pri-miR-195 would be anticipated given that Drosha-mediated cleavage occurs upon hybridization of two complementarity RNA molecules. Although it is critical for SCs to be able to maintain a quiescent state, in order to prevent the premature onset of myogenesis, it is equally important for the SCs to be primed for activation. As a consequence, quiescent SCs have been observed to transcribe Myf5, but utilize components of the miRNA-processing machinery to prevent it from being translated. Specifically, Myf5 is continuously targeted for degradation by miR-31 located within the (mRNP) granules in the cytoplasm [62]. It is only upon the dissociation of these mRNP granules from the SCs that cell quiescence is abolished and myogenesis initiated [62]. Interestingly, GW182 was also observed to colocalize with miR-31 and myf-5 at these mRNA granules [62]. The mRNA granules observed in the SCs were distinct from stress granules that have been characterized by the presence of the RNA-binding protein HuR [67,68], which happened to remain in the nucleus of quiescent SCs [62]. Additional evidence suggests that HuR continues to remain predominantly nuclear during the early phases of differentiation, subsequently translocating to the cytoplasm where it increases the stability of MyoD, myogenin (Myog) and p21 [69,70]. Although HuR stabilizes MyoD, Myog and p21, it can also exert its pro myogenic effect by destabilizing the mRNA-encoding nucleophosmin, a protein that promotes cell-cycle progression upon association with decay factor KSRP [71]. Both stabilizing and destabilizing properties of HuR are likely dependent on its associated complex subunits, which are of equal FEBS Journal (2014) ª 2014 FEBS

The role of ncRNAs in muscle regeneration

importance for HuR-mediated induction of muscle differentiation. Determining the potential importance of components of the mi-RISC complex in mediating translocation of miRNAs and lncRNAs would help elucidate whether specific components of the mi-RISC complex are critical towards ensuring quiescence and appropriate activation of SCs into muscle differentiation (see Fig. 2 for established roles of mi-RISC complexes in muscle). ncRNAs regulating the proliferation state of muscle stem cells Upon muscle damage, SCs become activated and undergo a first cell division between 24 and 48 h after injury [72]. During this cell division, the SC must decide between: (a) dividing asymmetrically to generate a SC and a committed muscle progenitor (ultimately maintaining the SC pool); (b) dividing symmetrically to generate two daughter SCs (ultimately expanding the SC pool); or (c) dividing symmetrically to generate two committed muscle progenitor cells (ultimately depleting the SC pool) (see Fig. 1) [57]. The population of committed muscle progenitors, defined as being Pax7+/ Myf5+/MyoD+/Myog (Step 2) then undergoes a rapid expansion during which cells divide every 8–10 h [72]. Thus, the epigenetic machinery must make the cellfate decision to either replenish the SC compartment or commit to generate a large number of progenitors committed to repair the damaged muscle fibers. This stage of the regeneration process has been well characterized for miRNA expression in both mice [73,74] and humans [75]. Assuming that miRNAs that decrease their expression during differentiation are important for maintaining the proliferative environment, 17 miRNAs are predominantly expressed in proliferating myoblasts with predicted target genes that function in a wide range of cellular processes during myogenesis [75]. Although many of the differentially expressed miRNAs have been previously investigated for their role in myogenesis, there are many whose role remains uncharacterized. An initial stage in the transition to the activated SC state is the downregulation of miR-31 and the release of the Myf5 mRNA from the mRNP granules to permit its translation for protein expression. The expression of Myf5 and MyoD in the activated SC leads to the activation of miR-133a/miR-133b [76], a pair of muscle-enriched miRNAs that are a part of the myomiR family of short ncRNAs (see Table 1). The expression of miR-133 plays a dual role in the proliferating myoblasts: (a) it prevents commitment to the alternate brown adipose cell fate by inhibiting transla5

The role of ncRNAs in muscle regeneration

G. Sohi and F. J. Dilworth

Fig. 2. Modes of miRNA action and the role of the mi-RISC complex in muscle differentiation. The illustration depicts the mi-RISC complex and its involvement in four major mechanisms, whereby most miRNAs function to repress their mRNA targets. These include: (1) translation repression at the initiation stage, (2) translation repression at the post-initiation stage, (3) degredation via mRNA cleavage, and (4) degredation via deadenylation. The three major components of the mi-RISC complex, Ago proteins, GW182 proteins and accessory proteins whose role in muscle differentiation have been evaluated are listed below. Ago, argonaute, GW182, glycine–tryptophan repeat-containing protein of 182 kDa.

tion of the adipogenic transcriptional regulator PRDM16 [77,78]; and (b) it blocks terminal differentiation of the muscle progenitor cells by preventing translation of SRF [76] and IGF1R [79]. Although MyoD is expressed in proliferating myoblasts, its protein level in this cellular state remains lower than that observed in cells induced to undergo terminal differentiation. This lower level of MyoD is regulated through transforming growth factor-beta signaling which acts to repress miR-29, through a mechanism where the DNA-binding transcription factor YY1 mediates recruitment of the PRC2 protein Ezh2 to the miR-29 locus [80]. The inhibition of miR-29 allows for upregulation of HDAC4 protein levels that, in turn, has been implicated in reducing MyoD gene expression, although the precise mechanism of action has not been investigated [81]. Meanwhile, it has recently been demonstrated that HDAC4 promotes SC proliferation by upregulating Pax7 expression, whereas MyoD expression remained unchanged in SC-specific HDAC4 knockout mice [82]. Moreover, HDAC4 inactivation led to a significant decrease in the MyoD target, miR133, and a concomitant increase in the adipogenic transcription regulator PRDM16 [82]. However, the precise mechanism by which transforming growth factor-beta signaling mediates the repression of miR-29 transcription through YY1 requires further clarification. We note that YY1-mediated recruitment of PRC2 to target loci has also been demonstrated to 6

restrict expression of the myomiRs (see Table 1) miR1, miR-133 and miR-206 during myoblast proliferation [83]. Moreover, YY1 binding has also been detected at the genomic loci that give rise to lncRNAs. Using genome-wide chromatin immunoprecipitation-sequence analysis, Lu et al. identified 63 potential YY1-associated muscle (Yam) lncRNAs [84]. Although most of the Yam lncRNAs remain uncharacterized, Yam-1 was shown to act in cis to upregulate neighboring gene expression leading to a block in muscle differentiation [84]. Among these neighboring genes, miR-715 expression is being activated in cis, where the miRNA is contributing to prevent terminal muscle differentiation by inhibiting translation of Wnt7b [84]. Although YY1 DNA binding has previously been shown to influence the myogenic process via direct regulation of miRNA expression [83,85], YY1–Yam-1–miR-715 appears to represents a novel circuitry in which YY1 indirectly influences muscle differentiation through expression of a lncRNA that activates expression of the miR-715 miRNA. It is important to note that not all Yam lncRNAs follow similar expression patterns during myogenesis. Whereas Yam-1 and Yam-3 promote expansion of muscle progenitor cells, Yam-2 and Yam-4 are promyogenic and display increased expression during muscle differentiation [84]. Further work is required to determine the specific mechanisms through which YY1 can mediate opposing expression patterns for FEBS Journal (2014) ª 2014 FEBS

G. Sohi and F. J. Dilworth

activation. Interestingly, miR-489 was found to posttranscriptionally suppress the oncogene Dek, a protein that promotes the proliferative expansion of quiescent SCs [65]. More recently, another study identified miR195 and miR-497 as novel regulators of SC selfrenewal through the attenuation of cell-cycle progression in muscle progenitor cells. In this case, it was shown that miR-195 and miR-497 cause cell-cycle exit by targeting Cdc25a, Cdc25b and Ccnd2 to establish the quiescent state [66]. The lncRNA Uc.283 + A has been demonstrated to complement with the immature pri-miR-195 transcript and prevent its cleavage by Drosha, which is required for the formation of a functional miR-195 molecule [39]. Consistent with mechanisms existing to maintain quiescent cells that are poised for activation, lncRNA Uc.283 + A may turn out to be a key regulator that can quickly disrupt the availability of functional miR-195 required for maintaining the quiescence of SCs. Under such conditions, rapid degradation of pri-miR-195 would be anticipated given that Drosha-mediated cleavage occurs upon hybridization of two complementarity RNA molecules. Although it is critical for SCs to be able to maintain a quiescent state, in order to prevent the premature onset of myogenesis, it is equally important for the SCs to be primed for activation. As a consequence, quiescent SCs have been observed to transcribe Myf5, but utilize components of the miRNA-processing machinery to prevent it from being translated. Specifically, Myf5 is continuously targeted for degradation by miR-31 located within the (mRNP) granules in the cytoplasm [62]. It is only upon the dissociation of these mRNP granules from the SCs that cell quiescence is abolished and myogenesis initiated [62]. Interestingly, GW182 was also observed to colocalize with miR-31 and myf-5 at these mRNA granules [62]. The mRNA granules observed in the SCs were distinct from stress granules that have been characterized by the presence of the RNA-binding protein HuR [67,68], which happened to remain in the nucleus of quiescent SCs [62]. Additional evidence suggests that HuR continues to remain predominantly nuclear during the early phases of differentiation, subsequently translocating to the cytoplasm where it increases the stability of MyoD, myogenin (Myog) and p21 [69,70]. Although HuR stabilizes MyoD, Myog and p21, it can also exert its pro myogenic effect by destabilizing the mRNA-encoding nucleophosmin, a protein that promotes cell-cycle progression upon association with decay factor KSRP [71]. Both stabilizing and destabilizing properties of HuR are likely dependent on its associated complex subunits, which are of equal FEBS Journal (2014) ª 2014 FEBS

The role of ncRNAs in muscle regeneration

importance for HuR-mediated induction of muscle differentiation. Determining the potential importance of components of the mi-RISC complex in mediating translocation of miRNAs and lncRNAs would help elucidate whether specific components of the mi-RISC complex are critical towards ensuring quiescence and appropriate activation of SCs into muscle differentiation (see Fig. 2 for established roles of mi-RISC complexes in muscle). ncRNAs regulating the proliferation state of muscle stem cells Upon muscle damage, SCs become activated and undergo a first cell division between 24 and 48 h after injury [72]. During this cell division, the SC must decide between: (a) dividing asymmetrically to generate a SC and a committed muscle progenitor (ultimately maintaining the SC pool); (b) dividing symmetrically to generate two daughter SCs (ultimately expanding the SC pool); or (c) dividing symmetrically to generate two committed muscle progenitor cells (ultimately depleting the SC pool) (see Fig. 1) [57]. The population of committed muscle progenitors, defined as being Pax7+/ Myf5+/MyoD+/Myog (Step 2) then undergoes a rapid expansion during which cells divide every 8–10 h [72]. Thus, the epigenetic machinery must make the cellfate decision to either replenish the SC compartment or commit to generate a large number of progenitors committed to repair the damaged muscle fibers. This stage of the regeneration process has been well characterized for miRNA expression in both mice [73,74] and humans [75]. Assuming that miRNAs that decrease their expression during differentiation are important for maintaining the proliferative environment, 17 miRNAs are predominantly expressed in proliferating myoblasts with predicted target genes that function in a wide range of cellular processes during myogenesis [75]. Although many of the differentially expressed miRNAs have been previously investigated for their role in myogenesis, there are many whose role remains uncharacterized. An initial stage in the transition to the activated SC state is the downregulation of miR-31 and the release of the Myf5 mRNA from the mRNP granules to permit its translation for protein expression. The expression of Myf5 and MyoD in the activated SC leads to the activation of miR-133a/miR-133b [76], a pair of muscle-enriched miRNAs that are a part of the myomiR family of short ncRNAs (see Table 1). The expression of miR-133 plays a dual role in the proliferating myoblasts: (a) it prevents commitment to the alternate brown adipose cell fate by inhibiting transla5

The role of ncRNAs in muscle regeneration

G. Sohi and F. J. Dilworth

Fig. 3. Feedback mechanisms between myogenic regulatory factors MRFs and myomiRs during muscle differentiation. Specifically, the figure highlights the positive and negative feedback mechanisms that exist between MRFs, MyoD and myogenin, and the following myomiRs, miR-1, miR-206 and miR-486. MyoD-mediated induction of miR-1, miR-206 and miR-486 represses a unique and common subset of downstream target genes such as HDAC4, PTEN and Foxo1a. Repressing HDAC4 leads to further derepression of MyoD, whereas PTEN and FoxO1 repression enables further activation of the IGF signaling pathway and downstream Myogenin expression. The net result is an increase in muscle differentiation. MyoD mediated direct and indirect activation of Myogenin, in turn results in the induction of miR-133 which suppresses IGF signaling pathway and its own expression via targeting IGF-1R mRNA. The net result is a decrease in muscle differentiation.

been shown to be regulated by lncRNAs. Studies from the Sartorelli laboratory have reported that the binding of MyoD and Myog in extragenic enhancer regions result in the expression of a class of lncRNAs that are termed enhancer RNAs (eRNAs) [94]. In particular, they identified eRNAs that are transcribed from the two developmentally important MyoD enhancers, the core enhancer (CE) and the distal regulatory region (DRR), located upstream of the MyoD coding sequence. Specifically, CEeRNA functioned in cis to activate expression of MyoD, whereas DRReRNA worked in trans to upregulate myogenin transcription by increasing chromatin accessibility and RNA Pol II recruitment [94]. However, the molecular mechanism whereby these lncRNAs increase chromatin accessibility and the mechanisms dictating whether CERNA and DRR RNA act in cis or trans require further elucidation. In yet another example of transcription factors being regulated by lncRNAs, a recent study by Dey et al. demonstrated that the imprinted gene encoding the H19 lncRNA can be processed to give rise to miR675-3p and miR-675-5p that promote muscle differentiation [95]. Specifically, miR-675-3p and miR-675-5p induced differentiation by downregulating the proliferation promoting Smad transcription factors [95]. Finally, lncRNAs can also serve as coactivators to increase output from MyoD target genes. In this case, the lncRNA SRA serves as a molecular scaffold for the assembly of a co-activator complex consisting of the RNA helicases p68 and p72 at MyoD-bound genes 8

to mediate transcriptional activation [96]. Although RNase treatment demonstrated that MyoD association with p68 was not dependent on interaction with SRA during differentiation, a combination of SRA, p68, p72 was required to get maximum MyoD dependent transactivation of the muscle-specific creatine kinase enhancer linked to the luciferase gene [96]. Thus, lncRNAs have been shown to act in through multiple different mechanisms to regulate the ability of muscle transcription factors to induce the muscle gene expression program. Cell-cycle exit Exiting cell cycle is a critical step for commitment towards myoblast differentiation. miRNAs play an important role in regulating the cell cycle during myoblast differentiation. For instance, in Myog-mediated induction of myoblast differentiation through cell-cycle exit, Myog upregulates transcription of miR-17–92 cluster, amongst which miR-20a targets the repression of transcription factors E2F1, E2F2 and E2F3 critical for activation of the cell cycle [86,97–99]. Myog can also upregulate Lats2 kinase, which cooperates with pRB to form repressive complexes with E2F-regulated promoters [86,100]. Activation of p38alpha MAPK signaling pathway during myoblast differentiation may also contribute to the upregulation of Lats2 kinase because MAPK signaling has been demonstrate to repress the expression of miR-31, a miRNA that has FEBS Journal (2014) ª 2014 FEBS

G. Sohi and F. J. Dilworth

been demonstrated to directly target the 30 -UTR of the Lats2 kinase [101]. Indeed, activation of p38 MAPK signaling is critical for the onset of myoblast differentiation, in part through induction of cell-cycle exit. It has previously been demonstrated to downregulate cyclin D1 by antagonizing the JNK pathway, leading to cell-cycle exit [102]. Recently, p38 signaling has been shown to be required for the transcription of miR-1/ miR-133a clusters [103]. Interestingly, miR-1 directly suppresses cyclin D1 via initiating its mRNA degradation, whereas miR-133 indirectly suppresses cyclin D1 by targeting the transcription factor Sp1 [103,104]. Given the evidence for Myog to bind to miR-1 and miR-133 promoters, it would be interesting to evaluate how p38 signaling, Myog and miRNAs converge to induce cell-cycle exit leading to the onset of myoblast differentiation. Muscle gene expression The identity of mature muscle fibers is partially defined by the complement of myosin proteins expressed in the cell [105]. Work from the Olson group has shown that miRNAs present in the introns of myosin genes permit the coexpression of genes that help influence muscle function [106]. In this context, miR-208a and miR-208b are processed from the intronic RNA of the Myh6 and Myh7 transcripts respectively, whereas miR-499 is processed from the intronic RNA of the Myh7b gene. Specifically, they demonstrated that expression of the slow muscle fiber myosin genes Myh7 and Myh7b genes help reinforce the slow muscle gene expression program through the expression of miR-208b and miR-499 [106]. This reinforcement of the slow-muscle gene expression program is at least partially due to translational inhibition of the transcriptional repressors Sox6, Purb, Sp3 and HP1b whose 30 -UTR contain target sites for miR-499 and/or miR-208 [106]. By contrast, the fast muscle fiber phenotype has been shown to be activated through the function of a specific linc-RNA, termed linc-MYH, which is transcribed from an enhancer within the Myh locus [107]. The linc-MYH transcript has been shown to upregulate expression of fast fibertype genes while downregulating expression of slow fiber-type genes [107]. However, the mechanism through which linc-MYH is functioning to regulate expression of these fiber-type specific genes has not yet been established. In yet another example of how ncRNA can influence the set of genes expressed in muscle cells, the Malat-1 lncRNA acts as a regulator of alternative splicing for a specific set of pre-mRNAs by interacting FEBS Journal (2014) ª 2014 FEBS

The role of ncRNAs in muscle regeneration

with and influencing distribution of the Serine/Arginine-rich (SR) family of splicing factors in nuclear speckle domains [37]. In the skeletal muscle cells, Malat1 is upregulated during differentiation and may potentially serve as a cue for differentiation as a knockdown of Malat1-suppressed myoblast proliferation due to cell-growth arrest in the G0/G1 phase [108]. Recently, miR-9 has been demonstrated to target Malat-1 for degradation in the nucleus, however, its impact on muscle differentiation remains to be investigated [109]. The miR-9-mediated downregulation of Malat-1 adds to the function of miRNA to serve as post-transcriptional regulators within the nuclear speckles. Investigating the novel roles of microRNAs in regulating levels of other nuclear lncRNAs involved in myogenesis would provide further insight into the functional interplay between these two types of ncRNAs in collectively controlling this robust cellular process (see Fig. 4 for established roles of lncRNAs during muscle differentiation).

Conclusions Research over the past decade has revealed an essential role for ncRNAs in the regulation of muscle regeneration. These studies have established miRNAmediated translation inhibition as a key mechanism modulating the fate of SCs during muscle regeneration. This well-characterized mechanism of miRNA function, combined with the availability of bioinformatic tools to identify target genes, has provided ‘lowhanging fruit’ for understanding how miRNAs expressed in muscle are contributing to regeneration. However, we expect that our understanding of the mechanisms through which ncRNAs contribute to muscle regeneration is in its infancy. Studies in yeast and Caenorhabditis elegans have suggested important roles for miRNAs and their related short dsRNA-containing Ago complexes in a broad range of cellular processes [110]. Many of these processes appear to be conserved in mammalian systems, including targeting of Ago complexes to specific genomic loci for transcriptional activation [111], transcriptional repression [112], facilitation of DNA repair [113] and the modulation of alternative splicing [114]. In addition, we lack a good understanding of the molecular mechanisms through which most of the lncRNAs contribute to altered gene expression – both in cis or in trans – during muscle regeneration. The development of technologies such as CRISPR [115] that allow for efficient deletion of specific genetic loci should facilitate studies to examine the effects of individual ncRNAs on the muscle regeneration process. Characterization of these 9

G. Sohi and F. J. Dilworth

these lncRNAs and how specific Yam RNAs contribute to expansion of the muscle progenitor cell population. ncRNAs Regulating the differentiation state of muscle stem cells As the muscle progenitor population continues to expand, a subset of these cells turn off Pax7 and Myf5 expression, and begin to express the muscle-specific transcription factor Myog. The expression of Myog (Pax7/Myf5/MyoD+/Myog+) activates a gene expression program that induces terminal muscle differentiation through cell-cycle exit and cell fusion to form multinucleated myofibers that reconstitute the damaged muscle [86,87]. Importantly, the expression of Myog represents a point of no return that commits the muscle progenitor to exit cell cycle [86,87]. This stage of the regeneration process has been the most well-characterized because of the availability of a greater number of cells, and our ability to control the timing of differentiation using alternate culture conditions. Regulation of the muscle transcriptional machinery The activation of pro-myogenic signals is sufficient to permit MyoD-bound target genes to be expressed [88]. However, muscle-gene expression is facilitated by increased MyoD accumulation in the differentiating cells. The suppression of transforming growth factorbeta signaling in differentiating progenitors allows for strong expression of miR-29, which reduces levels of the repressive HDAC4 and results in a subsequent increase in transcription of the MyoD gene [81]. We note that MyoD protein levels have also been shown to increase in the absence of miR-203b that targets the 30 -UTR of MyoD in the fish species Nile tilapia [89]. It remains to be determined whether this regulation of MyoD transcripts by miR-203b is conserved in mammals, and how levels of this miRNA change during differentiation. Among MyoD target genes activated during differentiation, Myog expression is a key step in the commitment of myogenic progenitors to differentiate because the two myogenic regulatory factors work synergistically to allow for high-level expression of the muscle gene expression program. Among pro-myogenic signaling pathways, activation of insulin-like growth factor (IGF) signaling has also been demonstrated to play a critical role in driving muscle differentiation in part via the up regulation of Myog [79]. FEBS Journal (2014) ª 2014 FEBS

The role of ncRNAs in muscle regeneration

The activation of this signaling pathway is enabled during differentiation, in part by the timely repression of miR-125b, which has been demonstrated to target IGF-II [90]. Interestingly, a negative feedback loop exists in which subsequent upregulation of Myog leads to upregulation of another miRNA, miR-133, which downregulates expression of the IGF-I receptor – an event that eventually attenuates the Myog-driven early muscle differentiation program to permit myofiber maturation [79]. In addition to activating Myog expression, MyoD also upregulates the expression of miR-378, a miRNA that targets the 30 -UTR of the inhibitory transcription factor MyoR that competes with MyoD for binding at target genes [91]. Thus the miR-378-mediated inhibition of MyoR translation allows for increased expression of MyoD target genes. These MyoD target genes include a series of important ncRNAs that are activated simultaneously to facilitate myotube formation. Among this group, MyoD turns on a series of muscle-enriched miRNAs that have been termed the myomiRs. This group includes miR-1, miR-206, miR-486 and miR-133 (see Fig. 3). Important to the differentiation process, the SC regulatory transcription factors Pax7 and Pax3 need to be turned off. Several of the myomiRs play an essential role in this repression, where miR-1 and miR-206 target the 30 -UTRs of both Pax3 and Pax7 for translational repression, and miR-486 can also block translation of Pax7 [92,93]. This myomiR-mediated reduction of Pax7 is essential to differentiation as expression of a miRNA-resistant form of Pax7 in differentiating muscle progenitors is sufficient to inhibit the differentiation process [92]. In addition, miR-1 expression is able to target the 30 -UTR of the transcriptional repressor YY1 to prevent protein translation. This reduction in YY1 levels is speculated to relieve PRC2-mediated repression of miR-206 and miR-133 as well [83]. Transcription factor levels in muscle cells are also controlled by the expression of lncRNAs. In the case of the pro-myogenic transcription factors Mef2C and MAML1, the 30 -UTR of their mRNA is a target for inhibition by the miRNAs miR-135, and miR-133 that continue to be expressed in differentiating muscle progenitors. To evade this miRNA-mediated block in Mef2C and MAML1 mRNA translation, the differentiating muscle cells express the lncRNA linc-MD1 [29]. This lncRNA contains a sequence complementary to miR-133 and miR-135 that acts as a molecular ‘sponge’ to sequester the miRNA that would otherwise block terminal differentiation [29]. Thus linc-MD1 acts to promote differentiation by preventing miRNA-mediated repression of pro-myogenic transcription factors. Recently, the expression of MyoD and Myog has also 7

G. Sohi and F. J. Dilworth

expression. Only then will we begin to appreciate the full extent to which miRNAs and lncRNAs are contributing to the global changes occurring as SCs undergo changes in cell fate to repair injured muscle.

Acknowledgements We would like to thank Yuefeng Li for helpful discussions. Work in the Dilworth laboratory is funded by the Canadian Institutes of Health Research and Muscular Dystrophy Canada. FJD is a Canada Research Chair in Epigenetic Regulation of Transcription.

Author contributions G.S. and F.J.D. wrote the manuscript.

References 1 Crick F (1970) Central dogma of molecular biology. Nature 227, 561–563. 2 Michalak P (2006) RNA world - the dark matter of evolutionary genomics. J Evol Biol 19, 1768–1774. 3 Velculescu VE, Zhang L, Vogelstein B & Kinzler KW (1995) Serial analysis of gene expression. Science 270, 484–487. 4 Wei CL, Ng P, Chiu KP, Wong CH, Ang CC, Lipovich L, Liu ET & Ruan Y (2004) 50 Long serial analysis of gene expression (LongSAGE) and 30 LongSAGE for transcriptome characterization and genome annotation. Proc Natl Acad Sci USA 101, 11701–11706. 5 Matsumura H, Reuter M, Kruger DH, Winter P, Kahl G & Terauchi R (2008) SuperSAGE. Methods Mol Biol 387, 55–70. 6 Shiraki T, Kondo S, Katayama S, Waki K, Kasukawa T, Kawaji H, Kodzius R, Watahiki A, Nakamura M, Arakawa T et al. (2003) Cap analysis gene expression for high-throughput analysis of transcriptional starting point and identification of promoter usage. Proc Natl Acad Sci USA 100, 15776–15781. 7 Mortazavi A, Williams BA, McCue K, Schaeffer L & Wold B (2008) Mapping and quantifying mammalian transcriptomes by RNA-Seq. Nat Methods 5, 621–628. 8 Djebali S, Davis CA, Merkel A, Dobin A, Lassmann T, Mortazavi A, Tanzer A, Lagarde J, Lin W, Schlesinger F et al. (2012) Landscape of transcription in human cells. Nature 489, 101–108. 9 Bushati N & Cohen SM (2007) microRNA functions. Annu Rev Cell Dev Biol 23, 175–205. 10 Ha M & Kim VN (2014) Regulation of microRNA biogenesis. Nat Rev Mol Cell Biol 15, 509–524. 11 Bartel DP (2004) MicroRNAs: genomics, biogenesis, mechanism, and function. Cell 116, 281–297.

FEBS Journal (2014) ª 2014 FEBS

The role of ncRNAs in muscle regeneration

12 Lee Y, Ahn C, Han J, Choi H, Kim J, Yim J, Lee J, Provost P, Radmark O, Kim S et al. (2003) The nuclear RNase III Drosha initiates microRNA processing. Nature 425, 415–419. 13 Lund E, Guttinger S, Calado A, Dahlberg JE & Kutay U (2004) Nuclear export of microRNA precursors. Science 303, 95–98. 14 Ruby JG, Jan CH & Bartel DP (2007) Intronic microRNA precursors that bypass Drosha processing. Nature 448, 83–86. 15 Schraivogel D & Meister G (2014) Import routes and nuclear functions of Argonaute and other small RNAsilencing proteins. Trends Biochem Sci 39, 420–431. 16 Liu J, Carmell MA, Rivas FV, Marsden CG, Thomson JM, Song JJ, Hammond SM, Joshua-Tor L & Hannon GJ (2004) Argonaute2 is the catalytic engine of mammalian RNAi. Science 305, 1437–1441. 17 Diederichs S & Haber DA (2007) Dual role for argonautes in microRNA processing and posttranscriptional regulation of microRNA expression. Cell 131, 1097–1108. 18 Matsui M, Chu Y, Zhang H, Gagnon KT, Shaikh S, Kuchimanchi S, Manoharan M, Corey DR & Janowski BA (2013) Promoter RNA links transcriptional regulation of inflammatory pathway genes. Nucleic Acids Res 41, 10086–10109. 19 Kim DH, Saetrom P, Snove O Jr & Rossi JJ (2008) MicroRNA-directed transcriptional gene silencing in mammalian cells. Proc Natl Acad Sci USA 105, 16230–16235. 20 O’Rourke JR, Georges SA, Seay HR, Tapscott SJ, McManus MT, Goldhamer DJ, Swanson MS & Harfe BD (2007) Essential role for Dicer during skeletal muscle development. Dev Biol 311, 359–368. 21 Guttman M, Amit I, Garber M, French C, Lin MF, Feldser D, Huarte M, Zuk O, Carey BW, Cassady JP et al. (2009) Chromatin signature reveals over a thousand highly conserved large non-coding RNAs in mammals. Nature 458, 223–227. 22 Mercer TR & Mattick JS (2013) Structure and function of long noncoding RNAs in epigenetic regulation. Nat Struct Mol Biol 20, 300–307. 23 Kapranov P, Drenkow J, Cheng J, Long J, Helt G, Dike S & Gingeras TR (2005) Examples of the complex architecture of the human transcriptome revealed by RACE and high-density tiling arrays. Genome Res 15, 987–997. 24 Dieci G, Fiorino G, Castelnuovo M, Teichmann M & Pagano A (2007) The expanding RNA polymerase III transcriptome. Trends Genet 23, 614–622. 25 Kapranov P, Cheng J, Dike S, Nix DA, Duttagupta R, Willingham AT, Stadler PF, Hertel J, Hackermuller J, Hofacker IL et al. (2007) RNA maps reveal new RNA classes and a possible function for pervasive transcription. Science 316, 1484–1488.

11

The role of ncRNAs in muscle regeneration

26 Cabili MN, Trapnell C, Goff L, Koziol M, TazonVega B, Regev A & Rinn JL (2011) Integrative annotation of human large intergenic noncoding RNAs reveals global properties and specific subclasses. Genes Dev 25, 1915–1927. 27 Batista PJ & Chang HY (2013) Long noncoding RNAs: cellular address codes in development and disease. Cell 152, 1298–1307. 28 van Heesch S, van Iterson M, Jacobi J, Boymans S, Essers PB, de Bruijn E, Hao W, MacInnes AW, Cuppen E & Simonis M (2014) Extensive localization of long noncoding RNAs to the cytosol and mono- and polyribosomal complexes. Genome Biol 15, R6. 29 Cesana M, Cacchiarelli D, Legnini I, Santini T, Sthandier O, Chinappi M, Tramontano A & Bozzoni I (2011) A long noncoding RNA controls muscle differentiation by functioning as a competing endogenous RNA. Cell 147, 358–369. 30 Wang Y, Xu Z, Jiang J, Xu C, Kang J, Xiao L, Wu M, Xiong J, Guo X & Liu H (2013) Endogenous miRNA sponge lincRNA-RoR regulates Oct4, Nanog, and Sox2 in human embryonic stem cell self-renewal. Dev Cell 25, 69–80. 31 Penny GD, Kay GF, Sheardown SA, Rastan S & Brockdorff N (1996) Requirement for Xist in X chromosome inactivation. Nature 379, 131–137. 32 Nagano T, Mitchell JA, Sanz LA, Pauler FM, Ferguson-Smith AC, Feil R & Fraser P (2008) The Air noncoding RNA epigenetically silences transcription by targeting G9a to chromatin. Science 322, 1717–1720. 33 Chu C, Qu K, Zhong FL, Artandi SE & Chang HY (2011) Genomic maps of long noncoding RNA occupancy reveal principles of RNA-chromatin interactions. Mol Cell 44, 667–678. 34 Plath K, Fang J, Mlynarczyk-Evans SK, Cao R, Worringer KA, Wang H, de la Cruz CC, Otte AP, Panning B & Zhang Y (2003) Role of histone H3 lysine 27 methylation in X inactivation. Science 300, 131–135. 35 Bertani S, Sauer S, Bolotin E & Sauer F (2011) The noncoding RNA Mistral activates Hoxa6 and Hoxa7 expression and stem cell differentiation by recruiting MLL1 to chromatin. Mol Cell 43, 1040–1046. 36 Zhao J, Ohsumi TK, Kung JT, Ogawa Y, Grau DJ, Sarma K, Song JJ, Kingston RE, Borowsky M & Lee JT (2010) Genome-wide identification of polycombassociated RNAs by RIP-seq. Mol Cell 40, 939–953. 37 Tripathi V, Ellis JD, Shen Z, Song DY, Pan Q, Watt AT, Freier SM, Bennett CF, Sharma A, Bubulya PA et al. (2010) The nuclear-retained noncoding RNA MALAT1 regulates alternative splicing by modulating SR splicing factor phosphorylation. Mol Cell 39, 925–938.

12

G. Sohi and F. J. Dilworth

38 Barry G, Briggs JA, Vanichkina DP, Poth EM, Beveridge NJ, Ratnu VS, Nayler SP, Nones K, Hu J, Bredy TW et al. (2014) The long non-coding RNA Gomafu is acutely regulated in response to neuronal activation and involved in schizophreniaassociated alternative splicing. Mol Psychiatry 19, 486–494. 39 Liz J, Portela A, Soler M, Gomez A, Ling H, Michlewski G, Calin GA, Guil S & Esteller M (2014) Regulation of pri-miRNA processing by a long noncoding RNA transcribed from an ultraconserved region. Mol Cell 55, 138–147. 40 Xiang JF, Yin QF, Chen T, Zhang Y, Zhang XO, Wu Z, Zhang S, Wang HB, Ge J, Lu X et al. (2014) Human colorectal cancer-specific CCAT1-L lncRNA regulates long-range chromatin interactions at the MYC locus. Cell Res 24, 513–531. 41 Cabianca DS, Casa V, Bodega B, Xynos A, Ginelli E, Tanaka Y & Gabellini D (2012) A long ncRNA links copy number variation to a polycomb/trithorax epigenetic switch in FSHD muscular dystrophy. Cell 149, 819–831. 42 Wang KC, Yang YW, Liu B, Sanyal A, CorcesZimmerman R, Chen Y, Lajoie BR, Protacio A, Flynn RA, Gupta RA et al. (2011) A long noncoding RNA maintains active chromatin to coordinate homeotic gene expression. Nature 472, 120–124. 43 Mao YS, Sunwoo H, Zhang B & Spector DL (2011) Direct visualization of the co-transcriptional assembly of a nuclear body by noncoding RNAs. Nat Cell Biol 13, 95–101. 44 Yang L, Lin C, Liu W, Zhang J, Ohgi KA, Grinstein JD, Dorrestein PC & Rosenfeld MG (2011) ncRNAand Pc2 methylation-dependent gene relocation between nuclear structures mediates gene activation programs. Cell 147, 773–788. 45 Beltran M, Puig I, Pena C, Garcia JM, Alvarez AB, Pena R, Bonilla F & de Herreros AG (2008) A natural antisense transcript regulates Zeb2/Sip1 gene expression during Snail1-induced epithelialmesenchymal transition. Genes Dev 22, 756–769. 46 Carrieri C, Cimatti L, Biagioli M, Beugnet A, Zucchelli S, Fedele S, Pesce E, Ferrer I, Collavin L, Santoro C et al. (2012) Long non-coding antisense RNA controls Uchl1 translation through an embedded SINEB2 repeat. Nature 491, 454–457. 47 Yoon JH, Abdelmohsen K, Srikantan S, Yang X, Martindale JL, De S, Huarte M, Zhan M, Becker KG & Gorospe M (2012) LincRNA-p21 suppresses target mRNA translation. Mol Cell 47, 648–655. 48 Kretz M, Siprashvili Z, Chu C, Webster DE, Zehnder A, Qu K, Lee CS, Flockhart RJ, Groff AF, Chow J et al. (2013) Control of somatic tissue differentiation by the long non-coding RNA TINCR. Nature 493, 231–235.

FEBS Journal (2014) ª 2014 FEBS

G. Sohi and F. J. Dilworth

49 Gong C & Maquat LE (2011) lncRNAs transactivate STAU1-mediated mRNA decay by duplexing with 30 UTRs via Alu elements. Nature 470, 284–288. 50 Poliseno L, Salmena L, Zhang J, Carver B, Haveman WJ & Pandolfi PP (2010) A coding-independent function of gene and pseudogene mRNAs regulates tumour biology. Nature 465, 1033–1038. 51 Hansen TB, Wiklund ED, Bramsen JB, Villadsen SB, Statham AL, Clark SJ & Kjems J (2011) miRNAdependent gene silencing involving Ago2-mediated cleavage of a circular antisense RNA. EMBO J 30, 4414–4422. 52 Memczak S, Jens M, Elefsinioti A, Torti F, Krueger J, Rybak A, Maier L, Mackowiak SD, Gregersen LH, Munschauer M et al. (2013) Circular RNAs are a large class of animal RNAs with regulatory potency. Nature 495, 333–338. 53 Franco-Zorrilla JM, Valli A, Todesco M, Mateos I, Puga MI, Rubio-Somoza I, Leyva A, Weigel D, Garcia JA & Paz-Ares J (2007) Target mimicry provides a new mechanism for regulation of microRNA activity. Nat Genet 39, 1033–1037. 54 Hansen TB, Jensen TI, Clausen BH, Bramsen JB, Finsen B, Damgaard CK & Kjems J (2013) Natural RNA circles function as efficient microRNA sponges. Nature 495, 384–388. 55 Mauro A (1961) Satellite cell of skeletal muscle fibers. J Biophys Biochem Cytol 9, 493–495. 56 Montarras D, L’Honore A & Buckingham M (2013) Lying low but ready for action: the quiescent muscle satellite cell. FEBS J 280, 4036–4050. 57 Kuang S, Kuroda K, Le Grand F & Rudnicki MA (2007) Asymmetric self-renewal and commitment of satellite stem cells in muscle. Cell 129, 999–1010. 58 Le Grand F, Jones AE, Seale V, Scime A & Rudnicki MA (2009) Wnt7a activates the planar cell polarity pathway to drive the symmetric expansion of satellite stem cells. Cell Stem Cell 4, 535–547. 59 Segales J, Perdiguero E & Mu~ noz-Canoves P (2014) Epigenetic control of adult skeletal muscle stem cell functions. FEBS J. doi:10.1111/febs.13065. 60 Aziz A, Sebastian S & Dilworth FJ (2012) The origin and fate of muscle satellite cells. Stem Cell Rev 8, 609–622. 61 McKinnell IW, Ishibashi J, Le Grand F, Punch VG, Addicks GC, Greenblatt JF, Dilworth FJ & Rudnicki MA (2008) Pax7 activates myogenic genes by recruitment of a histone methyltransferase complex. Nat Cell Biol 10, 77–84. 62 Crist CG, Montarras D & Buckingham M (2012) 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.

FEBS Journal (2014) ª 2014 FEBS

The role of ncRNAs in muscle regeneration

63 Liu L, Cheung TH, Charville GW, Hurgo BM, Leavitt T, Shih J, Brunet A & Rando TA (2013) Chromatin modifications as determinants of muscle stem cell quiescence and chronological aging. Cell Rep 4, 189–204. 64 Reimann J, Brimah K, Schroder R, Wernig A, Beauchamp JR & Partridge TA (2004) Pax7 distribution in human skeletal muscle biopsies and myogenic tissue cultures. Cell Tissue Res 315, 233–242. 65 Cheung TH, Quach NL, Charville GW, Liu L, Park L, Edalati A, Yoo B, Hoang P & Rando TA (2012) Maintenance of muscle stem-cell quiescence by microRNA-489. Nature 482, 524–528. 66 Sato T, Yamamoto T & Sehara-Fujisawa A (2014) miR–195/497 induce postnatal quiescence of skeletal muscle stem cells. Nat Commun 5, 4597. 67 Kedersha N & Anderson P (2007) Mammalian stress granules and processing bodies. Methods Enzymol 431, 61–81. 68 Buchan JR & Parker R (2009) Eukaryotic stress granules: the ins and outs of translation. Mol Cell 36, 932–941. 69 Legnini I, Morlando M, Mangiavacchi A, Fatica A & Bozzoni I (2014) A feedforward regulatory loop between HuR and the long noncoding RNA linc–MD1 controls early phases of myogenesis. Mol Cell 53, 506–514. 70 Figueroa A, Cuadrado A, Fan J, Atasoy U, Muscat GE, Munoz-Canoves P, Gorospe M & Munoz A (2003) Role of HuR in skeletal myogenesis through coordinate regulation of muscle differentiation genes. Mol Cell Biol 23, 4991–5004. 71 Cammas A, Sanchez BJ, Lian XJ, Dormoy-Raclet V, van der Giessen K, de Silanes IL, Ma J, Wilusz C, Richardson J, Gorospe M et al. (2014) Destabilization of nucleophosmin mRNA by the HuR/KSRP complex is required for muscle fibre formation. Nat Commun 5, 4190. 72 Siegel AL, Kuhlmann PK & Cornelison DD (2011) Muscle satellite cell proliferation and association: new insights from myofiber time-lapse imaging. Skelet Muscle 1, 7. 73 Ciarapica R, Russo G, Verginelli F, Raimondi L, Donfrancesco A, Rota R & Giordano A (2009) Deregulated expression of miR–26a and Ezh2 in rhabdomyosarcoma. Cell Cycle 8, 172–175. 74 Granjon A, Gustin MP, Rieusset J, Lefai E, Meugnier E, Guller I, Cerutti C, Paultre C, Disse E, RabasaLhoret R et al. (2009) The microRNA signature in response to insulin reveals its implication in the transcriptional action of insulin in human skeletal muscle and the role of a sterol regulatory elementbinding protein-1c/myocyte enhancer factor 2C pathway. Diabetes 58, 2555–2564.

13

The role of ncRNAs in muscle regeneration

75 Dmitriev P, Barat A, Polesskaya A, O’Connell MJ, Robert T, Dessen P, Walsh TA, Lazar V, Turki A, Carnac G et al. (2013) Simultaneous miRNA and mRNA transcriptome profiling of human myoblasts reveals a novel set of myogenic differentiationassociated miRNAs and their target genes. BMC Genom 14, 265. 76 Chen JF, Mandel EM, Thomson JM, Wu Q, Callis TE, Hammond SM, Conlon FL & Wang DZ (2006) The role of microRNA-1 and microRNA-133 in skeletal muscle proliferation and differentiation. Nat Genet 38, 228–233. 77 Yin H, Pasut A, Soleimani VD, Bentzinger CF, Antoun G, Thorn S, Seale P, Fernando P, van Ijcken W, Grosveld F et al. (2013) MicroRNA-133 controls brown adipose determination in skeletal muscle satellite cells by targeting Prdm16. Cell Metab 17, 210–224. 78 Trajkovski M, Ahmed K, Esau CC & Stoffel M (2012) MyomiR–133 regulates brown fat differentiation through Prdm16. Nat Cell Biol 14, 1330–1335. 79 Huang MB, Xu H, Xie SJ, Zhou H & Qu LH (2011) Insulin-like growth factor-1 receptor is regulated by microRNA-133 during skeletal myogenesis. PLoS One 6, e29173. 80 Zhou L, Wang L, Lu L, Jiang P, Sun H & Wang H (2012) A novel target of microRNA-29, Ring1 and YY1-binding protein (Rybp), negatively regulates skeletal myogenesis. J Biol Chem 287, 25255–25265. 81 Winbanks CE, Wang B, Beyer C, Koh P, White L, Kantharidis P & Gregorevic P (2011) TGF-beta regulates miR–206 and miR–29 to control myogenic differentiation through regulation of HDAC4. J Biol Chem 286, 13805–13814. 82 Choi MC, Ryu S, Hao R, Wang B, Kapur M, Fan CM & Yao TP (2014) HDAC4 promotes Pax7dependent satellite cell activation and muscle regeneration. EMBO Rep 15, 1175–1183. 83 Lu L, Zhou L, Chen EZ, Sun K, Jiang P, Wang L, Su X, Sun H & Wang H (2012) A Novel YY1-miR–1 regulatory circuit in skeletal myogenesis revealed by genome-wide prediction of YY1-miRNA network. PLoS One 7, e27596. 84 Lu L, Sun K, Chen X, Zhao Y, Wang L, Zhou L, Sun H & Wang H (2013) Genome-wide survey by ChIP-seq reveals YY1 regulation of lincRNAs in skeletal myogenesis. EMBO J 32, 2575–2588. 85 Wang H, Garzon R, Sun H, Ladner KJ, Singh R, Dahlman J, Cheng A, Hall BM, Qualman SJ, Chandler DS et al. (2008) NF-kappaB-YY1-miR–29 regulatory circuitry in skeletal myogenesis and rhabdomyosarcoma. Cancer Cell 14, 369–381. 86 Liu QC, Zha XH, Faralli H, Yin H, Louis-Jeune C, Perdiguero E, Pranckeviciene E, Munoz-Canoves P, Rudnicki MA, Brand M et al. (2012) Comparative

14

G. Sohi and F. J. Dilworth

87

88

89

90

91

92

93

94

95

96

97

98

99

expression profiling identifies differential roles for Myogenin and p38alpha MAPK signaling in myogenesis. J Mol Cell Biol 4, 386–397. Singh K & Dilworth FJ (2013) Differential modulation of cell cycle progression distinguishes members of the myogenic regulatory factor family of transcription factors. FEBS J 280, 3991–4003. Cao Y, Yao Z, Sarkar D, Lawrence M, Sanchez GJ, Parker MH, MacQuarrie KL, Davison J, Morgan MT, Ruzzo WL et al. (2010) Genome-wide MyoD binding in skeletal muscle cells: a potential for broad cellular reprogramming. Dev Cell 18, 662–674. Yan B, Guo JT, Zhu CD, Zhao LH & Zhao JL (2013) miR–203b: a novel regulator of MyoD expression in tilapia skeletal muscle. J Exp Biol 216, 447–451. Ge Y, Sun Y & Chen J (2011) IGF-II is regulated by microRNA-125b in skeletal myogenesis. J Cell Biol 192, 69–81. Gagan J, Dey BK, Layer R, Yan Z & Dutta A (2011) MicroRNA-378 targets the myogenic repressor MyoR during myoblast differentiation. J Biol Chem 286, 19431–19438. Dey BK, Gagan J & Dutta A (2011) miR–206 and -486 induce myoblast differentiation by downregulating Pax7. Mol Cell Biol 31, 203–214. Hirai H, Verma M, Watanabe S, Tastad C, Asakura Y & Asakura A (2010) MyoD regulates apoptosis of myoblasts through microRNA-mediated downregulation of Pax3. J Cell Biol 191, 347–365. Mousavi K, Zare H, Dell’orso S, Grontved L, Gutierrez-Cruz G, Derfoul A, Hager GL & Sartorelli V (2013) eRNAs promote transcription by establishing chromatin accessibility at defined genomic loci. Mol Cell 51, 606–617. Dey BK, Pfeifer K & Dutta A (2014) The H19 long noncoding RNA gives rise to microRNAs miR–675-3p and miR–675-5p to promote skeletal muscle differentiation and regeneration. Genes Dev 28, 491–501. Caretti G, Schiltz RL, Dilworth FJ, Di Padova M, Zhao P, Ogryzko V, Fuller-Pace FV, Hoffman EP, Tapscott SJ & Sartorelli V (2006) The RNA helicases p68/p72 and the noncoding RNA SRA are coregulators of MyoD and skeletal muscle differentiation. Dev Cell 11, 547–560. O’Donnell KA, Wentzel EA, Zeller KI, Dang CV & Mendell JT (2005) c-Myc-regulated microRNAs modulate E2F1 expression. Nature 435, 839–843. Sylvestre Y, De Guire V, Querido E, Mukhopadhyay UK, Bourdeau V, Major F, Ferbeyre G & Chartrand P (2007) An E2F/miR–20a autoregulatory feedback loop. J Biol Chem 282, 2135–2143. Nagel S, Venturini L, Przybylski GK, Grabarczyk P, Schmidt CA, Meyer C, Drexler HG, Macleod RA & Scherr M (2009) Activation of miR–17-92 by NK-like homeodomain proteins suppresses apoptosis via

FEBS Journal (2014) ª 2014 FEBS

G. Sohi and F. J. Dilworth

100

101

102

103

104

105

106

107

108

109

110 111

reduction of E2F1 in T-cell acute lymphoblastic leukemia. Leuk Lymphoma 50, 101–108. Tschop K, Conery AR, Litovchick L, Decaprio JA, Settleman J, Harlow E & Dyson N (2011) A kinase shRNA screen links LATS2 and the pRB tumor suppressor. Genes Dev 25, 814–830. Liu X, Cheng Y, Chen X, Yang J, Xu L & Zhang C (2011) MicroRNA-31 regulated by the extracellular regulated kinase is involved in vascular smooth muscle cell growth via large tumor suppressor homolog 2. J Biol Chem 286, 42371–42380. Perdiguero E, Ruiz-Bonilla V, Gresh L, Hui L, Ballestar E, Sousa-Victor P, Baeza-Raja B, Jardi M, Bosch-Comas A, Esteller M et al. (2007) Genetic analysis of p38 MAP kinases in myogenesis: fundamental role of p38alpha in abrogating myoblast proliferation. EMBO J 26, 1245–1256. Zhang D, Li X, Chen C, Li Y, Zhao L, Jing Y, Liu W, Wang X, Zhang Y, Xia H et al. (2012) Attenuation of p38-mediated miR–1/133 expression facilitates myoblast proliferation during the early stage of muscle regeneration. PLoS One 7, e41478. Jash S, Dhar G, Ghosh U & Adhya S (2014) Role of the mTORC1 complex in satellite cell activation by RNA-induced mitochondrial restoration: dual control of cyclin D1 through microRNAs. Mol Cell Biol 34, 3594–3606. Schiaffino S & Reggiani C (2011) Fiber types in mammalian skeletal muscles. Physiol Rev 91, 1447–1531. van Rooij E, Quiat D, Johnson BA, Sutherland LB, Qi X, Richardson JA, Kelm RJ Jr & Olson EN (2009) A family of microRNAs encoded by myosin genes governs myosin expression and muscle performance. Dev Cell 17, 662–673. Sakakibara I, Santolini M, Ferry A, Hakim V & Maire P (2014) Six homeoproteins and a Iinc-RNA at the fast MYH locus lock fast myofiber terminal phenotype. PLoS Genet 10, e1004386. Watts R, Johnsen VL, Shearer J & Hittel DS (2013) Myostatin-induced inhibition of the long noncoding RNA Malat1 is associated with decreased myogenesis. Am J Physiol Cell Physiol 304, C995–C1001. Leucci E, Patella F, Waage J, Holmstrom K, Lindow M, Porse B, Kauppinen S & Lund AH (2013) microRNA-9 targets the long non-coding RNA MALAT1 for degradation in the nucleus. Sci Rep 3, 2535. Grishok A (2013) Biology and mechanisms of short RNAs in Caenorhabditis elegans. Adv Genet 83, 1–69. Place RF, Li LC, Pookot D, Noonan EJ & Dahiya R (2008) MicroRNA-373 induces expression of genes with complementary promoter sequences. Proc Natl Acad Sci USA 105, 1608–1613.

FEBS Journal (2014) ª 2014 FEBS

The role of ncRNAs in muscle regeneration

112 Zardo G, Ciolfi A, Vian L, Starnes LM, Billi M, Racanicchi S, Maresca C, Fazi F, Travaglini L, Noguera N et al. (2012) Polycombs and microRNA223 regulate human granulopoiesis by transcriptional control of target gene expression. Blood 119, 4034–4046. 113 Francia S, Michelini F, Saxena A, Tang D, de Hoon M, Anelli V, Mione M, Carninci P & d’Adda di Fagagna F (2012) Site-specific DICER and DROSHA RNA products control the DNA-damage response. Nature 488, 231–235. 114 Ameyar-Zazoua M, Rachez C, Souidi M, Robin P, Fritsch L, Young R, Morozova N, Fenouil R, Descostes N, Andrau JC et al. (2012) Argonaute proteins couple chromatin silencing to alternative splicing. Nat Struct Mol Biol 19, 998–1004. 115 Gilbert LA, Larson MH, Morsut L, Liu Z, Brar GA, Torres SE, Stern-Ginossar N, Brandman O, Whitehead EH, Doudna JA et al. (2013) CRISPRmediated modular RNA-guided regulation of transcription in eukaryotes. Cell 154, 442–451. 116 Kwon C, Han Z, Olson EN & Srivastava D (2005) MicroRNA1 influences cardiac differentiation in Drosophila and regulates Notch signaling. Proc Natl Acad Sci USA 102, 18986–18991. 117 Sayed D, Hong C, Chen IY, Lypowy J & Abdellatif M (2007) MicroRNAs play an essential role in the development of cardiac hypertrophy. Circ Res 100, 416–424. 118 Pan Z, Sun X, Ren J, Li X, Gao X, Lu C, Zhang Y, Sun H, Wang Y, Wang H et al. (2012) miR–1 exacerbates cardiac ischemia-reperfusion injury in mouse models. PLoS One 7, e50515. 119 Zhao Y, Samal E & Srivastava D (2005) Serum response factor regulates a muscle-specific microRNA that targets Hand2 during cardiogenesis. Nature 436, 214–220. 120 Chen JF, Tao Y, Li J, Deng Z, Yan Z, Xiao X & Wang DZ (2010) microRNA-1 and microRNA-206 regulate skeletal muscle satellite cell proliferation and differentiation by repressing Pax7. J Cell Biol 190, 867–879. 121 Lu TY, Lin B, Li Y, Arora A, Han L, Cui C, Coronnello C, Sheng Y, Benos PV & Yang L (2013) Overexpression of microRNA-1 promotes cardiomyocyte commitment from human cardiovascular progenitors via suppressing WNT and FGF signaling pathways. J Mol Cell Cardiol 63, 146–154. 122 Tritsch E, Mallat Y, Lefebvre F, Diguet N, Escoubet B, Blanc J, De Windt LJ, Catalucci D, Vandecasteele G, Li Z et al. (2013) An SRF/miR–1 axis regulates NCX1 and annexin A5 protein levels in the normal and failing heart. Cardiovasc Res 98, 372–380.

15

The role of ncRNAs in muscle regeneration

123 Kukreti H, Amuthavalli K, Harikumar A, Sathiyamoorthy S, Feng PZ, Anantharaj R, Tan SL, Lokireddy S, Bonala S, Sriram S et al. (2013) Musclespecific microRNA1 (miR1) targets heat shock protein 70 (HSP70) during dexamethasone-mediated atrophy. J Biol Chem 288, 6663–6678. 124 Ai J, Zhang R, Gao X, Niu HF, Wang N, Xu Y, Li Y, Ma N, Sun LH, Pan ZW et al. (2012) Overexpression of microRNA-1 impairs cardiac contractile function by damaging sarcomere assembly. Cardiovasc Res 95, 385–393. 125 Ali R, Huang Y, Maher SE, Kim RW, Giordano FJ, Tellides G & Geirsson A (2012) miR–1 mediated suppression of Sorcin regulates myocardial contractility through modulation of Ca2+ signaling. J Mol Cell Cardiol 52, 1027–1037. 126 Qian L, Wythe JD, Liu J, Cartry J, Vogler G, Mohapatra B, Otway RT, Huang Y, King IN, Maillet M et al. (2011) Tinman/Nk2–5 acts via miR–1 and upstream of Cdc42 to regulate heart function across species. J Cell Biol 193, 1181–1196. 127 Xie C, Huang H, Sun X, Guo Y, Hamblin M, Ritchie RP, Garcia-Barrio MT, Zhang J & Chen YE (2011) MicroRNA-1 regulates smooth muscle cell differentiation by repressing Kruppel-like factor 4. Stem Cells Dev 20, 205–210. 128 Li Q, Song XW, Zou J, Wang GK, Kremneva E, Li XQ, Zhu N, Sun T, Lappalainen P, Yuan WJ et al. (2010) Attenuation of microRNA-1 derepresses the cytoskeleton regulatory protein twinfilin-1 to provoke cardiac hypertrophy. J Cell Sci 123, 2444–2452. 129 Elia L, Contu R, Quintavalle M, Varrone F, Chimenti C, Russo MA, Cimino V, De Marinis L, Frustaci A, Catalucci D et al. (2009) Reciprocal regulation of microRNA-1 and insulin-like growth factor-1 signal transduction cascade in cardiac and skeletal muscle in physiological and pathological conditions. Circulation 120, 2377–2385. 130 Yan D, Dong Xda E, Chen X, Wang L, Lu C, Wang J, Qu J & Tu L (2009) MicroRNA-1/206 targets c-Met and inhibits rhabdomyosarcoma development. J Biol Chem 284, 29596–29604. 131 Tang Y, Zheng J, Sun Y, Wu Z, Liu Z & Huang G (2009) MicroRNA-1 regulates cardiomyocyte apoptosis by targeting Bcl-2. Int Heart J 50, 377–387. 132 Ikeda S, He A, Kong SW, Lu J, Bejar R, Bodyak N, Lee KH, Ma Q, Kang PM, Golub TR et al. (2009) MicroRNA-1 negatively regulates expression of the hypertrophy-associated calmodulin and Mef2a genes. Mol Cell Biol 29, 2193–2204. 133 Terentyev D, Belevych AE, Terentyeva R, Martin MM, Malana GE, Kuhn DE, Abdellatif M, Feldman DS, Elton TS & Gyorke S (2009) miR–1 overexpression enhances Ca(2+) release and promotes cardiac arrhythmogenesis by targeting PP2A

16

G. Sohi and F. J. Dilworth

134

135

136

137

138

139

140

141

142

143

144

regulatory subunit B56alpha and causing CaMKIIdependent hyperphosphorylation of RyR2. Circ Res 104, 514–521. Gagan J, Dey BK, Layer R, Yan Z & Dutta A (2012) Notch3 and Mef2c proteins are mutually antagonistic via Mkp1 protein and miR–1/206 microRNAs in differentiating myoblasts. J Biol Chem 287, 40360–40370. Liu N, Bezprozvannaya S, Williams AH, Qi X, Richardson JA, Bassel-Duby R & Olson EN (2008) microRNA-133a regulates cardiomyocyte proliferation and suppresses smooth muscle gene expression in the heart. Genes Dev 22, 3242–3254. Care A, Catalucci D, Felicetti F, Bonci D, Addario A, Gallo P, Bang ML, Segnalini P, Gu Y, Dalton ND et al. (2007) MicroRNA-133 controls cardiac hypertrophy. Nat Med 13, 613–618. Feng Y, Niu LL, Wei W, Zhang WY, Li XY, Cao JH & Zhao SH (2013) A feedback circuit between miR– 133 and the ERK1/2 pathway involving an exquisite mechanism for regulating myoblast proliferation and differentiation. Cell Death Dis 4, e934. Liu W, Bi P, Shan T, Yang X, Yin H, Wang YX, Liu N, Rudnicki MA & Kuang S (2013) miR–133a regulates adipocyte browning in vivo. PLoS Genet 9, e1003626. Li Q, Lin X, Yang X & Chang J (2010) NFATc4 is negatively regulated in miR–133a-mediated cardiomyocyte hypertrophic repression. Am J Physiol Heart Circ Physiol 298, H1340–H1347. Horie T, Ono K, Nishi H, Iwanaga Y, Nagao K, Kinoshita M, Kuwabara Y, Takanabe R, Hasegawa K, Kita T et al. (2009) MicroRNA-133 regulates the expression of GLUT4 by targeting KLF15 and is involved in metabolic control in cardiac myocytes. Biochem Biophys Res Commun 389, 315–320. Duisters RF, Tijsen AJ, Schroen B, Leenders JJ, Lentink V, der van Made I, Herias V, van Leeuwen RE, Schellings MW, Barenbrug P et al. (2009) miR– 133 and miR–30 regulate connective tissue growth factor: implications for a role of microRNAs in myocardial matrix remodeling. Circ Res 104, 170–178, 6p following 178. Boutz PL, Chawla G, Stoilov P & Black DL (2007) MicroRNAs regulate the expression of the alternative splicing factor nPTB during muscle development. Genes Dev 21, 71–84. Anderson C, Catoe H & Werner R (2006) MIR–206 regulates connexin43 expression during skeletal muscle development. Nucleic Acids Res 34, 5863–5871. Amirouche A, Tadesse H, Miura P, Belanger G, Lunde JA, Cote J & Jasmin BJ (2014) Converging pathways involving microRNA-206 and the RNAbinding protein KSRP control post-transcriptionally utrophin A expression in skeletal muscle. Nucleic Acids Res 42, 3982–3997.

FEBS Journal (2014) ª 2014 FEBS

G. Sohi and F. J. Dilworth

145 Winbanks CE, Beyer C, Hagg A, Qian H, Sepulveda PV & Gregorevic P (2013) miR–206 represses hypertrophy of myogenic cells but not muscle fibers via inhibition of HDAC4. PLoS One 8, e73589. 146 Maciotta S, Meregalli M, Cassinelli L, Parolini D, Farini A, Fraro GD, Gandolfi F, Forcato M, Ferrari S, Gabellini D et al. (2012) Hmgb3 is regulated by microRNA-206 during muscle regeneration. PLoS One 7, e43464. 147 Limana F, Esposito G, D’Arcangelo D, Di Carlo A, Romani S, Melillo G, Mangoni A, Bertolami C, Pompilio G, Germani A et al. (2011) HMGB1 attenuates cardiac remodelling in the failing heart via enhanced cardiac regeneration and miR–206mediated inhibition of TIMP-3. PLoS One 6, e19845. 148 Shan ZX, Lin QX, Deng CY, Zhu JN, Mai LP, Liu JL, Fu YH, Liu XY, Li YX, Zhang YY et al. (2010) miR–1/miR–206 regulate Hsp60 expression contributing to glucose-mediated apoptosis in cardiomyocytes. FEBS Lett 584, 3592–3600. 149 Zhang Y, Wang Y, Wang X, Zhang Y, Eisner GM, Asico LD, Jose PA & Zeng C (2011) Insulin promotes vascular smooth muscle cell proliferation via

FEBS Journal (2014) ª 2014 FEBS

The role of ncRNAs in muscle regeneration

150

151

152

153

microRNA-208-mediated downregulation of p21. J Hypertens 29, 1560–1568. Small EM, O’Rourke JR, Moresi V, Sutherland LB, McAnally J, Gerard RD, Richardson JA & Olson EN (2010) Regulation of PI3-kinase/Akt signaling by muscle-enriched microRNA-486. Proc Natl Acad Sci USA 107, 4218–4223. Wang J, Jia Z, Zhang C, Sun M, Wang W, Chen P, Ma K, Zhang Y, Li X & Zhou C (2014) miR–499 protects cardiomyocytes from H 2O 2-induced apoptosis via its effects on Pdcd4 and Pacs2. RNA Biol 11, 339–350. Li X, Wang J, Jia Z, Cui Q, Zhang C, Wang W, Chen P, Ma K & Zhou C (2013) MiR–499 regulates cell proliferation and apoptosis during late-stage cardiac differentiation via Sox6 and cyclin D1. PLoS One 8, e74504. Ling TY, Wang XL, Chai Q, Lau TW, Koestler CM, Park SJ, Daly RC, Greason KL, Jen J, Wu LQ et al. (2013) Regulation of the SK3 channel by microRNA499–potential role in atrial fibrillation. Heart Rhythm 10, 1001–1009.

17

Noncoding RNAs as epigenetic mediators of skeletal muscle regeneration.

Skeletal muscle regeneration is a well-characterized biological process in which resident adult stem cells must undertake a series of cell-fate decisi...
595KB Sizes 4 Downloads 11 Views