Next Article in Journal
A First Insight into North American Plant Pathogenic Fungi Armillaria sinapina Transcriptome
Next Article in Special Issue
Interplay among ATP-Dependent Chromatin Remodelers Determines Chromatin Organisation in Yeast
Previous Article in Journal
A Selective PPARγ Modulator Reduces Hepatic Fibrosis
Previous Article in Special Issue
Structural Insights into the Evolutionarily Conserved BAF Chromatin Remodeling Complex
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Regulation of the Mammalian SWI/SNF Family of Chromatin Remodeling Enzymes by Phosphorylation during Myogenesis

by
Teresita Padilla-Benavides
1,2,*,
Pablo Reyes-Gutierrez
1 and
Anthony N. Imbalzano
1,*
1
Department of Biochemistry and Molecular Pharmacology, University of Massachusetts Medical School, Worcester, MA 01605, USA
2
Department of Molecular Biology and Biochemistry, Wesleyan University, Middletown, CT 06459, USA
*
Authors to whom correspondence should be addressed.
Biology 2020, 9(7), 152; https://doi.org/10.3390/biology9070152
Submission received: 30 April 2020 / Revised: 24 June 2020 / Accepted: 1 July 2020 / Published: 3 July 2020
(This article belongs to the Special Issue ATP-dependent Chromatin Remodeler)

Abstract

:
Myogenesis is the biological process by which skeletal muscle tissue forms. Regulation of myogenesis involves a variety of conventional, epigenetic, and epigenomic mechanisms that control chromatin remodeling, DNA methylation, histone modification, and activation of transcription factors. Chromatin remodeling enzymes utilize ATP hydrolysis to alter nucleosome structure and/or positioning. The mammalian SWItch/Sucrose Non-Fermentable (mSWI/SNF) family of chromatin remodeling enzymes is essential for myogenesis. Here we review diverse and novel mechanisms of regulation of mSWI/SNF enzymes by kinases and phosphatases. The integration of classic signaling pathways with chromatin remodeling enzyme function impacts myoblast viability and proliferation as well as differentiation. Regulated processes include the assembly of the mSWI/SNF enzyme complex, choice of subunits to be incorporated into the complex, and sub-nuclear localization of enzyme subunits. Together these processes influence the chromatin remodeling and gene expression events that control myoblast function and the induction of tissue-specific genes during differentiation.

1. Introduction

1.1. The Family of SWItch/Sucrose Non-Fermentable (SWI/SNF) Chromatin Remodeling Enzymes

Chromatin remodeling enzymes use the energy of ATP hydrolysis to alter histone:DNA contacts in nucleosomal DNA to modify the position or to disrupt the structure of nucleosomes. The consequences of these changes include permitting or restricting access of DNA binding regulatory factors to the chromatin, which can lead to changes in the ability of these factors to regulate DNA-templated processes such as transcription, replication, repair, and recombination. Most chromatin remodeling enzymes are multi-subunit complexes consisting of an ATPase and auxiliary proteins that help specify function.
Several families of chromatin-remodeling complexes have been described to date. These include the SWI/SNF, ISWI, INO80, and CHD complexes [1,2,3,4,5,6,7]. Components of the SWI/SNF (SWItch/Sucrose Non-Fermentable) complex were initially discovered in Saccharomyces cerevisiae as essential for transcription of genes involved in the mating-type switching and sucrose fermentation pathways [8,9,10]. Some of these SWI and SNF gene products were shown to form a multi-subunit complex that worked to oppose the repressive effects of chromatin [11,12]. Subsequent in vitro work demonstrated that purified SWI/SNF complexes from yeast and human cells altered nucleosome structure in an ATP-dependent manner and, as a consequence, facilitated nucleosome binding by transcription factors (TFs) [11,12,13,14,15]. Conclusive evidence of catalysis by SWI/SNF complexes followed [16].
In Drosophila melanogaster, a complex called BAP, which is homologous to the yeast SWI/SNF complex, is required for expression of homeotic and segmentation genes and regulates the wingless signaling pathway [17,18,19,20,21]. These functions are essential for development since deletion of BAP subunits during embryogenesis is lethal [22,23]. Not surprisingly, depletion of many of the mammalian SWI/SNF (mSWI/SNF) subunits in mice also results in embryonic or peri-natal lethality [23,24].
In mammalian cells, the first evidence for SWI/SNF enzyme function showed a role as a cofactor for transcriptional activation by nuclear hormone receptors [25,26,27]. Subsequent studies demonstrated that its co-activator function was mediated by direct chromatin remodeling activity on integrated viral genes or endogenous cellular genes [28,29]. Additional studies implicated the mSWI/SNF chromatin remodeling enzyme activity in promoting transcriptional repression, as well as other DNA transactions [30,31,32,33,34,35,36].
The mSWI/SNF complex is present in multiple divergent forms [13,15,37,38,39,40]. Two homologous, mutually exclusive ATPases, called BRM (Brahma) and BRG1 (Brahma related gene 1) can act as the catalytic subunit for mSWI/SNF enzyme complexes [25,26,27]. At least 11 additional subunits, or BRG1/BRM-associated factors (BAFs), are structural and functional components of the mSWI/SNF enzyme complexes. Many of these subunits are orthologous to subunits of yeast SWI/SNF and Drosophila BAP complexes [25,26,41,42,43]. The existence of splice variants, multiple isoforms of some of the subunits, and tissue-specific gene expression results in the possibility of hundreds of different potential combinations of proteins in any given enzyme complex, and it is widely believed that the combination of subunits that are present determines the functional specificity of the enzyme [44,45]. Recent efforts have significantly advanced understanding of enzyme complex assembly by demonstrating step-wise assembly of three main sub-families of mSWI/SNF complexes [40]. The links between enzyme composition and function, however, remain poorly understood.
Subunit composition is not the only mechanism by which chromatin remodeling enzyme function can be regulated. Post-transcriptional modifications such as acetylation, methylation, sumoylation, phosphorylation, among others, modulate the activity of the mSWI/SNF complex [46,47,48,49,50,51,52,53,54]. Signaling pathways involve a fine-tuned, differential regulation of kinases and phosphatases that are essential for lineage determination and tissue development and maturation [55,56,57]. Protein kinases and phosphatases direct cell fate through the reversible processes of phosphorylation and dephosphorylation, respectively. Hundreds of these enzymes and specific targets are known in higher eukaryotes, and have been reviewed elsewhere [58,59,60,61]. However, our understanding of regulated phosphorylation to control chromatin remodeling processes in the context of lineage determination and differentiation is limited. Emerging evidence has shown that signal transduction pathways are involved in the phosphorylation of subunits of the mSWI/SNF complex, modulating its catalytic activity and its cofactor function during tissue differentiation. Here we summarize work examining modulation of phosphorylation of mSWI/SNF subunits that functionally impacts myoblast proliferation and differentiation.

1.2. Myogenesis

The development of skeletal muscle is a complex, multi-step process in which mesoderm-derived structures form somites that then serve as the source for all skeletal muscles in the body. Somites are transient paired structures that align on either side of the neural tube around day 8 of mouse embryogenesis. Somitic cells will commit to specific lineages, including skeletal muscle, due to the influence of signaling molecules produced around the periphery [62,63,64,65,66,67,68]. Initial events give rise to the embryonic or primary muscle fibers of the organism, which lead to the subsequent development of fetal (secondary) fibers that will be the foundation of future adult muscles [65,66,69,70,71,72,73]. Embryonic development of skeletal muscle is established in a multi-step cellular process that involves the activation of gene expression programs to generate the different precursor cell types [62,66,74,75,76]. In mice, skeletal muscle generation begins from embryonic day 8.5 to 9 (E8.5–E9) to E18.5; muscle maturation continues for 2–3 weeks after birth.
Post-natal skeletal muscle stem cells, called satellite cells, exist in a quiescent state under the basal lamina of skeletal muscle fibers and are activated upon muscle injury or hypertrophy signaling. Homeostasis of mature skeletal muscle tissue also requires the activation and mitotic expansion of satellite cells as a mechanism to maintain terminally differentiated myofibers [69,71,77,78,79,80,81,82,83,84]. Maintenance of a functional satellite cell pool relies on a specific transcriptional program. Pax7 is the TF required for maintenance and proliferation of satellite cells [80,82,85,86]. Deletion of Pax7 in murine models led to death within 2–3 weeks after birth [87]. Mice depleted of Pax7 have a reduced satellite cell pool, and fail to regenerate muscle [88,89,90].
A wide variety of signaling molecules drive the development of skeletal muscle from embryonic stages throughout post-natal growth [91,92]. Cell signaling events are translated into the activation of myogenic TFs and chromatin remodelers in an organized manner during the maturation of the lineage. The mechanisms by which the cells integrate these signals to define and establish the myogenic lineage are the subject of intense investigation. Here, we will discuss emerging roles for different kinases and phosphatases in the regulation of transcription and chromatin remodeling mechanisms associated with growth and differentiation of the skeletal muscle lineage.

2. Cell Signaling Pathways Modulate Chromatin Remodeling Enzyme Function during Skeletal Muscle Myogenesis

Multiple groups determined many years ago that most, if not all, of the mSWI/SNF subunits were phosphoproteins [93,94,95,96]. Despite this knowledge, there has been little progress in determining the functional role of mSWI/SNF subunit phosphorylation, nor has there been significant attention given to the identification of the kinases and phosphatases that generate and regulate phosphorylation. Indeed, the vast majority of our current understanding of mSWI/SNF phosphorylation is limited to modification of the two ATPase subunits, BRG1 and BRM, and of BAF60c, one of three isoforms of the BAF60 protein.
The BRG1 ATPase is essential for viability, proliferation, and differentiation of myoblasts [29,97]. Functional distinctions between the BRG1 and BRM ATPases in the regulation of myogenesis have been identified [98], suggesting non-overlapping roles, but a full understanding of the divergence in the roles of these ATPases is lacking. The BAF60c protein is the BAF60 isoform most highly expressed in skeletal muscle [41], and it is required for normal skeletal muscle development [99]. BAF60 proteins have long been implicated as core mSWI/SNF subunits [40,100], though their molecular role as part of the complex is poorly defined.

2.1. The Stress Responsive P38 Kinase Is a Regulator of Chromatin Remodeling Enzymes during Myogenesis

Mitogen-activated protein kinases (MAPKs) phosphorylate the OH- groups of serine/threonine residues to regulate cellular processes such as proliferation, differentiation, survival, and apoptosis. Some members of this kinase family influence the development of skeletal muscle. For instance, extracellular signal-regulated kinase ERK1/2 is critical for myoblast proliferation, and ERK2 is necessary for myocyte fusion [101,102]. ERK activity also promotes the transcriptional activity of MyoD by an undefined mechanism [103]. Maintenance of quiescent satellite cells also relies on the ERK pathway [104].
P38 is the only MAPK that has been demonstrated to directly affect the activity of the mSWI/SNF complex. P38 responds to mitogens as well as a wide range of extracellular stress stimuli [105]. There are four widely expressed, highly homologous p38 isoforms (α, β, γ, and δ). P38α is ubiquitously expressed, whereas the other isoforms show differential expression across tissue types (reviewed in [105,106,107,108]). Over 100 direct targets for p38-mediated phosphorylation have been identified [105,109], demonstrating the broad effects of p38 function. P38 kinases are well-known for regulating pro-inflammatory signaling networks and the biosynthesis of cytokines, such as the tumor necrosis factor-α (TNF-α) and interleukin-1β (IL-1β) in immune cells [110,111,112].
P38 is a key regulator of skeletal muscle differentiation, with extensive investigation identifying multiple mechanisms of action in satellite cells, in immortalized myoblasts, and in cells of other origins that have been reprogrammed to recapitulate skeletal muscle differentiation [113,114,115,116,117,118,119,120,121,122,123,124]. During myogenesis, there is a sustained increase in the amount of activated p38α and β [125,126,127]. Activation of p38α induces cell cycle withdrawal by upregulating the activity of a different class of MAPK, the c-Jun N-terminal kinase (JNK) and cyclin D1 expression [119]. p38 phosphorylation of the Polycomb group repressor Ezh2 facilitates its interaction with the Pax7 locus, thereby promoting silencing of this essential satellite cell marker gene [128]. In contrast to functions that inhibit precursor cell properties, p38 also directly promotes myogenic differentiation by phosphorylating E proteins to increasing their dimerization with lineage-determining MyoD and MyoD-related TFs [115], a step necessary for the activation of skeletal muscle-specific gene expression. In addition, p38 phosphorylates MEF2 TFs that cooperate with MyoD family members to activate myogenic gene expression (Figure 1) [122,126,127,129,130,131].
The link between p38 and chromatin remodeling enzyme function derives from work showing that p38 kinase is recruited to myogenic promoters at the onset of differentiation and that its activity is necessary for the recruitment of the mSWI/SNF ATPases BRG1 and BRM to myogenic promoters [114]. These steps are essential for the activation of transcription at muscle-specific genes [29,114,132]. Interestingly, the mSWI/SNF subunit initially identified as the target of p38-mediated phosphorylation was not one of the ATPases but instead the BAF60c subunit [114]. Subsequent efforts showed that the BAF60c subunit, but not the ATPase subunits, formed a pre-assembled complex with MyoD that marks the promoters of myogenic target genes prior to the onset of differentiation signaling. Importantly, knockdown of BAF60c reduced MyoD binding in undifferentiated cells. During myogenesis, p38 phosphorylates the BAF60c subunit at threonine 229, permitting the recruitment of the mSWI/SNF complex ATPases to myogenic promoters and allowing the formation of a functional mSWI/SNF enzyme [114,133]. Chromatin remodeling then permits stable access of MyoD to its cognate binding sites [132] and further recruitment of additional MyoD-associated co-activators (Figure 1) [133]. This work identified function for an mSWI/SNF subunit independent of the ATPase subunits and promoted the idea of physical separation of different mSWI/SNF subunits in myoblasts when biochemical purification of mSWI/SNF enzymes generally have not revealed the existence of sub-complexes [13,14,15,94]. Recent characterization of mSWI/SNF assembly showed that the ATPase subunits are among the last components to be added to the complex [40], which provides further support for this concept. Intriguingly, BAF60c facilitates the recruitment of BRG1 and TFs such as GATA4 to target genes in cardiac progenitors [99,134], but whether phosphorylation by p38 is involved in this process remains to be determined.
The work establishing p38 as a regulator of the mSWI/SNF chromatin remodeling enzymes marked the first evidence that phosphorylation of subunits could contribute to the assembly of the enzyme complex. The work also provided further evidence and support for the idea that myogenic genes are marked for expression in precursor cells, even before the onset of differentiation. Thus, p38 signaling contributes to the activation of gene expression by modulating the activity of chromatin remodelers in addition to regulating TFs.

2.2. The Pleiotropic Network of AKT Converges on BAF60c to Promote Myogenesis

AKT, also known as protein kinase B, is a serine/threonine kinase broadly required for cell proliferation and migration, gene expression, and apoptosis [135]. Three isoforms of AKT (1, 2, and 3) have been reported. AKT1 is required for myoblast proliferation but not for differentiation. AKT2 expression and activity increase during differentiation, which promotes myogenesis. AKT2 is not required for myoblast proliferation and is unable to rescue Akt1 deficiency in proliferating cells [136,137,138,139,140,141]. A recent study has connected the cell adhesion molecule-related downregulated by oncogene (Cdo) and Protein kinase C-related kinase 2 (PKN2) with activation of the AKT signaling pathway [142]. Depletion of Cdo or PKN2 decreased AKT activation and inhibited myogenesis while overexpression of these proteins increased AKT signaling and enhanced differentiation. Studies revealed that regulation of differentiation correlated with the recruitment of BAF60c and MyoD to myogenic promoters [142], though the specific target molecule was not determined. Thus, it remains to be determined whether AKT signaling directly or indirectly affects mSWI/SNF assembly and function and which AKT isoform is responsible. Nevertheless, these data suggest that AKT signaling is influencing the ability of BAF60c to bind to promoters of myogenic genes, whereas p38 signaling activates chromatin-bound BAF60c via phosphorylation. AKT signaling may therefore precede p38 signaling in regulation of mSWI/SNF activity. However, other work using a pharmacological inhibitor showed that BRG1 binding to myogenic promoters was unaffected [143]. The reason for this apparent discrepancy is unclear. It may result from different modes of AKT activation, or the consequences of PKN2 manipulation on BAF60c binding may not be significant enough to preclude BRG1 binding.

2.3. Myoblast Cell Cycle Progression and Viability Are Regulated by Casein Kinase 2 (CK2)-Mediated Phosphorylation of BRG1

Recent studies have demonstrated that casein kinase 2 (CK2) phosphorylates BRG1 in proliferating primary myoblasts derived from mouse satellite cells [50,51]. CK2 is a serine/threonine kinase that exists as a tetramer of two catalytic subunits, CK2α or CK2α’, and two regulatory subunits (CK2β). This kinase is ubiquitously expressed and has more than 300 known substrates [144,145,146]. Experiments performed in diverse cultured cell types showed that CK2 inhibition leads to cell cycle inhibition and death [147,148,149,150]. Thus, CK2 has been associated with proliferation, survival, and apoptosis as well as transcriptional regulation of these processes. Additional work links CK2 to lineage determination and differentiation of many tissues [151]. In skeletal muscle, CK2 regulates the activity of Pax3, Pax7, MyoD, and MyoD-related lineage determining TFs [50,152,153,154,155,156,157,158,159]. Studies using immortalized myoblasts showed that each of the CK2 subunits have differential roles in the determination of the skeletal muscle lineage. For instance, CK2α contributes to the activation of the muscle-specific gene program [160], while CK2α’ contributes to the membrane translocation of fusogenic proteins, thus regulating myoblast fusion [160]. CK2β was shown to be essential for MyoD expression in proliferating myoblasts [160].
In proliferating primary myoblasts, BRG1 is phosphorylated by CK2, and CK2 inhibition or mutation of CK2 target sites in BRG1 resulted in altered myoblast proliferation. Myoblasts expressing a phosphomimetic mutant BRG1 did not proliferate and died. Thus, appropriate regulation of BRG1 phosphorylation by CK2 is required for myoblast survival and normal proliferation. The mechanistic explanation for these results is tied to compromised chromatin remodeling and transcriptional activity of BRG1 at the Pax7 locus in cells with dysregulated CK2 phosphorylation of BRG1 [50] (Figure 2). Pax7 is an essential transcriptional regulator that maintains proliferation of muscle satellite cells [80,81,85,86]. Deletion of Pax7 in mice resulted in a decreased number of satellite cells that were progressively lost with aging, and the animals had an impaired ability to regenerate muscle tissue in response to injury [88,89,90,161]. Recent studies demonstrated that the physical interaction between CK2 and BRG1 occurs specifically during mitosis [51]. Confocal microscopy and co-immunoprecipitation analyses showed that CK2 and BRG1 co-localize in cells undergoing mitosis in developing somites of mouse embryos and in primary myoblasts isolated from satellite cells [51]. Importantly, the mitotic CK2-dependent phosphorylation of BRG1 was conserved across different cell lineages.
Phosphorylation of BRG1 by CK2 also contributes to the localization of BRG1 in the nucleus [50]. Sub-nuclear fractionation of proliferating primary myoblasts showed that endogenous BRG1, as well as an ectopically expressed wild type version of BRG1, were associated with both the nuclear matrix and the chromatin [50], consistent with earlier work demonstrating BRG1 can associate with the nuclear matrix [31,162]. Interestingly, a phosphomimetic mutant form of BRG1, was unable to associate with either of these sub-nuclear fractions, and presented higher intra-nuclear mobility, as measured by FRAP, than wild type BRG1 proteins [50]. A mutant version of BRG1 that prevents BRG1 phosphorylation at the CK2 target sites associated only with the nuclear matrix and not with the chromatin fraction. Enzymatic inhibition of CK2 gave the same results. These findings suggest that sub-nuclear BRG1 localization is dynamic and is regulated by CK2 phosphorylation. Whether other mSWI/SNF subunits are similarly regulated in their sub-nuclear distribution remains an open question. Given the hierarchical assembly of mSWI/SNF subunits into functional enzyme complexes [40], sequestration of different subunits in distinct sub-nuclear compartments provides a potential mechanism to achieve regulated assembly (Figure 2).
Another consequence of BRG1 phosphorylation by CK2 is the different subunit composition of the mSWI/SNF complex [50]. BAF155 and BAF170 are structurally related subunits of mSWI/SNF enzymes [29]. Analysis of mSWI/SNF complex assembly and genetic knockout studies indicate BAFs 155 and 170 are essential for the formation of all mSWI/SNF complexes and for the stability of all of the other mSWI/SNF subunits [40,163]. Endogenous BRG1 or an ectopically expressed wild type version of BRG1 preferentially associates with BAF170, while a non-phosphorylatable BRG1 mutant as well as the endogenous BRG1 from myoblasts treated with CK2 inhibitor preferentially interacts with BAF155. A phosphomimetic BRG1 mutant interacted poorly with both BAF170 and BAF155. Thus, CK2 phosphorylation of BRG1 contributes to the regulation of the subunit composition of the mSWI/SNF complex (Figure 2, [50]). Variation in subunit composition has been associated with specialization in the function of the complex [44,45]. For instance, studies using mouse ESCs showed that the expression of the BAF155 subunit is favored over the expression of BAF170, and BAF155 is preferentially incorporated into BRG1-based SWI/SNF complexes. Upon ESC differentiation into the neuronal lineage, expression of BAF155 is downregulated, while BAF170 is induced and incorporated preferentially into the complex [164]. The consequences of incorporating one or the other subunit are not well understood, but presumably relate to specific chromatin remodeling functions at specific loci during developmental or other signaling.
Although there is no direct link to myogenesis, studies of BRG1 and BRM during the cell cycle showed that both ATPases are hyperphosphorylated upon initiation of mitosis and are excluded from the condensed mitotic chromosomes. The levels of BRG1 remained unaffected, whereas the BRM protein may be degraded [93,94]. ERK1, another member of the MAPK family, was implicated as the kinase that phosphorylates BRG1 and BRM and, consequently, inactivates both the ATPase and the chromatin remodeling activities of the mSWI/SNF complex during mitosis [94]. Mitotic condensation of the chromosomes leads to changes in the activities of TFs, cofactors, and RNA polymerase, causing a general repression of transcription and the removal of many of these factors from the chromatin [165,166,167,168,169]. Whether or not phosphorylation of mSWI/SNF ATPases is causal or is a consequence of mitotic chromosome condensation remains to be determined. It also remains to be seen whether evidence supporting or refuting the idea that CK2 and ERK1, the two kinases implicated in modifying the phosphorylation state of the mSWI/SNF complex during cell cycle progression, function independently or in a dependent manner. In either case, it appears that dynamic regulation of BRG1 and mSWI/SNF enzyme phosphorylation and function occurs during mitosis.

2.4. The PKCβ1 Kinase and the Calcineurin (Cn) Phosphatase Act in Opposition to Regulate BRG1 Phosphorylation and Myogenic Differentiation

In the context of skeletal muscle differentiation, a fine-tuned balance of phosphorylation and dephosphorylation of BRG1 driven by PKCβ1 and calcineurin controls myoblast fate [49]. It is well known that the calcium-sensitive serine/threonine phosphatase calcineurin [170] contributes to the differentiation of immortalized and primary myoblasts, and for muscle regeneration upon damage in vivo [171,172]. Calcineurin is a heterodimer formed by a catalytic and a regulatory subunit [173,174]. The mechanism of action of calcineurin was characterized in lymphocytes. Active calcineurin dephosphorylates the TF nuclear factor of activated T-cell (NFAT), leading to its nuclear translocation, where it regulates gene expression by binding to promoter regions of target genes [175,176,177,178]. During myogenesis, NFAT binds to target promoters in a calcineurin-dependent manner and contributes to differentiation, the determination of fiber type, the number of primary fibers, and the growth of multinucleated muscle cells (Figure 3) [162,179,180,181,182,183,184]. Calcineurin also triggers skeletal muscle differentiation by mechanisms that are independent of NFAT [185,186]. Other work showed that calmodulin, a calcium-binding second messenger protein, also interacts with the mSWI/SNF complex in other cell types and promotes its chromatin remodeling activities in a calcium-dependent manner [187]. Together, the data supported the hypothesis that the mSWI/SNF complex might be a target of calcineurin.
Studies in primary and immortalized myoblasts demonstrated that calcineurin inhibition blocked chromatin remodeling and expression of myogenic genes without affecting the expression of the lineage determining master regulators for myogenesis, MyoD and Myf5. As a result, differentiation was inhibited [49]. Importantly, calcineurin inhibition also led to hyperphosphorylation of the BRG1 ATPase of mSWI/SNF enzymes (Figure 3). Efforts to identify the kinase that opposes calcineurin function relied on software predictors of kinase activity and screening of inhibitors against those kinases. In a key experiment, inhibition of PKCβ kinases rescued the effect of calcineurin inhibition on myogenic differentiation. The simplest interpretation of such results is that in the absence of PKCβ1 activity, there was no consequence of calcineurin inhibition because there were no phosphate groups to remove from the substrate. Knockdown studies identified PKCβ1 as the specific isoform responsible. In vitro work identified serine residues N- and C-terminal to the bromodomain of BRG1 as targets of PKCβ1 and calcineurin [49]. Site directed mutagenesis studies demonstrated that mutation of these sites to phosphomimetic residues prevented myogenesis because the phosphomimetic BRG1 mutant was unable to bind to myogenic promoters while mutation of PKCβ1 target sites to non-phosphorylatable (alanine) residues had no effect on chromatin binding or differentiation [49,188]. Calcineurin inhibition also blocked the interaction of other subunits of the mSWI/SNF complex with myogenic promoters [188]. This raises the possibility of regulated dephosphorylation of other mSWI/SNF subunits or an indirect effect due to the failure to dephosphorylate BRG1. The work identifies a kinase and phosphatase working in opposition to each other to dynamically regulate the activity of a chromatin remodeling enzyme as part of the overall regulatory mechanism governing the initiation of myogenic differentiation (Figure 3).

3. Conclusions

The effect of kinases and phosphatases on mSWI/SNF chromatin remodeling enzymes during myoblast proliferation and differentiation are intriguing and represent another mechanism for cells to regulate differentiation. To date, each characterized pathway has been studied separately, and the crosstalk between these different regulatory signaling remains an important aspect to address. Similarly, evidence for the regulation of mSWI/SNF enzyme function by phosphorylation due to specific kinases exists in contexts other than myogenesis. For example, BAF60c is phosphorylated by PKCζ/λ in response to insulin, resulting in lipogenic gene transcription in liver [52]. BRG1 phosphorylation dependent on the kinase ataxia telangiectasia mutated (ATM) is a mechanism by which the ATPase contributes to double-strand break repair, as the phosphorylation enhances BRG1 binding to γ-H2AX-containing nucleosomes [189]. However, there are a limited number of cell types in which multiple signaling pathways have been implicated in the regulation of mSWI/SNF-dependent processes. Consequently, myoblast proliferation and differentiation represent an exceptional opportunity to take studies of regulated phosphorylation to the next level, which would be to understand how different signaling pathways converge on the same set of target proteins. Whether signaling pathways function independently, are temporally regulated, are dependent on other pathways, or work in a cooperative manner is largely unknown. Given that most, if not all, of the mSWI/SNF component proteins are phosphoproteins, we believe it is highly likely that at least some signaling pathways work in a concerted manner (Figure 4), and that additional signaling pathways affecting mSWI/SNF proteins phosphorylation are yet to be discovered.
In summary, mSWI/SNF enzymes are essential chromatin remodelers that modulate the proper growth and differentiation of mammalian cells. Dysregulation of mSWI/SNF complex activity can lead to consequences in transcriptional regulation, development, and metabolic homeostasis that may progress into severe pathologies such as cancer. Therefore, it is necessary to continue to advance our knowledge of the regulatory events that control the physical and functional states of individual subunit proteins and the enzyme complex. Reversible cell signaling events that cause phosphorylation and dephosphorylation play major roles in most aspects of cell and developmental biology. The work summarized in this review demonstrates that regulation of the mSWI/SNF enzyme assembly, chromatin binding, and chromatin remodeling activities are also targets for these kinases and phosphatases (Figure 4). Further dissection of these regulatory mechanisms will reveal insights into the functionality of the mSWI/SNF complex as well as the impact of chromatin remodeling enzyme function on human development and health.

Author Contributions

Conceptualization, T.P.-B. and A.N.I.; resources, A.N.I.; writing—original draft preparation, T.P.-B. and P.R.-G.; writing—review and editing, T.P.-B. and A.N.I.; visualization, T.P.-B.; supervision, A.N.I.; project administration, T.P.-B. and A.N.I.; funding acquisition, A.N.I. All authors have read and agreed to the published version of the manuscript.

Funding

The authors were supported by NIH grants GM56244 and GM136393.

Acknowledgments

The authors are thankful to Sabriya Syed and Tapan Sharma for critical reviews of this manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mills, A.A. The chromodomain helicase DNA-binding chromatin remodelers: Family traits that protect from and promote cancer. Cold Spring Harb. Perspect. Med. 2017, 7, a026450. [Google Scholar] [CrossRef] [PubMed]
  2. Marfella, C.G.; Imbalzano, A.N. The chd family of chromatin remodelers. Mutat. Res. 2007, 618, 30–40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Poli, J.; Gasser, S.M.; Papamichos-Chronakis, M. The ino80 remodeller in transcription, replication and repair. Philos. Trans. R. Soc. B Biol. Sci. 2017, 372, 20160290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Gerhold, C.B.; Hauer, M.H.; Gasser, S.M. Ino80-c and swr-c: Guardians of the genome. J. Mol. Biol. 2015, 427, 637–651. [Google Scholar] [CrossRef] [PubMed]
  5. Goodwin, L.R.; Picketts, D.J. The role of iswi chromatin remodeling complexes in brain development and neurodevelopmental disorders. Mol. Cell. Neurosci. 2018, 87, 55–64. [Google Scholar] [CrossRef]
  6. Kadoch, C.; Copeland, R.A.; Keilhack, H. Prc2 and swi/snf chromatin remodeling complexes in health and disease. Biochemistry 2016, 55, 1600–1614. [Google Scholar] [CrossRef]
  7. Euskirchen, G.; Auerbach, R.K.; Snyder, M. Swi/snf chromatin-remodeling factors: Multiscale analyses and diverse functions. J. Biol. Chem. 2012, 287, 30897–30905. [Google Scholar] [CrossRef] [Green Version]
  8. Neigeborn, L.; Carlson, M. Genes affecting the regulation of suc2 gene expression by glucose repression in saccharomyces cerevisiae. Genetics 1984, 108, 845–858. [Google Scholar]
  9. Peterson, C.L.; Herskowitz, I. Characterization of the yeast swi1, swi2, and swi3 genes, which encode a global activator of transcription. Cell 1992, 68, 573–583. [Google Scholar] [CrossRef]
  10. Stern, M.; Jensen, R.; Herskowitz, I. Five swi genes are required for expression of the ho gene in yeast. J. Mol. Biol. 1984, 178, 853–868. [Google Scholar] [CrossRef]
  11. Cote, J.; Quinn, J.; Workman, J.L.; Peterson, C.L. Stimulation of gal4 derivative binding to nucleosomal DNA by the yeast swi/snf complex. Science 1994, 265, 53–60. [Google Scholar] [CrossRef] [PubMed]
  12. Cairns, B.R.; Kim, Y.J.; Sayre, M.H.; Laurent, B.C.; Kornberg, R.D. A multisubunit complex containing the swi1/adr6, swi2/snf2, swi3, snf5, and snf6 gene products isolated from yeast. Proc. Natl. Acad. Sci. USA 1994, 91, 1950–1954. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Kwon, H.; Imbalzano, A.N.; Khavari, P.A.; Kingston, R.E.; Green, M.R. Nucleosome disruption and enhancement of activator binding by a human sw1/snf complex. Nature 1994, 370, 477–481. [Google Scholar] [CrossRef] [PubMed]
  14. Imbalzano, A.N.; Kwon, H.; Green, M.R.; Kingston, R.E. Facilitated binding of tata-binding protein to nucleosomal DNA. Nature 1994, 370, 481–485. [Google Scholar] [CrossRef]
  15. Wang, W.; Cote, J.; Xue, Y.; Zhou, S.; Khavari, P.A.; Biggar, S.R.; Muchardt, C.; Kalpana, G.V.; Goff, S.P.; Yaniv, M.; et al. Purification and biochemical heterogeneity of the mammalian swi-snf complex. EMBO J. 1996, 15, 5370–5382. [Google Scholar] [CrossRef]
  16. Logie, C.; Peterson, C.L. Catalytic activity of the yeast swi/snf complex on reconstituted nucleosome arrays. EMBO J. 1997, 16, 6772–6782. [Google Scholar] [CrossRef] [Green Version]
  17. Papoulas, O.; Beek, S.J.; Moseley, S.L.; McCallum, C.M.; Sarte, M.; Shearn, A.; Tamkun, J.W. The drosophila trithorax group proteins brm, ash1 and ash2 are subunits of distinct protein complexes. Development 1998, 125, 3955–3966. [Google Scholar]
  18. Dingwall, A.K.; Beek, S.J.; McCallum, C.M.; Tamkun, J.W.; Kalpana, G.V.; Goff, S.P.; Scott, M.P. The drosophila snr1 and brm proteins are related to yeast swi/snf proteins and are components of a large protein complex. Mol. Biol. Cell 1995, 6, 777–791. [Google Scholar] [CrossRef]
  19. Elfring, L.K.; Daniel, C.; Papoulas, O.; Deuring, R.; Sarte, M.; Moseley, S.; Beek, S.J.; Waldrip, W.R.; Daubresse, G.; DePace, A.; et al. Genetic analysis of brahma: The drosophila homolog of the yeast chromatin remodeling factor swi2/snf2. Genetics 1998, 148, 251–265. [Google Scholar]
  20. Tamkun, J.W.; Deuring, R.; Scott, M.P.; Kissinger, M.; Pattatucci, A.M.; Kaufman, T.C.; Kennison, J.A. Brahma: A regulator of drosophila homeotic genes structurally related to the yeast transcriptional activator snf2/swi2. Cell 1992, 68, 561–572. [Google Scholar] [CrossRef]
  21. Treisman, J.E.; Luk, A.; Rubin, G.M.; Heberlein, U. Eyelid antagonizes wingless signaling during drosophila development and has homology to the bright family of DNA-binding proteins. Genes Dev. 1997, 11, 1949–1962. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Brown, E.; Malakar, S.; Krebs, J.E. How many remodelers does it take to make a brain? Diverse and cooperative roles of atp-dependent chromatin-remodeling complexes in development. Biochem. Cell Biol. 2007, 85, 444–462. [Google Scholar] [CrossRef] [PubMed]
  23. Ho, L.; Crabtree, G.R. Chromatin remodelling during development. Nature 2010, 463, 474–484. [Google Scholar] [CrossRef] [PubMed]
  24. De la Serna, I.L.; Ohkawa, Y.; Imbalzano, A.N. Chromatin remodelling in mammalian differentiation: Lessons from atp-dependent remodellers. Nat. Rev. Genet. 2006, 7, 461–473. [Google Scholar] [CrossRef]
  25. Khavari, P.A.; Peterson, C.L.; Tamkun, J.W.; Mendel, D.B.; Crabtree, G.R. Brg1 contains a conserved domain of the swi2/snf2 family necessary for normal mitotic growth and transcription. Nature 1993, 366, 170–174. [Google Scholar] [CrossRef]
  26. Muchardt, C.; Yaniv, M. A human homologue of saccharomyces cerevisiae snf2/swi2 and drosophila brm genes potentiates transcriptional activation by the glucocorticoid receptor. EMBO J. 1993, 12, 4279–4290. [Google Scholar] [CrossRef]
  27. Chiba, H.; Muramatsu, M.; Nomoto, A.; Kato, H. Two human homologues of saccharomyces cerevisiae swi2/snf2 and drosophila brahma are transcriptional coactivators cooperating with the estrogen receptor and the retinoic acid receptor. Nucleic Acids Res. 1994, 22, 1815–1820. [Google Scholar] [CrossRef] [Green Version]
  28. Fryer, C.J.; Archer, T.K. Chromatin remodelling by the glucocorticoid receptor requires the brg1 complex. Nature 1998, 393, 88–91. [Google Scholar] [CrossRef]
  29. De la Serna, I.L.; Carlson, K.A.; Imbalzano, A.N. Mammalian swi/snf complexes promote myod-mediated muscle differentiation. Nat. Genet. 2001, 27, 187–190. [Google Scholar] [CrossRef]
  30. Lee, D.; Sohn, H.; Kalpana, G.V.; Choe, J. Interaction of e1 and hsnf5 proteins stimulates replication of human papillomavirus DNA. Nature 1999, 399, 487–491. [Google Scholar] [CrossRef]
  31. Murphy, D.J.; Hardy, S.; Engel, D.A. Human swi-snf component brg1 represses transcription of the c-fos gene. Mol. Cell. Biol. 1999, 19, 2724–2733. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Patenge, N.; Elkin, S.K.; Oettinger, M.A. Atp-dependent remodeling by swi/snf and iswi proteins stimulates v(d)j cleavage of 5 s arrays. J. Biol. Chem. 2004, 279, 35360–35367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Cai, S.; Lee, C.C.; Kohwi-Shigematsu, T. Satb1 packages densely looped, transcriptionally active chromatin for coordinated expression of cytokine genes. Nat. Genet. 2006, 38, 1278–1288. [Google Scholar] [CrossRef] [PubMed]
  34. Morshead, K.B.; Ciccone, D.N.; Taverna, S.D.; Allis, C.D.; Oettinger, M.A. Antigen receptor loci poised for v(d)j rearrangement are broadly associated with brg1 and flanked by peaks of histone h3 dimethylated at lysine 4. Proc. Natl. Acad. Sci. USA 2003, 100, 11577–11582. [Google Scholar] [CrossRef] [Green Version]
  35. Zhao, Q.; Wang, Q.E.; Ray, A.; Wani, G.; Han, C.; Milum, K.; Wani, A.A. Modulation of nucleotide excision repair by mammalian swi/snf chromatin-remodeling complex. J. Biol. Chem. 2009, 284, 30424–30432. [Google Scholar] [CrossRef] [Green Version]
  36. Zhang, L.; Zhang, Q.; Jones, K.; Patel, M.; Gong, F. The chromatin remodeling factor brg1 stimulates nucleotide excision repair by facilitating recruitment of xpc to sites of DNA damage. Cell Cycle 2009, 8, 3953–3959. [Google Scholar] [CrossRef] [Green Version]
  37. Wang, X.; Wang, S.; Troisi, E.C.; Howard, T.P.; Haswell, J.R.; Wolf, B.K.; Hawk, W.H.; Ramos, P.; Oberlick, E.M.; Tzvetkov, E.P.; et al. Brd9 defines a swi/snf sub-complex and constitutes a specific vulnerability in malignant rhabdoid tumors. Nat. Commun. 2019, 10, 1881. [Google Scholar] [CrossRef] [Green Version]
  38. Alpsoy, A.; Dykhuizen, E.C. Glioma tumor suppressor candidate region gene 1 (gltscr1) and its paralog gltscr1-like form swi/snf chromatin remodeling subcomplexes. J. Biol. Chem. 2018, 293, 3892–3903. [Google Scholar] [CrossRef] [Green Version]
  39. Michel, B.C.; D’Avino, A.R.; Cassel, S.H.; Mashtalir, N.; McKenzie, Z.M.; McBride, M.J.; Valencia, A.M.; Zhou, Q.; Bocker, M.; Soares, L.M.M.; et al. A non-canonical swi/snf complex is a synthetic lethal target in cancers driven by baf complex perturbation. Nat. Cell Biol. 2018, 20, 1410–1420. [Google Scholar] [CrossRef]
  40. Mashtalir, N.; D’Avino, A.R.; Michel, B.C.; Luo, J.; Pan, J.; Otto, J.E.; Zullow, H.J.; McKenzie, Z.M.; Kubiak, R.L.; St Pierre, R.; et al. Modular organization and assembly of swi/snf family chromatin remodeling complexes. Cell 2018, 175, 1272–1288.e1220. [Google Scholar] [CrossRef] [Green Version]
  41. Wang, W.; Xue, Y.; Zhou, S.; Kuo, A.; Cairns, B.R.; Crabtree, G.R. Diversity and specialization of mammalian swi/snf complexes. Genes Dev. 1996, 10, 2117–2130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Muchardt, C.; Sardet, C.; Bourachot, B.; Onufryk, C.; Yaniv, M. A human protein with homology to saccharomyces cerevisiae snf5 interacts with the potential helicase hbrm. Nucleic Acids Res. 1995, 23, 1127–1132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Kalpana, G.V.; Marmon, S.; Wang, W.; Crabtree, G.R.; Goff, S.P. Binding and stimulation of hiv-1 integrase by a human homolog of yeast transcription factor snf5. Science 1994, 266, 2002–2006. [Google Scholar] [CrossRef] [PubMed]
  44. Wu, J.I. Diverse functions of atp-dependent chromatin remodeling complexes in development and cancer. Acta Biochim. Biophys. Sin. 2012, 44, 54–69. [Google Scholar] [CrossRef] [Green Version]
  45. Hargreaves, D.C.; Crabtree, G.R. Atp-dependent chromatin remodeling: Genetics, genomics and mechanisms. Cell Res. 2011, 21, 396–420. [Google Scholar] [CrossRef]
  46. Choudhary, C.; Kumar, C.; Gnad, F.; Nielsen, M.L.; Rehman, M.; Walther, T.C.; Olsen, J.V.; Mann, M. Lysine acetylation targets protein complexes and co-regulates major cellular functions. Science 2009, 325, 834–840. [Google Scholar] [CrossRef] [Green Version]
  47. Hendriks, I.A.; Lyon, D.; Young, C.; Jensen, L.J.; Vertegaal, A.C.; Nielsen, M.L. Site-specific mapping of the human sumo proteome reveals co-modification with phosphorylation. Nat. Struct. Mol. Biol. 2017, 24, 325–336. [Google Scholar] [CrossRef]
  48. Guo, A.; Gu, H.; Zhou, J.; Mulhern, D.; Wang, Y.; Lee, K.A.; Yang, V.; Aguiar, M.; Kornhauser, J.; Jia, X.; et al. Immunoaffinity enrichment and mass spectrometry analysis of protein methylation. Mol. Cell. Proteom. 2014, 13, 372–387. [Google Scholar] [CrossRef] [Green Version]
  49. Nasipak, B.T.; Padilla-Benavides, T.; Green, K.M.; Leszyk, J.D.; Mao, W.; Konda, S.; Sif, S.; Shaffer, S.A.; Ohkawa, Y.; Imbalzano, A.N. Opposing calcium-dependent signalling pathways control skeletal muscle differentiation by regulating a chromatin remodelling enzyme. Nat. Commun. 2015, 6, 7441. [Google Scholar] [CrossRef] [Green Version]
  50. Padilla-Benavides, T.; Nasipak, B.T.; Paskavitz, A.L.; Haokip, D.T.; Schnabl, J.M.; Nickerson, J.A.; Imbalzano, A.N. Casein kinase 2-mediated phosphorylation of brahma-related gene 1 controls myoblast proliferation and contributes to swi/snf complex composition. J. Biol. Chem. 2017, 292, 18592–18607. [Google Scholar] [CrossRef] [Green Version]
  51. Padilla-Benavides, T.; Haokip, D.T.; Yoon, Y.; Reyes-Gutierrez, P.; Rivera-Perez, J.A.; Imbalzano, A.N. Ck2-dependent phosphorylation of the brg1 chromatin remodeling enzyme occurs during mitosis. Int. J. Mol. Sci. 2020, 21, 923. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Wang, Y.; Wong, R.H.; Tang, T.; Hudak, C.S.; Yang, D.; Duncan, R.E.; Sul, H.S. Phosphorylation and recruitment of baf60c in chromatin remodeling for lipogenesis in response to insulin. Mol. Cell 2013, 49, 283–297. [Google Scholar] [CrossRef] [Green Version]
  53. Puri, P.L.; Mercola, M. Baf60 a, b, and cs of muscle determination and renewal. Genes Dev. 2012, 26, 2673–2683. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Wang, X.; Roberts, C.W. Carma: Carm1 methylation of swi/snf in breast cancer. Cancer Cell 2014, 25, 3–4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Basson, M.A. Signaling in cell differentiation and morphogenesis. Cold Spring Harb. Perspect. Biol. 2012, 4, a008151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Li, P.; Elowitz, M.B. Communication codes in developmental signaling pathways. Development 2019, 146. [Google Scholar] [CrossRef] [Green Version]
  57. Perrimon, N.; Pitsouli, C.; Shilo, B.Z. Signaling mechanisms controlling cell fate and embryonic patterning. Cold Spring Harb. Perspect. Biol. 2012, 4, a005975. [Google Scholar] [CrossRef] [Green Version]
  58. Manning, G.; Whyte, D.B.; Martinez, R.; Hunter, T.; Sudarsanam, S. The protein kinase complement of the human genome. Science 2002, 298, 1912–1934. [Google Scholar] [CrossRef] [Green Version]
  59. Caenepeel, S.; Charydczak, G.; Sudarsanam, S.; Hunter, T.; Manning, G. The mouse kinome: Discovery and comparative genomics of all mouse protein kinases. Proc. Natl. Acad. Sci. USA 2004, 101, 11707–11712. [Google Scholar] [CrossRef] [Green Version]
  60. Wilson, L.J.; Linley, A.; Hammond, D.E.; Hood, F.E.; Coulson, J.M.; MacEwan, D.J.; Ross, S.J.; Slupsky, J.R.; Smith, P.D.; Eyers, P.A.; et al. New perspectives, opportunities, and challenges in exploring the human protein kinome. Cancer Res. 2018, 78, 15–29. [Google Scholar] [CrossRef] [Green Version]
  61. Fleuren, E.D.; Zhang, L.; Wu, J.; Daly, R.J. The kinome ‘at large’ in cancer. Nat. Rev. Cancer 2016, 16, 83–98. [Google Scholar] [CrossRef] [PubMed]
  62. Tajbakhsh, S.; Buckingham, M. The birth of muscle progenitor cells in the mouse: Spatiotemporal considerations. Curr. Top. Dev. Biol. 2000, 48, 225–268. [Google Scholar] [PubMed]
  63. Bryson-Richardson, R.J.; Currie, P.D. The genetics of vertebrate myogenesis. Nat. Rev. Genet. 2008, 9, 632–646. [Google Scholar] [CrossRef] [PubMed]
  64. Borycki, A.G.; Emerson, C.P., Jr. Multiple tissue interactions and signal transduction pathways control somite myogenesis. Curr. Top. Dev. Biol. 2000, 48, 165–224. [Google Scholar]
  65. Sambasivan, R.; Tajbakhsh, S. Skeletal muscle stem cell birth and properties. Semin. Cell Dev. Biol. 2007, 18, 870–882. [Google Scholar] [CrossRef]
  66. Chal, J.; Pourquie, O. Making muscle: Skeletal myogenesis in vivo and in vitro. Development 2017, 144, 2104–2122. [Google Scholar] [CrossRef] [Green Version]
  67. Yabe, T.; Takada, S. Molecular mechanism for cyclic generation of somites: Lessons from mice and zebrafish. Dev. Growth Differ. 2016, 58, 31–42. [Google Scholar] [CrossRef] [Green Version]
  68. Mallo, M. Revisiting the involvement of signaling gradients in somitogenesis. FEBS J. 2016, 283, 1430–1437. [Google Scholar] [CrossRef] [Green Version]
  69. Shi, X.; Garry, D.J. Muscle stem cells in development, regeneration, and disease. Genes Dev. 2006, 20, 1692–1708. [Google Scholar] [CrossRef]
  70. Zammit, P.S.; Relaix, F.; Nagata, Y.; Ruiz, A.P.; Collins, C.A.; Partridge, T.A.; Beauchamp, J.R. Pax7 and myogenic progression in skeletal muscle satellite cells. J. Cell Sci. 2006, 119, 1824–1832. [Google Scholar] [CrossRef] [Green Version]
  71. Collins, C.A.; Olsen, I.; Zammit, P.S.; Heslop, L.; Petrie, A.; Partridge, T.A.; Morgan, J.E. Stem cell function, self-renewal, and behavioral heterogeneity of cells from the adult muscle satellite cell niche. Cell 2005, 122, 289–301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Biressi, S.; Tagliafico, E.; Lamorte, G.; Monteverde, S.; Tenedini, E.; Roncaglia, E.; Ferrari, S.; Cusella-De Angelis, M.G.; Tajbakhsh, S.; Cossu, G. Intrinsic phenotypic diversity of embryonic and fetal myoblasts is revealed by genome-wide gene expression analysis on purified cells. Dev. Biol. 2007, 304, 633–651. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Parker, M.H.; Seale, P.; Rudnicki, M.A. Looking back to the embryo: Defining transcriptional networks in adult myogenesis. Nat. Rev. Genet. 2003, 4, 497–507. [Google Scholar] [CrossRef] [PubMed]
  74. Biressi, S.; Molinaro, M.; Cossu, G. Cellular heterogeneity during vertebrate skeletal muscle development. Dev. Biol. 2007, 308, 281–293. [Google Scholar] [CrossRef] [Green Version]
  75. Messina, G.; Cossu, G. The origin of embryonic and fetal myoblasts: A role of pax3 and pax7. Genes Dev. 2009, 23, 902–905. [Google Scholar] [CrossRef] [Green Version]
  76. Bentzinger, C.F.; Wang, Y.X.; Rudnicki, M.A. Building muscle: Molecular regulation of myogenesis. Cold Spring Harb. Perspect. Biol. 2012, 4, a008342. [Google Scholar] [CrossRef]
  77. Dhawan, J.; Rando, T.A. Stem cells in postnatal myogenesis: Molecular mechanisms of satellite cell quiescence, activation and replenishment. Trends Cell Biol. 2005, 15, 666–673. [Google Scholar] [CrossRef]
  78. Relaix, F. Skeletal muscle progenitor cells: From embryo to adult. Cell. Mol. Life Sci. CMLS 2006, 63, 1221–1225. [Google Scholar] [CrossRef]
  79. Zammit, P.S. All muscle satellite cells are equal, but are some more equal than others? J. Cell Sci. 2008, 121, 2975–2982. [Google Scholar] [CrossRef] [Green Version]
  80. Motohashi, N.; Asakura, A. Muscle satellite cell heterogeneity and self-renewal. Front Cell Dev. Biol. 2014, 2, 1. [Google Scholar] [CrossRef]
  81. Chang, N.C.; Rudnicki, M.A. Satellite cells: The architects of skeletal muscle. Curr. Top. Dev. Biol. 2014, 107, 161–181. [Google Scholar]
  82. Sambasivan, R.; Tajbakhsh, S. Adult skeletal muscle stem cells. Results Probl. Cell Differ. 2015, 56, 191–213. [Google Scholar] [PubMed]
  83. Rudnicki, M.A.; Le Grand, F.; McKinnell, I.; Kuang, S. The molecular regulation of muscle stem cell function. Cold Spring Harbor Symp. Quant. Biol. 2008, 73, 323–331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Mauro, A. Satellite cell of skeletal muscle fibers. J. Biophys. Biochem. Cytol. 1961, 9, 493–495. [Google Scholar] [CrossRef] [PubMed]
  85. Montarras, D.; L’Honore, A.; Buckingham, M. Lying low but ready for action: The quiescent muscle satellite cell. FEBS J. 2013, 280, 4036–4050. [Google Scholar] [CrossRef] [PubMed]
  86. Brack, A.S.; Rando, T.A. Tissue-specific stem cells: Lessons from the skeletal muscle satellite cell. Cell Stem Cell 2012, 10, 504–514. [Google Scholar] [CrossRef] [Green Version]
  87. Mansouri, A.; Stoykova, A.; Torres, M.; Gruss, P. Dysgenesis of cephalic neural crest derivatives in pax7-/-mutant mice. Development 1996, 122, 831–838. [Google Scholar]
  88. Seale, P.; Sabourin, L.A.; Girgis-Gabardo, A.; Mansouri, A.; Gruss, P.; Rudnicki, M.A. Pax7 is required for the specification of myogenic satellite cells. Cell 2000, 102, 777–786. [Google Scholar] [CrossRef] [Green Version]
  89. Oustanina, S.; Hause, G.; Braun, T. Pax7 directs postnatal renewal and propagation of myogenic satellite cells but not their specification. EMBO J. 2004, 23, 3430–3439. [Google Scholar] [CrossRef] [Green Version]
  90. Relaix, F.; Montarras, D.; Zaffran, S.; Gayraud-Morel, B.; Rocancourt, D.; Tajbakhsh, S.; Mansouri, A.; Cumano, A.; Buckingham, M. Pax3 and pax7 have distinct and overlapping functions in adult muscle progenitor cells. J Cell Biol. 2006, 172, 91–102. [Google Scholar] [CrossRef]
  91. Kuang, S.; Gillespie, M.A.; Rudnicki, M.A. Niche regulation of muscle satellite cell self-renewal and differentiation. Cell Stem Cell 2008, 2, 22–31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Bentzinger, C.F.; von Maltzahn, J.; Rudnicki, M.A. Extrinsic regulation of satellite cell specification. Stem Cell Res. Ther. 2010, 1, 27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Muchardt, C.; Reyes, J.C.; Bourachot, B.; Leguoy, E.; Yaniv, M. The hbrm and brg-1 proteins, components of the human snf/swi complex, are phosphorylated and excluded from the condensed chromosomes during mitosis. EMBO J. 1996, 15, 3394–3402. [Google Scholar] [CrossRef] [PubMed]
  94. Sif, S.; Stukenberg, P.T.; Kirschner, M.W.; Kingston, R.E. Mitotic inactivation of a human swi/snf chromatin remodeling complex. Genes Dev. 1998, 12, 2842–2851. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Kimura, A.; Arakawa, N.; Hirano, H. Mass spectrometric analysis of the phosphorylation levels of the swi/snf chromatin remodeling/tumor suppressor proteins arid1a and brg1 in ovarian clear cell adenocarcinoma cell lines. J. Proteome Res. 2014, 13, 4959–4969. [Google Scholar] [CrossRef] [PubMed]
  96. Dallas, P.B.; Yaciuk, P.; Moran, E. Characterization of monoclonal antibodies raised against p300: Both p300 and cbp are present in intracellular tbp complexes. J. Virol. 1997, 71, 1726–1731. [Google Scholar] [CrossRef] [Green Version]
  97. Padilla-Benavides, T.; Nasipak, B.T.; Imbalzano, A.N. Brg1 controls the expression of pax7 to promote viability and proliferation of mouse primary myoblasts. J. Cell. Physiol. 2015, 230, 2990–2997. [Google Scholar] [CrossRef] [Green Version]
  98. Albini, S.; Coutinho Toto, P.; Dall’Agnese, A.; Malecova, B.; Cenciarelli, C.; Felsani, A.; Caruso, M.; Bultman, S.J.; Puri, P.L. Brahma is required for cell cycle arrest and late muscle gene expression during skeletal myogenesis. EMBO Rep. 2015, 16, 1037–1050. [Google Scholar] [CrossRef]
  99. Lickert, H.; Takeuchi, J.K.; Von Both, I.; Walls, J.R.; McAuliffe, F.; Adamson, S.L.; Henkelman, R.M.; Wrana, J.L.; Rossant, J.; Bruneau, B.G. Baf60c is essential for function of baf chromatin remodelling complexes in heart development. Nature 2004, 432, 107–112. [Google Scholar] [CrossRef]
  100. Phelan, M.L.; Sif, S.; Narlikar, G.J.; Kingston, R.E. Reconstitution of a core chromatin remodeling complex from swi/snf subunits. Mol. Cell 1999, 3, 247–253. [Google Scholar] [CrossRef]
  101. Adi, S.; Bin-Abbas, B.; Wu, N.Y.; Rosenthal, S.M. Early stimulation and late inhibition of extracellular signal-regulated kinase 1/2 phosphorylation by igf-i: A potential mechanism mediating the switch in igf-i action on skeletal muscle cell differentiation. Endocrinology 2002, 143, 511–516. [Google Scholar] [CrossRef] [PubMed]
  102. Coolican, S.A.; Samuel, D.S.; Ewton, D.Z.; McWade, F.J.; Florini, J.R. The mitogenic and myogenic actions of insulin-like growth factors utilize distinct signaling pathways. J. Biol. Chem. 1997, 272, 6653–6662. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Gredinger, E.; Gerber, A.N.; Tamir, Y.; Tapscott, S.J.; Bengal, E. Mitogen-activated protein kinase pathway is involved in the differentiation of muscle cells. J. Biol. Chem. 1998, 273, 10436–10444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Reed, S.A.; Ouellette, S.E.; Liu, X.; Allen, R.E.; Johnson, S.E. E2f5 and lek1 translocation to the nucleus is an early event demarcating myoblast quiescence. J. Cell. Biochem. 2007, 101, 1394–1408. [Google Scholar] [CrossRef]
  105. Cuenda, A.; Rousseau, S. P38 map-kinases pathway regulation, function and role in human diseases. Biochim. Biophys. Acta 2007, 1773, 1358–1375. [Google Scholar] [CrossRef] [Green Version]
  106. Lee, J.K.; Kim, N.J. Recent advances in the inhibition of p38 mapk as a potential strategy for the treatment of alzheimer’s disease. Molecules 2017, 22, 1287. [Google Scholar] [CrossRef] [Green Version]
  107. Risco, A.; Cuenda, A. New insights into the p38gamma and p38delta mapk pathways. J. Signal Transduct. 2012, 2012, 520289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Escos, A.; Risco, A.; Alsina-Beauchamp, D.; Cuenda, A. P38gamma and p38delta mitogen activated protein kinases (mapks), new stars in the mapk galaxy. Front Cell Dev. Biol. 2016, 4, 31. [Google Scholar] [CrossRef] [Green Version]
  109. Cuadrado, A.; Nebreda, A.R. Mechanisms and functions of p38 mapk signalling. Biochem. J. 2010, 429, 403–417. [Google Scholar] [CrossRef] [Green Version]
  110. Kumar, S.; Boehm, J.; Lee, J.C. P38 map kinases: Key signalling molecules as therapeutic targets for inflammatory diseases. Nat. Rev. Drug Discov. 2003, 2, 717–726. [Google Scholar] [CrossRef]
  111. Kyriakis, J.M.; Avruch, J. Protein kinase cascades activated by stress and inflammatory cytokines. BioEssays News Rev. Mol. Cell. Dev. Biol. 1996, 18, 567–577. [Google Scholar] [CrossRef] [PubMed]
  112. Kim, E.K.; Choi, E.J. Compromised mapk signaling in human diseases: An update. Arch. Toxicol. 2015, 89, 867–882. [Google Scholar] [CrossRef] [PubMed]
  113. Xu, Q.; Yu, L.; Liu, L.; Cheung, C.F.; Li, X.; Yee, S.P.; Yang, X.J.; Wu, Z. P38 mitogen-activated protein kinase-, calcium-calmodulin-dependent protein kinase-, and calcineurin-mediated signaling pathways transcriptionally regulate myogenin expression. Mol. Biol. Cell 2002, 13, 1940–1952. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Simone, C.; Forcales, S.V.; Hill, D.A.; Imbalzano, A.N.; Latella, L.; Puri, P.L. P38 pathway targets swi-snf chromatin-remodeling complex to muscle-specific loci. Nat. Genet. 2004, 36, 738–743. [Google Scholar] [CrossRef] [Green Version]
  115. Lluis, F.; Ballestar, E.; Suelves, N.; Esteller, M.; Munoz-Canoves, P. E47 phosphorylation by p38 mapk promotes myod/e47 association and muscle-specific gene transcription. EMBO J. 2005, 24, 974–984. [Google Scholar] [CrossRef] [Green Version]
  116. Lluis, F.; Perdiguero, E.; Nebreda, A.R.; Munoz-Canoves, P. Regulation of skeletal muscle gene expression by p38 map kinases. Trends Cell. Biol. 2006, 16, 36–44. [Google Scholar] [CrossRef]
  117. Keren, A.; Tamir, Y.; Bengal, E. The p38 mapk signaling pathway: A major regulator of skeletal muscle development. Mol. Cell. Endocrinol. 2006, 252, 224–230. [Google Scholar] [CrossRef] [PubMed]
  118. Rampalli, S.; Li, L.; Mak, E.; Ge, K.; Brand, M.; Tapscott, S.J.; Dilworth, F.J. P38 mapk signaling regulates recruitment of ash2l-containing methyltransferase complexes to specific genes during differentiation. Nat. Struct. Mol. Biol. 2007, 14, 1150–1156. [Google Scholar] [CrossRef]
  119. Perdiguero, E.; Ruiz-Bonilla, V.; Serrano, A.L.; Munoz-Canoves, P. Genetic deficiency of p38alpha reveals its critical role in myoblast cell cycle exit: The p38alpha-jnk connection. Cell Cycle 2007, 6, 1298–1303. [Google Scholar] [CrossRef] [Green Version]
  120. 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. Genetic analysis of p38 map kinases in myogenesis: Fundamental role of p38alpha in abrogating myoblast proliferation. EMBO J. 2007, 26, 1245–1256. [Google Scholar] [CrossRef] [Green Version]
  121. Ruiz-Bonilla, V.; Perdiguero, E.; Gresh, L.; Serrano, A.L.; Zamora, M.; Sousa-Victor, P.; Jardi, M.; Wagner, E.F.; Munoz-Canoves, P. Efficient adult skeletal muscle regeneration in mice deficient in p38beta, p38gamma and p38delta map kinases. Cell Cycle 2008, 7, 2208–2214. [Google Scholar] [CrossRef] [PubMed]
  122. Penn, B.H.; Bergstrom, D.A.; Dilworth, F.J.; Bengal, E.; Tapscott, S.J. A myod-generated feed-forward circuit temporally patterns gene expression during skeletal muscle differentiation. Genes Dev. 2004, 18, 2348–2353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Bergstrom, D.A.; Penn, B.H.; Strand, A.; Perry, R.L.; Rudnicki, M.A.; Tapscott, S.J. Promoter-specific regulation of myod binding and signal transduction cooperate to pattern gene expression. Mol. Cell 2002, 9, 587–600. [Google Scholar] [CrossRef]
  124. Gillespie, M.A.; Le Grand, F.; Scime, A.; Kuang, S.; von Maltzahn, J.; Seale, V.; Cuenda, A.; Ranish, J.A.; Rudnicki, M.A. P38-[190]-dependent gene silencing restricts entry into the myogenic differentiation program. J. Cell Biol. 2009, 187, 991–1005. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Cuenda, A.; Cohen, P. Stress-activated protein kinase-2/p38 and a rapamycin-sensitive pathway are required for c2c12 myogenesis. J. Biol. Chem. 1999, 274, 4341–4346. [Google Scholar] [CrossRef] [Green Version]
  126. Wu, Z.; Woodring, P.J.; Bhakta, K.S.; Tamura, K.; Wen, F.; Feramisco, J.R.; Karin, M.; Wang, J.Y.; Puri, P.L. P38 and extracellular signal-regulated kinases regulate the myogenic program at multiple steps. Mol. Cell. Biol. 2000, 20, 3951–3964. [Google Scholar] [CrossRef] [Green Version]
  127. Zetser, A.; Gredinger, E.; Bengal, E. P38 mitogen-activated protein kinase pathway promotes skeletal muscle differentiation. Participation of the mef2c transcription factor. J. Biol. Chem. 1999, 274, 5193–5200. [Google Scholar] [CrossRef] [Green Version]
  128. Palacios, D.; Mozzetta, C.; Consalvi, S.; Caretti, G.; Saccone, V.; Proserpio, V.; Marquez, V.E.; Valente, S.; Mai, A.; Forcales, S.V.; et al. Tnf/p38alpha/polycomb signaling to pax7 locus in satellite cells links inflammation to the epigenetic control of muscle regeneration. Cell Stem Cell 2010, 7, 455–469. [Google Scholar] [CrossRef] [Green Version]
  129. Ornatsky, O.I.; Cox, D.M.; Tangirala, P.; Andreucci, J.J.; Quinn, Z.A.; Wrana, J.L.; Prywes, R.; Yu, Y.T.; McDermott, J.C. Post-translational control of the mef2a transcriptional regulatory protein. Nucleic Acids Res. 1999, 27, 2646–2654. [Google Scholar] [CrossRef] [Green Version]
  130. Zhao, M.; New, L.; Kravchenko, V.V.; Kato, Y.; Gram, H.; di Padova, F.; Olson, E.N.; Ulevitch, R.J.; Han, J. Regulation of the mef2 family of transcription factors by p38. Mol. Cell. Biol. 1999, 19, 21–30. [Google Scholar] [CrossRef] [Green Version]
  131. Taylor, M.V.; Hughes, S.M. Mef2 and the skeletal muscle differentiation program. Semin. Cell Dev. Biol. 2017, 72, 33–44. [Google Scholar] [CrossRef]
  132. de la Serna, I.L.; Ohkawa, Y.; Berkes, C.A.; Bergstrom, D.A.; Dacwag, C.S.; Tapscott, S.J.; Imbalzano, A.N. Myod targets chromatin remodeling complexes to the myogenin locus prior to forming a stable DNA-bound complex. Mol. Cell. Biol. 2005, 25, 3997–4009. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Forcales, S.V.; Albini, S.; Giordani, L.; Malecova, B.; Cignolo, L.; Chernov, A.; Coutinho, P.; Saccone, V.; Consalvi, S.; Williams, R.; et al. Signal-dependent incorporation of myod-baf60c into brg1-based swi/snf chromatin-remodelling complex. EMBO J. 2012, 31, 301–316. [Google Scholar] [CrossRef] [PubMed]
  134. Takeuchi, J.K.; Bruneau, B.G. Directed transdifferentiation of mouse mesoderm to heart tissue by defined factors. Nature 2009, 459, 708–711. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Manning, B.D.; Toker, A. Akt/pkb signaling: Navigating the network. Cell 2017, 169, 381–405. [Google Scholar] [CrossRef] [Green Version]
  136. Gonzalez, I.; Tripathi, G.; Carter, E.J.; Cobb, L.J.; Salih, D.A.; Lovett, F.A.; Holding, C.; Pell, J.M. Akt2, a novel functional link between p38 mitogen-activated protein kinase and phosphatidylinositol 3-kinase pathways in myogenesis. Mol. Cell. Biol. 2004, 24, 3607–3622. [Google Scholar] [CrossRef] [Green Version]
  137. Fujio, Y.; Mitsuuchi, Y.; Testa, J.R.; Walsh, K. Activation of akt2 inhibits anoikis and apoptosis induced by myogenic differentiation. Cell Death Differ. 2001, 8, 1207–1212. [Google Scholar] [CrossRef] [Green Version]
  138. Vandromme, M.; Rochat, A.; Meier, R.; Carnac, G.; Besser, D.; Hemmings, B.A.; Fernandez, A.; Lamb, N.J. Protein kinase b beta/akt2 plays a specific role in muscle differentiation. J. Biol. Chem. 2001, 276, 8173–8179. [Google Scholar] [CrossRef] [Green Version]
  139. Heron-Milhavet, L.; Franckhauser, C.; Rana, V.; Berthenet, C.; Fisher, D.; Hemmings, B.A.; Fernandez, A.; Lamb, N.J. Only akt1 is required for proliferation, while akt2 promotes cell cycle exit through p21 binding. Mol. Cell. Biol. 2006, 26, 8267–8280. [Google Scholar] [CrossRef] [Green Version]
  140. Heron-Milhavet, L.; Mamaeva, D.; Rochat, A.; Lamb, N.J.; Fernandez, A. Akt2 is implicated in skeletal muscle differentiation and specifically binds prohibitin2/rea. J. Cell. Physiol. 2008, 214, 158–165. [Google Scholar] [CrossRef]
  141. Sumitani, S.; Goya, K.; Testa, J.R.; Kouhara, H.; Kasayama, S. Akt1 and akt2 differently regulate muscle creatine kinase and myogenin gene transcription in insulin-induced differentiation of c2c12 myoblasts. Endocrinology 2002, 143, 820–828. [Google Scholar] [CrossRef] [PubMed]
  142. Lee, S.J.; Hwang, J.; Jeong, H.J.; Yoo, M.; Go, G.Y.; Lee, J.R.; Leem, Y.E.; Park, J.W.; Seo, D.W.; Kim, Y.K.; et al. Pkn2 and cdo interact to activate akt and promote myoblast differentiation. Cell Death Dis. 2016, 7, e2431. [Google Scholar] [CrossRef] [Green Version]
  143. Serra, C.; Palacios, D.; Mozzetta, C.; Forcales, S.V.; Morantte, I.; Ripani, M.; Jones, D.R.; Du, K.; Jhala, U.S.; Simone, C.; et al. Functional interdependence at the chromatin level between the mkk6/p38 and igf1/pi3k/akt pathways during muscle differentiation. Mol. Cell 2007, 28, 200–213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Meggio, F.; Pinna, L.A. One-thousand-and-one substrates of protein kinase ck2? FASEB J. 2003, 17, 349–368. [Google Scholar] [CrossRef] [PubMed]
  145. Litchfield, D.W. Protein kinase ck2: Structure, regulation and role in cellular decisions of life and death. Biochem. J. 2003, 369, 1–15. [Google Scholar] [CrossRef]
  146. Nunez de Villavicencio-Diaz, T.; Rabalski, A.J.; Litchfield, D.W. Protein kinase ck2: Intricate relationships within regulatory cellular networks. Pharmaceuticals 2017, 10, 27. [Google Scholar] [CrossRef] [Green Version]
  147. Pepperkok, R.; Lorenz, P.; Jakobi, R.; Ansorge, W.; Pyerin, W. Cell growth stimulation by egf: Inhibition through antisense-oligodeoxynucleotides demonstrates important role of casein kinase ii. Exp. Cell Res. 1991, 197, 245–253. [Google Scholar] [CrossRef]
  148. Lorenz, P.; Pepperkok, R.; Ansorge, W.; Pyerin, W. Cell biological studies with monoclonal and polyclonal antibodies against human casein kinase ii subunit beta demonstrate participation of the kinase in mitogenic signaling. J. Biol. Chem. 1993, 268, 2733–2739. [Google Scholar]
  149. Pepperkok, R.; Lorenz, P.; Ansorge, W.; Pyerin, W. Casein kinase ii is required for transition of g0/g1, early g1, and g1/s phases of the cell cycle. J. Biol. Chem. 1994, 269, 6986–6991. [Google Scholar]
  150. Lorenz, P.; Pepperkok, R.; Pyerin, W. Requirement of casein kinase 2 for entry into and progression through early phases of the cell cycle. Cell. Mol. Biol. Res. 1994, 40, 519–527. [Google Scholar]
  151. Gotz, C.; Montenarh, M. Protein kinase ck2 in development and differentiation. Biomed. Rep. 2017, 6, 127–133. [Google Scholar] [CrossRef] [PubMed]
  152. Winter, B.; Kautzner, I.; Issinger, O.G.; Arnold, H.H. Two putative protein kinase ck2 phosphorylation sites are important for myf-5 activity. Biol. Chem. 1997, 378, 1445–1456. [Google Scholar] [CrossRef] [PubMed]
  153. Johnson, S.E.; Wang, X.; Hardy, S.; Taparowsky, E.J.; Konieczny, S.F. Casein kinase ii increases the transcriptional activities of mrf4 and myod independently of their direct phosphorylation. Mol. Cell. Biol. 1996, 16, 1604–1613. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Dietz, K.N.; Miller, P.J.; Hollenbach, A.D. Phosphorylation of serine 205 by the protein kinase ck2 persists on pax3-foxo1, but not pax3, throughout early myogenic differentiation. Biochemistry 2009, 48, 11786–11795. [Google Scholar] [CrossRef] [Green Version]
  155. Dietz, K.N.; Miller, P.J.; Iyengar, A.S.; Loupe, J.M.; Hollenbach, A.D. Identification of serines 201 and 209 as sites of pax3 phosphorylation and the altered phosphorylation status of pax3-foxo1 during early myogenic differentiation. Int. J. Biochem. Cell Biol. 2011, 43, 936–945. [Google Scholar] [CrossRef] [Green Version]
  156. Iyengar, A.S.; Loupe, J.M.; Miller, P.J.; Hollenbach, A.D. Identification of ck2 as the kinase that phosphorylates pax3 at ser209 in early myogenic differentiation. Biochem. Biophys. Res. Commun. 2012, 428, 24–30. [Google Scholar] [CrossRef] [Green Version]
  157. Dick, S.A.; Chang, N.C.; Dumont, N.A.; Bell, R.A.; Putinski, C.; Kawabe, Y.; Litchfield, D.W.; Rudnicki, M.A.; Megeney, L.A. Caspase 3 cleavage of pax7 inhibits self-renewal of satellite cells. Proc. Natl. Acad. Sci. USA 2015, 112, E5246–E5252. [Google Scholar] [CrossRef] [Green Version]
  158. Gonzalez, N.; Moresco, J.J.; Cabezas, F.; de la Vega, E.; Bustos, F.; Yates, J.R., 3rd; Olguin, H.C. Ck2-dependent phosphorylation is required to maintain pax7 protein levels in proliferating muscle progenitors. PLoS ONE 2016, 11, e0154919. [Google Scholar] [CrossRef]
  159. Miller, P.J.; Dietz, K.N.; Hollenbach, A.D. Identification of serine 205 as a site of phosphorylation on pax3 in proliferating but not differentiating primary myoblasts. Protein Sci. 2008, 17, 1979–1986. [Google Scholar] [CrossRef] [Green Version]
  160. Salizzato, V.; Zanin, S.; Borgo, C.; Lidron, E.; Salvi, M.; Rizzuto, R.; Pallafacchina, G.; Donella-Deana, A. Protein kinase ck2 subunits exert specific and coordinated functions in skeletal muscle differentiation and fusogenic activity. FASEB J. 2019, fj201801833RR. [Google Scholar] [CrossRef]
  161. Kuang, S.; Charge, S.B.; Seale, P.; Huh, M.; Rudnicki, M.A. Distinct roles for pax7 and pax3 in adult regenerative myogenesis. J. Cell Biol. 2006, 172, 103–113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Chin, E.R.; Olson, E.N.; Richardson, J.A.; Yang, Q.; Humphries, C.; Shelton, J.M.; Wu, H.; Zhu, W.; Bassel-Duby, R.; Williams, R.S. A calcineurin-dependent transcriptional pathway controls skeletal muscle fiber type. Genes Dev. 1998, 12, 2499–2509. [Google Scholar] [CrossRef] [Green Version]
  163. Narayanan, R.; Pirouz, M.; Kerimoglu, C.; Pham, L.; Wagener, R.J.; Kiszka, K.A.; Rosenbusch, J.; Seong, R.H.; Kessel, M.; Fischer, A.; et al. Loss of baf (mswi/snf) complexes causes global transcriptional and chromatin state changes in forebrain development. Cell Rep. 2015, 13, 1842–1854. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Ho, L.; Ronan, J.L.; Wu, J.; Staahl, B.T.; Chen, L.; Kuo, A.; Lessard, J.; Nesvizhskii, A.I.; Ranish, J.; Crabtree, G.R. An embryonic stem cell chromatin remodeling complex, esbaf, is essential for embryonic stem cell self-renewal and pluripotency. Proc. Natl. Acad. Sci. USA 2009, 106, 5181–5186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Gottesfeld, J.M.; Forbes, D.J. Mitotic repression of the transcriptional machinery. Trends Biochem. Sci. 1997, 22, 197–202. [Google Scholar] [CrossRef]
  166. Hartl, P.; Gottesfeld, J.; Forbes, D.J. Mitotic repression of transcription in vitro. J. Cell Biol. 1993, 120, 613–624. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Gottesfeld, J.M.; Wolf, V.J.; Dang, T.; Forbes, D.J.; Hartl, P. Mitotic repression of rna polymerase iii transcription in vitro mediated by phosphorylation of a tfiiib component. Science 1994, 263, 81–84. [Google Scholar] [CrossRef]
  168. Segil, N.; Guermah, M.; Hoffmann, A.; Roeder, R.G.; Heintz, N. Mitotic regulation of tfiid: Inhibition of activator-dependent transcription and changes in subcellular localization. Genes Dev. 1996, 10, 2389–2400. [Google Scholar] [CrossRef] [Green Version]
  169. Martinez-Balbas, M.A.; Dey, A.; Rabindran, S.K.; Ozato, K.; Wu, C. Displacement of sequence-specific transcription factors from mitotic chromatin. Cell 1995, 83, 29–38. [Google Scholar] [CrossRef]
  170. Stewart, A.A.; Ingebritsen, T.S.; Manalan, A.; Klee, C.B.; Cohen, P. Discovery of a Ca2+- and calmodulin-dependent protein phosphatase: Probable identity with calcineurin (cam-bp80). FEBS Lett. 1982, 137, 80–84. [Google Scholar] [CrossRef] [Green Version]
  171. Mitchell, P.O.; Pavlath, G.K. Multiple roles of calcineurin in skeletal muscle growth. Clin. Orthop. Relat. Res. 2002, S197–S202. [Google Scholar] [CrossRef] [PubMed]
  172. Bassel-Duby, R.; Olson, E.N. Role of calcineurin in striated muscle: Development, adaptation, and disease. Biochem. Biophys. Res. Commun. 2003, 311, 1133–1141. [Google Scholar] [CrossRef] [PubMed]
  173. Rusnak, F.; Mertz, P. Calcineurin: Form and function. Physiol. Rev. 2000, 80, 1483–1521. [Google Scholar] [CrossRef] [PubMed]
  174. Manalan, A.S.; Krinks, M.H.; Klee, C.B. Calcineurin: A member of a family of calmodulin-stimulated protein phosphatases. Proc. Soc. Exp. Biol. Med. 1984, 177, 12–16. [Google Scholar] [CrossRef]
  175. Liu, J.; Farmer, J.D., Jr.; Lane, W.S.; Friedman, J.; Weissman, I.; Schreiber, S.L. Calcineurin is a common target of cyclophilin-cyclosporin a and fkbp-fk506 complexes. Cell 1991, 66, 807–815. [Google Scholar] [CrossRef]
  176. Shaw, K.T.; Ho, A.M.; Raghavan, A.; Kim, J.; Jain, J.; Park, J.; Sharma, S.; Rao, A.; Hogan, P.G. Immunosuppressive drugs prevent a rapid dephosphorylation of transcription factor nfat1 in stimulated immune cells. Proc. Natl. Acad. Sci. USA 1995, 92, 11205–11209. [Google Scholar] [CrossRef] [Green Version]
  177. Flanagan, W.M.; Corthesy, B.; Bram, R.J.; Crabtree, G.R. Nuclear association of a t-cell transcription factor blocked by fk-506 and cyclosporin a. Nature 1991, 352, 803–807. [Google Scholar] [CrossRef]
  178. Ruff, V.A.; Leach, K.L. Direct demonstration of nfatp dephosphorylation and nuclear localization in activated ht-2 cells using a specific nfatp polyclonal antibody. J. Biol. Chem. 1995, 270, 22602–22607. [Google Scholar] [CrossRef] [Green Version]
  179. Ravel-Chapuis, A.; Belanger, G.; Cote, J.; Michel, R.N.; Jasmin, B.J. Misregulation of calcium-handling proteins promotes hyperactivation of calcineurin-nfat signaling in skeletal muscle of dm1 mice. Hum. Mol. Genet. 2017, 26, 2192–2206. [Google Scholar] [CrossRef]
  180. Schiaffino, S.; Sandri, M.; Murgia, M. Activity-dependent signaling pathways controlling muscle diversity and plasticity. Physiol. Bethesda 2007, 22, 269–278. [Google Scholar] [CrossRef] [Green Version]
  181. Horsley, V.; Friday, B.B.; Matteson, S.; Kegley, K.M.; Gephart, J.; Pavlath, G.K. Regulation of the growth of multinucleated muscle cells by an nfatc2-dependent pathway. J. Cell Biol. 2001, 153, 329–338. [Google Scholar] [CrossRef] [PubMed]
  182. Dunn, S.E.; Burns, J.L.; Michel, R.N. Calcineurin is required for skeletal muscle hypertrophy. J. Biol. Chem. 1999, 274, 21908–21912. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Musaro, A.; McCullagh, K.J.; Naya, F.J.; Olson, E.N.; Rosenthal, N. Igf-1 induces skeletal myocyte hypertrophy through calcineurin in association with gata-2 and nf-atc1. Nature 1999, 400, 581–585. [Google Scholar] [CrossRef] [PubMed]
  184. Semsarian, C.; Wu, M.J.; Ju, Y.K.; Marciniec, T.; Yeoh, T.; Allen, D.G.; Harvey, R.P.; Graham, R.M. Skeletal muscle hypertrophy is mediated by a ca2+-dependent calcineurin signalling pathway. Nature 1999, 400, 576–581. [Google Scholar] [CrossRef] [PubMed]
  185. Friday, B.B.; Horsley, V.; Pavlath, G.K. Calcineurin activity is required for the initiation of skeletal muscle differentiation. J. Cell Biol. 2000, 149, 657–666. [Google Scholar] [CrossRef]
  186. Friday, B.B.; Mitchell, P.O.; Kegley, K.M.; Pavlath, G.K. Calcineurin initiates skeletal muscle differentiation by activating mef2 and myod. Differ. Res. Biol. Divers. 2003, 71, 217–227. [Google Scholar] [CrossRef]
  187. Lai, D.; Wan, M.; Wu, J.; Preston-Hurlburt, P.; Kushwaha, R.; Grundstrom, T.; Imbalzano, A.N.; Chi, T. Induction of tlr4-target genes entails calcium/calmodulin-dependent regulation of chromatin remodeling. Proc. Natl. Acad. Sci. USA 2009, 106, 1169–1174. [Google Scholar] [CrossRef] [Green Version]
  188. Witwicka, H.; Nogami, J.; Syed, S.A.; Maehara, K.; Padilla-Benavides, T.; Ohkawa, Y.; Imbalzano, A.N. Calcineurin broadly regulates the initiation of skeletal muscle-specific gene expression by binding target promoters and facilitating the interaction of the swi/snf chromatin remodeling enzyme. Mol. Cell. Biol. 2019, 39, e00063-19. [Google Scholar] [CrossRef] [Green Version]
  189. Kwon, S.J.; Park, J.H.; Park, E.J.; Lee, S.A.; Lee, H.S.; Kang, S.W.; Kwon, J. Atm-mediated phosphorylation of the chromatin remodeling enzyme brg1 modulates DNA double-strand break repair. Oncogene 2015, 34, 303–313. [Google Scholar] [CrossRef]
Figure 1. P38 has pleiotropic effects during skeletal muscle differentiation. Phosphorylation of transcription factors (TFs) (E-proteins, Mef2) and the mSWI/SNF subunit BAF60c by p38 enables expression of myogenic genes and differentiation. P38-dependent phosphorylation of Ezh2 downregulates the expression of Pax7. P38 also induces cell cycle withdrawal by activating c-Jun N-terminal kinase (JNK) and Cyclin D1.
Figure 1. P38 has pleiotropic effects during skeletal muscle differentiation. Phosphorylation of transcription factors (TFs) (E-proteins, Mef2) and the mSWI/SNF subunit BAF60c by p38 enables expression of myogenic genes and differentiation. P38-dependent phosphorylation of Ezh2 downregulates the expression of Pax7. P38 also induces cell cycle withdrawal by activating c-Jun N-terminal kinase (JNK) and Cyclin D1.
Biology 09 00152 g001
Figure 2. Casein kinase 2 (CK2) modulates mSWI/SNF activity in proliferating myoblasts. CK2-dependent phosphorylation of Brahma related gene 1 (BRG1) regulates the sub-nuclear localization and the subunit composition of the mSWI/SNF complex, viability and cell cycle progression, and the ability to remodel promoter chromatin and promote gene expression that allows myoblast proliferation. The diagram shows the presence or absence of phosphorylation of BRG1 for illustrative purposes. The exact number of phosphorylation sites in the presence of the different mSWI/SNF subunits has not been determined.
Figure 2. Casein kinase 2 (CK2) modulates mSWI/SNF activity in proliferating myoblasts. CK2-dependent phosphorylation of Brahma related gene 1 (BRG1) regulates the sub-nuclear localization and the subunit composition of the mSWI/SNF complex, viability and cell cycle progression, and the ability to remodel promoter chromatin and promote gene expression that allows myoblast proliferation. The diagram shows the presence or absence of phosphorylation of BRG1 for illustrative purposes. The exact number of phosphorylation sites in the presence of the different mSWI/SNF subunits has not been determined.
Biology 09 00152 g002
Figure 3. Antagonistic roles of PKCβ1 and calcineurin in myogenesis. PKCβ1 interacts with and phosphorylates BRG1 to block differentiation. Upon myogenic stimulus, calcineurin dephosphorylates BRG1 and the TF nuclear factor of activated T-cell (NFAT), enabling chromatin remodeling at myogenic loci, myogenic gene expression, and skeletal muscle differentiation.
Figure 3. Antagonistic roles of PKCβ1 and calcineurin in myogenesis. PKCβ1 interacts with and phosphorylates BRG1 to block differentiation. Upon myogenic stimulus, calcineurin dephosphorylates BRG1 and the TF nuclear factor of activated T-cell (NFAT), enabling chromatin remodeling at myogenic loci, myogenic gene expression, and skeletal muscle differentiation.
Biology 09 00152 g003
Figure 4. Kinases and phosphatases that modulate the chromatin remodeling activity of the mSWI/SNF complex during myogenesis. Schematic representation of post-natal myogenesis and indications of when specific kinases and phosphatases that control the activity of the mSWI/SNF complex act.
Figure 4. Kinases and phosphatases that modulate the chromatin remodeling activity of the mSWI/SNF complex during myogenesis. Schematic representation of post-natal myogenesis and indications of when specific kinases and phosphatases that control the activity of the mSWI/SNF complex act.
Biology 09 00152 g004

Share and Cite

MDPI and ACS Style

Padilla-Benavides, T.; Reyes-Gutierrez, P.; Imbalzano, A.N. Regulation of the Mammalian SWI/SNF Family of Chromatin Remodeling Enzymes by Phosphorylation during Myogenesis. Biology 2020, 9, 152. https://doi.org/10.3390/biology9070152

AMA Style

Padilla-Benavides T, Reyes-Gutierrez P, Imbalzano AN. Regulation of the Mammalian SWI/SNF Family of Chromatin Remodeling Enzymes by Phosphorylation during Myogenesis. Biology. 2020; 9(7):152. https://doi.org/10.3390/biology9070152

Chicago/Turabian Style

Padilla-Benavides, Teresita, Pablo Reyes-Gutierrez, and Anthony N. Imbalzano. 2020. "Regulation of the Mammalian SWI/SNF Family of Chromatin Remodeling Enzymes by Phosphorylation during Myogenesis" Biology 9, no. 7: 152. https://doi.org/10.3390/biology9070152

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop