Next Article in Journal
Clinical Characteristics of 6102 Asymptomatic and Mild Cases for Patients with COVID-19 in Indonesia
Next Article in Special Issue
Long Non-Coding RNAs as “MYC Facilitators”
Previous Article in Journal
Pathophysiology of Red Blood Cell Dysfunction in Diabetes and Its Complications
Previous Article in Special Issue
Shift of N-MYC Oncogene Expression in AML Patients Carrying the FLT3-ITD Mutation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Myc beyond Cancer: Regulation of Mammalian Tissue Regeneration

Institute of Molecular Biology and Pathology, National Research Council, c/o Department of Biology and Biotechnology “Charles Darwin”, Sapienza University of Rome, 00185 Rome, Italy
*
Authors to whom correspondence should be addressed.
Pathophysiology 2023, 30(3), 346-365; https://doi.org/10.3390/pathophysiology30030027
Submission received: 5 July 2023 / Revised: 28 July 2023 / Accepted: 31 July 2023 / Published: 2 August 2023
(This article belongs to the Special Issue MYC in Regeneration and Tumorigenesis)

Abstract

:
Myc is one of the most well-known oncogenes driving tumorigenesis in a wide variety of tissues. From the brain to blood, its deregulation derails physiological pathways that grant the correct functioning of the cell. Its action is carried out at the gene expression level, where Myc governs basically every aspect of transcription. Indeed, in addition to its role as a canonical, chromatin-bound transcription factor, Myc rules RNA polymerase II (RNAPII) transcriptional pause–release, elongation and termination and mRNA capping. For this reason, it is evident that minimal perturbations of Myc function mirror malignant cell behavior and, consistently, a large body of literature mainly focuses on Myc malfunctioning. In healthy cells, Myc controls molecular mechanisms involved in pivotal functions, such as cell cycle (and proliferation thereof), apoptosis, metabolism and cell size, angiogenesis, differentiation and stem cell self-renewal. In this latter regard, Myc has been found to also regulate tissue regeneration, a hot topic in the research fields of aging and regenerative medicine. Indeed, Myc appears to have a role in wound healing, in peripheral nerves and in liver, pancreas and even heart recovery. Herein, we discuss the state of the art of Myc’s role in tissue regeneration, giving an overview of its potent action beyond cancer.

1. Introduction

Regeneration is a complex, energy-demanding process that is essential for organisms’ survival, as it normally occurs in homeostasis or during the replacement of damaged or lost tissues. This process strictly resembles embryonic development, but it is not identical, as some molecular features diverge [1]. In humans, different tissues have diverse regeneration capacity, which depends mainly on the loss and damage entity [2,3]. Liver, blood, gut, skin, the endothelium and skeletal muscle are examples of tissues that regenerate throughout life. Conversely, the nervous (both central and peripheral) system and the heart possess limited reparative capacity. This characteristic of terminally differentiated tissues provides the background for the onset of diseases typical in aged individuals, such as heart failure and neurodegeneration (when this latter lacks genetic bases).
Regeneration requires the execution of complex gene expression programs able to orchestrate reconstitution from cell precursors, which differ in their origin according to the tissue to which they belong. Indeed, skin, blood and skeletal muscle are provided with stem cells niches which undergo cell division and differentiation upon homeostatic or injury signals. Other tissues, such as liver and the endothelium, rely on tissue-specific cell types for self-duplication and repair.
Myc is a basic helix–loop–helix–lecine zipper (bHLHZip) transcription factor (TF) [4,5] that controls pivotal cell functions, such as proliferation, apoptosis [6,7,8,9], metabolism [10,11,12] and senescence [13,14]. Of note, it is highly important for embryonic stem cells’ (ES) self-renewal [15] and somatic cell reprogramming. Indeed, it is one of the four Yamanaka factors [16,17], which are able to revert a differentiated phenotype into an ES-like ground state. Myc exerts its function at the gene expression level, but it is no longer considered as a canonical transcription factor which is bound to chromatin at promoters/enhancers, recruiting transcriptional co-activators and, thus, allowing RNA Polymerases (RNAPs)-dependent transcription to start. Surprisingly, Myc has been found to regulate all the three RNAPs; it has been demonstrated to control RNAPII-dependent activities, such as pause–release [18], elongation [19] and termination [20], as well as to ensure splicing fidelity [21], to influence chromatin topology [22] and to indirectly regulate transcription controlling the expression of microRNAs (miRNAs) [23]. For these reasons, any perturbation of Myc function has a profound impact on cell behavior. This is particularly evident in cancer cells, where Myc is frequently deregulated and overexpressed [24]. In healthy cells, Myc is tightly controlled and essential for survival; indeed, Myc-targeted anti-cancer therapies have to face the detrimental effects that decreased Myc activity may produce in normal cells. Tissue-specific resident stem cells need Myc to maintain their self-renewal and to proliferate; furthermore, reactivation of Myc-dependent transcriptional programs has been found as of major importance for the regeneration of tissues and for strategies aimed at counteracting the tissue degeneration typical of aging. In this review paper, we discuss the role Myc has in stem cell biology and, consequently, in tissue regeneration and repair, further underpinning its importance as a master regulator of cell functions beyond cancer.

2. Overview of Myc-Dependent Transcription

Myc is represented by three paralogs: c-myc, n-myc [25] and l-myc [26]. The c-myc (hereafter simply myc) gene is embedded in chromosome 8 long arm (8q24) and encodes a 439 amino-acid protein with a molecular mass of 64KD. Myc structure contains two highly conserved Myc polypeptides at the N-terminal and C-terminal regions. The N-terminal transactivating domain (TAD) holds three conserved sub-domains called Myc boxes (MB0, MBI and MBII) [27,28]. MBI recruits the cyclin CDK-complex positive transcription elongation factor b (P-TEFb), which phosphorylates RNAPII, allowing transcriptional elongation [18]. MBII rules the activation and repression of Myc target genes, promotes cellular transformation and is involved in Myc turnover. MBIII binds histone deacetylase 3 (HDAC3), promoting transcriptional repression, whilst MBIV is supposed to govern Myc pro-apoptotic activity. The C-terminal domain is provided with a 100-amino-acid bHLH-LZ domain, which represents the docking site for another small bHLH-LZ protein called Max, [29]. The Myc/Max interaction is stable. The heterodimer directly binds DNA “Enhancer boxes” (E-boxes) to induce transcription [30]. A proline-, glutamic acid-, serine- and threonine-rich region (PEST) is located across MBIII and MBIV boxes and two nuclear localization sequences are present within the MBIV and C-terminal basic region (Figure 1).
Interestingly, Myc has an intrinsically disordered structure. Intrinsically disordered-structure proteins (iDPs) are a challenge for scientists for a number of reasons [31]. Indeed, they cannot be studied with the standard methods used for folded proteins and present consistent deviation from the paradigm that structure = function. However, it is widely accepted that iDPs have biological roles that depend on their structure/dynamics and on their interaction with folded domains. Myc is basically organized in loose modules with unpacked secondary structure [32]. This is the main reason why the pharmacological inhibition of Myc is barely attainable, as it lacks pocket domains which may allocate small molecules. Specifically, the Myc bHLH-LZ domain is highly disordered and a conventional globular structure is only assumed when bound to Max (see below) [33] and DNA [34]. When looking at cancer cells, the only chance to interfere with Myc function is to perturb its binding with protein partners, especially with Max [35]. Indeed, a number of papers reported the screening and the selection of a number of Myc-/Max-interfering small molecules [36,37,38], whereas the Myc mutant termed Omomyc [39] has been proven to be highly effective in inhibiting Myc/Max heterodimerization and Myc binding to cognate DNA elements [40]. Most recently, other approaches have been attempted, although less efficient, such as the use of bromodomain and extraterminal domain inhibitors (BETi) to switch off Myc expression [41] and promote inhibition of Myc-interacting kinases (e.g., Aurora Kinase A) [42].

2.1. The Myc (Max) Network

Given the requirement of Max for Myc to exert its transcriptional activation properties, Max has to be considered as a hub for the so-called Myc network. Max is a ubiquitous, stable TF. It has a short half-life, which makes its activity strictly dependent on the amount of associated TFs. Max forms homodimers, competes with Myc/Max heterodimers for the binding to E-boxes and associates with the bHLH-LZ domain, which contains families of TFs Mad1-4 (Mad1, Mxi1, Mad3 and Mad4), Mnt and Mga. These latter are very similar to Myc, as they poorly homodimerize and bind DNA, whereas they form strong heterodimers with Max. The Mad1 and Mxi1 families of transcriptional repressors compete with Myc/Max for DNA binding. These TFs are able to interact with co-repressors, such as Sin3a and Sin3b, and recruit HDACs and other chromatin remodelers to repress transcription. A shift in the balance between Myc/Max and Mad/Max heterodimers is dictated by the amount of Mads within the cell, determining cell fate decision and a repositioning from proliferation/transformation to differentiation/quiescence [43]. From this evidence, it appears that Myc and Mad act in opposite ways to activate/repress transcription by competing for Max and for the binding to E-boxes [44] (Figure 2). In spite of the repressor function of Mad/Max heterodimers, Myc itself can specifically repress transcription by replacing p300 and binding the transcriptional activator Miz-1 [45].

2.2. Myc Transcription beyond DNA Binding

In recent years, the extensive knowledge about novel Myc functions led researchers to reconsider its role as a “classical” TF, promoting transcription through DNA binding and histone lysine acetyltransferases (KATs) recruitment, making chromatin permissive to the loading of the basal transcriptional machinery. Myc regulates all three RNAPs and controls transcripts output of mRNAs, non-coding(nc)RNAs, tRNAs and rRNAs. Importantly, Myc tightly controls RNAPII. In the last decade, the finding that Myc controls transcriptional pause–release was a milestone in the comprehension of some aspects of Myc-dependent transcription. By ChIPseq analyses, Rahl et al. were able to demonstrate that inhibiting Myc/Max heterodimer formation with the 10058-F4 small molecule caused a reduction in RNAPII in the transcribed regions of Myc-dependent gene in ES, with a consequent increase in RNAPII travelling ratio (TR) [18]. Immediately after, by high-throuhgput ChIPseq and RNAseq analyses, a number of works have revealed Myc as an amplifier of transcriptional programs already active or poised in normal and cancer cells [46,47,48]. This activity also depends on the ability of Myc to invade chromatin, even when the binding affinity is low. The more Myc is expressed, the more binding sites (both high- and low-affinity) are saturated [49]. Open chromatin at active promoters is a pivotal feature in Myc-bound DNA elements [50]. Further, enhancer loops adjacent to gene core promoters may put Myc in close proximity to other enhancers, when promoters are already saturated. More recently, Myc has been found to even associate with and activate topoisomerases 1 and 2 to sustain torsional stress due to high rates of transcription occurring during the amplification process [51]. Myc also controls RNAPII elongation [19], maintaisn the fidelity of splicing [21] and is involved in transcription termination through the Protein Arginine Methyltrasferases 5 (PRMT5) [20]. Myc also regulates mRNA capping by TFIIH kinase recruitment, RNAPII phosphorylation and cap RNA methyltransferase (RNMT) activation [52]. Moreover, Myc controls transcription of indirect target genes by regulating the expression of specific miRNAs, as also revealed by miRnome analysis of Myc-interfered cells [23,53].

3. The Role of Myc in Pluripotent Stem Cells

The fundamental importance of Myc for the maintenance of pluripotent cells stemness is proven by the evidence that c-myc knock out is embryonic lethal at days 9.5–10.5 of gestation, due to major deficiency in hematopoiesis and vascular ontogeny [54]. Furthermore, although Myc and N-Myc play redundant roles during pre- and early post-implantation stages of embryogenesis [55,56], at least one of them is required for the correct execution of early developmental programs in mouse ES [57,58].
Myc action in pluripotent stem cells is not limited to the control of self-renewal and proliferation potential, but is exerted also at the metabolic level and controls diverse signaling pathways. Overall, all these Myc activities may be attributed to the impact Myc has on the epigenome as a whole. In this section, highlights as to the multiple roles Myc plays in pluripotent stem cells are provided.

3.1. Myc and the Stem Cells Epigenome

Pluripotent stem cells’ epigenome is highly dynamic and characterized by an open chromatin conformation, which responds to two fundamental requirements: it has to be malleable to activate any developmental transcriptional program and has to possess a memory to recall its pluripotent state [59,60,61]. Stem cells’ chromatin typically shows histone marks of gene activation—i.e., acetylated histone H3 and H4 (H3Ac, H4Ac) and methylated histone H3 on lysine 4 (H3K4me3)—and has the unique feature to simultaneously present, on developmental gene promoters, histone marks of both transcriptional activation and repression, specifically, H3K4me3 and H3K27me3, respectively. This specific signature, the so-called bivalent mark, makes the downstream gene “poised” for activation or repression according to the chromatin reader recruited upon the appropriate stimulus [62,63]. Myc accounts for stem cells’ epigenome plasticity in a number of different ways. Its genomic location mirrors chromatin hyperacetylation and H3K4me3-enriched domains; indeed, Myc depletion results in a 70% decrease in H3K4me3 [64,65,66]. Furthermore, Myc directly controls the expression of chromatin remodelers such as members of the Polycomb repressive 1 and 2 (PRC1/2) complexes and Switch/Sucrose-non fermentable (SWI/SNF) complexes [67,68], interacts with KATs [66,69] and co-localizes with histone demethylases [70]. At the same time, it also represses the expression of developmental genes, interacting with co-repressors such as Sin3a and 3b, HDAC1/3 and LSD1 [71,72,73]. More importantly, by interacting with PRC2 and due to the overexpression of Suz12—a component of the PRC2 complex—it allows a massive deposition of H3K27me3 repressive marks on differentiation-related genes [74].
Inducible, long term overexpression of exogenous Myc makes ES independent from leukemia inhibitory factor (LIF). In these cells, wnt signaling induces endogenous myc and nmyc genes to maintain their pluripotency-activated programs, even after exogenous Myc deprivation, in a positive feedback loop, allowing ES cells to fully rely only on Myc for their stemness [74]. According to the reversible nature of epigenetic mechanisms [75], Myc-/Wnt-dependent ES fully reprogram to a JAK-/STAT3-regulated cascade upon LIF exposure, with a global change in the epigenetic state of chromatin regions regulating the expression of genes related to both pathways. These also evidence the role of Myc in the regulation and maintenance of stem cells’ epigenetic memory.

3.2. Myc in ES

As anticipated above, myc and n-myc are required for early development in ES. Simultaneous myc and n-myc deficiency causes premature ES endoderm and mesoderm differentiation [76]. Indeed, it is now widely accepted that Myc-dependent self-renewal of ES relies on its ability to suppress early differentiation programs. Myc exerts this function both directly and indirectly, regulating both Polycomb complexes and the expression of differentiation-related genes [66,74,77]. Myc-deficient cells arrest their development to the early progenitors’ stage and do not express markers of terminal differentiation [58]. This function of Myc is also bolstered by the finding that pluripotency markers, such as Oct-4 and Nanog, are Myc-independent. In fact, they continue to be expressed during early stages of development when they co-exist with differentiation markers [78], constituting a transcriptional module—the “core module”—well-defined with respect to the “Myc-module” [69]. The latter has been identified by using the in vivo metabolic biotin tagging method—used to tag well-known Myc and both direct and indirect Myc interactors—followed by mass spectrometry coupled with bioChIP-chip experiments, performed to underscore the genomic targets of the whole module. This Myc-centered cluster of TFs and chromatin remodelers, which includes the Tip60–Ep400 complex [79], does not interact with the core module and is located onto distinct, acetylated histone H3 (H3Ac)- and H4 (H4Ac)-enriched genomic loci. Of note, the Myc module occupies more genomic targets in ES than the core module, indicating Myc as a master regulator of ES transcriptional programs [69]. Specifically, 8000 chromatin elements have been found to be bound by Myc in ES, clearly different from genomic loci loaded with the core module factors [72]. This huge number of Myc-enriched chromatin sites suggest that, in addition to its direct role related to chromatin unwinding ability and RNAPs recruitment, in leading to transcriptional activation, Myc may act as a hot spot for the integration of different gene networks which ultimately grant ES pluripotency. Indeed, genes involved in cell cycle, metabolic and growth regulation belong to the Myc module [69,72]. However, an exception is represented by Myc direct regulation of Sox2 [66] (a component of the core module) and Gata6 (a determinant TF of endoderm differentiation) [57].
Other direct Myc targets in ES cells are cell cycle regulators. ES usually present a fast cell cycle progression and spend 50–60% of their time in S phase. This characteristic is consistent with the permanent expression of cycline-dependent kinases (CDKs) [80,81], with lack of retinoblastoma (RB) [82] and CDKs inhibitors (CDKi) such as p21 and INK4A [80,83]. Cell cycle regulators—including cyclins, CDKs and E2F—are regulated by Myc in ES either at the transcriptional or post-transcriptional level [48,68,72,77,84]. Further, Myc inhibits the expression of CDKi, competing with Miz-1 for chromatin binding, recruiting chromatin-remodeling enzymes or modulating the expression of other regulators [84,85]. In this latter regard, Myc binds to the promoter and upregulates mir17-92 cluster, which target p21, RB2 and cyclin D1 [57,77], forcing ES to proliferate.
All these Myc activities account, at least in part, for the maintenance of ES’ typical cell cycle structure, with very short G1 phase, impeding the start of differentiation by limiting chromatin access to chromatin remodelers [86,87,88].
As for the cell cycle, Myc also controls ES metabolism, which mainly depends on aerobic glycolysis [89], also known as the Warburg effect, a phenomenon extensively investigated in cancer cells [90]. Myc basically regulates any metabolic pathway, from glycolysis and glutaminolysis to nucleotides, ‘lipids’ and amino acids’ synthesis [91,92]. It has been demonstrated that, upon mir290-mediated derepression, Myc directly binds the promoters of pyruvate kinase M2 (PKM2) and lactate dehydrogenase A (LDHA) genes, allowing the glycolytic metabolic switch in ES [93]. Furthermore, since Myc modulates the level of metabolic intermediates such as α-ketogluatarate, acetyl-coenzyme A, nicotinamide adenine dinucleotide, flavin adenine dinucleotide and S-adenosylmethionine [10,11,12,94], all known to be associated with epigenetic activity [95,96,97,98], it is supposed to orchestrate epigenomic changes—represented by both histone acetylation and histone/DNA methylation—in ES, determining their fate across stemness maintenance or differentiation.

3.3. Myc and Somatic Cell Reprogramming

In the early 2000s, the work of Yamanaka’s lab [16] opened a new research avenue with large potential and enormous impact in the field of regenerative medicine. In fact, the generation of pluripotent stem cells from human fibroblasts theoretically became the basis for the reconstruction of any kind of damaged tissue. From then on, induced pluripotent stem cells (iPS) have been generated from many adult, tissue-specific cells [99,100] and have been induced to follow a wide variety of differentiation routes [101,102,103,104,105,106]. Substantially, Yamanaka’s protocol relies on the ability of defined TFs to revert the heterochromatic somatic cell epigenome into a euchromatic state typical of pluripotency (see Section 3.1). Oct4, Sox2, KLF-4 and Myc (OSKM) are sufficient to produce stem cells from somatic cells through a multistep program in which Myc and OSK modulate each other [107]. During this process, Myc, on the one hand, makes chromatin permissive to OSK binding and, on the other hand, represses fibroblast-specific genes and binds to pluripotency genes enhancers, strengthening reprogramming [108,109]. However, Myc requires OSK chromatin binding to exert its function in a positive feedback loop [109]. Early reports on the many attempts to substitute Myc in the Yamanaka’s cocktail, due to its oncogenic potential [110], displayed that reprogramming results were not very efficient [111,112]. However, a few years later, the use of L-Myc and of Myc mutants lacking transforming activity, but retaining Max binding properties, was found to be even more efficient than wild type Myc in generating iPSC colonies [113]. Very recently, the precise role of Myc during somatic cell reprogramming has been defined. Homogenous populations of mouse fibroblasts were depleted of methyl-CpG binding domain protein 3 (Mbd3) and GATA zinc finger domain-containing 2a (Gata2da), both participating in the Nucleosome remodeling and deacetylase (NuRD) repressor complex [114]. In this cell system, whereas OSK factors dynamically bind overlapping enhancers throughout the process, Myc preferentially binds to promoters. Specifically, Myc binds constitutively active promoters’ genes (CAPGs), igniting biosynthetic pathways, such as DNA synthesis and repair, cell proliferation and chromatin dynamics. In this system, OSKs are sufficient for somatic reprogramming to occur until endogenous Myc is expressed. Upon depletion of all three paralogs, OSKs fail to reprogram mouse fibroblasts. However, the opposite—i.e., OSK deprivation and Myc overexpression—promotes the expression of genes driven by CAPGs and the repression of somatic genes’, but not pluripotency genes’, induction. Of note, although Myc has been invariably associated with histone modifications related to gene activation (see above), CAPGs do not present the chromatin-switching typical of pluripotency genes upon reprogramming, but, as in cancer cells [115], are regulated by changes in the tRNA pool and their preferential codon usage. During reprogramming, a shift in the codon usage by tRNA is observed. In particular, CAPG-activated genes present a shift from G/C ending codon to A/T, whereas repressed genes shift A/T ending codon to G/C. This phenomenon corresponds to the enrichment of activating histone marks (i.e., H3K4me3) in chromatin regions next to tRNA encoding genes, fulfilling the request of enhanced elongation and translation of transcripts during reprogramming [116]. Given Myc-dependent regulation of RNAPIII activity [117,118], it may be hypothesized that Myc controls somatic cell reprogramming by directly binding to CAPGs and amplifying elongation and translation of the corresponding gene products.

4. Mammalian Tissue Regeneration: A Transcriptional Point of View

Transcriptional profiles activated during tissue regeneration closely resemble embryogenesis-related programs which drive proliferation, migration and cell morphogenesis in different organisms, including humans [119,120,121]. Of course, transcriptional signatures in regenerating cells are the result of the intersection of multiple, orchestrated layers of control; indeed, DNA elements, transcription factors and epigenetic machineries altogether participate in the constitution of networks that rule the execution of programs which may diverge from one tissue to another. In some organisms, as stated above, embryonic-like profiles are found [122,123], but this is not a dogma, as, in other species, transcriptional signatures mostly mirror those activated during injury and/or damage [124,125]. Furthermore, as expected, these regeneration-specific signatures are affected by environmental cues, mainly inflammation and immunometabolism [126].
It has to be noted that higher vertebrates, with the exception of African spiny mice [127,128], possess limited regenerating capacity. This has led to the hypothesis that, during evolution, some regeneration-specific chromatin regulatory elements have been lost. In particular, tissue regeneration enhancer elements (TREEs) [129], which control regeneration in zebrafish, salamanders and Drosophila melanogaster [130,131,132], have not been identified yet in mammalians. Indeed, strategies to transfer zebrafish TREEs, combined with CRISPR/Cas-dependent genome editing, to enhance regeneration capacity upon tissue injury have been recently described in small and large mammalians [133]. However, intronic elements have been found to potentially substitute TREEs in mammalians. In fact they control the expression of GATA2 [134] and SAMD14 [135] during hematopoietic cells’ regeneration following anemia in mice.
Trans-acting elements, that is, TFs devoted to the maintenance of pluripotency and/or acting during mammalian development, also act as master regulators of regeneration. In particular, SOX2 and KLF4, two of the OSKM factors, have been involved in the regeneration of olfactory neurons and retinal ganglion in mice [136,137]. ETS, SOX and GATA factors have been found to regulate neo-osteogenesis upon jaw distraction in mice, whereas developmental genes such as RUNX2 and DLX5 are repressed [121]. The same occurs in the mouse heart: a number of tissue-specific TFs, including Meis1 [138], Yes-associated protein (YAP) [139] and the Thyroid Hormone Receptor α (THRA) [140], constrain heart regeneration.
Strictly related to TFs’ action, chromatin accessibility represents a major characteristic of genomic regions involved in regeneration. This is also in line with a phenomenon known as “hypertranscription”, characterized by an increase in the nascent transcript output of the majority of the transcriptome, in particular, of housekeeping and ribosomal genes, which specifically respond to the biosynthetic needs of highly replicating cells [141,142]. Although not investigated in detail in mammalian regenerating tissues, it is highly probable that this phenomenon occurs during vertebrate regeneration, as an increase in total RNA content has been found in hematopoietic [143] and small intestine progenitors [144] during physiological regeneration. However, chromatin marks of regeneration are different from those appearing during embryogenesis, regardless of the similar final biological outcome. For example, during zebrafish cardiogenesis, the KAT p300 deposit H3K27Ac marks throughout cardio-specific genes [145], but the same phenomenon does not occur during regeneration [146]. Notably, mammalian chromatin de-compaction due to KATs’ activity has been mostly reported in pluripotency- and stemness-related genomic loci that are not directly involved in regeneration processes [147,148,149]. Nevertheless, chromatin acetylation has been found to be involved in liver, muscle and neuron regeneration [150,151,152], whereas histone deacetylases and their inhibition (HDACs) have been implicated in bone and dental pulp mineralization [153,154], axon remyelination [155,156] and neuronal regeneration [157,158].
A central role in re-shaping chromatin topology during regeneration seems to be played by the Polycomb group (PcG) and SWI/SNF remodelers, which may be considered as two sides of the same coin.
Erasure of PcG-dependent H3K27me3 histone marks from enhancers devoted to regeneration has been found both in Drosophila melanogaster and zebrafish [132,159], and also in mouse Schwann cells’ proliferation and axon regeneration upon the occurrence of damage [160]. A direct role of PcG complex members has been found in neurons [161], bladder [162], heart [163] and skeletal muscle [164].
SWI/SNF activity has been found to be essential for liver regeneration [165]. However, in certain conditions, SWI/SNF negatively controls the regeneration process. Indeed, the SWI/SNF component AT-rich interactive domain-containing protein 1A (ARID1A) has been found to impair Hippo-dependent regeneration [166,167] by sequestering YAP and TAZ from their target, TEAD1, the lead factor regulating regeneration downstream from the Hippo pathway [167]. Under mechanical stress, which may occur during regeneration [168], nuclear F-actin prevents the binding of AIRD1A to YAP/TAZ, allowing the latter to activate TEAD1 and Hippo-dependent transcription [169].

5. Myc Action in Tissue Regeneration

That Myc may exert pivotal functions in the regeneration of tissues is almost obvious. Indeed, both quiescent stem cells’ activation and dedifferentiation of adult, tissue-specific cells foresee the ignition of silent transcriptional programs which are largely dependent on Myc protein activity. As occurs during somatic cell re-programming [116], Myc modules have also been identified in physiologically regenerating cells and upon tissue injury. In contrast to what is observed during carcinogenesis [69], Myc action is tightly regulated in this context and turned off once no longer necessary. In this paragraph, examples of Myc function in the regeneration of different tissues are provided.

5.1. Myc in Liver Regeneration

Myc is rapidly induced upon partial hepatectomy (PH), as it is necessary for the G0/G1-S transition of the cell cycle in hepatocytes. Elegant experiments performed in transgenic mice where myc was conditionally knocked-down have also shown that Myc controls hepatocyte ploidy and size, being involved in multiple sequential cell cycles not followed by cytokinesis, and, although dispensable for adult hepatocytes proliferation, it is essential for liver regeneration after PH [170]. Knock-out of myc and of its partner mlx completely blunts liver regenerative capacity, affecting signaling pathways involved in mRNA translation, hepatic steatosis and mitochondrial function; KO livers show characteristics typical of non-fatty alcoholic liver disease (NFALD) [171]. At the molecular level, it has been described that PH-induced Inhibitor of DNA binding (or inhibition of differentiation) 2 (Id2)—another well-known Myc functional partner [172,173,174]—detaches and sequesters transcriptional co-repressors from myc promoter, leaving E2F4 and p130 pocket protein on, activating Myc transcription. In turn, after a few hours, Myc activates Id2 transcription, allowing the repositioning of this protein together with co-repressors onto myc promoter, in a tightly regulated feedback loop [175]. Furthermore, a Jun/Myc/BLC2/CCND1 axis has been found to regulate liver regeneration upon PH due to the downregulation of mir429 [176].
It has to be mentioned that Myc may exert a detrimental role in the liver, further to its well-established tumorigenic effect. Indeed, it may induce liver fibrosis upon PH in gata4-deficient hepatocytes by inducing the expression of the platelet-derived growth factor-β (PDGFRβ), transdifferentiation of endothelial cells from sinusoidal to continuous and activation of hepatocyte stellate cells [177].

5.2. Myc Role in Pancreatic β-Cells’ Replacement

In pancreatic β-cells, Myc is regulated by glucose and is probably part of the proliferation adaptive machinery in these cells [178]. Further, it is required for neonatal β-cells’ specification [179], leading to the exciting hypothesis of using Myc overexpression to counteract β-cells’ loss. However, years ago, the possibility of recovering pancreatic β-cells’ dysfunction by Myc overexpression was discarded, as supraphysiological levels of Myc induced cell death and differentiation [180,181,182]. Nonetheless, although Myc is dispensable for adult β-cells’ proper function in healthy conditions, in situations of metabolic stress, Myc has been revealed to be necessary to counteract the impairment in glucose tolerance, hyperglycemia and ipoinsulinemia [179,183]. Therefore, whereas high levels of Myc are detrimental, a fine-tuning of its expression may be beneficial to dysfunctional β-cells [184].

5.3. Myc in the Regulation of Intestinal Crypts’ Stemness and Regeneration

Wnt-dependent Myc expression has been found to be essential for the proliferation of progenitor cells in the intestine and is expressed in the amplifying compartment of intestinal crypts during development; however, it appeared to be dispensable for juvenile and adult intestine homeostasis [185,186]. Myc-depleted crypt progenitors do not exit the cell cycle but progress more slowly and are smaller in size, indicating the fundamental biosynthetic role of Myc in progenitor cells [187]. In addition to this developmental role, Wnt-/Myc-signaling plays a role during intestine regeneration upon DNA damage via upregulation of Focal Adhesion Kinase (FAK). Indeed, apc deletion produces an upregulation of FAK, which reverts to its normal level upon apc/myc double depletion. Moreover, upon myc conditional knock-out, intestinal crypts’ regeneration after gamma irradiation is severely impaired, with few enlarged regenerating crypts and a huge number of cystic areas. Highly similar results are obtained upon fak deletion [188]. In agreement with these findings, a role of the PCNA Associated Factor (PAF)-mediated Myc activity in intestinal progenitor cells upon injury has been found. Specifically, PAF has been found located onto the myc promoter, leading to its transactivation and intestinal regeneration due to progenitor cells’ self-renewal. Remarkably, PAF knock-out impairs both intestinal regeneration and Apc depletion-induced tumorigenesis, along with Myc downregulation [189].

5.4. The Role of Myc in Cardiomyocytes’ Proliferation

Terminally differentiated tissues—such as the heart and the central nervous system (CNS)—are refractory to Myc-dependent proliferation signals. Indeed, it has been demonstrated that, whereas in permissive tissues, such as liver, Myc-bound promoters can be activated by mitogenic stimuli, in non-permissive tissues, Myc activity is locked even in the presence of such signals, probably because of limited availability of components of the basal transcriptional machinery in which Myc is deployed to exert its function [190]. Supporting this hypothesis, overexpression of cyclin T1, which, together with cyclin-dependent kinase 9 (CDK9), constitutes the P-TEFb complex [18], in cardiomyocytes, unlocks the proliferative capacity of Myc in vivo, inducing the transcription of a subset of Myc targets normally expressed in the liver but not in the heart and cardiomyocytes (CMs). More importantly, when Myc was transiently induced in a cyclin T1 overexpressing context, no consistent and detrimental effects on long-term survival were detected [191]. In the heart, Myc is also engaged in a positive feedback loop with the lncRNA small nucleolar RNA host gene 1 Snhg1 (Snhg1), promoting cardiomyocytes’ proliferation and regeneration upon both myocardial infarction and ischemia/reperfusion, with reduced CMs’ apoptosis and size of the infarcted area. Mechanistically, Snhg1 induces phosphatase and tensin homolog (PTEN) degradation and phosphatydilinositol 3-kinase (PI3K)/protein kinase B (AKT) activation which, in turn, via Glycogen synthase kinase-3β (GSK3β), promote Myc stabilization. Myc occupies and activates Snhg1 promoter in a self-sustaining loop, leading to CMs’ proliferation [192].
Moreover, the ability of Myc to grant reprogramming of adult somatic cells has also been successfully tested in cardiomyocytes, with different outcomes according to the timing of its expression. Indeed, whereas doxycycline-inducible, short-term expression of OSKM opens a transient transcriptional window of dedifferentiation with the expression of α-smooth muscle actin (α-SMA) and slightly reduces left ventricular ejection fraction (LVEF) and stroke volume (SV), prolonged OSKM exposure leads to an extensive heart remodeling, to a deterioration of cardiac function in the middle period and to the onset of cardiac tumors in the long period, even after doxycycline withdrawal. These results indicate the full epigenetic reprogramming of CMs in vivo [193].

5.5. Myc and the Regeneration of the Nervous System

The nervous system (NS) has evolved as a protected closed district—unresponsive to proliferative signals [194]—to preserve synapses and strengthen memory and character. These features are dependent, at least in part, on changes in the NS transcriptome, which is epigenetically preserved to express neuron-specific transcripts and is refractory even to OSKM reprogramming [195]. Accordingly, Myc overexpression in the NS provides opposite results when compared to other tissues, where it leads to carcinogenesis. Indeed, its expression is blunted in a number of ways in the NS and it, surprisingly, induces apoptosis and neuronal loss preceding and during neurodegeneration. Phosphorylated Myc has been detected in brain tissues affected by tauthopathies, including supranuclear palsy, Alzheimer’s Disease (AD) and corticobasal degeneration [196,197] and has been linked to both Ras/MAPK [196] and PI3K/mTOR pathways [198,199]. Derangement of the cell cycle is typical of both cancer and neurodegeneration, but, whereas in the first condition, hyperproliferation is the major outcome, neurons abort the cell cycle before DNA synthesis [200] or, rarely, after G2/M phase [194]. These two opposite phenomena, despite a common molecular background, seem to rely on the entity of the stress to which neuronal cells are exposed. Upon acute and high-energy stress exposure, neurons activate mTOR and ROS production, with the consequent engagement of the Forkhead box (FOXO) family of transcription factors, leading to the production of anti-oxidant enzymes, protecting neurons from death [201,202]. FOXO proteins compete with Myc for binding to cell cycle gene promoters [203,204]. Upon chronic stress stimuli, the FOXO protective role is bypassed and Myc activities are not repressed [205]. Furthermore, Myc is downstream from an NMDA receptor-dependent apoptotic cascade, which also comprises NF-kB and p53 activation as the final event [206].
Thus, it may seem that Myc exerts a potent detrimental role in the NS. This is not true, as Myc plays an important function during neuronal cells’ repair upon injury. However, there is a clear dichotomy regarding Myc reparative action between the peripheral and the central NS (PNS and CNS). In the PNS, Myc is re-activated upon axotomy in adult dorsal root ganglion [207] and supports the regeneration of sensory axons. In this latter regard, mechanistically, Myc activates TERT subunit of telomerase which, in turn, activates p53. This Myc/TERT/p53 pathway appears to play a major role in axon regeneration, acting both on metabolism and mitochondria, regulating ROS production [208]. Notably, when axotomy is performed on CNS elements, such as the optic nerve, Myc levels decrease even in uninjured neurons which undergo cell death [207]. The nature of the mechanism controlling the differential expression of Myc in PNS and CNS has to be elucidated, but probably depends on the strict relationship between the levels of FOXO protein and Myc [209,210].

6. Conclusions and Translational Perspectives

Far from being exhaustive, this review article provides highlights on the role of Myc in both the physiological regeneration and recovery upon damage in main mammalian tissues, such as liver, intestine, pancreas, heart and the CNS. As expected, other tissue districts deploy Myc expression as a tool to regenerate, repair and counteract cellular senescence. Indeed, Myc has been found to promote basal epithelial cells’ self-renewal in the esophagus, counteracting ROS-dependent aging processes [211], and to participate in the recovery of impaired diabetic wound healing downstream from the silent information regulator 1 (SIRT1) signal [212]. When looking at the possibility to employ Myc as a rejuvenating factor, somatic cell reprogramming represents the most logical strategy to obtain tissue regeneration even in high-proliferation signals’ refractory tissues. Partial OSKM-dependent reprogramming of myofibers, for example, has been employed to modify satellite stem cells’ niche epigenetic state through downregulation of Wnt4 signaling and to induce muscle regeneration [213], whereas co-activation of exogenous Myc and Notch results in the reprogramming of different inner ear cell types, including hair cells, whose loss is responsible for hearing loss in humans. It has to be noted, however, that, in this latter case, only a transient expression of Myc/Notch promotes hair cells’ transdifferentiation from supporting cochlear cells [214].
More importantly, it has been recently reported that cycles of transient expression of the OSKM factors slows the appearance of senescence markers and increases life-span in a mouse model of premature aging by modifying the cell epigenome. Moreover, OSKM factors improve cell recovery upon metabolic disease and muscle injury in old mice, indicating a comprehensive amelioration of the age-associated phenotype based on epigenetic mechanisms [215].
Although exciting, this perspective is far from being the elixir of life. In fact, the expression of OSKM factors—especially Myc—has to be tightly controlled in space and time to avoid highly dangerous side effects, such as the development of tumors (Figure 3).
Associated with this consideration, it has to be underlined that the field of cell-based regenerative medicine presents a number of challenges that have to be solved. First, the choice of vectors to be used has to be carefully weighed [216]. Currently, viral non-integrating episomal vectors and transposons are used to safely generate iPS [216,217]. Furthermore, each step to translate a cell-based therapy proof-of-concept into the clinic has to be tightly controlled. First of all, standard criteria and associated information should come with the release of expanded cells (e.g., stage of differentiation, phenotype, delivery system, factors to provide and granting the maintenance of the phenotype). Linked to this issue is manufacturing. Bulk, automated manufacturing arising from small pilot labs is highly expensive and requires bioreactors which, however, get into the game only when evidence of clinical efficacy has been provided and when extensive funds are available. Further, the pace of science in the field frequently makes the current manufacturing protocols and technologies obsolete. This requires continuous updates and advancements to satisfy the latest innovations, such as the direct expansion of cells on 3D-porous scaffolds or their assembly through bioprinting. Biomaterials, including starting fetal bovine serum-deprived cell culture media, represent another important issue. Extracellular matrix-derived substances, biopolymers and hydrogels are the most relevant, likely providing environmental responsiveness [218]. Financial profit is a major problem, as it underlies the commercialization of poorly characterized cell products with uncertain benefits by companies which leverage patients with no expectancy of cure. This highlights the importance of robust regulatory procedures and guidelines, ensuring evidence-based production of cell-based products for therapeutic purposes [219]. In spite of this evidence, somatic cell reprogramming experiments have opened a new avenue in the search of platforms—constituted by TFs and small-molecule modulators of chromatin remodeling enzymes, for example—which are suitable to epigenetically remodel adult and old genomes to fight age-associated and degenerative diseases.

Author Contributions

B.I. and S.N. wrote and revised the entire manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work has been partially funded by the Italian Association for Cancer Research grant IG15927 to B.I.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Goldman, J.A.; Poss, K.D. Gene regulatory programmes of tissue regeneration. Nat. Rev. Genet. 2020, 21, 511–525. [Google Scholar] [CrossRef] [PubMed]
  2. Mokalled, M.H.; Poss, K.D. A Regeneration Toolkit. Dev. Cell 2018, 47, 267–280. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Wells, J.M.; Watt, F.M. Diverse mechanisms for endogenous regeneration and repair in mammalian organs. Nature 2018, 557, 322–328. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Colby, W.W.; Chen, E.Y.; Smith, D.H.; Levinson, A.D. Identification and nucleotide sequence of a human locus homologous to the v-myc oncogene of avian myelocytomatosis virus MC29. Nature 1983, 301, 722–725. [Google Scholar] [CrossRef]
  5. Blackwood, E.M.; Eisenman, R.N. Max: A helix-loop-helix zipper protein that forms a sequence-specific DNA-binding complex with Myc. Science 1991, 251, 1211–1217. [Google Scholar] [CrossRef]
  6. Kanazawa, S.; Soucek, L.; Evan, G.; Okamoto, T.; Peterlin, B.M. c-Myc recruits P-TEFb for transcription, cellular proliferation and apoptosis. Oncogene 2003, 22, 5707–5711. [Google Scholar] [CrossRef] [Green Version]
  7. Evan, G.I.; Wyllie, A.H.; Gilbert, C.S.; Littlewood, T.D.; Land, H.; Brooks, M.; Waters, C.M.; Penn, L.Z.; Hancock, D.C. Induction of apoptosis in fibroblasts by c-myc protein. Cell 1992, 69, 119–128. [Google Scholar] [CrossRef]
  8. Shi, Y.; Glynn, J.M.; Guilbert, L.J.; Cotter, T.G.; Bissonnette, R.P.; Green, D.R. Role for c-myc in activation-induced apoptotic cell death in T cell hybridomas. Science 1992, 257, 212–214. [Google Scholar] [CrossRef]
  9. Soucie, E.L.; Annis, M.G.; Sedivy, J.; Filmus, J.; Leber, B.; Andrews, D.W.; Penn, L.Z. Myc potentiates apoptosis by stimulating Bax activity at the mitochondria. Mol. Cell Biol. 2001, 21, 4725–4736. [Google Scholar] [CrossRef] [Green Version]
  10. Shim, H.; Dolde, C.; Lewis, B.C.; Wu, C.S.; Dang, G.; Jungmann, R.A.; Dalla-Favera, R.; Dang, C.V. c-Myc transactivation of LDH-A: Implications for tumor metabolism and growth. Proc. Natl. Acad. Sci. USA 1997, 94, 6658–6663. [Google Scholar] [CrossRef]
  11. Gao, P.; Tchernyshyov, I.; Chang, T.C.; Lee, Y.S.; Kita, K.; Ochi, T.; Zeller, K.I.; De Marzo, A.M.; Van Eyk, J.E.; Mendell, J.T.; et al. c-Myc suppression of miR-23a/b enhances mitochondrial glutaminase expression and glutamine metabolism. Nature 2009, 458, 762–765. [Google Scholar] [CrossRef] [Green Version]
  12. Bello-Fernandez, C.; Packham, G.; Cleveland, J.L. The ornithine decarboxylase gene is a transcriptional target of c-Myc. Proc. Natl. Acad. Sci. USA 1993, 90, 7804–7808. [Google Scholar] [CrossRef]
  13. Hydbring, P.; Larsson, L.G. Cdk2: A key regulator of the senescence control function of Myc. Aging 2010, 2, 244–250. [Google Scholar] [CrossRef] [Green Version]
  14. Singh, V.P.; Hassan, H.; Deng, F.; Tsuchiya, D.; McKinney, S.; Ferro, K.; Gerton, J.L. c-Myc promotes polyploidy in murine trophoblast cells and suppresses senescence. Development 2023, 150, 11. [Google Scholar] [CrossRef]
  15. Cartwright, P.; McLean, C.; Sheppard, A.; Rivett, D.; Jones, K.; Dalton, S. LIF/STAT3 controls ES cell self-renewal and pluripotency by a Myc-dependent mechanism. Development 2005, 132, 885–896. [Google Scholar] [CrossRef] [Green Version]
  16. Takahashi, K.; Tanabe, K.; Ohnuki, M.; Narita, M.; Ichisaka, T.; Tomoda, K.; Yamanaka, S. Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell 2007, 131, 861–872. [Google Scholar] [CrossRef] [Green Version]
  17. Takahashi, K.; Yamanaka, S. A decade of transcription factor-mediated reprogramming to pluripotency. Nat. Rev. Mol. Cell Biol. 2016, 17, 183–193. [Google Scholar] [CrossRef]
  18. Rahl, P.B.; Lin, C.Y.; Seila, A.C.; Flynn, R.A.; McCuine, S.; Burge, C.B.; Sharp, P.A.; Young, R.A. c-Myc regulates transcriptional pause release. Cell 2010, 141, 432–445. [Google Scholar] [CrossRef] [Green Version]
  19. Baluapuri, A.; Hofstetter, J.; Dudvarski Stankovic, N.; Endres, T.; Bhandare, P.; Vos, S.M.; Adhikari, B.; Schwarz, J.D.; Narain, A.; Vogt, M.; et al. MYC Recruits SPT5 to RNA Polymerase II to Promote Processive Transcription Elongation. Mol. Cell 2019, 74, 674–687.e611. [Google Scholar] [CrossRef] [Green Version]
  20. Scagnoli, F.; Palma, A.; Favia, A.; Scuoppo, C.; Illi, B.; Nasi, S. A New Insight into MYC Action: Control of RNA Polymerase II Methylation and Transcription Termination. Biomedicines 2023, 11, 412. [Google Scholar] [CrossRef]
  21. Koh, C.M.; Bezzi, M.; Low, D.H.; Ang, W.X.; Teo, S.X.; Gay, F.P.; Al-Haddawi, M.; Tan, S.Y.; Osato, M.; Sabo, A.; et al. MYC regulates the core pre-mRNA splicing machinery as an essential step in lymphomagenesis. Nature 2015, 523, 96–100. [Google Scholar] [CrossRef] [PubMed]
  22. See, Y.X.; Chen, K.; Fullwood, M.J. MYC overexpression leads to increased chromatin interactions at super-enhancers and MYC binding sites. Genome Res. 2022, 32, 629–642. [Google Scholar] [CrossRef] [PubMed]
  23. Psathas, J.N.; Thomas-Tikhonenko, A. MYC and the art of microRNA maintenance. Cold Spring Harb. Perspect. Med. 2014, 4, 8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Dang, C.V. MYC on the path to cancer. Cell 2012, 149, 22–35. [Google Scholar] [CrossRef] [Green Version]
  25. Kohl, N.E.; Kanda, N.; Schreck, R.R.; Bruns, G.; Latt, S.A.; Gilbert, F.; Alt, F.W. Transposition and amplification of oncogene-related sequences in human neuroblastomas. Cell 1983, 35, 359–367. [Google Scholar] [CrossRef]
  26. Nau, M.M.; Brooks, B.J.; Battey, J.; Sausville, E.; Gazdar, A.F.; Kirsch, I.R.; McBride, O.W.; Bertness, V.; Hollis, G.F.; Minna, J.D. L-myc, a new myc-related gene amplified and expressed in human small cell lung cancer. Nature 1985, 318, 69–73. [Google Scholar] [CrossRef]
  27. Kalkat, M.; Resetca, D.; Lourenco, C.; Chan, P.K.; Wei, Y.; Shiah, Y.J.; Vitkin, N.; Tong, Y.; Sunnerhagen, M.; Done, S.J.; et al. MYC Protein Interactome Profiling Reveals Functionally Distinct Regions that Cooperate to Drive Tumorigenesis. Mol. Cell 2018, 72, 836–848.e837. [Google Scholar] [CrossRef] [Green Version]
  28. Beaulieu, M.E.; Castillo, F.; Soucek, L. Structural and Biophysical Insights into the Function of the Intrinsically Disordered Myc Oncoprotein. Cells 2020, 9, 1038. [Google Scholar] [CrossRef]
  29. Cole, M.D. Myc meets its Max. Cell 1991, 65, 715–716. [Google Scholar] [CrossRef]
  30. Grandori, C.; Cowley, S.M.; James, L.P.; Eisenman, R.N. The Myc/Max/Mad network and the transcriptional control of cell behavior. Annu. Rev. Cell Dev. Biol. 2000, 16, 653–699. [Google Scholar] [CrossRef]
  31. Forman-Kay, J.D.; Mittag, T. From sequence and forces to structure, function, and evolution of intrinsically disordered proteins. Structure 2013, 21, 1492–1499. [Google Scholar] [CrossRef] [Green Version]
  32. McDuff, F.O.; Naud, J.F.; Montagne, M.; Sauve, S.; Lavigne, P. The Max homodimeric b-HLH-LZ significantly interferes with the specific heterodimerization between the c-Myc and Max b-HLH-LZ in absence of DNA: A quantitative analysis. J. Mol. Recognit. 2009, 22, 261–269. [Google Scholar] [CrossRef]
  33. Sammak, S.; Hamdani, N.; Gorrec, F.; Allen, M.D.; Freund, S.M.V.; Bycroft, M.; Zinzalla, G. Crystal Structures and Nuclear Magnetic Resonance Studies of the Apo Form of the c-MYC:MAX bHLHZip Complex Reveal a Helical Basic Region in the Absence of DNA. Biochemistry 2019, 58, 3144–3154. [Google Scholar] [CrossRef] [Green Version]
  34. Nair, S.K.; Burley, S.K. X-ray structures of Myc-Max and Mad-Max recognizing DNA. Molecular bases of regulation by proto-oncogenic transcription factors. Cell 2003, 112, 193–205. [Google Scholar] [CrossRef] [Green Version]
  35. Whitfield, J.R.; Beaulieu, M.E.; Soucek, L. Strategies to Inhibit Myc and Their Clinical Applicability. Front. Cell Dev. Biol. 2017, 5, 10. [Google Scholar] [CrossRef] [Green Version]
  36. Berg, T.; Cohen, S.B.; Desharnais, J.; Sonderegger, C.; Maslyar, D.J.; Goldberg, J.; Boger, D.L.; Vogt, P.K. Small-molecule antagonists of Myc/Max dimerization inhibit Myc-induced transformation of chicken embryo fibroblasts. Proc. Natl. Acad. Sci. USA 2002, 99, 3830–3835. [Google Scholar] [CrossRef]
  37. Yin, X.; Giap, C.; Lazo, J.S.; Prochownik, E.V. Low molecular weight inhibitors of Myc-Max interaction and function. Oncogene 2003, 22, 6151–6159. [Google Scholar] [CrossRef] [Green Version]
  38. Han, H.; Jain, A.D.; Truica, M.I.; Izquierdo-Ferrer, J.; Anker, J.F.; Lysy, B.; Sagar, V.; Luan, Y.; Chalmers, Z.R.; Unno, K.; et al. Small-Molecule MYC Inhibitors Suppress Tumor Growth and Enhance Immunotherapy. Cancer Cell 2019, 36, 483–497.e415. [Google Scholar] [CrossRef]
  39. Soucek, L.; Helmer-Citterich, M.; Sacco, A.; Jucker, R.; Cesareni, G.; Nasi, S. Design and properties of a Myc derivative that efficiently homodimerizes. Oncogene 1998, 17, 2463–2472. [Google Scholar] [CrossRef] [Green Version]
  40. Savino, M.; Annibali, D.; Carucci, N.; Favuzzi, E.; Cole, M.D.; Evan, G.I.; Soucek, L.; Nasi, S. The action mechanism of the Myc inhibitor termed Omomyc may give clues on how to target Myc for cancer therapy. PLoS ONE 2011, 6, e22284. [Google Scholar] [CrossRef] [Green Version]
  41. Delmore, J.E.; Issa, G.C.; Lemieux, M.E.; Rahl, P.B.; Shi, J.; Jacobs, H.M.; Kastritis, E.; Gilpatrick, T.; Paranal, R.M.; Qi, J.; et al. BET bromodomain inhibition as a therapeutic strategy to target c-Myc. Cell 2011, 146, 904–917. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Dauch, D.; Rudalska, R.; Cossa, G.; Nault, J.C.; Kang, T.W.; Wuestefeld, T.; Hohmeyer, A.; Imbeaud, S.; Yevsa, T.; Hoenicke, L.; et al. A MYC-aurora kinase A protein complex represents an actionable drug target in p53-altered liver cancer. Nat. Med. 2016, 22, 744–753. [Google Scholar] [CrossRef] [PubMed]
  43. Grandori, C.; Eisenman, R.N. Myc target genes. Trends Biochem. Sci. 1997, 22, 177–181. [Google Scholar] [CrossRef] [PubMed]
  44. Farhana, L.; Dawson, M.I.; Fontana, J.A. Down regulation of miR-202 modulates Mxd1 and Sin3A repressor complexes to induce apoptosis of pancreatic cancer cells. Cancer Biol. Ther. 2015, 16, 115–124. [Google Scholar] [CrossRef]
  45. Moroy, T.; Saba, I.; Kosan, C. The role of the transcription factor Miz-1 in lymphocyte development and lymphomagenesis-Binding Myc makes the difference. Semin. Immunol. 2011, 23, 379–387. [Google Scholar] [CrossRef]
  46. Nie, Z.; Hu, G.; Wei, G.; Cui, K.; Yamane, A.; Resch, W.; Wang, R.; Green, D.R.; Tessarollo, L.; Casellas, R.; et al. c-Myc is a universal amplifier of expressed genes in lymphocytes and embryonic stem cells. Cell 2012, 151, 68–79. [Google Scholar] [CrossRef] [Green Version]
  47. Lin, C.Y.; Loven, J.; Rahl, P.B.; Paranal, R.M.; Burge, C.B.; Bradner, J.E.; Lee, T.I.; Young, R.A. Transcriptional amplification in tumor cells with elevated c-Myc. Cell 2012, 151, 56–67. [Google Scholar] [CrossRef] [Green Version]
  48. Nie, Z.; Guo, C.; Das, S.K.; Chow, C.C.; Batchelor, E.; Simons, S.S.J.; Levens, D. Dissecting transcriptional amplification by MYC. eLife 2020, 9, e52483. [Google Scholar] [CrossRef]
  49. Walz, S.; Lorenzin, F.; Morton, J.; Wiese, K.E.; von Eyss, B.; Herold, S.; Rycak, L.; Dumay-Odelot, H.; Karim, S.; Bartkuhn, M.; et al. Activation and repression by oncogenic MYC shape tumour-specific gene expression profiles. Nature 2014, 511, 483–487. [Google Scholar] [CrossRef]
  50. Guccione, E.; Martinato, F.; Finocchiaro, G.; Luzi, L.; Tizzoni, L.; Dall’ Olio, V.; Zardo, G.; Nervi, C.; Bernard, L.; Amati, B. Myc-binding-site recognition in the human genome is determined by chromatin context. Nat. Cell Biol. 2006, 8, 764–770. [Google Scholar] [CrossRef]
  51. Das, S.K.; Kuzin, V.; Cameron, D.P.; Sanford, S.; Jha, R.K.; Nie, Z.; Rosello, M.T.; Holewinski, R.; Andresson, T.; Wisniewski, J.; et al. MYC assembles and stimulates topoisomerases 1 and 2 in a “topoisome”. Mol. Cell 2022, 82, 140–158.e112. [Google Scholar] [CrossRef]
  52. Cowling, V.H.; Cole, M.D. The Myc transactivation domain promotes global phosphorylation of the RNA polymerase II carboxy-terminal domain independently of direct DNA binding. Mol. Cell Biol. 2007, 27, 2059–2073. [Google Scholar] [CrossRef] [Green Version]
  53. Galardi, S.; Savino, M.; Scagnoli, F.; Pellegatta, S.; Pisati, F.; Zambelli, F.; Illi, B.; Annibali, D.; Beji, S.; Orecchini, E.; et al. Resetting cancer stem cell regulatory nodes upon MYC inhibition. EMBO Rep. 2016, 17, 1872–1889. [Google Scholar] [CrossRef] [Green Version]
  54. Davis, A.C.; Wims, M.; Spotts, G.D.; Hann, S.R.; Bradley, A. A null c-myc mutation causes lethality before 10.5 days of gestation in homozygotes and reduced fertility in heterozygous female mice. Genes Dev. 1993, 7, 671–682. [Google Scholar] [CrossRef] [Green Version]
  55. Zimmerman, K.A.; Yancopoulos, G.D.; Collum, R.G.; Smith, R.K.; Kohl, N.E.; Denis, K.A.; Nau, M.M.; Witte, O.N.; Toran-Allerand, D.; Gee, C.E.; et al. Differential expression of myc family genes during murine development. Nature 1986, 319, 780–783. [Google Scholar] [CrossRef]
  56. Malynn, B.A.; de Alboran, I.M.; O’Hagan, R.C.; Bronson, R.; Davidson, L.; DePinho, R.A.; Alt, F.W. N-myc can functionally replace c-myc in murine development, cellular growth, and differentiation. Genes Dev. 2000, 14, 1390–1399. [Google Scholar] [CrossRef]
  57. Smith, K.N.; Singh, A.M.; Dalton, S. Myc represses primitive endoderm differentiation in pluripotent stem cells. Cell Stem Cell 2010, 7, 343–354. [Google Scholar] [CrossRef] [Green Version]
  58. Varlakhanova, N.V.; Cotterman, R.F.; deVries, W.N.; Morgan, J.; Donahue, L.R.; Murray, S.; Knowles, B.B.; Knoepfler, P.S. myc maintains embryonic stem cell pluripotency and self-renewal. Differentiation 2010, 80, 9–19. [Google Scholar] [CrossRef] [Green Version]
  59. Gaspar-Maia, A.; Alajem, A.; Meshorer, E.; Ramalho-Santos, M. Open chromatin in pluripotency and reprogramming. Nat. Rev. Mol. Cell Biol. 2011, 12, 36–47. [Google Scholar] [CrossRef] [Green Version]
  60. Tee, W.W.; Reinberg, D. Chromatin features and the epigenetic regulation of pluripotency states in ESCs. Development 2014, 141, 2376–2390. [Google Scholar] [CrossRef] [Green Version]
  61. Zhu, J.; Adli, M.; Zou, J.Y.; Verstappen, G.; Coyne, M.; Zhang, X.; Durham, T.; Miri, M.; Deshpande, V.; De Jager, P.L.; et al. Genome-wide chromatin state transitions associated with developmental and environmental cues. Cell 2013, 152, 642–654. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Bernstein, B.E.; Mikkelsen, T.S.; Xie, X.; Kamal, M.; Huebert, D.J.; Cuff, J.; Fry, B.; Meissner, A.; Wernig, M.; Plath, K.; et al. A bivalent chromatin structure marks key developmental genes in embryonic stem cells. Cell 2006, 125, 315–326. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Pan, G.; Tian, S.; Nie, J.; Yang, C.; Ruotti, V.; Wei, H.; Jonsdottir, G.A.; Stewart, R.; Thomson, J.A. Whole-genome analysis of histone H3 lysine 4 and lysine 27 methylation in human embryonic stem cells. Cell Stem Cell 2007, 1, 299–312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Frye, M.; Fisher, A.G.; Watt, F.M. Epidermal stem cells are defined by global histone modifications that are altered by Myc-induced differentiation. PLoS ONE 2007, 2, e763. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Cotterman, R.; Jin, V.X.; Krig, S.R.; Lemen, J.M.; Wey, A.; Farnham, P.J.; Knoepfler, P.S. N-Myc regulates a widespread euchromatic program in the human genome partially independent of its role as a classical transcription factor. Cancer Res. 2008, 68, 9654–9662. [Google Scholar] [CrossRef] [Green Version]
  66. Lin, C.H.; Lin, C.; Tanaka, H.; Fero, M.L.; Eisenman, R.N. Gene regulation and epigenetic remodeling in murine embryonic stem cells by c-Myc. PLoS ONE 2009, 4, e7839. [Google Scholar] [CrossRef]
  67. Krepelova, A.; Neri, F.; Maldotti, M.; Rapelli, S.; Oliviero, S. Myc and max genome-wide binding sites analysis links the Myc regulatory network with the polycomb and the core pluripotency networks in mouse embryonic stem cells. PLoS ONE 2014, 9, e88933. [Google Scholar] [CrossRef]
  68. Kim, J.; Chu, J.; Shen, X.; Wang, J.; Orkin, S.H. An extended transcriptional network for pluripotency of embryonic stem cells. Cell 2008, 132, 1049–1061. [Google Scholar] [CrossRef] [Green Version]
  69. Kim, J.; Woo, A.J.; Chu, J.; Snow, J.W.; Fujiwara, Y.; Kim, C.G.; Cantor, A.B.; Orkin, S.H. A Myc network accounts for similarities between embryonic stem and cancer cell transcription programs. Cell 2010, 143, 313–324. [Google Scholar] [CrossRef] [Green Version]
  70. Das, P.P.; Shao, Z.; Beyaz, S.; Apostolou, E.; Pinello, L.; De Los Angeles, A.; O’Brien, K.; Atsma, J.M.; Fujiwara, Y.; Nguyen, M.; et al. Distinct and combinatorial functions of Jmjd2b/Kdm4b and Jmjd2c/Kdm4c in mouse embryonic stem cell identity. Mol. Cell 2014, 53, 32–48. [Google Scholar] [CrossRef] [Green Version]
  71. Amente, S.; Lania, L.; Majello, B. The histone LSD1 demethylase in stemness and cancer transcription programs. Biochim. Biophys. Acta 2013, 1829, 981–986. [Google Scholar] [CrossRef] [Green Version]
  72. Smith, K.N.; Lim, J.M.; Wells, L.; Dalton, S. Myc orchestrates a regulatory network required for the establishment and maintenance of pluripotency. Cell Cycle 2011, 10, 592–597. [Google Scholar] [CrossRef]
  73. Hu, G.; Wade, P.A. NuRD and pluripotency: A complex balancing act. Cell Stem Cell 2012, 10, 497–503. [Google Scholar] [CrossRef] [Green Version]
  74. Fagnocchi, L.; Mazzoleni, S.; Zippo, A. Integration of Signaling Pathways with the Epigenetic Machinery in the Maintenance of Stem Cells. Stem Cells Int. 2016, 2016, 8652748. [Google Scholar] [CrossRef] [Green Version]
  75. Bonasio, R.; Tu, S.; Reinberg, D. Molecular signals of epigenetic states. Science 2010, 330, 612–616. [Google Scholar] [CrossRef] [Green Version]
  76. Laurenti, E.; Varnum-Finney, B.; Wilson, A.; Ferrero, I.; Blanco-Bose, W.E.; Ehninger, A.; Knoepfler, P.S.; Cheng, P.F.; MacDonald, H.R.; Eisenman, R.N.; et al. Hematopoietic stem cell function and survival depend on c-Myc and N-Myc activity. Cell Stem Cell 2008, 3, 611–624. [Google Scholar] [CrossRef] [Green Version]
  77. Smith, K.; Dalton, S. Myc transcription factors: Key regulators behind establishment and maintenance of pluripotency. Regen. Med. 2010, 5, 947–959. [Google Scholar] [CrossRef] [Green Version]
  78. Gibson, J.D.; Jakuba, C.M.; Boucher, N.; Holbrook, K.A.; Carter, M.G.; Nelson, C.E. Single-cell transcript analysis of human embryonic stem cells. Integr. Biol. 2009, 1, 540–551. [Google Scholar] [CrossRef]
  79. Doyon, Y.; Cote, J. The highly conserved and multifunctional NuA4 HAT complex. Curr. Opin. Genet. Dev. 2004, 14, 147–154. [Google Scholar] [CrossRef]
  80. Faast, R.; White, J.; Cartwright, P.; Crocker, L.; Sarcevic, B.; Dalton, S. Cdk6-cyclin D3 activity in murine ES cells is resistant to inhibition by p16(INK4a). Oncogene 2004, 23, 491–502. [Google Scholar] [CrossRef]
  81. Stead, E.; White, J.; Faast, R.; Conn, S.; Goldstone, S.; Rathjen, J.; Dhingra, U.; Rathjen, P.; Walker, D.; Dalton, S. Pluripotent cell division cycles are driven by ectopic Cdk2, cyclin A/E and E2F activities. Oncogene 2002, 21, 8320–8333. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Savatier, P.; Huang, S.; Szekely, L.; Wiman, K.G.; Samarut, J. Contrasting patterns of retinoblastoma protein expression in mouse embryonic stem cells and embryonic fibroblasts. Oncogene 1994, 9, 809–818. [Google Scholar] [PubMed]
  83. White, J.; Stead, E.; Faast, R.; Conn, S.; Cartwright, P.; Dalton, S. Developmental activation of the Rb-E2F pathway and establishment of cell cycle-regulated cyclin-dependent kinase activity during embryonic stem cell differentiation. Mol. Biol. Cell 2005, 16, 2018–2027. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Bretones, G.; Delgado, M.D.; Leon, J. Myc and cell cycle control. Biochim. Biophys. Acta 2015, 1849, 506–516. [Google Scholar] [CrossRef]
  85. Brenner, C.; Deplus, R.; Didelot, C.; Loriot, A.; Vire, E.; De Smet, C.; Gutierrez, A.; Danovi, D.; Bernard, D.; Boon, T.; et al. Myc represses transcription through recruitment of DNA methyltransferase corepressor. EMBO J. 2005, 24, 336–346. [Google Scholar] [CrossRef]
  86. Sela, Y.; Molotski, N.; Golan, S.; Itskovitz-Eldor, J.; Soen, Y. Human embryonic stem cells exhibit increased propensity to differentiate during the G1 phase prior to phosphorylation of retinoblastoma protein. Stem Cells 2012, 30, 1097–1108. [Google Scholar] [CrossRef]
  87. Pauklin, S.; Vallier, L. The cell-cycle state of stem cells determines cell fate propensity. Cell 2013, 155, 135–147. [Google Scholar] [CrossRef] [Green Version]
  88. Singh, A.M.; Dalton, S. The cell cycle and Myc intersect with mechanisms that regulate pluripotency and reprogramming. Cell Stem Cell 2009, 5, 141–149. [Google Scholar] [CrossRef] [Green Version]
  89. Gu, W.; Gaeta, X.; Sahakyan, A.; Chan, A.B.; Hong, C.S.; Kim, R.; Braas, D.; Plath, K.; Lowry, W.E.; Christofk, H.R. Glycolytic Metabolism Plays a Functional Role in Regulating Human Pluripotent Stem Cell State. Cell Stem Cell 2016, 19, 476–490. [Google Scholar] [CrossRef] [Green Version]
  90. Epstein, T.; Gatenby, R.A.; Brown, J.S. The Warburg effect as an adaptation of cancer cells to rapid fluctuations in energy demand. PLoS ONE 2017, 12, e0185085. [Google Scholar] [CrossRef] [Green Version]
  91. Stine, Z.E.; Walton, Z.E.; Altman, B.J.; Hsieh, A.L.; Dang, C.V. MYC, Metabolism, and Cancer. Cancer Discov. 2015, 5, 1024–1039. [Google Scholar] [CrossRef] [Green Version]
  92. Cencioni, C.; Scagnoli, F.; Spallotta, F.; Nasi, S.; Illi, B. The “Superoncogene” Myc at the Crossroad between Metabolism and Gene Expression in Glioblastoma Multiforme. Int. J. Mol. Sci. 2023, 24, 4217. [Google Scholar] [CrossRef]
  93. Cao, Y.; Guo, W.T.; Tian, S.; He, X.; Wang, X.W.; Liu, X.; Gu, K.L.; Ma, X.; Huang, D.; Hu, L.; et al. miR-290/371-Mbd2-Myc circuit regulates glycolytic metabolism to promote pluripotency. EMBO J. 2015, 34, 609–623. [Google Scholar] [CrossRef] [Green Version]
  94. Edmunds, L.R.; Sharma, L.; Kang, A.; Lu, J.; Vockley, J.; Basu, S.; Uppala, R.; Goetzman, E.S.; Beck, M.E.; Scott, D.; et al. c-Myc programs fatty acid metabolism and dictates acetyl-CoA abundance and fate. J. Biol. Chem. 2014, 289, 25382–25392. [Google Scholar] [CrossRef] [Green Version]
  95. Shyh-Chang, N.; Locasale, J.W.; Lyssiotis, C.A.; Zheng, Y.; Teo, R.Y.; Ratanasirintrawoot, S.; Zhang, J.; Onder, T.; Unternaehrer, J.J.; Zhu, H.; et al. Influence of threonine metabolism on S-adenosylmethionine and histone methylation. Science 2013, 339, 222–226. [Google Scholar] [CrossRef] [Green Version]
  96. Moussaieff, A.; Rouleau, M.; Kitsberg, D.; Cohen, M.; Levy, G.; Barasch, D.; Nemirovski, A.; Shen-Orr, S.; Laevsky, I.; Amit, M.; et al. Glycolysis-mediated changes in acetyl-CoA and histone acetylation control the early differentiation of embryonic stem cells. Cell Metab. 2015, 21, 392–402. [Google Scholar] [CrossRef] [Green Version]
  97. Shiraki, N.; Shiraki, Y.; Tsuyama, T.; Obata, F.; Miura, M.; Nagae, G.; Aburatani, H.; Kume, K.; Endo, F.; Kume, S. Methionine metabolism regulates maintenance and differentiation of human pluripotent stem cells. Cell Metab. 2014, 19, 780–794. [Google Scholar] [CrossRef] [Green Version]
  98. Carey, B.W.; Finley, L.W.; Cross, J.R.; Allis, C.D.; Thompson, C.B. Intracellular alpha-ketoglutarate maintains the pluripotency of embryonic stem cells. Nature 2015, 518, 413–416. [Google Scholar] [CrossRef] [Green Version]
  99. Spitalieri, P.; Talarico, V.R.; Murdocca, M.; Novelli, G.; Sangiuolo, F. Human induced pluripotent stem cells for monogenic disease modelling and therapy. World J. Stem Cells 2016, 8, 118–135. [Google Scholar] [CrossRef]
  100. Streckfuss-Bomeke, K.; Wolf, F.; Azizian, A.; Stauske, M.; Tiburcy, M.; Wagner, S.; Hubscher, D.; Dressel, R.; Chen, S.; Jende, J.; et al. Comparative study of human-induced pluripotent stem cells derived from bone marrow cells, hair keratinocytes, and skin fibroblasts. Eur. Heart J. 2013, 34, 2618–2629. [Google Scholar] [CrossRef] [Green Version]
  101. Rao, L.; Qian, Y.; Khodabukus, A.; Ribar, T.; Bursac, N. Engineering human pluripotent stem cells into a functional skeletal muscle tissue. Nat. Commun. 2018, 9, 126. [Google Scholar] [CrossRef] [Green Version]
  102. Cyganek, L.; Tiburcy, M.; Sekeres, K.; Gerstenberg, K.; Bohnenberger, H.; Lenz, C.; Henze, S.; Stauske, M.; Salinas, G.; Zimmermann, W.H.; et al. Deep phenotyping of human induced pluripotent stem cell-derived atrial and ventricular cardiomyocytes. JCI Insight 2018, 3, 12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Lee, J.; Böscke, R.; Tang, P.C.; Hartman, B.H.; Heller, S.; Koehler, K.R. Hair Follicle Development in Mouse Pluripotent Stem Cell-Derived Skin Organoids. Cell Rep. 2018, 22, 242–254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Yamamoto, Y.; Gotoh, S.; Korogi, Y.; Seki, M.; Konishi, S.; Ikeo, S.; Sone, N.; Nagasaki, T.; Matsumoto, H.; Muro, S.; et al. Long-term expansion of alveolar stem cells derived from human iPS cells in organoids. Nat. Methods 2017, 14, 1097–1106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Ardeshirylajimi, A. Applied Induced Pluripotent Stem Cells in Combination with Biomaterials in Bone Tissue Engineering. J. Cell. Biochem. 2017, 118, 3034–3042. [Google Scholar] [CrossRef]
  106. Wang, A.; Tang, Z.; Park, I.H.; Zhu, Y.; Patel, S.; Daley, G.Q.; Li, S. Induced pluripotent stem cells for neural tissue engineering. Biomaterials 2011, 32, 5023–5032. [Google Scholar] [CrossRef] [Green Version]
  107. Brambrink, T.; Foreman, R.; Welstead, G.G.; Lengner, C.J.; Wernig, M.; Suh, H.; Jaenisch, R. Sequential expression of pluripotency markers during direct reprogramming of mouse somatic cells. Cell Stem Cell 2008, 2, 151–159. [Google Scholar] [CrossRef] [Green Version]
  108. Sridharan, R.; Tchieu, J.; Mason, M.J.; Yachechko, R.; Kuoy, E.; Horvath, S.; Zhou, Q.; Plath, K. Role of the murine reprogramming factors in the induction of pluripotency. Cell 2009, 136, 364–377. [Google Scholar] [CrossRef] [Green Version]
  109. Soufi, A.; Donahue, G.; Zaret, K.S. Facilitators and impediments of the pluripotency reprogramming factors’ initial engagement with the genome. Cell 2012, 151, 994–1004. [Google Scholar] [CrossRef] [Green Version]
  110. Okita, K.; Ichisaka, T.; Yamanaka, S. Generation of germline-competent induced pluripotent stem cells. Nature 2007, 448, 313–317. [Google Scholar] [CrossRef]
  111. Nakagawa, M.; Koyanagi, M.; Tanabe, K.; Takahashi, K.; Ichisaka, T.; Aoi, T.; Okita, K.; Mochiduki, Y.; Takizawa, N.; Yamanaka, S. Generation of induced pluripotent stem cells without Myc from mouse and human fibroblasts. Nat. Biotechnol. 2008, 26, 101–106. [Google Scholar] [CrossRef]
  112. Wernig, M.; Meissner, A.; Cassady, J.P.; Jaenisch, R. c-Myc is dispensable for direct reprogramming of mouse fibroblasts. Cell Stem Cell 2008, 2, 10–12. [Google Scholar] [CrossRef] [Green Version]
  113. Nakagawa, M.; Takizawa, N.; Narita, M.; Ichisaka, T.; Yamanaka, S. Promotion of direct reprogramming by transformation-deficient Myc. Proc. Natl. Acad. Sci. USA 2010, 107, 14152–14157. [Google Scholar] [CrossRef]
  114. Torchy, M.P.; Hamiche, A.; Klaholz, B.P. Structure and function insights into the NuRD chromatin remodeling complex. Cell. Mol. Life Sci. 2015, 72, 2491–2507. [Google Scholar] [CrossRef]
  115. Gingold, H.; Tehler, D.; Christoffersen, N.R.; Nielsen, M.M.; Asmar, F.; Kooistra, S.M.; Christophersen, N.S.; Christensen, L.L.; Borre, M.; Sorensen, K.D.; et al. A dual program for translation regulation in cellular proliferation and differentiation. Cell 2014, 158, 1281–1292. [Google Scholar] [CrossRef] [Green Version]
  116. Zviran, A.; Mor, N.; Rais, Y.; Gingold, H.; Peles, S.; Chomsky, E.; Viukov, S.; Buenrostro, J.D.; Scognamiglio, R.; Weinberger, L.; et al. Deterministic Somatic Cell Reprogramming Involves Continuous Transcriptional Changes Governed by Myc and Epigenetic-Driven Modules. Cell Stem Cell 2019, 24, 328–341.e329. [Google Scholar] [CrossRef] [Green Version]
  117. Felton-Edkins, Z.A.; Kenneth, N.S.; Brown, T.R.; Daly, N.L.; Gomez-Roman, N.; Grandori, C.; Eisenman, R.N.; White, R.J. Direct regulation of RNA polymerase III transcription by RB, p53 and c-Myc. Cell Cycle 2003, 2, 181–184. [Google Scholar] [CrossRef] [Green Version]
  118. Campbell, K.J.; White, R.J. MYC regulation of cell growth through control of transcription by RNA polymerases I and III. Cold Spring Harb. Perspect. Med. 2014, 4, 5. [Google Scholar] [CrossRef]
  119. Lepilina, A.; Coon, A.N.; Kikuchi, K.; Holdway, J.E.; Roberts, R.W.; Burns, C.G.; Poss, K.D. A dynamic epicardial injury response supports progenitor cell activity during zebrafish heart regeneration. Cell 2006, 127, 607–619. [Google Scholar] [CrossRef] [Green Version]
  120. Fausett, B.V.; Goldman, D. A role for alpha1 tubulin-expressing Muller glia in regeneration of the injured zebrafish retina. J. Neurosci. 2006, 26, 6303–6313. [Google Scholar] [CrossRef] [Green Version]
  121. Ransom, R.C.; Carter, A.C.; Salhotra, A.; Leavitt, T.; Marecic, O.; Murphy, M.P.; Lopez, M.L.; Wei, Y.; Marshall, C.D.; Shen, E.Z.; et al. Mechanoresponsive stem cells acquire neural crest fate in jaw regeneration. Nature 2018, 563, 514–521. [Google Scholar] [CrossRef] [PubMed]
  122. Gupta, V.; Gemberling, M.; Karra, R.; Rosenfeld, G.E.; Evans, T.; Poss, K.D. An injury-responsive gata4 program shapes the zebrafish cardiac ventricle. Curr. Biol. 2013, 23, 1221–1227. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Schindler, Y.L.; Garske, K.M.; Wang, J.; Firulli, B.A.; Firulli, A.B.; Poss, K.D.; Yelon, D. Hand2 elevates cardiomyocyte production during zebrafish heart development and regeneration. Development 2014, 141, 3112–3122. [Google Scholar] [CrossRef] [Green Version]
  124. Oka, T.; Xu, J.; Molkentin, J.D. Re-employment of developmental transcription factors in adult heart disease. Semin. Cell Dev. Biol. 2007, 18, 117–131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Bisping, E.; Ikeda, S.; Kong, S.W.; Tarnavski, O.; Bodyak, N.; McMullen, J.R.; Rajagopal, S.; Son, J.K.; Ma, Q.; Springer, Z.; et al. Gata4 is required for maintenance of postnatal cardiac function and protection from pressure overload-induced heart failure. Proc. Natl. Acad. Sci. USA 2006, 103, 14471–14476. [Google Scholar] [CrossRef]
  126. Eming, S.A.; Wynn, T.A.; Martin, P. Inflammation and metabolism in tissue repair and regeneration. Science 2017, 356, 1026–1030. [Google Scholar] [CrossRef] [Green Version]
  127. Matias Santos, D.; Rita, A.M.; Casanellas, I.; Brito Ova, A.; Araujo, I.M.; Power, D.; Tiscornia, G. Ear wound regeneration in the African spiny mouse Acomys cahirinus. Regeneration 2016, 3, 52–61. [Google Scholar] [CrossRef]
  128. Simkin, J.; Gawriluk, T.R.; Gensel, J.C.; Seifert, A.W. Macrophages are necessary for epimorphic regeneration in African spiny mice. eLife 2017, 6, e24623. [Google Scholar] [CrossRef]
  129. Yang, K.; Kang, J. Tissue Regeneration Enhancer Elements: A Way to Unlock Endogenous Healing Power. Dev. Dyn. 2019, 248, 34–42. [Google Scholar] [CrossRef] [Green Version]
  130. Kang, J.; Hu, J.; Karra, R.; Dickson, A.L.; Tornini, V.A.; Nachtrab, G.; Gemberling, M.; Goldman, J.A.; Black, B.L.; Poss, K.D. Modulation of tissue repair by regeneration enhancer elements. Nature 2016, 532, 201–206. [Google Scholar] [CrossRef] [Green Version]
  131. Da Silva, S.M.; Gates, P.B.; Brockes, J.P. The newt ortholog of CD59 is implicated in proximodistal identity during amphibian limb regeneration. Dev. Cell 2002, 3, 547–555. [Google Scholar] [CrossRef]
  132. Vizcaya-Molina, E.; Klein, C.C.; Serras, F.; Mishra, R.K.; Guigo, R.; Corominas, M. Damage-responsive elements in Drosophila regeneration. Genome Res. 2018, 28, 1852–1866. [Google Scholar] [CrossRef] [Green Version]
  133. Yan, R.; Cigliola, V.; Oonk, K.A.; Petrover, Z.; DeLuca, S.; Wolfson, D.W.; Vekstein, A.; Mendiola, M.A.; Devlin, G.; Bishawi, M.; et al. An enhancer-based gene-therapy strategy for spatiotemporal control of cargoes during tissue repair. Cell Stem Cell 2023, 30, 96–111.e116. [Google Scholar] [CrossRef]
  134. Soukup, A.A.; Zheng, Y.; Mehta, C.; Wu, J.; Liu, P.; Cao, M.; Hofmann, I.; Zhou, Y.; Zhang, J.; Johnson, K.D.; et al. Single-nucleotide human disease mutation inactivates a blood-regenerative GATA2 enhancer. J. Clin. Investig. 2019, 129, 1180–1192. [Google Scholar] [CrossRef] [Green Version]
  135. Hewitt, K.J.; Katsumura, K.R.; Matson, D.R.; Devadas, P.; Tanimura, N.; Hebert, A.S.; Coon, J.J.; Kim, J.S.; Dewey, C.N.; Keles, S.; et al. GATA Factor-Regulated Samd14 Enhancer Confers Red Blood Cell Regeneration and Survival in Severe Anemia. Dev. Cell 2017, 42, 213–225.e214. [Google Scholar] [CrossRef] [Green Version]
  136. Gadye, L.; Das, D.; Sanchez, M.A.; Street, K.; Baudhuin, A.; Wagner, A.; Cole, M.B.; Choi, Y.G.; Yosef, N.; Purdom, E.; et al. Injury Activates Transient Olfactory Stem Cell States with Diverse Lineage Capacities. Cell Stem Cell 2017, 21, 775–790.e779. [Google Scholar] [CrossRef] [Green Version]
  137. Rocha-Martins, M.; de Toledo, B.C.; Santos-Franca, P.L.; Oliveira-Valenca, V.M.; Vieira-Vieira, C.H.; Matos-Rodrigues, G.E.; Linden, R.; Norden, C.; Martins, R.A.P.; Silveira, M.S. De novo genesis of retinal ganglion cells by targeted expression of Klf4 in vivo. Development 2019, 146, 176586. [Google Scholar] [CrossRef] [Green Version]
  138. Mahmoud, A.I.; Kocabas, F.; Muralidhar, S.A.; Kimura, W.; Koura, A.S.; Thet, S.; Porrello, E.R.; Sadek, H.A. Meis1 regulates postnatal cardiomyocyte cell cycle arrest. Nature 2013, 497, 249–253. [Google Scholar] [CrossRef] [Green Version]
  139. Xin, M.; Kim, Y.; Sutherland, L.B.; Murakami, M.; Qi, X.; McAnally, J.; Porrello, E.R.; Mahmoud, A.I.; Tan, W.; Shelton, J.M.; et al. Hippo pathway effector Yap promotes cardiac regeneration. Proc. Natl. Acad. Sci. USA 2013, 110, 13839–13844. [Google Scholar] [CrossRef]
  140. Patterson, M.; Barske, L.; Van Handel, B.; Rau, C.D.; Gan, P.; Sharma, A.; Parikh, S.; Denholtz, M.; Huang, Y.; Yamaguchi, Y.; et al. Frequency of mononuclear diploid cardiomyocytes underlies natural variation in heart regeneration. Nat. Genet. 2017, 49, 1346–1353. [Google Scholar] [CrossRef]
  141. Guzman-Ayala, M.; Sachs, M.; Koh, F.M.; Onodera, C.; Bulut-Karslioglu, A.; Lin, C.J.; Wong, P.; Nitta, R.; Song, J.S.; Ramalho-Santos, M. Chd1 is essential for the high transcriptional output and rapid growth of the mouse epiblast. Development 2015, 142, 118–127. [Google Scholar] [CrossRef] [Green Version]
  142. Mattout, A.; Meshorer, E. Chromatin plasticity and genome organization in pluripotent embryonic stem cells. Curr. Opin. Cell Biol. 2010, 22, 334–341. [Google Scholar] [CrossRef] [PubMed]
  143. Percharde, M.; Bulut-Karslioglu, A.; Ramalho-Santos, M. Hypertranscription in Development, Stem Cells, and Regeneration. Dev. Cell 2017, 40, 9–21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Jao, C.Y.; Salic, A. Exploring RNA transcription and turnover in vivo by using click chemistry. Proc. Natl. Acad. Sci. USA 2008, 105, 15779–15784. [Google Scholar] [CrossRef] [PubMed]
  145. Dai, Y.S.; Cserjesi, P.; Markham, B.E.; Molkentin, J.D. The transcription factors GATA4 and dHAND physically interact to synergistically activate cardiac gene expression through a p300-dependent mechanism. J. Biol. Chem. 2002, 277, 24390–24398. [Google Scholar] [CrossRef] [Green Version]
  146. Goldman, J.A.; Kuzu, G.; Lee, N.; Karasik, J.; Gemberling, M.; Foglia, M.J.; Karra, R.; Dickson, A.L.; Sun, F.; Tolstorukov, M.Y.; et al. Resolving Heart Regeneration by Replacement Histone Profiling. Dev. Cell 2017, 40, 392–404.e395. [Google Scholar] [CrossRef] [Green Version]
  147. Li, W.; Long, Q.; Wu, H.; Zhou, Y.; Duan, L.; Yuan, H.; Ding, Y.; Huang, Y.; Wu, Y.; Huang, J.; et al. Nuclear localization of mitochondrial TCA cycle enzymes modulates pluripotency via histone acetylation. Nat. Commun. 2022, 13, 7414. [Google Scholar] [CrossRef]
  148. Li, L.; Chen, K.; Wang, T.; Wu, Y.; Xing, G.; Chen, M.; Hao, Z.; Zhang, C.; Zhang, J.; Ma, B.; et al. Glis1 facilitates induction of pluripotency via an epigenome-metabolome-epigenome signalling cascade. Nat. Metab. 2020, 2, 882–892. [Google Scholar] [CrossRef]
  149. Yucel, N.; Wang, Y.X.; Mai, T.; Porpiglia, E.; Lund, P.J.; Markov, G.; Garcia, B.A.; Bendall, S.C.; Angelo, M.; Blau, H.M. Glucose Metabolism Drives Histone Acetylation Landscape Transitions that Dictate Muscle Stem Cell Function. Cell Rep. 2019, 27, 3939–3955.e3936. [Google Scholar] [CrossRef] [Green Version]
  150. He, Y.; Wang, S.; Liu, S.; Qin, D.; Liu, Z.; Wang, L.; Chen, X.; Zhang, L. MSL1 Promotes Liver Regeneration by Driving Phase Separation of STAT3 and Histone H4 and Enhancing Their Acetylation. Adv. Sci. 2023, e2301094. [Google Scholar] [CrossRef]
  151. Das, S.; Morvan, F.; Morozzi, G.; Jourde, B.; Minetti, G.C.; Kahle, P.; Rivet, H.; Brebbia, P.; Toussaint, G.; Glass, D.J.; et al. ATP Citrate Lyase Regulates Myofiber Differentiation and Increases Regeneration by Altering Histone Acetylation. Cell Rep. 2017, 21, 3003–3011. [Google Scholar] [CrossRef] [Green Version]
  152. Sun, J.; Ji, Y.; Liang, Q.; Ming, M.; Chen, Y.; Zhang, Q.; Zhou, S.; Shen, M.; Ding, F. Expression of Protein Acetylation Regulators During Peripheral Nerve Development, Injury, and Regeneration. Front. Mol. Neurosci. 2022, 15, 888523. [Google Scholar] [CrossRef]
  153. Huynh, N.C.; Everts, V.; Ampornaramveth, R.S. Histone deacetylases and their roles in mineralized tissue regeneration. Bone Rep. 2017, 7, 33–40. [Google Scholar] [CrossRef]
  154. Yamauchi, Y.; Cooper, P.R.; Shimizu, E.; Kobayashi, Y.; Smith, A.J.; Duncan, H.F. Histone Acetylation as a Regenerative Target in the Dentine-Pulp Complex. Front. Genet. 2020, 11, 1. [Google Scholar] [CrossRef]
  155. Yadav, A.; Huang, T.C.; Chen, S.H.; Ramasamy, T.S.; Hsueh, Y.Y.; Lin, S.P.; Lu, F.I.; Liu, Y.H.; Wu, C.C. Sodium phenylbutyrate inhibits Schwann cell inflammation via HDAC and NFkappaB to promote axonal regeneration and remyelination. J. Neuroinflammation 2021, 18, 238. [Google Scholar] [CrossRef]
  156. Cho, Y.; Cavalli, V. HDAC signaling in neuronal development and axon regeneration. Curr. Opin. Neurobiol. 2014, 27, 118–126. [Google Scholar] [CrossRef] [Green Version]
  157. Ganai, S.A.; Ramadoss, M.; Mahadevan, V. Histone Deacetylase (HDAC) Inhibitors–emerging roles in neuronal memory, learning, synaptic plasticity and neural regeneration. Curr. Neuropharmacol. 2016, 14, 55–71. [Google Scholar] [CrossRef] [Green Version]
  158. Fessler, E.B.; Chibane, F.L.; Wang, Z.; Chuang, D.M. Potential roles of HDAC inhibitors in mitigating ischemia-induced brain damage and facilitating endogenous regeneration and recovery. Curr. Pharm. Des. 2013, 19, 5105–5120. [Google Scholar] [CrossRef]
  159. Stewart, S.; Tsun, Z.Y.; Izpisua Belmonte, J.C. A histone demethylase is necessary for regeneration in zebrafish. Proc. Natl. Acad. Sci. USA 2009, 106, 19889–19894. [Google Scholar] [CrossRef]
  160. Ma, K.H.; Duong, P.; Moran, J.J.; Junaidi, N.; Svaren, J. Polycomb repression regulates Schwann cell proliferation and axon regeneration after nerve injury. Glia 2018, 66, 2487–2502. [Google Scholar] [CrossRef]
  161. Duan, R.S.; Tang, G.B.; Du, H.Z.; Hu, Y.W.; Liu, P.P.; Xu, Y.J.; Zeng, Y.Q.; Zhang, S.F.; Wang, R.Y.; Teng, Z.Q.; et al. Polycomb protein family member CBX7 regulates intrinsic axon growth and regeneration. Cell Death Differ. 2018, 25, 1598–1611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Guo, C.; Balsara, Z.R.; Hill, W.G.; Li, X. Stage- and subunit-specific functions of polycomb repressive complex 2 in bladder urothelial formation and regeneration. Development 2017, 144, 400–408. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Ai, S.; Yu, X.; Li, Y.; Peng, Y.; Li, C.; Yue, Y.; Tao, G.; Li, C.; Pu, W.T.; He, A. Divergent Requirements for EZH1 in Heart Development Versus Regeneration. Circ. Res. 2017, 121, 106–112. [Google Scholar] [CrossRef] [PubMed]
  164. 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]
  165. Li, W.; Yang, L.; He, Q.; Hu, C.; Zhu, L.; Ma, X.; Ma, X.; Bao, S.; Li, L.; Chen, Y.; et al. A Homeostatic Arid1a-Dependent Permissive Chromatin State Licenses Hepatocyte Responsiveness to Liver-Injury-Associated YAP Signaling. Cell Stem Cell 2019, 25, 54–68.e55. [Google Scholar] [CrossRef]
  166. Wang, J.; Liu, S.; Heallen, T.; Martin, J.F. The Hippo pathway in the heart: Pivotal roles in development, disease, and regeneration. Nat. Rev. Cardiol. 2018, 15, 672–684. [Google Scholar] [CrossRef]
  167. Moya, I.M.; Halder, G. Hippo-YAP/TAZ signalling in organ regeneration and regenerative medicine. Nat. Rev. Mol. Cell Biol. 2019, 20, 211–226. [Google Scholar] [CrossRef]
  168. Benham-Pyle, B.W.; Pruitt, B.L.; Nelson, W.J. Cell adhesion. Mechanical strain induces E-cadherin-dependent Yap1 and beta-catenin activation to drive cell cycle entry. Science 2015, 348, 1024–1027. [Google Scholar] [CrossRef] [Green Version]
  169. Chang, L.; Azzolin, L.; Di Biagio, D.; Zanconato, F.; Battilana, G.; Lucon Xiccato, R.; Aragona, M.; Giulitti, S.; Panciera, T.; Gandin, A.; et al. The SWI/SNF complex is a mechanoregulated inhibitor of YAP and TAZ. Nature 2018, 563, 265–269. [Google Scholar] [CrossRef]
  170. Baena, E.; Gandarillas, A.; Vallespinos, M.; Zanet, J.; Bachs, O.; Redondo, C.; Fabregat, I.; Martinez, A.C.; de Alboran, I.M. c-Myc regulates cell size and ploidy but is not essential for postnatal proliferation in liver. Proc. Natl. Acad. Sci. USA 2005, 102, 7286–7291. [Google Scholar] [CrossRef]
  171. Wang, H.; Lu, J.; Alencastro, F.; Roberts, A.; Fiedor, J.; Carroll, P.; Eisenman, R.N.; Ranganathan, S.; Torbenson, M.; Duncan, A.W.; et al. Coordinated Cross-Talk Between the Myc and Mlx Networks in Liver Regeneration and Neoplasia. Cell Mol. Gastroenterol. Hepatol. 2022, 13, 1785–1804. [Google Scholar] [CrossRef]
  172. Lasorella, A.; Noseda, M.; Beyna, M.; Yokota, Y.; Iavarone, A. Id2 is a retinoblastoma protein target and mediates signalling by Myc oncoproteins. Nature 2000, 407, 592–598. [Google Scholar] [CrossRef]
  173. Lasorella, A.; Boldrini, R.; Dominici, C.; Donfrancesco, A.; Yokota, Y.; Inserra, A.; Iavarone, A. Id2 is critical for cellular proliferation and is the oncogenic effector of N-myc in human neuroblastoma. Cancer Res. 2002, 62, 301–306. [Google Scholar] [PubMed]
  174. Cotta, C.V.; Leventaki, V.; Atsaves, V.; Vidaki, A.; Schlette, E.; Jones, D.; Medeiros, L.J.; Rassidakis, G.Z. The helix-loop-helix protein Id2 is expressed differentially and induced by myc in T-cell lymphomas. Cancer 2008, 112, 552–561. [Google Scholar] [CrossRef] [PubMed]
  175. Rodriguez, J.L.; Sandoval, J.; Serviddio, G.; Sastre, J.; Morante, M.; Perrelli, M.G.; Martinez-Chantar, M.L.; Vina, J.; Vina, J.R.; Mato, J.M.; et al. Id2 leaves the chromatin of the E2F4-p130-controlled c-myc promoter during hepatocyte priming for liver regeneration. Biochem. J. 2006, 398, 431–437. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Zhang, C.; Chang, C.; Gao, H.; Wang, Q.; Zhang, F.; Xu, C. MiR-429 regulates rat liver regeneration and hepatocyte proliferation by targeting JUN/MYC/BCL2/CCND1 signaling pathway. Cell. Signal. 2018, 50, 80–89. [Google Scholar] [CrossRef]
  177. Winkler, M.; Staniczek, T.; Kurschner, S.W.; Schmid, C.D.; Schonhaber, H.; Cordero, J.; Kessler, L.; Mathes, A.; Sticht, C.; Nessling, M.; et al. Endothelial GATA4 controls liver fibrosis and regeneration by preventing a pathogenic switch in angiocrine signaling. J. Hepatol. 2021, 74, 380–393. [Google Scholar] [CrossRef]
  178. Jonas, J.C.; Laybutt, D.R.; Steil, G.M.; Trivedi, N.; Pertusa, J.G.; Van de Casteele, M.; Weir, G.C.; Henquin, J.C. High glucose stimulates early response gene c-Myc expression in rat pancreatic beta cells. J. Biol. Chem. 2001, 276, 35375–35381. [Google Scholar] [CrossRef] [Green Version]
  179. Puri, S.; Roy, N.; Russ, H.A.; Leonhardt, L.; French, E.K.; Roy, R.; Bengtsson, H.; Scott, D.K.; Stewart, A.F.; Hebrok, M. Replication confers beta cell immaturity. Nat. Commun. 2018, 9, 485. [Google Scholar] [CrossRef] [Green Version]
  180. Laybutt, D.R.; Weir, G.C.; Kaneto, H.; Lebet, J.; Palmiter, R.D.; Sharma, A.; Bonner-Weir, S. Overexpression of c-Myc in beta-cells of transgenic mice causes proliferation and apoptosis, downregulation of insulin gene expression, and diabetes. Diabetes 2002, 51, 1793–1804. [Google Scholar] [CrossRef] [Green Version]
  181. Cheung, L.; Zervou, S.; Mattsson, G.; Abouna, S.; Zhou, L.; Ifandi, V.; Pelengaris, S.; Khan, M. c-Myc directly induces both impaired insulin secretion and loss of beta-cell mass, independently of hyperglycemia in vivo. Islets 2010, 2, 37–45. [Google Scholar] [CrossRef] [Green Version]
  182. Pelengaris, S.; Khan, M.; Evan, G.I. Suppression of Myc-induced apoptosis in beta cells exposes multiple oncogenic properties of Myc and triggers carcinogenic progression. Cell 2002, 109, 321–334. [Google Scholar] [CrossRef] [Green Version]
  183. Rosselot, C.; Kumar, A.; Lakshmipathi, J.; Zhang, P.; Lu, G.; Katz, L.S.; Prochownik, E.V.; Stewart, A.F.; Lambertini, L.; Scott, D.K.; et al. Myc Is Required for Adaptive beta-Cell Replication in Young Mice but Is Not Sufficient in One-Year-Old Mice Fed with a High-Fat Diet. Diabetes 2019, 68, 1934–1949. [Google Scholar] [CrossRef]
  184. Rosselot, C.; Baumel-Alterzon, S.; Li, Y.; Brill, G.; Lambertini, L.; Katz, L.S.; Lu, G.; Garcia-Ocana, A.; Scott, D.K. The many lives of Myc in the pancreatic beta-cell. J. Biol. Chem. 2021, 296, 100122. [Google Scholar] [CrossRef]
  185. Bettess, M.D.; Dubois, N.; Murphy, M.J.; Dubey, C.; Roger, C.; Robine, S.; Trumpp, A. c-Myc is required for the formation of intestinal crypts but dispensable for homeostasis of the adult intestinal epithelium. Mol. Cell. Biol. 2005, 25, 7868–7878. [Google Scholar] [CrossRef] [Green Version]
  186. Konsavage, W.M., Jr.; Jin, G.; Yochum, G.S. The Myc 3’ Wnt-responsive element regulates homeostasis and regeneration in the mouse intestinal tract. Mol. Cell. Biol. 2012, 32, 3891–3902. [Google Scholar] [CrossRef] [Green Version]
  187. Muncan, V.; Sansom, O.J.; Tertoolen, L.; Phesse, T.J.; Begthel, H.; Sancho, E.; Cole, A.M.; Gregorieff, A.; de Alboran, I.M.; Clevers, H.; et al. Rapid loss of intestinal crypts upon conditional deletion of the Wnt/Tcf-4 target gene c-Myc. Mol. Cell. Biol. 2006, 26, 8418–8426. [Google Scholar] [CrossRef] [Green Version]
  188. Ashton, G.H.; Morton, J.P.; Myant, K.; Phesse, T.J.; Ridgway, R.A.; Marsh, V.; Wilkins, J.A.; Athineos, D.; Muncan, V.; Kemp, R.; et al. Focal adhesion kinase is required for intestinal regeneration and tumorigenesis downstream of Wnt/c-Myc signaling. Dev. Cell 2010, 19, 259–269. [Google Scholar] [CrossRef] [Green Version]
  189. Kim, M.J.; Xia, B.; Suh, H.N.; Lee, S.H.; Jun, S.; Lien, E.M.; Zhang, J.; Chen, K.; Park, J.I. PAF-Myc-Controlled Cell Stemness Is Required for Intestinal Regeneration and Tumorigenesis. Dev. Cell 2018, 44, 582–596.e584. [Google Scholar] [CrossRef]
  190. De Pretis, S.; Kress, T.R.; Morelli, M.J.; Sabo, A.; Locarno, C.; Verrecchia, A.; Doni, M.; Campaner, S.; Amati, B.; Pelizzola, M. Integrative analysis of RNA polymerase II and transcriptional dynamics upon MYC activation. Genome Res. 2017, 27, 1658–1664. [Google Scholar] [CrossRef] [Green Version]
  191. Bywater, M.J.; Burkhart, D.L.; Straube, J.; Sabo, A.; Pendino, V.; Hudson, J.E.; Quaife-Ryan, G.A.; Porrello, E.R.; Rae, J.; Parton, R.G.; et al. Reactivation of Myc transcription in the mouse heart unlocks its proliferative capacity. Nat. Commun. 2020, 11, 1827. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Li, M.; Zheng, H.; Han, Y.; Chen, Y.; Li, B.; Chen, G.; Chen, X.; Huang, S.; He, X.; Wei, G.; et al. LncRNA Snhg1-driven self-reinforcing regulatory network promoted cardiac regeneration and repair after myocardial infarction. Theranostics 2021, 11, 9397–9414. [Google Scholar] [CrossRef] [PubMed]
  193. Chen, Y.; Luttmann, F.F.; Schoger, E.; Scholer, H.R.; Zelarayan, L.C.; Kim, K.P.; Haigh, J.J.; Kim, J.; Braun, T. Reversible reprogramming of cardiomyocytes to a fetal state drives heart regeneration in mice. Science 2021, 373, 1537–1540. [Google Scholar] [CrossRef] [PubMed]
  194. Klein, J.A.; Ackerman, S.L. Oxidative stress, cell cycle, and neurodegeneration. J. Clin. Investig. 2003, 111, 785–793. [Google Scholar] [CrossRef] [Green Version]
  195. Yang, S.G.; Wang, X.W.; Qian, C.; Zhou, F.Q. Reprogramming neurons for regeneration: The fountain of youth. Prog. Neurobiol. 2022, 214, 102284. [Google Scholar] [CrossRef]
  196. Ferrer, I.; Blanco, R.; Carmona, M.; Puig, B. Phosphorylated c-MYC expression in Alzheimer disease, Pick’s disease, progressive supranuclear palsy and corticobasal degeneration. Neuropathol. Appl. Neurobiol. 2001, 27, 343–351. [Google Scholar] [CrossRef]
  197. Ferrer, I.; Blanco, R. N-myc and c-myc expression in Alzheimer disease, Huntington disease and Parkinson disease. Brain Res. Mol. Brain Res. 2000, 77, 270–276. [Google Scholar] [CrossRef]
  198. Porta, C.; Paglino, C.; Mosca, A. Targeting PI3K/Akt/mTOR Signaling in Cancer. Front. Oncol. 2014, 4, 64. [Google Scholar] [CrossRef] [Green Version]
  199. Pei, J.J.; Hugon, J. mTOR-dependent signalling in Alzheimer’s disease. J. Cell. Mol. Med. 2008, 12, 2525–2532. [Google Scholar] [CrossRef] [Green Version]
  200. Arendt, T. Synaptic plasticity and cell cycle activation in neurons are alternative effector pathways: The ‘Dr. Jekyll and Mr. Hyde concept’ of Alzheimer’s disease or the yin and yang of neuroplasticity. Prog. Neurobiol. 2003, 71, 83–248. [Google Scholar] [CrossRef]
  201. Hay, N. Interplay between FOXO, TOR, and Akt. Biochim. Biophys. Acta 2011, 1813, 1965–1970. [Google Scholar] [CrossRef] [Green Version]
  202. Gomez-Crisostomo, N.P.; Rodriguez Martinez, E.; Rivas-Arancibia, S. Oxidative stress activates the transcription factors FoxO 1a and FoxO 3a in the hippocampus of rats exposed to low doses of ozone. Oxidative Med. Cell. Longev. 2014, 2014, 805764. [Google Scholar] [CrossRef] [Green Version]
  203. Chandramohan, V.; Jeay, S.; Pianetti, S.; Sonenshein, G.E. Reciprocal control of Forkhead box O 3a and c-Myc via the phosphatidylinositol 3-kinase pathway coordinately regulates p27Kip1 levels. J. Immunol. 2004, 172, 5522–5527. [Google Scholar] [CrossRef] [Green Version]
  204. Riddell, M.; Nakayama, A.; Hikita, T.; Mirzapourshafiyi, F.; Kawamura, T.; Pasha, A.; Li, M.; Masuzawa, M.; Looso, M.; Steinbacher, T.; et al. aPKC controls endothelial growth by modulating c-Myc via FoxO1 DNA-binding ability. Nat. Commun. 2018, 9, 5357. [Google Scholar] [CrossRef] [Green Version]
  205. Majd, S.; Power, J.; Majd, Z. Alzheimer’s Disease and Cancer: When Two Monsters Cannot Be Together. Front. Neurosci. 2019, 13, 155. [Google Scholar] [CrossRef] [Green Version]
  206. Qin, Z.H.; Chen, R.W.; Wang, Y.; Nakai, M.; Chuang, D.M.; Chase, T.N. Nuclear factor kappaB nuclear translocation upregulates c-Myc and p53 expression during NMDA receptor-mediated apoptosis in rat striatum. J. Neurosci. 1999, 19, 4023–4033. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Belin, S.; Nawabi, H.; Wang, C.; Tang, S.; Latremoliere, A.; Warren, P.; Schorle, H.; Uncu, C.; Woolf, C.J.; He, Z.; et al. Injury-induced decline of intrinsic regenerative ability revealed by quantitative proteomics. Neuron 2015, 86, 1000–1014. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Ma, J.J.; Ju, X.; Xu, R.J.; Wang, W.H.; Luo, Z.P.; Liu, C.M.; Yang, L.; Li, B.; Chen, J.Q.; Meng, B.; et al. Telomerase Reverse Transcriptase and p53 Regulate Mammalian Peripheral Nervous System and CNS Axon Regeneration Downstream of c-Myc. J. Neurosci. 2019, 39, 9107–9118. [Google Scholar] [CrossRef]
  209. Peck, B.; Ferber, E.C.; Schulze, A. Antagonism between FOXO and MYC Regulates Cellular Powerhouse. Front. Oncol. 2013, 3, 96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Oli, V.; Gupta, R.; Kumar, P. FOXO and related transcription factors binding elements in the regulation of neurodegenerative disorders. J. Chem. Neuroanat. 2021, 116, 102012. [Google Scholar] [CrossRef] [PubMed]
  211. Hishida, T.; Vazquez-Ferrer, E.; Hishida-Nozaki, Y.; Takemoto, Y.; Hatanaka, F.; Yoshida, K.; Prieto, J.; Sahu, S.K.; Takahashi, Y.; Reddy, P.; et al. Myc Supports Self-Renewal of Basal Cells in the Esophageal Epithelium. Front. Cell Dev. Biol. 2022, 10, 786031. [Google Scholar] [CrossRef]
  212. Huang, X.; Sun, J.; Chen, G.; Niu, C.; Wang, Y.; Zhao, C.; Sun, J.; Huang, H.; Huang, S.; Liang, Y.; et al. Resveratrol Promotes Diabetic Wound Healing via SIRT1-FOXO1-c-Myc Signaling Pathway-Mediated Angiogenesis. Front. Pharmacol. 2019, 10, 421. [Google Scholar] [CrossRef] [Green Version]
  213. Wang, C.; Rabadan Ros, R.; Martinez-Redondo, P.; Ma, Z.; Shi, L.; Xue, Y.; Guillen-Guillen, I.; Huang, L.; Hishida, T.; Liao, H.K.; et al. In vivo partial reprogramming of myofibers promotes muscle regeneration by remodeling the stem cell niche. Nat. Commun. 2021, 12, 3094. [Google Scholar] [CrossRef]
  214. Shu, Y.; Li, W.; Huang, M.; Quan, Y.Z.; Scheffer, D.; Tian, C.; Tao, Y.; Liu, X.; Hochedlinger, K.; Indzhykulian, A.A.; et al. Renewed proliferation in adult mouse cochlea and regeneration of hair cells. Nat. Commun. 2019, 10, 5530. [Google Scholar] [CrossRef] [Green Version]
  215. Ocampo, A.; Reddy, P.; Martinez-Redondo, P.; Platero-Luengo, A.; Hatanaka, F.; Hishida, T.; Li, M.; Lam, D.; Kurita, M.; Beyret, E.; et al. In Vivo Amelioration of Age-Associated Hallmarks by Partial Reprogramming. Cell 2016, 167, 1719–1733.e1712. [Google Scholar] [CrossRef] [Green Version]
  216. Kaufmann, K.B.; Buning, H.; Galy, A.; Schambach, A.; Grez, M. Gene therapy on the move. EMBO Mol. Med. 2013, 5, 1642–1661. [Google Scholar] [CrossRef]
  217. Deng, X.Y.; Wang, H.; Wang, T.; Fang, X.T.; Zou, L.L.; Li, Z.Y.; Liu, C.B. Non-viral methods for generating integration-free, induced pluripotent stem cells. Curr. Stem. Cell. Res. Ther. 2015, 10, 153–158. [Google Scholar] [CrossRef] [Green Version]
  218. Martin, I.; Simmons, P.J.; Williams, D.F. Manufacturing challenges in regenerative medicine. Sci. Transl. Med. 2014, 6, 232fs216. [Google Scholar] [CrossRef]
  219. Cossu, G.; Fears, R.; Griffin, G.; Ter Meulen, V. Regenerative medicine: Challenges and opportunities. Lancet 2020, 395, 1746–1747. [Google Scholar] [CrossRef]
Figure 1. Domain structure of Myc protein. Schematic representation of Myc and its major interacting partners. Abbreviations: MB = Myc box; KAT = lysine (K) acetyltransferase; TF = transcription factor; HDAC3 = histone deacetylase 3; P-TEFb = positive transcription elongation factor b.
Figure 1. Domain structure of Myc protein. Schematic representation of Myc and its major interacting partners. Abbreviations: MB = Myc box; KAT = lysine (K) acetyltransferase; TF = transcription factor; HDAC3 = histone deacetylase 3; P-TEFb = positive transcription elongation factor b.
Pathophysiology 30 00027 g001
Figure 2. The Myc/Max network. Picture depicting Max as the central hub for the Myc/Max network. According to the relative balance between Myc levels and Mad/Mx1/Mnt/Mga within a given cell, different heterodimers are formed, dictating chromatin state (condensed vs. relaxed), repression/transcription of different gene sets and cell fate.
Figure 2. The Myc/Max network. Picture depicting Max as the central hub for the Myc/Max network. According to the relative balance between Myc levels and Mad/Mx1/Mnt/Mga within a given cell, different heterodimers are formed, dictating chromatin state (condensed vs. relaxed), repression/transcription of different gene sets and cell fate.
Pathophysiology 30 00027 g002
Figure 3. Biological outcomes of increased Myc protein levels. Myc levels are tightly controlled within the cell. It may increase in narrow windows of expression upon mitogenic stimuli ruling the physiological homeostasis of highly regenerating permissive tissues and upon tissue damage. In highly terminally differentiated tissues, refractory to proliferation signals, Myc levels increased upon both acute and chronic tissue damage, with different outcomes. Indeed, either repair or tissue degeneration may occur. As well established, sustained and uncontrolled Myc levels promote its oncogenic properties.
Figure 3. Biological outcomes of increased Myc protein levels. Myc levels are tightly controlled within the cell. It may increase in narrow windows of expression upon mitogenic stimuli ruling the physiological homeostasis of highly regenerating permissive tissues and upon tissue damage. In highly terminally differentiated tissues, refractory to proliferation signals, Myc levels increased upon both acute and chronic tissue damage, with different outcomes. Indeed, either repair or tissue degeneration may occur. As well established, sustained and uncontrolled Myc levels promote its oncogenic properties.
Pathophysiology 30 00027 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Illi, B.; Nasi, S. Myc beyond Cancer: Regulation of Mammalian Tissue Regeneration. Pathophysiology 2023, 30, 346-365. https://doi.org/10.3390/pathophysiology30030027

AMA Style

Illi B, Nasi S. Myc beyond Cancer: Regulation of Mammalian Tissue Regeneration. Pathophysiology. 2023; 30(3):346-365. https://doi.org/10.3390/pathophysiology30030027

Chicago/Turabian Style

Illi, Barbara, and Sergio Nasi. 2023. "Myc beyond Cancer: Regulation of Mammalian Tissue Regeneration" Pathophysiology 30, no. 3: 346-365. https://doi.org/10.3390/pathophysiology30030027

Article Metrics

Back to TopTop