Next Article in Journal
In Silico Analysis of Ferroptosis-Related Genes and Its Implication in Drug Prediction against Fluorosis
Next Article in Special Issue
Brain Calcifications: Genetic, Molecular, and Clinical Aspects
Previous Article in Journal
Addressing Critical Issues Related to Storage and Stability of the Vault Nanoparticle Expressed and Purified from Komagataella phaffi
Previous Article in Special Issue
The SMN Complex at the Crossroad between RNA Metabolism and Neurodegeneration
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The “Superoncogene” Myc at the Crossroad between Metabolism and Gene Expression in Glioblastoma Multiforme

1
Institute of System Analysis and Informatics—National Research Council (IASI-CNR), 00185 Rome, Italy
2
Axxam SpA, Bresso, 20091 Milan, Italy
3
Department of Biology and Biotechnology Charles Darwin, Sapienza University, 00185 Rome, Italy
4
Istituto Pasteur Italia-Fondazione Cenci Bolognetti, Sapienza University, 00185 Rome, Italy
5
Institute of Molecular Biology and Pathology—National Research Council (IBPM-CNR), 00185 Rome, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(4), 4217; https://doi.org/10.3390/ijms24044217
Submission received: 30 December 2022 / Revised: 10 February 2023 / Accepted: 16 February 2023 / Published: 20 February 2023
(This article belongs to the Special Issue Neurogenetics in Neurology)

Abstract

:
The concept of the Myc (c-myc, n-myc, l-myc) oncogene as a canonical, DNA-bound transcription factor has consistently changed over the past few years. Indeed, Myc controls gene expression programs at multiple levels: directly binding chromatin and recruiting transcriptional coregulators; modulating the activity of RNA polymerases (RNAPs); and drawing chromatin topology. Therefore, it is evident that Myc deregulation in cancer is a dramatic event. Glioblastoma multiforme (GBM) is the most lethal, still incurable, brain cancer in adults, and it is characterized in most cases by Myc deregulation. Metabolic rewiring typically occurs in cancer cells, and GBM undergoes profound metabolic changes to supply increased energy demand. In nontransformed cells, Myc tightly controls metabolic pathways to maintain cellular homeostasis. Consistently, in Myc-overexpressing cancer cells, including GBM cells, these highly controlled metabolic routes are affected by enhanced Myc activity and show substantial alterations. On the other hand, deregulated cancer metabolism impacts Myc expression and function, placing Myc at the intersection between metabolic pathway activation and gene expression. In this review paper, we summarize the available information on GBM metabolism with a specific focus on the control of the Myc oncogene that, in turn, rules the activation of metabolic signals, ensuring GBM growth.

1. Introduction

Despite the great body of knowledge acquired over the past few years, especially at the molecular level, glioblastoma multiforme (GBM), a rare (but the most frequent) brain cancer in adults, still remains an incurable tumor. Survival prognosis is extremely unfavorable—between 15 and 18 months from diagnosis—and depends on both the lack of effective therapies and tumor recurrence [1]. A search of the US clinicaltrials.gov website annotated on 20 October 2022 yielded 1787 clinical studies on GBM, with most of them focused on testing novel or repositioned compounds, such as protein kinase inhibitors, monoclonal antibodies, and small molecules, or strategies, including intraoperative photodynamic therapy, the application of electric fields, and CAR-T immunotherapy. This proves the enormous effort the scientific community is lavishing on methods of novel therapeutic routes for the treatment of this lethal tumor. GBM is characterized by an elevated intratumoral heterogeneity at the single-cell level [2], infiltrative capacity of the surrounding brain tissue linked to neoangiogenesis [3], and the ability to suppress a patient’s immune response by TGF-β upregulation [4]. Three major GBM subtypes may be identified—proneural, classical, and mesenchymal—by converging transcription-based and transcriptomic studies (see below), with the proneural subtype showing the most favorable prognosis but less efficient response to therapy [5]. It is now well-established that, as with other types of cancer, GBM may stem from tumor-initiating cells—the so-called glioblastoma stem cells (GSCs) [6]. Nevertheless, a dedifferentiation process from astrocytes to neural stem cells that, in turn, may transform in GSCs cannot be excluded [7,8]. GSCs resemble the molecular characteristics of the GBM subtype they belong to, suggesting that differential therapeutic targets have to be taken into consideration when approaching a GBM-affected patient. However, within the same tumor, not only can GSCs with distinct molecular signatures be found, but they may evolve and shift from one subtype to another according to the tissue microenvironment (TME) and metabolic cues to which they are exposed [9]. These cells are now considered responsible for recurrency after surgery and resistance to the GBM gold-standard treatment consisting of radiotherapy, followed by adjuvant chemotherapy with temozolomide (TMZ) [1]. Therefore, an in-depth characterization of GSC molecular, biological, and metabolic properties may reasonably represent a promise for the design of more efficient and targeted treatments. Cancer metabolism has to face higher requests for energy supply to sustain enhanced proliferation, migration, and invasion capacity of tumor cells. These latter extract ready-to-use energy from nutrients, mainly glucose, pushing glycolysis, even in the presence of oxygen, known as the so-called Warburg effect [10]. However, tumor cells may also survive glucose starvation, rewiring their metabolism to oxidative phosphorylation (OXPHOS) and lipid oxidation [11,12]. This is also the case of GSCs, which are able to withstand glucose deprivation since they possess residual mitochondrial activity that allows the shift to the OXPHOS and fatty acids oxidation (FAO) pathways. GBM metabolism also relies on the genetic background of a tumor. Indeed, genes that are typically mutated in GBM, such as phosphatase and TENsin homolog deleted on chromosome 10 (PTEN), epidermal growth factor receptor (EGFR), isocitrate dehydrogenase (IDH1), p53, and v-raf murine sarcoma viral oncogene homolog B1 (BRAF), drive downstream metabolic pathways divergent from those found in healthy cells and are strictly dependent on the TME. One of the genes that controls and is controlled as a downstream effector by cell metabolism is the Myc oncogene, which is deregulated in the vast majority of cancers [13]. Myc deregulation may not only depend on genomic alterations in the myc locus (insertions, translocation, or amplification [13]) but also on indirect mechanisms, such as post-transcriptional control by microRNAs [14,15] or regulation of myc mRNA and protein stability [16,17,18]. Myc plays a pivotal role in glioblastomagenesis. At the transcriptional level, it is responsible for EGFR upregulation [19] and for the expression or transcriptional repression of a variety of miRs involved in chemoresistance [20,21]. Some Myc tumorigenic effects are exerted through the molecular partners belonging to its protein network [22,23]. More importantly, Myc represents one of the master genes governing the stem cell properties of GSCs [24]. Inhibiting Myc results in a profound resetting of GSC transcriptional nodes, which parallels changes in their biological properties [25].
In addition to Myc’s known response to growth factor stimulation [26], nutrient sensing and hypoxia may also regulate this potent oncogene in healthy cells, both transcriptionally and post-transcriptionally, usually repressing either Myc expression or activity when nutrient and oxygen levels are low [27]. However, this tight, metabolic-dependent regulation of Myc may be bypassed when Myc is overexpressed. Indeed, despite healthy and cancer cells sharing Myc metabolic target genes, upon Myc overexpression, they are deregulated and this leads, at least in part, to cancer cell metabolic rewiring.
Here, based on the available literature to date, we provide a comprehensive view of Myc-dependent control of GBM gene expression profiles and metabolism, highlighting Myc’s role as pivotal regulator of GBM metabolic genes, as well as a target of GBM cell metabolic pathways, in a bidirectional fashion that places Myc at the intersection between transcriptional output and metabolism in glioblastoma.

2. Glioblastoma Multiforme

GBM is a grade-IV astrocytoma and, similar to most tumors, is a genetic disease of a single cell. Histopathological analyses identified both the disease stage and the morphological characteristics of the tumor [28]. In the past few years, immunohistochemical analyses flanked, integrated, and supported GBM molecular classification [29], exploiting advanced technologies such as deep convolutional neural networks [30]. GBM is the most malignant brain tumor affecting people of >50 years of age [31], although pediatric GBM may also occur in children. In this latter case, the genetic background is highly different from that found in adults [32]. GBM is defined as rare; however, it represents 14.3% of all central nervous system (CNS) cancers and 49.1% of all malignant CNS tumors [31] in the US and may be classified as “primary” or “secondary”, depending on its “de novo” onset or development from a lower-grade (usually II or III) astrocytoma. Given the similarity of the transcript landscapes among human GBM, radial glial cells, and intermediate progenitors, it is now believed that GBM onset may be due to the reactivation of normal developmental programs [33]. A lack of effective therapies after maximal surgical resection makes this brain neoplasm a never-ending challenge for neuro-oncologists.

2.1. GBM Classification

Different GBM subtypes have been identified to date. The term “multiforme” clearly elucidates the morphological heterogeneity of the tumor [34], whose subtypes were first identified as variants by the World Health Organization (WHO) by histopathological characteristics as follows: conventional GBM, giant GBM cell, and gliosarcoma (GS). A number of further histotypes have been identified, as reviewed in Miller CR. 2007 [28].

2.1.1. Genetic Classification

An in-depth characterization in the early 2000s of primary and secondary GBM revealed that these tumors did not differ in their genetic background but, rather, in the frequency of genetic alterations, which was higher in primary GBM with respect to secondary GBM [35]. The major genetic abnormalities found were a loss of heterozygosity (LOH) of chromosome 10q, TP53 mutation, p16INK4a deletion, EGFR amplification, and PTEN mutation. About 10 years ago, the identification of point mutations in the isocitrate dehydrogenase (IDH) 1 (cytosolic) and 2 (mitochondrial) genes [36], provided a further level of classification and, importantly, helped to discriminate between primary (IDH1 wild-type, IDH1wt) and secondary (IDH1 mutated, IDH1mut) GBM. IDH2 mutation at R172 (R172K) residue is found only in 3% of WHO grade-II and -III astrocytomas, oligodendrogliomas, and secondary glioblastomas, whereas IDH1 mutation at R132 (R132H) affects 70% of these cancers [36]. Only 5% of primary GBM shows IDH1 mutation [37]. Highly aggressive IDHwt GBM typically shows EGFR amplification, which has to be considered as a consequence of copy number gains on chromosome 7q. Furthermore, half of EGFR-amplified GBM instances carry a truncated, constitutively active form of the receptor (EGFRvIII), resulting from the deletion of exons 2–7 and encoding a large part of the extracellular domain [38]. Other frequent chromosomal and genetic alterations detected in primary GBM are monosomy of chromosome 10, cyclin-dependent kinase inhibitor (CDKN) 2A and 2B deletion, and telomerase reverse transcriptase (TERT) mutation [39]. A subset of tumors also presents mutations in cyclin-dependent kinases (CDKs) 4 and 6, murine minute doubles (MDMs) 2 and 4, and hepatocyte growth factor receptor (MET) genes [37,40]. Secondary IDHmut GBM also carries α-thalassemia/mental retardation syndrome X-linked (ATRX) mutation [32], together with the mutations mentioned above [35].

2.1.2. Molecular Classification

The coming of high-throughput technologies in the last decades has further detailed the molecular features of GBM subtypes from transcriptional and epigenetic points of view. The term “molecular classification” covers different methodologies used to characterize GBM subtypes, from the analysis of transcriptional and transcriptomic signatures to epigenetic abnormality detection (including genomic methylation profiles) and sensitivity to BUB1 inhibition.

Transcriptional/Transcriptomic Classification

One of the first transcriptional classifications of GBM was provided by a microarray analysis of 21 GBM in the context of a larger study also including 45 astrocytic cancers and 19 pilocytic astrocytomas. Among the 6800 genes analyzed, 360 were found to discriminate between GBM and astrocytomas, with some of them (ZYX, SDC1, FLN1, FOXM1, and FOXGB1) previously unidentified as typical GBM genes and related to cell proliferation and migration [41]. A further microarray analysis of 76 grade-III astrocytomas and GBM established distinct, subtype-specific, molecular signatures associated to key stages of neurogenesis, giving emphasis to the presence of stem-like cells as extremely unfavorable from a prognostic point of view [42]. Different probe sets were strongly expressed by three tumor subtypes, and 35 genes for each subtype were identified as robust subtype determinants and used for clustering. From this analysis, three subtypes were identified: proneural (PN), proliferative (Prolif), and mesenchymal (Mes). These reflect both the molecular and the biological features of tumors, with the PN subtype consisting mostly of more differentiated cells and the Prolif and Mes subtypes characterized by genes involved in cell cycle progression control and angiogenesis, respectively. The average expression levels of these 35 genes, termed centroids, were used to evaluate other tumor samples in different databases, strengthening the prognostic value of this classification. Genes depicting Prolif and Mes subtype features were associated to poor prognosis, which was also connected to chromosome 10 LOH, gains in chromosome 7, and activation of the Akt and Notch pathways. Importantly, upon recurrence, a shift from the PN to the Mes subtype is observed with the typical occurrence of YKL-40 gene expression. This gene encodes for an extracellular matrix glycoprotein secreted by different solid tumors and is the most-expressed gene in GBM with respect to healthy brains [43,44,45,46].
The Cancer Genome Atlas network in 2008 provided a comprehensive, multidimensional analysis of copy number, gene expression, and DNA methylation alterations in 206 GBM samples [40]. Together with known genetic aberrations (LOH on chromosome 10, copy number acquisition on chromosome 7, and TP53 mutations), a number of new genomic abnormalities were found. Among them, NF1 inactivation, resulting either from NF1 gene deletions or point mutations, and mutations in proteins belonging to the PI3K complex were detected. Interestingly, a set of hypermutated tumors was uncovered. These GBM, treated with TMZ or lomustine alone or in combination, presented mutations in genes belonging to the mismatch repair (MMR) pathway. This hypermutator phenotype was also found associated to the promoter methylation of O [6]-methylguanine-DNA methyltransferase (MGMT), which removes alkyl residues from guanine residues in the DNA [47] and predicts GBM response to TMZ [48,49]. Starting from this analysis, Verhaak et al. [5] identified three major GBM subtypes that still represent a reference point for GBM classification at the genomic level: classical, mesenchymal, and proneural. A fourth subtype, called neural, was also described. Each subtype was defined by a set of typical genomic alterations, including single-gene point mutations, and each subtype resembled different neural cell types. Thus, the classical subtype was reminiscent of astrocytes, whereas the proneural subtype was associated with oligodendrocyte signatures. The mesenchymal subtype resembled cultured astroglia, while the neural subtype showed signatures of either astrocytic, oligodendrocytic, or neural differentiation. To simplify, we may say that classical GBM is characterized by EGFR amplification; deletion of the CDKN2A locus; and overexpression of the NES (a neural precursor and stem cell marker), Notch (NOTCH3, JAG1, and LFNG), and Sonic hedgehog (SMO, GAS1, and GLI2) signaling pathways. The mesenchymal subtype presents deletions in NF1, as well as overexpressions of YKL-40, MET, TRADD, RELB, and TNFRSF1A belonging to the NK-kB network, probably depending on the infiltrative, inflammatory, and necrotic features of this subtype. The proneural class is invariably identified by alteration in the PDGFRA gene and point mutations in IDH1. Furthermore, PDGFRA, NKX2-2, and OLIG2—all genes typical of oligodendrocyte development—are highly expressed, making proneural GBM an atypical tumor. The neural subtype is characterized by the expression of neuron-specific genes. However, this subtype is now supposed to derive from contaminated healthy brain cells, as resulted from subsequent analyses [9].
Single-cell RNA sequencing of 596 GBM cells identified after two levels of filtering (for details see Wang Q, Cancer Cell, 2017 [9]) bona fide glioma genes (BFGs). Among them, ~7000 genes matched with the Affymetrix U133A array already used to profile the TCGA cohort [5,40]. This filtered set was used to cluster 369 IDHwt GBM, and the proneural, classical, and mesenchymal subtypes emerged [9]. Further, mesenchymal GBM were transcriptionally characterized by tumor-associated glia and microglia, supporting the hypothesis that GBM subtypes may be shaped by the immune microenvironment. Specularly, NF1 inactivation was shown to attract microglia, highlighting the bidirectional influence between the tumor and the immune TME.

Epigenetic Classification

The term epigenetics refers to heritable—and reversible—changes in the expression of a gene not dependent on alterations in the corresponding DNA sequence. Epigenetic mechanisms are related to the chemical and structural modification of chromatin, including both DNA methylation and histone modifications (acetylation, methylation, ubiquitylation, succinylation, isomerization, phosphorylation, and sumoylation represent some of the 500 histone modifications found to date [50]), as well as histone variants (such as H3.3 [51,52]), RNA-based mechanisms [53], and loops of structurally organized chromatin termed topology-associated domains (TADs), which connect enhancers and gene promoters, coordinating gene expression programs in time and space [54,55]. Histone modifications are catalyzed by chromatin-remodeling enzymes and serve as docking sites for other coregulators, according to the so-called histone code [56], which rules gene activation and silencing. A recent multidimensional omics analysis revealed that genes involved in chromatin organization are mutated in gliomas, in particular in IDHmut gliomas not presenting 1p/19q codeletion [57]. Histone-coding genes have been mostly found mutated in pediatric high-grade gliomas [58,59], whereas the most common epigenetic mark in GBM is a change in the DNA methylation pattern, especially at position 5 of cytosine (5meC) in the context of CpG-rich loci. DNA methylation is invariably associated to gene silencing occurring by recruitment on 5meC residues of methyl-CpG-binding domain (MBD) proteins and, consequently, histone methyltransferases [60]. Both DNA hypo- and hypermethylation may characterize cancer cells. The first is typical of genomic loci that comprise intergenic regions, repetitive DNA sequences, and gene bodies, including oncogenes. The second usually occurs in tumor suppressor genes or the negative regulators of pivotal pathways leading to carcinogenesis. In GBM, for example, DNA hypermethylation is found in genes that negatively control the WNT, Frizzled, and Ras pathways [61,62,63].
IDH status profoundly affects glioma genome hypermethylation and is responsible for the glioma CpG island methylator phenotype (G-CIMP) [64]. Although IDH1 R132H mutation affects only 5% of primary GBM [37], IDH1mut proteins represent a striking example of how metabolism and epigenetics are interconnected. IDH is an enzyme of the tricaboxylic acid (TCA) cycle and converts isocitrate into α-ketoglutarate (or D-2-ketoglutarate, D-2-KG) [65]. D-2-KG serves as cofactor for many dioxygenases, including the ten-to-eleven translocation (TET) family of proteins, which demethylate DNA. IDHmut enzymes transform D-2-KG into D-2-hydroxy-ketoglutarate (D-2-HG) [66], which competes with D-2-KG for the active sites of TETs [67], leading to their inhibition and to the acquisition of the G-CIMP hypermethylator phenotype. Furthermore, the IDHmut genotype seems to not allow the proper organization of TADs, with consequent dangerous proximity between strong enhancers and oncogenes, leading to uncontrolled growth [67].
A further level of epigenetic classification has been recently proposed. It takes into account the N6-methyladenosine (m6A) modification landscapes of specific miRNAs in low-grade gliomas (LGGs; in particular, grades II and III) and appears to be more specific in discriminating between low- and high-risk gliomas than IDH status [68]. m6A is the most common RNA modification, affecting both mRNA and noncoding RNAs, including miRNAs, tRNAs, and rRNAs [69,70]. It is catalyzed by writers (such as METTL3/14, WTAP, RBM15/15B, and KIAA1429), recognized by readers (YTHDF1/2/3, IGF2BP1, and HNRNPA2B1), and deleted by erasers (FTO and ALKBH5) [71,72,73,74,75,76]. m6A impacts many aspects of RNA biology (from transcription to processing and from translation to degradation), as well as cell and organ functions. Indeed, it is involved in stress response, metabolism, infectious and metabolic diseases, neural system development, and even cancer [77]. Six m6A-related miRNAs have been identified to have individual prognostic value and to be able to discern between low- and high-risk LGGs. IDHmut and IDHwt LGGs were variably classified as low- or high-risk LGGs using these m6A-related miRNAs as a risk model [68].

Classification Based on BUB1B Inhibition Sensitivity

The BUB1B gene encodes for the Bub1-like pseudokinase, BubR1, involved in the control of the mitotic checkpoint, timing, and kinetochore–microtubule attachment [78]. BUB1B was revealed as essential for GSC expansion [79] and GBM displaying sensitivity to BUB1B inhibition (BUB1BS) had a worse prognosis when compared to tumors resistant to BUB1B inhibition (BUB1BR), regardless of molecular subtype. Molecular subnetworks belonging to the BUB1BS state comprise pathways related to cell cycle regulation, microtubule organization, and chromosome segregation. Furthermore, genes in these subnetworks are overexpressed in BUB1BS GBM, explaining the occurrence of mitotic catastrophe in BUB1BS GSCs upon BUB1B inhibition [80]. This classification is of particular importance when considering the great heterogeneity within a single GBM and the shift from one molecular subtype to another during tumor relapse.

2.2. Glioblastoma Stem Cells (GSCs)

Cancer stem cells (CSCs)—also termed tumor-initiating cells (TICs)—were first described in hematologic malignancies [81] but were found immediately after also in solid cancers [82], including GBM [83]. These cells represent a minor population within the tumor mass. Nevertheless, GSCs are responsible, in the vast majority of cases, for tumor recurrence and resistance to therapies. GSCs recapitulate most of the molecular characteristics of pluripotent stem cells, including the activation of stem factors such as Oct4, KLF4, Nanog, and SOX2, which constitute a self-fueled transcriptional circuit [84,85,86], and the Wnt/β-catenin, Sonic Hedgehog (Shh), and Notch pathways [87,88,89], granting self-renewal. They also possess the capacity to undergo asymmetric cell division, giving rise to both proliferating, self-renewing stem cells and to differentiated cells with low tumorigenic potential, which enrich the bulk of tumors. Nevertheless, this process is not so efficient due to the small amount of CSCs harboring this ability [90,91,92]. Given the infiltrative nature of GBM, surgical eradication of GSCs is always incomplete; further, a number of efficient mechanisms, including DNA repair systems, as well as the Notch, NF-κB, EZH2, and PARP pathways [89,93,94,95], are engaged by these cells when exposed to radiation and chemotherapeutic agents, escaping apoptosis [96]. By definition, GSCs present self-renewal and multipotency properties owned by neural stem cells [97] and recapitulated tumor heterogeneity when implanted in immunocompromised mice [98]. They may derive from neural stem or progenitor cells or from a dedifferentiation process occurring in glial cells [99], with this latter mechanism still an object of debate. Further, they resemble the tumor subtype of origin in terms of molecular features, harbored genetic lesions, and biological behavior [5,100]. How many subtypes of GSCs exist is still controversial. Three GBM subtypes have been described (see Transcriptional/Transcriptomic Classification), but some papers have reported that mainly the mesenchymal and proneural subtypes of GSCs may be detected within tumors [101,102,103]. Furthermore, although the neural subtype has been referred as a simple contamination from healthy surrounding brain tissue [9], a very recent work demonstrated that this specific GSC subtype may be present at the outer invading border of the tumor as a result of the activation of a neuronal developmental program elicited by signals originated from neighboring neurons [104].
Furthermore, GSCs have been considered responsible for GBM intratumoral genomic heterogeneity [105,106]. Indeed, genomic heterogeneity and evolution have been observed in both GBM patients and mouse xenografts [5,37,107,108]. Therefore, it is conceivable that these phenomena could depend, at least in part, on genomic variations in GBM cells of origin, leading to the acquisition of fitness advantages. Nevertheless, it was demonstrated that GSC heterogeneity acquired through serial transplantation in mice of homogenous populations did not depend on heritable epigenetic or genetic changes but rather on growth dynamics and cell fate decisions [109]. When looking at TMZ resistance, however, epigenetic variations have been found to distinguish between resistant vs. nonresistant clones. In this case, changes in histone H3.3 level have been found, corresponding to downregulation of the encoding gene [109,110].

2.2.1. GSCs and the Tumor Microenvironment

The TME is particularly important for tumors, as it provides nutrients. GBM evolution is strictly dependent on the interaction with the TME, especially TME immune components. It has been revealed that distinct GSC subtypes evoke transcriptional signatures typical of different immune cells. Indeed, by single-cell sequencing, the mesenchymal subtype was found associated with transcript outputs belonging to M1 and M2 macrophages and neutrophils, whilst natural killer (NK) transcripts were almost absent. The proneural subtype was depleted of memory CD4+ T lymphocytes, whereas the classical subtype (according to molecular classification in the TCGA network, Nature 2008 and in Verhaak, R.G., Cell, 2009 [5,40]) was enriched in dendritic gene signatures [9]. Importantly, the immune TME is subjected to GSC-subtype-specific variations upon recurrence. Thus, recurrent mesenchymal GBM showed decreased levels of associated monocytes but increased M2 macrophages. Classical and proneural GBM were characterized by a global decrease in immune cells defined by their specific transcriptomic landscapes, but when shifting to the mesenchymal subtype, an increase in the TME immune fraction was found. Notably, when hypermutation was sustained by TMZ treatment, enrichment in CD8+ T lymphocytes was detected [9].
Interaction with TME nonimmune components may activate surprisingly normal transcriptional programs in GSCs. This phenomenon is particularly evident at tumor relapse. Indeed, in IDHwt GBM where a stem-like state transcriptional signature is found, GSCs located at the outer invasive border present a typical neuronal gene expression pattern, which lacks GSCs at the core of the tumor. In fact, the neuronal marker NeuN, the stem factors SOX2 and SNAP25, a genes expressed by GSCs upon recurrence are present in the infiltrative region, whereas SNAP25 is absent and few neurons are present in the tumor core. This suggests that the neuronal-like phenotype, characterizing GSCs at the tumor infiltrative border, is the result of signals that hit GSCs from the surrounding neurons in the invaded brain. Conversely, IDHmut tumors—characterized by a differentiated-like state signature—differentially express genes belonging to cell cycle and mitosis categories [104].
These studies clearly show that GBM—and the GSCs within—and the immune microenvironment co-evolve. Indeed, by sequencing 61,062 single cells from eight IDHwt GBM cases, a natural evolution signature (NES) has been identified. This signature is a module of 12 genes and is associated to the activation of brain developmental genes, such as MYBL2 and FOSL2. The higher the presence of the NES in a tumor, the older the lesion. Furthermore, the NES is promoted by hypoxia and associated with the infiltration of bone-marrow-derived macrophages in the pseudopalisading cell region (PC) and microvascular proliferating region (MP). The infiltration of bone-marrow-derived macrophages matches with the expression of a specific gene expression signature. Among the genes differentially expressed, ANXA1 has been revealed to have a pivotal role in recruiting monocytes at tumor sites and inducing their differentiation into M2-like macrophages, which are typically immunosuppressive [111,112]. This resulted in a low proliferative rate and low INFγ production of CD8+ lymphocytes co-cultured in vitro with ANXA1+ tumor M2 macrophages, as well as in increased tumor mass in orthotopic mouse xenografts [113].

2.2.2. GSC Metabolism

GSC metabolism is highly divergent with respect to bulk tumors (see Section 3 below). Indeed, it has been reported that they are characterized by lower glycolysis and maximal respiratory capacity. Moreover, GSCs present globally lower levels of metabolites (such as GTP and ATP), nonessential amino acids, fructose 6-phosphate, and fructose 1,6-bisphosphate. However, metabolites belonging to the pentose phosphate and galactose metabolism, glutaminolysis, and the citric acid cycle are present at high levels in GSCs [114]. Notably, the metabolic profile reflects GSC subtype identity. Low mobile lipids and high glutamine consumption define the GSC restricted-stem (GSr) phenotype [115,116], whereas the GSC full-stem (GSf) phenotype presents the opposite [101]. These typical phenotypes correspond to the proneural and the mesenchymal subtypes, respectively, according to the molecular classification of Verhaak et al., Cell, 2009 [5]. Overall, two metabolic classes of GSCs may be identified: one dependent on glucose and aerobic glycolysis to support proliferation and the second sustaining OXPHOS through glutamine and activating glycolysis only when glutamine is limited [117]. Importantly, GSC metabolism reflects resistance to therapies and may shift upon treatment. Active lipid and glutamine metabolism, low glucose consumption, and OXPHOS have been found correlated to resistance to radio- and chemotherapies [118,119,120]. When irradiated, GSCs still showed their quiescent metabolic state, with little change in adenylate energy charge, glutamine/glutamate ratio, and amino acid profile [114].

3. GBM Metabolism

GBM presents an altered metabolic state at multiple levels. Glucose, lipid, and fatty acid metabolic pathways exhibit relevant alterations in GBM, which depend on different cues. Even cell location in a tumor mass may dictate metabolic rewiring [121], thus making GBM a heterogenous tumor also from a metabolic point of view. Moreover, GBM metabolism has to be considered as the result of a complex networking of different cellular components within the CNS [122,123]. This is particularly important for IDHmut GBM, which, deriving from a lower-grade neoplasia, is considered a metabolic disease.
Activation of distinct metabolic routes also depend on the GBM genomic and epigenomic landscape, as specific mutations are associated with typical metabolic pathways, whilst perturbations in chromatin-remodeling enzyme activity may impact metabolic changes and vice versa. These specific topics have been extensively reviewed elsewhere [124,125,126].

3.1. Glucose Metabolism

The brain almost entirely relies on glucose uptake as the main source of energy. However, glucose influx into the brain is hampered by the presence of the blood–brain barrier. To counteract limited glucose availability, neurons express GLUT3 (SLC2A3) glucose transporter, which has a five-fold higher affinity for glucose. Tumors, including GBM, mainly transport glucose through GLUT1. However, recently it was demonstrated that GSCs preferentially express GLUT3, which supports GSC tumorigenic properties [127]. As with other cancer types, GBM uses aerobic glycolysis to fulfill energy demand and is characterized by a high glucose uptake, which undergoes cytosolic fermentation resulting in an overproduction of lactate released in the intercellular space [128], stimulating angiogenesis and impairing immune surveillance [129,130]. High glycolytic rate—occurring especially in the tumor core—has been associated with dismal prognosis and poor survival of GBM patients [11,131]. However, glucose consumption through glycolysis is not a “dogma” and, depending on a multiplicity of factors (interaction with the TME, cell cycle phase, and oxygen levels), GBM metabolism may switch to OXPHOS, typically at the tumor outer layer [10]. Interactions with the immune components of a tumor mass play a pivotal role in GBM metabolic adaptation [132], pushing GBM through glycolysis. Among the molecular pathways that contribute to enhance glycolysis in GBM, there is the upregulation of human ribonucleoprotein (hRNP) A1, which promotes the alternative splicing of Myc-associated factor X (Max), generating Delta-Max, which is typical of EGFRvIII-mutated GBM. Delta-Max enhances Myc-dependent transformation and supports glycolytic gene expression. Indeed, expression of Delta-Max counterbalances EGFRvIII loss. It also induces EGFRvIII expression, originating a transcriptional positive feedback loop whose intermediate is hRNPA1, and fuels GBM cell proliferation [133]. Overexpression of isoform 2 of hexokinase (HK2), the first enzyme of the glycolytic pathway, characterizes GBM-carrying PTEN mutations, as well. HK2 is not an isoform typically expressed by the brain, which is characterized by HK1. Inhibition of HK2 restrains GBM proliferation, reversing the Warburg effect. Sensitization to radiation and TMZ also occur upon HK2 depletion [134]. Furthermore, the dissociation of HK2 from mitochondria in the presence of high glucose leads to the activation of the NK-kB pathway through IkBα binding, phosphorylation, and degradation by μ–calpain. Activated NF-kB induces the expression of PD-L1, connecting GBM metabolism with immune evasion [135]. Blockage of cytochrome oxidase 2 synthesis, which inhibits glycolysis, and upregulation of glucose transporters (SLC2A1 and A4) [136,137] have been found in p53-mutated GBM. The TCA cycle and downstream glycolysis also present large perturbations in GBM. In IDHwt tumors, IDH is overexpressed compared to healthy brain tissue, and this leads to the production of high levels of D-2-KG using NADP+ or NAD+ as cofactors. D-2-KG, in turn, serves as a cofactor for the ten-to-eleven translocation (TET) family of demethylating enzymes [67]. As a consequence, in IDHmut secondary GBM, a novel metabolite has been found called D-2-HG, whose detection requires specific methods that are able to distinguish between the “D” and the “L” isomers, which differ by an asymmetric carbon atom [138]. Only D-2-HG may compete with D-2-KG for TET binding, inhibiting its activity and leading to the “hypermethylator phenotype” associated with a better prognosis [64]. These findings are a clear demonstration of how metabolic alterations directly reflect the epigenome.

3.2. Glutamine Addiction

When glucose is limited, GBM cells utilize glutamine as a primary energy source. Glutamine is essential for the synthesis of final metabolic products (amino acids, proteins, purine/pyrimidines, and hexosamine), as well as cofactors, such as NAD. It is also required to refuel the TCA cycle of intermediates subtracted for the synthesis of compounds under nutrient deprivation conditions, activating catabolism through a series of reactions not requiring Acetyl-CoA. Glutamine also activates the mechanistic target of rapamycin (mTOR) and controls cellular pH [139,140]. Glutamine levels are very high in GBM [141], and this is due to the overexpression of glutamine synthetase (GS), which produces glutamine from glutamate and ammonia [142]. Conversely, glutaminase (GLS) hydrolyzes glutamine into glutamate and ammonia. This is the first step of glutaminolysis, leading to the production of D-2-KG from glutamate via glutamate dehydrogenase (GDH). It has to be mentioned that pyruvate carboxylation may be also used to resupply TCA cycle intermediates, specifically citrate resulting from the condensation of Acetyl-CoA and oxaloacetate [143]. However, this reaction is downregulated in most cancers [144], and glutaminolysis remains the major source of D-2-KG to push the TCA cycle. Glutamine may also be transformed via glutamate into proline, which has gained attention in recent years as a modulator of cancer biology [145]. Proline levels were found elevated in GBM because of its reduced catabolism due to decreased expression of proline hydroxylase (POX) [146]. In addition to D-2-KG, glutathione (GSH) and lactate are the end-products of glutamate. GSH is of major importance because, as a potent antioxidant, it is one of the main responsible factors for GBM resistance to therapies. Indeed, GSH depletion has been explored as a putative strategy for GBM treatment [147,148]. Glutamine-derived lactate essentially recapitulates the Warburg effect through the oxidative decarboxylation of malate to pyruvate CO2 and NADPH, which enter glycolysis via glyceraldehyde triphosphate dehydrogenase (GAPDH) [149].

3.3. Lipid Metabolism

The brain is the second organ, after adipose tissue, to rely on lipids to build its structure, extract energy, and maintain cell growth [150]. The lipidomic profiles of GBM cells and tissue derived from mouse xenografts were different [151], but human gliomas present in large proportions sphingomyelin, phosphatidylinositol, and lysophosphoglycerides, as well as increases in phospholipids, cholesterol esters, arachidonic acid, and oleic and linoleic acids, when compared to a healthy brain cortex [151,152]. Lipid biosynthesis requires fatty acids that are, indeed, elevated in malignant brain tissue [153]. Acetyl-CoA synthesis is a fundamental prerequisite to produce fatty acids, and most GBM tumors present an upregulation of Acetyl-CoA synthetase 2 (ACCS2), which converts acetate into Acetyl-CoA [154,155]. Activation of Acetyl-CoA carboxylase (ACACA) via the EGFR/PI3K/Akt pathway also enhances lipogenesis in GBM [156]. FAO is a pivotal process to foster GBM growth, as inhibiting fatty acid oxidation slows the proliferation rate of glioma cells both in vitro and in vivo [12,157]. Importantly, lipid metabolism is linked to a newly discovered form of programmed cell death, ferroptosis, which depends on enhanced iron-dependent lipid peroxidation and ROS production, a route that is being explored to treat GBM [158,159].

4. Myc

Almost 40 years ago, the transforming MC29 avian virus sequence causing myelocytomatosis, later termed myc, was identified [160]. The finding of the human homolog of the viral avian myc gene [161] served as the basis for the work of entire laboratories worldwide, which spent the following decades elucidating Myc biology, profoundly marking cancer research [13]. The finding that myc gene deregulation in tumors did not occur through mutations in the coding region—as with other oncogenes, such as ras—was surprising. Indeed, myc deregulation in cancer occurs through three different mechanisms: gene amplification [162], chromosomal translocation [163], and insertional mutagenesis [164]. Three myc paralogs have been identified, including c-myc (hereafter termed myc), n-myc [165], and l-myc [166], which are invariably deregulated in a wide variety of cancers in a tissue-specific manner, including GBM.

4.1. Brief Overview on Myc Structure, Post-Transcriptional Regulation, and Control of Gene Expression

Myc in all its three isoforms is a basic helix-loop-helix-leucine zipper (bHLH-LZ) transcription factor, an object of extensive review elsewhere [13,167]. It binds almost 15% of genomic loci and is deregulated in 70% of cancers. Here, we provide a snapshot of its structure and regulation and how it works in the control of gene expression, both directly and indirectly.

4.1.1. Myc Structure

Myc is a 439 amino acid protein. The encoding gene is located on the long arm of chromosome 8 (8q24) and contains three exons, which encode a translational product of 64KD. The two highly conserved Myc polypeptides consist of a N-terminal transactivating domain (TAD) and a C-terminal DNA-binding region. The TAD is constituted by three conserved Myc boxes (MB0, MBI, and MBII) [168,169]. MBI and MBII are required for transcriptional activation. In particular, MBI recruits P-TEFb, a cyclin CDK-complex promoting transcriptional elongation by RNA polymerase II (RNAPII) phosphorylation [170]. MBII is essential for Myc targeting of transcriptional activation and repression. MBII also plays a role in promoting cellular transformation and tumorigenesis; it also controls Myc turnover. Myc structure also presents a middle proline, glutamic acid, serine, and threonine (PEST)-rich region adjacent to the MBIII and MBIV conserved boxes and two nuclear localization sequences. The MBIII box, by recruiting histone deacetylase 3 (HDAC3), is engaged in transcriptional repression, whereas Myc pro-apoptotic function seems to rely on MBIV. The 100-amino-acid carboxyterminal region is a bHLH-LZ domain required for heterodimerization with Max, another small bHLH-LZ protein [171]. This interaction forms a stable Myc/Max heterodimer that directly binds specific DNA sequences named “enhancer boxes” (E-boxes) to stimulate transcription. Indeed, Myc homodimers are highly unstable [172].

4.1.2. Myc Protein Network

Myc-driven proliferation, transformation, and apoptosis processes require Myc/Max heterodimerization. Max is a stable, constitutively expressed transcription factor (TF) with a short half-life. This suggests that Max activity is highly dependent on the amount of Max-associated TFs. Max homodimers, by competition for E-box binding, may impair Myc biological activity. Other Max interactors are Mad1–4 (Mad1, Mxi1, Mad3, and Mad4), Mnt, and Mga, also provided with a bHLH-LZ domain. These proteins share some Myc properties: (a) scarce homodimerization and DNA-binding ability; (b) robust heterodimerization with Max and consequent E-box binding; and (c) interference with Myc/Max heterodimer activity. The family of Mad1, Mxi1, and their related members of transcriptional repressors competes for Myc/Max binding sites. The competition of Myc and Mad for Max defines a cell’s choice between proliferation and transformation and differentiation and quiescence [173], as increased expressions of Mads determine cellular differentiation and growth arrest. Indeed, Mad1–4 proteins possess a repression motif interacting with Sin3a and Sin3b corepressors, which in turn engage HDACs and other chromatin remodelers. In conclusion, the opposite functions of Myc and Mad are exerted at multiple levels: (i) competition with Max to form heterodimers; (ii) competition for E-box binding; and (iii) transcriptional activation and repression of target genes [174]. Interestingly, among Max partners, Mnt is a unique Myc antagonist. Its expression overlaps with the expression of Myc, but it is constitutively expressed [175,176]. Furthermore, Myc can also repress the transcription of specific subsets of genes competing with p300 for binding to the transcriptional activator Miz1 [177].

4.1.3. Myc Post-Translational Regulation

Myc undergoes extensive post-translational regulation, which is required to both switch on and off its activity. MBI, -II, and -III in its N-terminal TAD domain are involved in protein stabilization and transcriptional activity [178]. A series of phosphorylation and dephosphorylation events occurs at MBI, where serine 62 (S62) is phosphorylated by RAS/MEK/ERK/CDK2 [179], which precedes and induces the phosphorylation of threonine 58 (T58) by GSK3β [18]. Recently, S67 has been found to be the target of Aurora B kinase. This event blocks T58 phosphorylation by GSK3β, leading to enhanced protein stability [180]. Indeed, following T58 phosphorylation, S62 is dephosphorylated by Pin1/PP2A, and this promotes Myc degradation and turnover through the SCFFBXW7/Ub-proteasome pathway [18]. In addition to the SCFFBXW7/Ub-proteasome pathway, Myc may also be degraded by the SCF-SKP2 ubiquitin-ligase complex [181]. SKP2–Myc interaction is intriguing, as SKP2 and its related components of the SCF complex are first recruited at Myc target promoters and, after Myc binding and ubiquitination, transcriptional activation of Myc occurs before its degradation through the proteasome, finely tuning Myc transcription in a very narrow window [182]. Notably, a third ubiquitin ligase, named HECTH9, was found to promote Myc transcription rather than its degradation by polyubiquitination and polymerization of lysine (K) 63, enhancing K acetyltransferase (KAT) recruitment [183].
Myc ubiquitination is interconnected with its acetylation by KATs, such as p300, mammalian GCN5 (mGCN5), and Tip60 [184,185]. Myc holds many Ks that may be acetylated, and at least one of them, K323, is located in the NLS [186]. However, acetylation does not alter Myc localization or Max binding; rather, p300-dependent acetylation of Myc has been revealed to promote a more efficient recruitment of Myc-interacting zinc finger protein (Miz1) to Myc/Max heterodimers.
Most recently, it has been demonstrated that Myc may also be methylated on arginine (R) residues by the protein arginine methyltransferase (PRMT) family of enzymes. Class-I PRMT monomethylates or asymmetrically dimethylates R residues on target histone and nonhistone proteins. Class-II PRMT monomethylates and symmetrically dimethylates R residues on targets [187]. In particular, both asymmetrical and symmetrical Myc dimethylation occurs, having opposite roles on Myc properties. In M2 macrophages, Myc asymmetrical dimethylation by PRMT1 leads to the efficient recruitment of p300, enhancing transcription [188]. In GSCs, both asymmetrical and symmetrical Myc dimethylation has been detected. These modifications differently impact Myc stability in GSCs and rule differentiation and stemness, respectively [189]. Figure 1 summarizes Myc structure, post-translational modifications, and interactions with protein partners.

4.1.4. Control of Gene Expression by Myc

The concept of Myc as a “classical” transcription factor that binds to DNA and promotes transcription through the recruitment of KATs—which untangle compacted chromatin, acetylating histones—and components of the basal transcriptional machinery has consistently changed over the last decade. Indeed, Myc unusually controls all three RNA polymerases (RNAPs), thus controlling the synthesis not only of encoding and noncoding RNAs, but also of tRNAs and rRNA, which are pivotal to protein biosynthesis upregulation in tumor cells. Of particular note are the recent findings of a tight control of RNAPII by Myc. In fact, it has been demonstrated that Myc regulates RNAPII pause, release, and elongation [170,190] and also maintains the fidelity of splicing in cancer cells [191]. Furthermore, Myc was postulated as a global transcriptional amplifier of already active gene expression programs [192], especially depending on the amount of molecules within a given cell (Figure 2). Indeed, in cancer cells, high Myc levels drive transcriptional amplification, potentially invading chromatin regulatory elements (promoters, enhancers, and superenhancers) with both high and low affinity for Myc binding [193]. The degree of Myc occupancy determines the expression level of each active gene: high-affinity promoters are already saturated by Myc in proliferating cells, and a further increase in Myc molecules only enhances its occupancy at low-affinity binding sites [194].
This behavior of Myc is consistent with the evidence that open chromatin at active promoters is important for Myc binding [195] and that enhancer loops in the proximity of core promoters at active genes may facilitate Myc recruitment to close enhancer elements once binding sites in core promoters are saturated. It has to be noted that Myc-bound and Myc-regulated genes do not always overlap. An explanation of this phenomenon resides in subsets of microRNAs regulated by Myc and clustered according to the specific role their targets play in Myc-controlled cellular processes [196].
Myc also regulates mRNA cap methylation, an essential step of mRNA translation. In fact, Myc recruits TFIIH kinase, which phosphorylates RNAPII that, in turn, recruits and activates cap RNA methyltransferase (RNMT) [197]. Finally, although generally considered a transcriptional activator, Myc was also found to repress transcription in complex with Miz1 in a ratio near to 1 at Myc-repressed promoters [194].

4.1.5. Myc-Dependent Regulation of Apoptosis

In healthy cells, a high concentration of growth factors is responsible for cell proliferation in response to sustained Myc levels; when growth factors are limited, Myc primes cells to undergo apoptosis [198,199]. Conversely, deregulated Myc expression is responsible for resistance to apoptosis in transformed cells [200,201,202]. Although the pro-apoptotic function of Myc is not precisely defined at the mechanistic level, multiple pathways seem to be involved, and it is not clear whether Myc constitutively modulates downstream effectors or whether second stimuli are needed for the full activation of each apoptotic pathway. Myc was demonstrated to activate p53 [203], partially depending on p19ARF induction and activity and, consequently, MDM2 and p21 activation [200]. However, Myc was also found to repress p21, overcoming the G1-S cell cycle checkpoint by both directly binding its promoter in complex with Miz1 [204] and competing with Sp1/Sp3 transcription factors [205]. Myc may also alter the balance between anti-apoptotic and pro-apoptotic pathways in favor of the latter, when required. For example, Myc acts as a suppressor of Bcl2 and Bcl-XL [206,207], promoting apoptosis through Bax [208,209] and influencing cytochrome c release from mitochondria [210] with the consequent activation of caspases [211].
Further, under certain conditions, Myc may promote cellular senescence, a phenomenon blunting cancer development at the premalignant stage [212,213].

4.2. Myc and the Control of Gene Expression in Glioblastoma

As in other cancer types, Myc is usually overexpressed in GBM. Indeed, genes pivotal to glioblastoma progression and resistance to therapies are aberrantly regulated by Myc in GBM. Myc binds to the promoter of EGFR, which, in turn, regulates Myc in a positive feedback loop. An overexpressed EGFR/myc axis results in the activation of the TGF-β, Notch, and Hippo pathways through the downregulation of mir524 [19]. Furthermore, EGFRvIII induces glioma angiogenesis by Myc-dependent transcriptional activation of angiopoietin-4 (ANG4) (see Crosstalk between Myc-Dependent Angiogenesis and Metabolism in GBM below) [214]. At the epigenetic level, lysine demethylase 4 (KDM4) binds to the Myc promoter, inducing its expression [215]. The same finding was obtained for histone methyltransferase G9a, which was revealed to induce Myc gene expression and the proliferation and invasiveness of GBM cells [216].
Far-upstream element (FUSE)-binding protein 1 (FBP1), RANBP10, and Trip13 all regulate Myc in GBM. FBP1 directly binds to the Myc promoter [23], whereas RANBP10 and Trip13 both increase Myc stability by downregulating FBXW7 [22,217]. SWD3, a WD repeat domain 5 (WDR5) protein, also regulates MYC activity. The Myc/SWD3 complex is recruited to the PRMT4 (CARM1) promoter, and PRMT4 expression enhances glioblastoma proliferation [218]. Interestingly, in GBM Myc is regulated by the lncRNA HOXC13, which, acting as a sponge for mir122-5p, activates SATB1 and Myc. Myc, in turn, is able to bind to the HOXC13 promoter, generating a positive feedback loop sustaining GBM invasive properties [219].
A number of microRNAs are controlled by Myc in GBM. Interestingly, most of them are related to the acquisition of chemoresistance. Indeed, Myc negatively controls mir29c expression, an oncosuppressor whose expression is inversely correlated to tumor cell proliferation and invasion [220]. Myc-dependent mir29c downregulation is paralleled to the TMZ resistance of GBM cells [21]. Myc also enhances mir20a expression, pushing cell proliferation and chemoresistance [20]. More importantly, Myc directly binds to the MGMT promoter [221]. Myc activity also regulates GBM epithelial–mesenchymal transition through wnt/β-catenin signaling [222].

4.3. Myc and the Maintenance of GSCs

Myc exerts one of its major roles in glioblastomagenesis by controlling the proliferation and self-renewal of GSCs, being located at multiple GSC genomic loci involved in metabolism, protein synthesis, and the cell cycle. This is a feature of Myc not restricted to CSCs, as in embryonic stem (ES) cells, Myc exerts the same role, while pluripotency is governed by core transcription factors including OCT4, Nanog, and SOX2 [223]. Indeed, it has been found that, in ES cells, gene modules related to the core of pluripotency and to the Myc network are independent. The Myc module is centered on Myc–NuA4 interaction. NuA4 is a HAT complex comprising Tip60 and Ep400. In ES cells, the latter have been found associated to Myc/Max heterodimers, as well as with Dmap and Trrap, which keep chromatin in the open state, promoting massive histone acetylation. The Myc cluster includes transcription factors, such as E2F1 and 4, Rex1, Zfx, and N-Myc, which invariably are associated with hyperacetylated chromatin. Interestingly, this Myc module is found also in cancer stem cells, where it does not activate the core module [224], confirming the independency between self-renewal and proliferation potential and pluripotency. Consistently, Myc gene depletion blocks GSC proliferation and capacity to form neurospheres, whilst bulk tumor cells are less dependent on Myc expression [24]. Myc interactions with E2F3 and the chromatin regulator helicase, lymphoid-specific (HELLS) support the expression of genes critical to GSC maintenance [225].
Inhibiting Myc protein interactions and DNA-binding abilities results in a resetting of specific gene regulatory nodes in GSCs, with the highest impact on gene modules ruling stemness and neural development. This is consistently mirrored at the biological level with changes in prominent GSC features, such as enhanced proliferation and differentiation capacity [25].
Myc is regulated in GSCs in a number of ways. Very recently it was demonstrated that, in GSCs, Myc mRNA is methylated on adenosine (N6methyladenosine; m6A) and stabilized by the m6ARNA reader YTHDF2, which is itself indispensable for the expression of Myc target genes and GSC maintenance [226]. Myc-sustained expression is also regulated by Piwi1, which belongs to the family of small, RNA-binding argonaute proteins. Indeed, in the presence of Piwi1, the expression of FBXW7 is low, and this leads to a decrease in Myc degradation [227]. Consistently, deubiquitinases, such as USP13, maintain GSC-impairing proteasome-mediated Myc degradation [228]. In addition, cycline-dependent kinase 8 (CDK8) has a role in maintaining high Myc levels in GSCs, although the mechanisms still have to be elucidated [229].
The immune TME also plays an important role in regulating Myc activity. Indeed, it has been recently found that β2-microglobulin activates Myc through the PI3K/AKT/mTOR pathway, inducing the secretion of TGF-β. The latter, in turn, activates SMAD and PI3K/AKT signals in M2 macrophages, promoting the consolidation of a immunosuppressive tumor microenvironment [230].

4.4. Myc-Targeted Therapies in GBM

Being deregulated in >70% of tumor types, Myc targeting for anticancer purposes has been, and still represents, a challenge for both biologists and clinicians. Inhibiting Myc activity has to be balanced with the pivotal functions Myc exerts in healthy cells. Therefore, the complete abolishment of Myc activity through gene-silencing techniques would have detrimental side effects in the whole organism. Nevertheless, impacting Myc ability to form protein complexes with molecular partners and targeting Myc transcriptional coregulators have been explored as evaluable strategies to be translated to clinics (reviewed in Chen H, 2018 [231]; Llombart V, 2022 [232]). Clinical trials specifically related to Myc-overexpressing brain tumors are listed in Table 1 and show that limited actions are currently available to hit Myc in GBM. One of the most promising molecules affecting glioblastoma cell viability is TG02, a CDK inhibitor that is supposed to act through the CDK9-inhibition-dependent downregulation of oncoproteins, including Myc. TG02-treated cells undergo mainly apoptosis. However, blocking caspases does not result either in apoptosis impairment or in Myc expression recovery. Interestingly, TG02 activity is independent from MGMT expression, and repetitive exposure to TG02 does not induce resistance in GBM cells [233]. Inhibition of the bromodomain-containing chromatin modifier BRD4 is another exploited strategy to treat solid and hematologic malignancies characterized by high levels of Myc. This therapeutic route is based on the reciprocal control between Myc and BRD4. BRD4 destabilizes Myc by direct interaction [234]; conversely, Myc stimulates its own transcription via BRD4 recruitment and histone acetylation within its promoter [235]. JQ1 was one of the first bromodomain and extraterminal domain (BET) inhibitors used for clinical purposes and has been found effective in halting tumor growth in a variety of cancers by limiting myc expression [236,237]. However, its pharmacokinetic properties have precluded its use in clinical trials. Other BET inhibitors have been tested and have reached phase-I and -II clinical trials. INCB057643 is a novel BET inhibitor that is orally available and was already tested in phase-I clinical trials in patients with myelofibrosis and advanced malignancies [238], including GBM (Table 1; https://clinicaltrials.gov/ct2/results?cond=Glioblastoma&term=myc&cntry=&state=&city=&dist=&Search=Search; accessed on 12 January 2023). Recently, its application was extended to preclinical models of pancreatic cancer, where it was shown to modulate the immune TME [239].

5. Myc and Metabolism in Cancer

Independently of Myc expression levels, healthy and cancer cells share Myc target genes involved in many metabolic pathways, such as glycolysis, glutaminolysis, and lipid and nucleotide synthesis. In healthy cells, Myc and related metabolic pathways are tightly regulated, especially upon nutrient deprivation. In neoplastic cells, derangements at the genetic and epigenetic levels and loss of checkpoints (e.g., p53 and PTEN) that usually restrain dangerous events depending on Myc overexpression or activation allow the occurrence of metabolic Myc activities that promote cell growth.
In nontransformed mammalian cells, Myc metabolic activity is regulated by multiple factors. In optimal nutrient conditions, Myc drives metabolic pathways in order to provide cellular bricks and energy required to replicate DNA, undergo cell division, and increase cell mass. In addition growth factors, which are known to stimulate Myc [240], nutrient sensing controls Myc activity through mTOR, which modulates its translation and its stability [241,242]. Active FOXO3A inhibits Myc in a number of ways: it transactivates Myc antagonist max interactor-1 (MXI-1), which displaces Myc from Max, inhibiting Myc/Max transcriptional activation, and it also counteracts Myc-dependent transactivation of mitochondrial genes, affecting mitochondrial biogenesis [243,244]. Hypoxia and glucose deprivation are two other factors that negatively influence Myc expression through degradation or antagonist expression [245,246].
Cancer metabolism is one of the best proofs of Myc as a transcriptional amplifier. Indeed, metabolic genes are expressed at steady-state levels in almost all cells. To fulfill tumor cells’ enhanced energy demand, overexpressed Myc aberrantly amplifies genes devoted to glycolysis, glutaminolysis, and polyamine synthesis [247,248,249]. Consistently, most of the metabolic genes controlled by Myc are provided with E-boxes [250,251], and genes holding low-affinity E-boxes may be regulated by Myc as well when it is overexpressed [252]. Myc not only controls the expressions of enzymes devoted to glucose and glutamine metabolism—such as the glucose transporter GLUT1 and the glutamine transporter SLC1A5 [248,253]—but it also regulates the preferential splicing of variants [254]. Upregulation of the nutrient-sensing transcription factors MondoA and ChREBP, which control different aspects of cellular metabolism and accumulate in the nucleus depending on metabolic flux modifications, is Myc-dependent. Myc-driven metabolic reprogramming during tumor progression is modulated by these two proteins [255]. Furthermore, increases in MondoA and ChREBP protein levels lead to Mlx (Max-like protein X) sequestration, promoting the competition between Myc and Mxd (Max dimerization) proteins for Max. Therefore, an imbalance in the Myc protein network may generate metabolic alterations typical of cancer cells. Given the high levels of intermediates generated, Myc also allows shunting from one metabolic pathway to another. Indeed, the glycolytic intermediates glucose-6-phosphate and 3-phosphoglycerate may be used by the pentose phosphate pathway to generate NADPH and ribose for nucleotide biosynthesis [256,257] and serine, which is ultimately converted into NADPH in mitochondria [258].
Myc-enhanced nucleotide biosynthesis sustains the high rate of proliferation characterizing cancer cells. A number of studies have demonstrated Myc-dependent regulation of both purine and pyrimidine [259,260]. Through the pentose phosphate pathway, Myc produces ribose-5-phosphate and induces phosphoribosyl pyrophosphate synthetase 2 (PRPS2) and the synthesis of phosphoribosyl pyrophosphate, serving as scaffold for purine biosynthesis and for the pyrimidine rescue pathway [261]. Myc also supports the induction of genes of metabolic cycles indispensable for nucleotide metabolism, such as the folate cycle and one-carbon metabolism [257,262], orchestrating the control of the expressions of genes participating in different steps of nucleotide synthesis.
The uptake of essential amino acids (EEAs), both large neutral and branched-chain, which serve to build macromolecules and to activate mTOR, requires the expression of families of transporters and enzymes that are under the transcriptional control of Myc [263,264]. Myc is also involved in the alteration of tryptophan metabolism, driving its conversion into kynurenine by upregulating the transporters SLC7A5 and SLC1A5 and the arylformamidase enzyme [253]. Notably, kynurenine was related to enhanced proliferation, migration, and immune escape capacity of tumor cells [265]. Cancer cell lipid and fatty acid metabolism rewires to promote membrane biogenesis and energy storage in fasting, proliferating cells. Myc has been found to be involved in lipid metabolism, promoting both fatty acid–cholesterol synthesis and fatty acid oxidation (FAO). By upregulating enzymes of the TCA cycle, Myc induces the synthesis of citrate, the precursor of fatty acids, as well as a number of enzymes involved in fatty acid synthesis, such as ATP citrate lyase (ACLY), Acetyl-CoA carboxylase (ACACA), fatty acid synthase (FASN), and stearoyl-CoA desaturase (SCD) [266,267,268,269]. Myc also cooperates with regulators of fatty acid synthesis such as MondoA [270] and sterol response element-binding protein 1 (SREBP1) [266]. Furthermore, a positive Myc-dependent feedback loop has been hypothesized to sustain cholesterol metabolism and malignant transformation in cancer cells. Indeed, Myc was demonstrated to induce the expression of 3-hydroxy-3-methyl-glutarylcoenzyme A reductase (HMGCR) [271], which is essential for cholesterol synthesis during tumorigenesis. Conversely, HMGCR phosphorylates and activates Myc in at least some cancer models [272]. Surprisingly, Myc also participates in FAO, typically exploited by healthy cells to produce energy in mitochondria. In this context, Myc activates enzymes required for FAO [273,274], induces the expressions of receptors located both on the cell membrane and on the mitochondrial inner membrane necessary for the uptake of fatty acids undergoing oxidation in mitochondria, and altering Ca++ levels [274]. Suppression of FAO in cancer cells occurs through Myc downregulation of a series of enzymes critical for this process [269].

5.1. Metabolic Control of Myc Activity in Glioblastoma

Taking into account that the metabolic control of Myc expression in GBM and in other cancer types is largely similar [241,242,243,244,245,246], one of the best recognized and characterized metabolic pathways regulating Myc in GBM is glycolysis, which exerts its modulatory role on Myc expression especially through the mTORC2 component of the mTOR complex [275]. In cancer cells, as stated above, glycolysis and glutaminolysis may be linked through the hexosamine biosynthetic pathway, and glutamine may be converted to glucosamine-6-phosphate and glutamate starting from fructose-6-phosphate and glutamine, a reaction catalyzed by fructose-6-phosphate aminotransferase (GFAT) [276]. It was demonstrated that, in GBM, mTORC2 promotes GFAT1 activity, enhancing glucosamine-6-phosphate synthesis independently of PI3K/Akt signaling. Mechanistically, high glucose and glutamine levels promote mTORC2 activity, which supports Myc protein function as a transcriptional regulator of GFAT1, pushing glucosamine-6-phosphate synthesis [277]. Furthermore, mTORC2 inhibits the phosphorylation of class-IIa HDACs, leading to the acetylation of FOXO1 and 3, the release of Myc from a inhibitory mir34c-dependent pathway, and the activation of aerobic glycolytic genes [278]. Aurora kinase A is another regulator of Myc-dependent glycolytic gene expression in GBM. Indeed, Aurora kinase A inhibition rewires glycolytic metabolism to OXPHOS and FAO through Myc downregulation and the impaired expression of glycolytic genes [279]. Myc is also epigenetically regulated by pyruvate kinase M2 (PKM2) and phosphofructokinase 1 platelet isoform (PFKP), both catalyzing pivotal steps in the glycolytic cycle. In EGFR-mutated GBM, PKM2 translocates to the nucleus, causing the dissociation of HDAC3 from the Myc promoter through the phosphorylation of Threo11 on histone H3, inducing its transcription [280]. PFKP induces β–catenin translocation via EGFR-signaling activation and transcriptional enhancement of the downstream effectors CCND1 and Myc [281]. Both processes further promote glycolysis, glutaminolysis, and cell proliferation. Factors controlling FAO, such as peroxisome-proliferator-activated receptor α (PPRα), have also been reported to regulate Myc expression in GBM. Indeed, PPRα depletion led to a downregulation in Myc expression in GSCs [282].

5.2. Myc-Dependent Regulation of Metabolic Pathways in Glioblastoma

As in other tumors, in this malignant brain cancer Myc activity underlies the control of glycolysis and glutaminolysis. Recent studies have revealed novel partners and effectors of Myc in determining metabolic rewiring in GBM. A role for importin-α1 (also known as karyopherin α2, KNAP2) in the regulation of Myc activity and glycolysis has been found. Indeed, importin-α1 leads to E2F1 nuclear translocation and Myc transcription activation in GBM cells. This, in turn, promotes the expression of Myc-regulated enzymes involved in glucose metabolism, such as Glut1, HK2, PKM2, and PFK1 [283]. As stated above, Myc induces the expressions of the glycolytic genes Glut1, Glut3, PDK1, and HK2 in EGFRvIII+GBM in a self-sustaining loop based on the Myc-dependent upregulation of splicing factor hnRNPA1. The latter provides the production of a spliced isoform of Max called Delta-Max, which depends on the availability of glucose. By forming functional heterodimers with Myc, Delta-Max binds to the Glut1, Glut3, PDK1, and HK2 promoters, enhancing the expressions of respective enzymes. Intriguingly, while Delta-Max is able to rescue Glut1, Glut3, PDK1, and HK2 expressions upon EGFRVIII depletion, wild-type Max is not, and consistently does not support EGFRvIII GBM cell proliferation in the presence of glucose. In parallel, Delta-Max knockout inhibits the expressions of glycolytic genes and reduces the size of GBM tumors in vivo [133]. By cooperating with the aryl hydrocarbon receptor (AHR), a detoxifying cytoplasmic receptor [284,285], Myc regulates glycolysis and pyrimidine biosynthesis in GBM cells. In fact, AHR knockdown altered the levels of 26 metabolites belonging to redox equilibrium and fatty acid and nucleotide metabolism. Binding of the Myc/Max and AHR/aryl hydrocarbon receptor nuclear translocator (ARNT) heterodimers on the promoter regions of CAD, DHODH, UMPS, and LDHA has been postulated [286], and in vivo lactate labeling overlaps Myc expression in patient-derived GBM samples [287]. At the epigenetic level, Myc binds to superenhancers in GBM glycolytic genes. The disruption of superenhancers by HDAC inhibitors (HDACis) impairs glycolysis, with a shift toward OXPHOS and FAO driven by the PGC1α and PPARD genes and paralleled by a decrease in Myc activity (due to transcriptional repression), which in normal conditions blunts HDAC2, PGC1α, and PPARD expression. These metabolic alterations lead to a decrease in ATP production and a compensatory increase in oxygen consumption rate (OCR). U-13C-glucose carbon tracing shows that not only glycolytic intermediates and other associated metabolites are low in HDACi-treated GBM cells, but also PPP, ribose, and serine production, as well as the hexosamine pathway and lipid synthesis, all of them related to the Warburg effect [288].
Myc also controls the expressions of the kidney-type isoforms of glutaminase KGA and CAG encoded by the GLS gene associated with cell proliferation, whereas the liver-type isoforms GAB and LGA encoded by the GLS2 gene and related to quiescence and are not detectable in GBM. Indeed, forcing GLS2 expression in GBM cells slowed their proliferation and potentiated the antiproliferative effect of GLS silencing [289]. In other tumor types, Myc upregulates mitochondrial glutaminase through the transcriptional repression of mir23a/b and the consequent upregulation of its target, mitochondrial glutaminase [248]. Although the mechanisms have not been yet established in glioblastoma, neurons express mir23a/b, and mir23a controls EMT in GBM by targeting homeobox D10 (HOXD10) [290], whereas mir23b is downregulated by Myc in ischemic neurons, leading to nuclear factor erythroid 2-related factor 2 (Nrf2) upregulation and reductions in apoptosis and the infarcted area [291]. Furthermore, Myc has been found located on the mir23b promoter by chromatin immunoprecipitation, suggesting a direct Myc-dependent transcriptional mechanism [292]. Therefore, we may argue that similar transcriptional pathways could occur in GBM, leading to mitochondrial glutaminase upregulation.
The Myc-dependent synthesis of different metabolic products has a profound impact on the maintenance of GSCs. Indeed, Myc has been found to regulate mevalonate signaling in GSCs. Coenzyme-Q and cholesterol are the end products of the mevalonate pathway, together with isoprenoid intermediates, serving as signal transducers [293]. GSCs typically overexpress mevalonate pathway enzymes and silencing of the rate-limiting enzyme HMGCR, which converts HMG-CoA to mevalonic acid, dramatically impairs GSCs proliferation ability and viability. Affecting protein prenylation with farnesyl and geranylgeranyl transferase inhibitors blocks GSC viability and neurosphere formation. Myc has been found recruited at the promoters of the six major enzymes of the mevalonate pathway, and its depletion leads to decreases in their expressions that are paralleled with reduced GSC viability, self-renewal, and proliferation, indicating the Myc-promoted mevalonate pathway as pivotal for GSC maintenance [294]. Purine synthesis, which is strictly connected to glucose availability and uptake through the GLUT3 transporter [127], is another metabolic pathway controlled by Myc in GSCs. Myc KO in GSCs results in decreased expressions of purine biosynthetic enzymes and the IMP, AMP, and GMP metabolites, whilst in GSCs committed to differentiation, high levels of Myc enhance the expressions of purine biosynthetic enzymes and related metabolites. Mechanistically, Myc has been found located on promoters of IMPDH1, PPAT, PRPS1, ADSS, ADSL, and GMPS and interfering with the expressions of enzymes devoted to purine biosynthesis, strongly limiting GSC self-renewal ability [295]. Glucose is not the only source of carbon in cancer metabolism. Indeed, glutamine partially rescues GSC viability upon glucose deprivation and serves also as a source of nitrogen atoms for purine synthesis. By controlling glutaminolysis, Myc provides both intermediates and nitrogen atoms to fulfill GSC demands for the TCA cycle and purine synthesis, respectively.

Crosstalk between Myc-Dependent Angiogenesis and Metabolism in GBM

GBM is one of the most vascularized tumors, and GBM cells exploit different ways to ensure the adequate uptake of nutrients and oxygen from the microenvironment. In addition to the well-known process of neoangiogenesis, the formation of new blood vessels from pre-existing ones, GBM cells deploy vascular mimicry and vascular co-options to both increase blood supply and invade surrounding tissue (for details, see Rosiňska and Gavard, 2021 [296]). GBM is characterized by extremely hypoxic and necrotic areas. The hypoxic environment characterizing a tumor mass, with the activation of HIF1α and the downstream transcriptional cascade leading to the production of a plethora of pro-angiogenic factors, is one of the major determinants of GBM neoangiogenesis [297]. Once activated, this pathway not only promotes the remodeling of the vasculature through extracellular matrix degradation via metalloproteinases and the activation of stromal cells, as well as the recruitment, migration, and proliferation of endothelial cells [298], but it also induces GSCs to differentiate into endothelial-like cells, forming capillary-like structures [299]. The latter have a larger lumen with respect to their normal counterparts and present irregular branching [300]. Leakiness frequently occurs, causing plasma extravasation and focal edema [301]. At the molecular level, in addition to the well-established mechanism relying on the O2-deprivation-dependent induction of HIF1α [302], it has been recently established that, upstream of HIF1α, lncRNA H19 regulates glioma angiogenesis. Acting as a sponge for mir138, which targets HIF1α, lncRNA H19 promotes glioma angiogenesis in a HIF1α- and vascular endothelial growth factor (VEGF)-dependent manner [303]. Other noncoding RNAs control GBM angiogenesis [304], as well as transcription factors such as NK-kB [305] and STAT3 [306].
In this scenario, Myc participates in GBM angiogenesis regulation, even in normoxia. In this condition, wnt/β-catenin pathway activation leads to Myc overexpression and consequent HIF1α upregulation independently of O2 deprivation. β-catenin/TCF4 in complex with STAT3 also activates HIF1α, which in turn upregulates Myc in a positive feedback loop [307].
EGFRvIII+ GBM has been found to be more vascularized than EGFRwt tumors in mouse xenograft models. In this condition, Myc also plays a pivotal role in driving tumor angiogenesis. Indeed, the EGFRvIII-dependent upregulation of Myc promotes ANG4 transcription through direct binding to its promoter [214]. Myc also blunts the production of antiangiogenic factors in GBM, thus promoting angiogenesis indirectly. In fact, one of the most well-known Myc target loci, the mir17-92 cluster [308], was reported to downregulate the expressions of transforming growth factor receptor (TGFRβ) II and Smad4, impairing the production of antiangiogenic molecules such as thrombospondin and clusterin [309].
Very recently, Myc was also found to induce GBM angiogenesis through histamine production and the release of this metabolite in the tumor microenvironment from GSCs. Indeed, Myc binds the histidine decarboxylase (HDC) promoter, which in GSCs is in an open chromatin conformation, bearing a high amount of trimethylated histone H4 on lysine 4 (H3K4me3), as revealed by ATACseq experiments. Released histamine in the extracellular space binds to the histamine H1 receptor on endothelial cells, activating the NF-kB pathway and angiogenesis [310]. Furthermore, the activation of Myc/HIF1α-dependent aerobic glycolysis gene expression pushes the production of lactate, which is released by GBM cells in the extracellular space, inducing acidosis. The latter activates HIF1α in the surrounding endothelial cells with the production of proangiogenic molecules, such as VEGF [307]. These results indicate a strict interrelationship among gene expression, metabolism, and final biological outcomes (i.e., angiogenesis), with Myc as one of the most important molecular coordinators.

5.3. Targeting Myc-Dependent Metabolic Pathways in GBM

Targeting cancer metabolism for therapeutic purposes is a hot topic. Glucose metabolism is the pathway toward which a large effort has been made, and many antiglycolytic molecules have been investigated in preclinical studies or have entered phase-I and -II clinical trials [311]. These agents target glycolytic enzymes, such as HK2 [312,313], GAPDH [314,315], PDK [316,317,318], and LDHA [319,320], or glucose transporters [321,322,323]. Interestingly, most of them are under the transcriptional control of Myc, as well as GLS1, which has become a putative therapeutic target for cancer in the past few decades [324], with at least one inhibitor (CB-839 [325,326]) used in 21 clinical trials on a variety of tumors (including ovarian cancer, renal carcinoma, breast cancer, colorectal cancer, non-small-cell lung cancer, and IDH-mut grade-II and -III astrocytomas; https://www.clinicaltrials.gov/ct2/results?cond=Cancer&term=CB-839&cntry=&state=&city=&dist=&Search=Search; accessed on 12 January 2023), mainly in combination with other chemotherapeutic agents. The feasibility of targeting glycolysis in malignant brain tumors was demonstrated in the early 1990s [312] and validated in the past few years [316]. Given the dependency of GBM on Myc-driven enhanced glycolytic flux, recently Myc-deregulated glycolysis has been evaluated as a putative therapeutic target in preclinical models of GBM. Specifically, it has been demonstrated that forcing Myc expression in GBM cells characterized by low levels of Myc protein strongly enhances the production of glycolytic intermediates and induces glucose dependence for cell growth. Consistently, in patient-derived GBM cells typically expressing high levels of Myc, Myc KO led to downregulations in glycolytic enzymes and related products [327]. On these bases, small molecule inhibitors of nicotinamide phosphoribosyl-transferase (NAMPT) essential for the NAD+ requiring glycolytic GAPDH step have been employed both in vitro and in vivo as a proof-of-concept for Myc-driven glycolytic targeting in GBM. Specifically, FK866 [328] and GMX1778 [329] each induced apoptotic cytotoxicity in GBM cells in a Myc-dependent manner by glycolysis inhibition, and orally administered GMX1778 significantly extended the survival of mice undergoing Myc-amplified patient-derived orthotopic xenografts [327]. These findings indicate Myc-controlled metabolic routes as valuable and significant therapeutic targets to take into account for GBM treatment.

6. Conclusions: Postulating a Myc-Centered Metabolic Circuit in Glioblastoma

The information provided within this work clearly demonstrates how Myc is central for almost all aspects of cell biology, either healthy or cancerous, as outlined below.
I.
Coordinating a plethora of transcriptional coregulators, especially RNAP I, II, and III, as well as controlling RNA splicing and capping, Myc activity is not limited to the activation or repression of directly bound genes at promoters. Furthermore, by binding to superenhancers and increasing chromatin contacts at transcriptionally active domains, Myc also governs chromatin topology [330]. Taking into account all this information, Myc may be defined as a “superoncogene”, orchestrating chromatin architecture, modulating RNAP activity, and directly inducing or repressing transcription, thus ruling the expression of gene expression programs and, ultimately, determining cell fate.
II.
Myc involvement in metabolic processes further underpins its importance in the maintenance of cellular homeostasis and in the metabolic rewiring of cancer cells where it is largely overexpressed. In GBM, multiple metabolic signals control Myc expression, which, in turn, governs the activation of cancer metabolic routes and the shunting from one pathway to another (Figure 3), suggesting a Myc-centered, self-sustaining metabolic circuit fulfilling GBM cell and, more importantly, GSC energy demand, fostering tumor growth.
III.
The launching and completion of many clinical trials targeting cancer cell metabolism in a variety of cancer types using small molecules that hit pivotal Myc-dependent metabolic enzymes and pathways, including astrocytomas and gliomas, highlights the importance of providing an in-depth characterization of this Myc-centered circuit in GBM as a way to design novel therapeutic strategies aimed at increasing the pool of weapons against this deadly type of tumor.

Author Contributions

C.C. wrote sections of the MS related to cancer metabolism; F.S. (Fiorella Scagnoli) wrote the sections of the MS related to Myc; F.S. (Francesco Spallotta) wrote the sections of the MS related to cancer metabolism; S.N. supervised and approved the MS; B.I. supervised and approved the MS and wrote the sections of the MS related to GBM, GSCs, and Myc. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by AIRC (Italian Association for Cancer Research); MFAG grant number 23099 to Francesco Spallotta.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Preusser, M.; de Ribaupierre, S.; Wohrer, A.; Erridge, S.C.; Hegi, M.; Weller, M.; Stupp, R. Current concepts and management of glioblastoma. Ann. Neurol. 2011, 70, 9–21. [Google Scholar] [CrossRef]
  2. Patel, A.P.; Tirosh, I.; Trombetta, J.J.; Shalek, A.K.; Gillespie, S.M.; Wakimoto, H.; Cahill, D.P.; Nahed, B.V.; Curry, W.T.; Martuza, R.L.; et al. Single-cell RNA-seq highlights intratumoral heterogeneity in primary glioblastoma. Science 2014, 344, 1396–1401. [Google Scholar] [CrossRef] [Green Version]
  3. Plate, K.H.; Breier, G.; Weich, H.A.; Risau, W. Vascular endothelial growth factor is a potential tumour angiogenesis factor in human gliomas in vivo. Nature 1992, 359, 845–848. [Google Scholar] [CrossRef]
  4. Frei, K.; Gramatzki, D.; Tritschler, I.; Schroeder, J.J.; Espinoza, L.; Rushing, E.J.; Weller, M. Transforming growth factor-beta pathway activity in glioblastoma. Oncotarget 2015, 6, 5963–5977. [Google Scholar] [CrossRef] [Green Version]
  5. Verhaak, R.G.; Hoadley, K.A.; Purdom, E.; Wang, V.; Qi, Y.; Wilkerson, M.D.; Miller, C.R.; Ding, L.; Golub, T.; Mesirov, J.P.; et al. Integrated genomic analysis identifies clinically relevant subtypes of glioblastoma characterized by abnormalities in PDGFRA, IDH1, EGFR, and NF1. Cancer Cell 2010, 17, 98–110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Gimple, R.C.; Bhargava, S.; Dixit, D.; Rich, J.N. Glioblastoma stem cells: Lessons from the tumor hierarchy in a lethal cancer. Genes Dev. 2019, 33, 591–609. [Google Scholar] [CrossRef] [PubMed]
  7. Friedmann-Morvinski, D.; Bushong, E.A.; Ke, E.; Soda, Y.; Marumoto, T.; Singer, O.; Ellisman, M.H.; Verma, I.M. Dedifferentiation of neurons and astrocytes by oncogenes can induce gliomas in mice. Science 2012, 338, 1080–1084. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Liu, C.; Sage, J.C.; Miller, M.R.; Verhaak, R.G.; Hippenmeyer, S.; Vogel, H.; Foreman, O.; Bronson, R.T.; Nishiyama, A.; Luo, L.; et al. Mosaic analysis with double markers reveals tumor cell of origin in glioma. Cell 2011, 146, 209–221. [Google Scholar] [CrossRef] [Green Version]
  9. Wang, Q.; Hu, B.; Hu, X.; Kim, H.; Squatrito, M.; Scarpace, L.; deCarvalho, A.C.; Lyu, S.; Li, P.; Li, Y.; et al. Tumor Evolution of Glioma-Intrinsic Gene Expression Subtypes Associates with Immunological Changes in the Microenvironment. Cancer Cell 2017, 32, 42–56.e46. [Google Scholar] [CrossRef] [Green Version]
  10. 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]
  11. Hoang-Minh, L.B.; Siebzehnrubl, F.A.; Yang, C.; Suzuki-Hatano, S.; Dajac, K.; Loche, T.; Andrews, N.; Schmoll Massari, M.; Patel, J.; Amin, K.; et al. Infiltrative and drug-resistant slow-cycling cells support metabolic heterogeneity in glioblastoma. EMBO J. 2018, 37, e98772. [Google Scholar] [CrossRef]
  12. Lin, H.; Patel, S.; Affleck, V.S.; Wilson, I.; Turnbull, D.M.; Joshi, A.R.; Maxwell, R.; Stoll, E.A. Fatty acid oxidation is required for the respiration and proliferation of malignant glioma cells. Neuro Oncol. 2017, 19, 43–54. [Google Scholar] [CrossRef] [Green Version]
  13. Meyer, N.; Penn, L.Z. Reflecting on 25 years with MYC. Nat. Rev. Cancer 2008, 8, 976–990. [Google Scholar] [CrossRef]
  14. Zhu, S.; Cheng, X.; Wang, R.; Tan, Y.; Ge, M.; Li, D.; Xu, Q.; Sun, Y.; Zhao, C.; Chen, S.; et al. Restoration of microRNA function impairs MYC-dependent maintenance of MLL leukemia. Leukemia 2020, 34, 2484–2488. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Mao, A.; Zhao, Q.; Zhou, X.; Sun, C.; Si, J.; Zhou, R.; Gan, L.; Zhang, H. MicroRNA-449a enhances radiosensitivity by downregulation of c-Myc in prostate cancer cells. Sci. Rep. 2016, 6, 27346. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Zhu, Y.; Zhou, B.; Hu, X.; Ying, S.; Zhou, Q.; Xu, W.; Feng, L.; Hou, T.; Wang, X.; Zhu, L.; et al. LncRNA LINC00942 promotes chemoresistance in gastric cancer by suppressing MSI2 degradation to enhance c-Myc mRNA stability. Clin. Transl. Med. 2022, 12, e703. [Google Scholar] [CrossRef] [PubMed]
  17. Zhu, P.; He, F.; Hou, Y.; Tu, G.; Li, Q.; Jin, T.; Zeng, H.; Qin, Y.; Wan, X.; Qiao, Y.; et al. A novel hypoxic long noncoding RNA KB-1980E6.3 maintains breast cancer stem cell stemness via interacting with IGF2BP1 to facilitate c-Myc mRNA stability. Oncogene 2021, 40, 1609–1627. [Google Scholar] [CrossRef]
  18. Farrell, A.S.; Sears, R.C. MYC degradation. Cold Spring Harb. Perspect. Med. 2014, 4, a014365. [Google Scholar] [CrossRef] [Green Version]
  19. Zhao, K.; Wang, Q.; Wang, Y.; Huang, K.; Yang, C.; Li, Y.; Yi, K.; Kang, C. EGFR/c-myc axis regulates TGFbeta/Hippo/Notch pathway via epigenetic silencing miR-524 in gliomas. Cancer Lett. 2017, 406, 12–21. [Google Scholar] [CrossRef]
  20. Xu, Q.; Ahmed, A.K.; Zhu, Y.; Wang, K.; Lv, S.; Li, Y.; Jiang, Y. Oncogenic MicroRNA-20a is downregulated by the HIF-1alpha/c-MYC pathway in IDH1 R132H-mutant glioma. Biochem. Biophys. Res. Commun. 2018, 499, 882–888. [Google Scholar] [CrossRef]
  21. Luo, H.; Chen, Z.; Wang, S.; Zhang, R.; Qiu, W.; Zhao, L.; Peng, C.; Xu, R.; Chen, W.; Wang, H.W.; et al. c-Myc-miR-29c-REV3L signalling pathway drives the acquisition of temozolomide resistance in glioblastoma. Brain 2015, 138, 3654–3672. [Google Scholar] [CrossRef] [Green Version]
  22. Zhang, G.; Zhu, Q.; Fu, G.; Hou, J.; Hu, X.; Cao, J.; Peng, W.; Wang, X.; Chen, F.; Cui, H. TRIP13 promotes the cell proliferation, migration and invasion of glioblastoma through the FBXW7/c-MYC axis. Br. J. Cancer 2019, 121, 1069–1078. [Google Scholar] [CrossRef]
  23. Ding, Z.; Liu, X.; Liu, Y.; Zhang, J.; Huang, X.; Yang, X.; Yao, L.; Cui, G.; Wang, D. Expression of far upstream element (FUSE) binding protein 1 in human glioma is correlated with c-Myc and cell proliferation. Mol. Carcinog. 2015, 54, 405–415. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, J.; Wang, H.; Li, Z.; Wu, Q.; Lathia, J.D.; McLendon, R.E.; Hjelmeland, A.B.; Rich, J.N. c-Myc is required for maintenance of glioma cancer stem cells. PLoS ONE 2008, 3, e3769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. 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]
  26. Dhanasekaran, R.; Deutzmann, A.; Mahauad-Fernandez, W.D.; Hansen, A.S.; Gouw, A.M.; Felsher, D.W. The MYC oncogene—The grand orchestrator of cancer growth and immune evasion. Nat. Rev. Clin. Oncol. 2022, 19, 23–36. [Google Scholar] [CrossRef] [PubMed]
  27. 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] [PubMed] [Green Version]
  28. Miller, C.R.; Perry, A. Glioblastoma. Arch. Pathol. Lab. Med. 2007, 131, 397–406. [Google Scholar] [CrossRef]
  29. Nagy, A.; Garzuly, F.; Padanyi, G.; Szucs, I.; Feldmann, A.; Murnyak, B.; Hortobagyi, T.; Kalman, B. Molecular Subgroups of Glioblastoma- an Assessment by Immunohistochemical Markers. Pathol. Oncol. Res. 2019, 25, 21–31. [Google Scholar] [CrossRef]
  30. Yonekura, A.; Kawanaka, H.; Prasath, V.B.S.; Aronow, B.J.; Takase, H. Automatic disease stage classification of glioblastoma multiforme histopathological images using deep convolutional neural network. BioMed. Eng. Lett. 2018, 8, 321–327. [Google Scholar] [CrossRef]
  31. Ostrom, Q.T.; Cioffi, G.; Waite, K.; Kruchko, C.; Barnholtz-Sloan, J.S. CBTRUS Statistical Report: Primary Brain and Other Central Nervous System Tumors Diagnosed in the United States in 2014–2018. Neuro. Oncol. 2021, 23, iii1–iii105. [Google Scholar] [CrossRef] [PubMed]
  32. Sturm, D.; Witt, H.; Hovestadt, V.; Khuong-Quang, D.A.; Jones, D.T.; Konermann, C.; Pfaff, E.; Tonjes, M.; Sill, M.; Bender, S.; et al. Hotspot mutations in H3F3A and IDH1 define distinct epigenetic and biological subgroups of glioblastoma. Cancer Cell 2012, 22, 425–437. [Google Scholar] [CrossRef] [Green Version]
  33. Pollen, A.A.; Nowakowski, T.J.; Chen, J.; Retallack, H.; Sandoval-Espinosa, C.; Nicholas, C.R.; Shuga, J.; Liu, S.J.; Oldham, M.C.; Diaz, A.; et al. Molecular identity of human outer radial glia during cortical development. Cell 2015, 163, 55–67. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Burger, P.C.; Kleihues, P. Cytologic composition of the untreated glioblastoma with implications for evaluation of needle biopsies. Cancer 1989, 63, 2014–2023. [Google Scholar] [CrossRef]
  35. Ohgaki, H.; Dessen, P.; Jourde, B.; Horstmann, S.; Nishikawa, T.; Di Patre, P.L.; Burkhard, C.; Schuler, D.; Probst-Hensch, N.M.; Maiorka, P.C.; et al. Genetic pathways to glioblastoma: A population-based study. Cancer Res. 2004, 64, 6892–6899. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Hartmann, C.; Meyer, J.; Balss, J.; Capper, D.; Mueller, W.; Christians, A.; Felsberg, J.; Wolter, M.; Mawrin, C.; Wick, W.; et al. Type and frequency of IDH1 and IDH2 mutations are related to astrocytic and oligodendroglial differentiation and age: A study of 1,010 diffuse gliomas. Acta Neuropathol. 2009, 118, 469–474. [Google Scholar] [CrossRef] [Green Version]
  37. Brennan, C.W.; Verhaak, R.G.; McKenna, A.; Campos, B.; Noushmehr, H.; Salama, S.R.; Zheng, S.; Chakravarty, D.; Sanborn, J.Z.; Berman, S.H.; et al. The somatic genomic landscape of glioblastoma. Cell 2013, 155, 462–477. [Google Scholar] [CrossRef]
  38. Yamazaki, H.; Ohba, Y.; Tamaoki, N.; Shibuya, M. A deletion mutation within the ligand binding domain is responsible for activation of epidermal growth factor receptor gene in human brain tumors. Jpn. J. Cancer Res. 1990, 81, 773–779. [Google Scholar] [CrossRef]
  39. Appin, C.L.; Brat, D.J. Molecular pathways in gliomagenesis and their relevance to neuropathologic diagnosis. Adv. Anat. Pathol. 2015, 22, 50–58. [Google Scholar] [CrossRef]
  40. Cancer Genome Atlas Research, N. Comprehensive genomic characterization defines human glioblastoma genes and core pathways. Nature 2008, 455, 1061–1068. [Google Scholar] [CrossRef] [Green Version]
  41. Rickman, D.S.; Bobek, M.P.; Misek, D.E.; Kuick, R.; Blaivas, M.; Kurnit, D.M.; Taylor, J.; Hanash, S.M. Distinctive molecular profiles of high-grade and low-grade gliomas based on oligonucleotide microarray analysis. Cancer Res. 2001, 61, 6885–6891. [Google Scholar]
  42. Phillips, H.S.; Kharbanda, S.; Chen, R.; Forrest, W.F.; Soriano, R.H.; Wu, T.D.; Misra, A.; Nigro, J.M.; Colman, H.; Soroceanu, L.; et al. Molecular subclasses of high-grade glioma predict prognosis, delineate a pattern of disease progression, and resemble stages in neurogenesis. Cancer Cell 2006, 9, 157–173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Kastrup, J. Can YKL-40 be a new inflammatory biomarker in cardiovascular disease? Immunobiology 2012, 217, 483–491. [Google Scholar] [CrossRef] [PubMed]
  44. Johansen, J.S.; Schultz, N.A.; Jensen, B.V. Plasma YKL-40: A potential new cancer biomarker? Future Oncol. 2009, 5, 1065–1082. [Google Scholar] [CrossRef]
  45. Nutt, C.L.; Betensky, R.A.; Brower, M.A.; Batchelor, T.T.; Louis, D.N.; Stemmer-Rachamimov, A.O. YKL-40 is a differential diagnostic marker for histologic subtypes of high-grade gliomas. Clin. Cancer Res. 2005, 11, 2258–2264. [Google Scholar] [CrossRef] [Green Version]
  46. Pelloski, C.E.; Mahajan, A.; Maor, M.; Chang, E.L.; Woo, S.; Gilbert, M.; Colman, H.; Yang, H.; Ledoux, A.; Blair, H.; et al. YKL-40 expression is associated with poorer response to radiation and shorter overall survival in glioblastoma. Clin. Cancer Res. 2005, 11, 3326–3334. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Pegg, A.E.; Dolan, M.E.; Moschel, R.C. Structure, function, and inhibition of O6-alkylguanine-DNA alkyltransferase. Prog. Nucleic Acid Res. Mol. Biol. 1995, 51, 167–223. [Google Scholar] [CrossRef]
  48. Esteller, M.; Garcia-Foncillas, J.; Andion, E.; Goodman, S.N.; Hidalgo, O.F.; Vanaclocha, V.; Baylin, S.B.; Herman, J.G. Inactivation of the DNA-repair gene MGMT and the clinical response of gliomas to alkylating agents. N. Engl. J. Med. 2000, 343, 1350–1354. [Google Scholar] [CrossRef]
  49. Stommel, J.M.; Kimmelman, A.C.; Ying, H.; Nabioullin, R.; Ponugoti, A.H.; Wiedemeyer, R.; Stegh, A.H.; Bradner, J.E.; Ligon, K.L.; Brennan, C.; et al. Coactivation of receptor tyrosine kinases affects the response of tumor cells to targeted therapies. Science 2007, 318, 287–290. [Google Scholar] [CrossRef]
  50. Kouzarides, T. Chromatin modifications and their function. Cell 2007, 128, 693–705. [Google Scholar] [CrossRef] [Green Version]
  51. Ng, R.K.; Gurdon, J.B. Epigenetic memory of an active gene state depends on histone H3.3 incorporation into chromatin in the absence of transcription. Nat. Cell Biol. 2008, 10, 102–109. [Google Scholar] [CrossRef]
  52. Pasque, V.; Halley-Stott, R.P.; Gillich, A.; Garrett, N.; Gurdon, J.B. Epigenetic stability of repressed states involving the histone variant macroH2A revealed by nuclear transfer to Xenopus oocytes. Nucleus 2011, 2, 533–539. [Google Scholar] [CrossRef] [Green Version]
  53. Lejeune, E.; Allshire, R.C. Common ground: Small RNA programming and chromatin modifications. Curr. Opin. Cell Biol. 2011, 23, 258–265. [Google Scholar] [CrossRef]
  54. Ong, C.T.; Corces, V.G. CTCF: An architectural protein bridging genome topology and function. Nat. Rev. Genet. 2014, 15, 234–246. [Google Scholar] [CrossRef] [Green Version]
  55. Neems, D.S.; Garza-Gongora, A.G.; Smith, E.D.; Kosak, S.T. Topologically associated domains enriched for lineage-specific genes reveal expression-dependent nuclear topologies during myogenesis. Proc. Natl. Acad. Sci. USA 2016, 113, E1691–E1700. [Google Scholar] [CrossRef] [Green Version]
  56. Jenuwein, T.; Allis, C.D. Translating the histone code. Science 2001, 293, 1074–1080. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Ceccarelli, M.; Barthel, F.P.; Malta, T.M.; Sabedot, T.S.; Salama, S.R.; Murray, B.A.; Morozova, O.; Newton, Y.; Radenbaugh, A.; Pagnotta, S.M.; et al. Molecular Profiling Reveals Biologically Discrete Subsets and Pathways of Progression in Diffuse Glioma. Cell 2016, 164, 550–563. [Google Scholar] [CrossRef] [Green Version]
  58. Mackay, A.; Burford, A.; Carvalho, D.; Izquierdo, E.; Fazal-Salom, J.; Taylor, K.R.; Bjerke, L.; Clarke, M.; Vinci, M.; Nandhabalan, M.; et al. Integrated Molecular Meta-Analysis of 1,000 Pediatric High-Grade and Diffuse Intrinsic Pontine Glioma. Cancer Cell 2017, 32, 520–537.e525. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Lewis, P.W.; Muller, M.M.; Koletsky, M.S.; Cordero, F.; Lin, S.; Banaszynski, L.A.; Garcia, B.A.; Muir, T.W.; Becher, O.J.; Allis, C.D. Inhibition of PRC2 activity by a gain-of-function H3 mutation found in pediatric glioblastoma. Science 2013, 340, 857–861. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Portela, A.; Esteller, M. Epigenetic modifications and human disease. Nat. Biotechnol. 2010, 28, 1057–1068. [Google Scholar] [CrossRef] [PubMed]
  61. Lambiv, W.L.; Vassallo, I.; Delorenzi, M.; Shay, T.; Diserens, A.C.; Misra, A.; Feuerstein, B.; Murat, A.; Migliavacca, E.; Hamou, M.F.; et al. The Wnt inhibitory factor 1 (WIF1) is targeted in glioblastoma and has a tumor suppressing function potentially by induction of senescence. Neuro Oncol. 2011, 13, 736–747. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Gotze, S.; Wolter, M.; Reifenberger, G.; Muller, O.; Sievers, S. Frequent promoter hypermethylation of Wnt pathway inhibitor genes in malignant astrocytic gliomas. Int. J. Cancer 2010, 126, 2584–2593. [Google Scholar] [CrossRef] [PubMed]
  63. Horiguchi, K.; Tomizawa, Y.; Tosaka, M.; Ishiuchi, S.; Kurihara, H.; Mori, M.; Saito, N. Epigenetic inactivation of RASSF1A candidate tumor suppressor gene at 3p21.3 in brain tumors. Oncogene 2003, 22, 7862–7865. [Google Scholar] [CrossRef] [Green Version]
  64. Noushmehr, H.; Weisenberger, D.J.; Diefes, K.; Phillips, H.S.; Pujara, K.; Berman, B.P.; Pan, F.; Pelloski, C.E.; Sulman, E.P.; Bhat, K.P.; et al. Identification of a CpG island methylator phenotype that defines a distinct subgroup of glioma. Cancer Cell 2010, 17, 510–522. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Ichimura, K. Molecular pathogenesis of IDH mutations in gliomas. Brain Tumor Pathol. 2012, 29, 131–139. [Google Scholar] [CrossRef]
  66. Dang, L.; White, D.W.; Gross, S.; Bennett, B.D.; Bittinger, M.A.; Driggers, E.M.; Fantin, V.R.; Jang, H.G.; Jin, S.; Keenan, M.C.; et al. Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature 2009, 462, 739–744. [Google Scholar] [CrossRef] [Green Version]
  67. Xu, W.; Yang, H.; Liu, Y.; Yang, Y.; Wang, P.; Kim, S.H.; Ito, S.; Yang, C.; Wang, P.; Xiao, M.T.; et al. Oncometabolite 2-hydroxyglutarate is a competitive inhibitor of alpha-ketoglutarate-dependent dioxygenases. Cancer Cell 2011, 19, 17–30. [Google Scholar] [CrossRef] [Green Version]
  68. Yuan, F.; Wang, Y.; Cai, X.; Du, C.; Zhu, J.; Tang, C.; Yang, J.; Ma, C. N6-methyladenosine-related microRNAs risk model trumps the isocitrate dehydrogenase mutation status as a predictive biomarker for the prognosis and immunotherapy in lower grade gliomas. Explor. Target. Antitumor Ther. 2022, 3, 553–569. [Google Scholar] [CrossRef]
  69. Pan, T. N6-methyl-adenosine modification in messenger and long non-coding RNA. Trends Biochem. Sci. 2013, 38, 204–209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Ma, S.; Chen, C.; Ji, X.; Liu, J.; Zhou, Q.; Wang, G.; Yuan, W.; Kan, Q.; Sun, Z. The interplay between m6A RNA methylation and noncoding RNA in cancer. J. Hematol. Oncol. 2019, 12, 121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Wang, Y.; Gao, M.; Zhu, F.; Li, X.; Yang, Y.; Yan, Q.; Jia, L.; Xie, L.; Chen, Z. METTL3 is essential for postnatal development of brown adipose tissue and energy expenditure in mice. Nat. Commun. 2020, 11, 1648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Ping, X.L.; Sun, B.F.; Wang, L.; Xiao, W.; Yang, X.; Wang, W.J.; Adhikari, S.; Shi, Y.; Lv, Y.; Chen, Y.S.; et al. Mammalian WTAP is a regulatory subunit of the RNA N6-methyladenosine methyltransferase. Cell Res. 2014, 24, 177–189. [Google Scholar] [CrossRef] [Green Version]
  73. Meyer, K.D.; Jaffrey, S.R. Rethinking m(6)A Readers, Writers, and Erasers. Annu. Rev. Cell Dev. Biol. 2017, 33, 319–342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Schwartz, S.; Mumbach, M.R.; Jovanovic, M.; Wang, T.; Maciag, K.; Bushkin, G.G.; Mertins, P.; Ter-Ovanesyan, D.; Habib, N.; Cacchiarelli, D.; et al. Perturbation of m6A writers reveals two distinct classes of mRNA methylation at internal and 5′ sites. Cell Rep. 2014, 8, 284–296. [Google Scholar] [CrossRef] [Green Version]
  75. Jia, G.; Fu, Y.; Zhao, X.; Dai, Q.; Zheng, G.; Yang, Y.; Yi, C.; Lindahl, T.; Pan, T.; Yang, Y.G.; et al. N6-methyladenosine in nuclear RNA is a major substrate of the obesity-associated FTO. Nat. Chem. Biol. 2011, 7, 885–887. [Google Scholar] [CrossRef] [PubMed]
  76. Zheng, G.; Dahl, J.A.; Niu, Y.; Fedorcsak, P.; Huang, C.M.; Li, C.J.; Vagbo, C.B.; Shi, Y.; Wang, W.L.; Song, S.H.; et al. ALKBH5 is a mammalian RNA demethylase that impacts RNA metabolism and mouse fertility. Mol. Cell 2013, 49, 18–29. [Google Scholar] [CrossRef] [Green Version]
  77. Chen, X.Y.; Zhang, J.; Zhu, J.S. The role of m(6)A RNA methylation in human cancer. Mol. Cancer 2019, 18, 103. [Google Scholar] [CrossRef] [Green Version]
  78. Musacchio, A.; Salmon, E.D. The spindle-assembly checkpoint in space and time. Nat. Rev. Mol. Cell Biol. 2007, 8, 379–393. [Google Scholar] [CrossRef]
  79. Ding, Y.; Hubert, C.G.; Herman, J.; Corrin, P.; Toledo, C.M.; Skutt-Kakaria, K.; Vazquez, J.; Basom, R.; Zhang, B.; Risler, J.K.; et al. Cancer-Specific requirement for BUB1B/BUBR1 in human brain tumor isolates and genetically transformed cells. Cancer Discov. 2013, 3, 198–211. [Google Scholar] [CrossRef] [Green Version]
  80. Lee, E.; Pain, M.; Wang, H.; Herman, J.A.; Toledo, C.M.; DeLuca, J.G.; Yong, R.L.; Paddison, P.; Zhu, J. Sensitivity to BUB1B Inhibition Defines an Alternative Classification of Glioblastoma. Cancer Res. 2017, 77, 5518–5529. [Google Scholar] [CrossRef] [Green Version]
  81. Bonnet, D.; Dick, J.E. Human acute myeloid leukemia is organized as a hierarchy that originates from a primitive hematopoietic cell. Nat. Med. 1997, 3, 730–737. [Google Scholar] [CrossRef]
  82. Vlashi, E.; Pajonk, F. Cancer stem cells, cancer cell plasticity and radiation therapy. Semin. Cancer Biol. 2015, 31, 28–35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Ignatova, T.N.; Kukekov, V.G.; Laywell, E.D.; Suslov, O.N.; Vrionis, F.D.; Steindler, D.A. Human cortical glial tumors contain neural stem-like cells expressing astroglial and neuronal markers in vitro. Glia 2002, 39, 193–206. [Google Scholar] [CrossRef]
  84. Guo, G.; Yang, J.; Nichols, J.; Hall, J.S.; Eyres, I.; Mansfield, W.; Smith, A. Klf4 reverts developmentally programmed restriction of ground state pluripotency. Development 2009, 136, 1063–1069. [Google Scholar] [CrossRef] [Green Version]
  85. Akberdin, I.R.; Omelyanchuk, N.A.; Fadeev, S.I.; Leskova, N.E.; Oschepkova, E.A.; Kazantsev, F.V.; Matushkin, Y.G.; Afonnikov, D.A.; Kolchanov, N.A. Pluripotency gene network dynamics: System views from parametric analysis. PLoS ONE 2018, 13, e0194464. [Google Scholar] [CrossRef] [PubMed]
  86. Ignatova, T.N.; Chaitin, H.J.; Kukekov, N.V.; Suslov, O.N.; Dulatova, G.I.; Hanafy, K.A.; Vrionis, F.D. Gliomagenesis is orchestrated by the Oct3/4 regulatory network. J. Neurosurg. Sci. 2021. [Google Scholar] [CrossRef]
  87. Kanwar, S.S.; Yu, Y.; Nautiyal, J.; Patel, B.B.; Majumdar, A.P. The Wnt/beta-catenin pathway regulates growth and maintenance of colonospheres. Mol. Cancer 2010, 9, 212. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Liu, Y.; Luan, Y.; Ma, K.; Zhang, Z.; Liu, Y.; Chen, X. ISL1 Promotes Human Glioblastoma-Derived Stem Cells’ Self-Renewal by Activation of Sonic Hedgehog/GLI1 Function. Stem Cells Dev. 2022, 31, 258–268. [Google Scholar] [CrossRef]
  89. Wang, J.; Wakeman, T.P.; Lathia, J.D.; Hjelmeland, A.B.; Wang, X.F.; White, R.R.; Rich, J.N.; Sullenger, B.A. Notch promotes radioresistance of glioma stem cells. Stem Cells 2010, 28, 17–28. [Google Scholar] [CrossRef] [Green Version]
  90. Adams, J.M.; Strasser, A. Is tumor growth sustained by rare cancer stem cells or dominant clones? Cancer Res. 2008, 68, 4018–4021. [Google Scholar] [CrossRef] [Green Version]
  91. Tamura, K.; Aoyagi, M.; Ando, N.; Ogishima, T.; Wakimoto, H.; Yamamoto, M.; Ohno, K. Expansion of CD133-positive glioma cells in recurrent de novo glioblastomas after radiotherapy and chemotherapy. J. Neurosurg. 2013, 119, 1145–1155. [Google Scholar] [CrossRef] [Green Version]
  92. Diehn, M.; Cho, R.W.; Clarke, M.F. Therapeutic implications of the cancer stem cell hypothesis. Semin. Radiat. Oncol. 2009, 19, 78–86. [Google Scholar] [CrossRef] [Green Version]
  93. Bao, S.; Wu, Q.; McLendon, R.E.; Hao, Y.; Shi, Q.; Hjelmeland, A.B.; Dewhirst, M.W.; Bigner, D.D.; Rich, J.N. Glioma stem cells promote radioresistance by preferential activation of the DNA damage response. Nature 2006, 444, 756–760. [Google Scholar] [CrossRef]
  94. Bhat, K.P.L.; Balasubramaniyan, V.; Vaillant, B.; Ezhilarasan, R.; Hummelink, K.; Hollingsworth, F.; Wani, K.; Heathcock, L.; James, J.D.; Goodman, L.D.; et al. Mesenchymal differentiation mediated by NF-kappaB promotes radiation resistance in glioblastoma. Cancer Cell 2013, 24, 331–346. [Google Scholar] [CrossRef] [Green Version]
  95. Venere, M.; Hamerlik, P.; Wu, Q.; Rasmussen, R.D.; Song, L.A.; Vasanji, A.; Tenley, N.; Flavahan, W.A.; Hjelmeland, A.B.; Bartek, J.; et al. Therapeutic targeting of constitutive PARP activation compromises stem cell phenotype and survival of glioblastoma-initiating cells. Cell Death Differ. 2014, 21, 258–269. [Google Scholar] [CrossRef] [Green Version]
  96. Barazzuol, L.; Jena, R.; Burnet, N.G.; Meira, L.B.; Jeynes, J.C.; Kirkby, K.J.; Kirkby, N.F. Evaluation of poly (ADP-ribose) polymerase inhibitor ABT-888 combined with radiotherapy and temozolomide in glioblastoma. Radiat. Oncol. 2013, 8, 65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Gage, F.H. Mammalian neural stem cells. Science 2000, 287, 1433–1438. [Google Scholar] [CrossRef] [PubMed]
  98. Eyme, K.M.; Sammarco, A.; Badr, C.E. Orthotopic brain tumor models derived from glioblastoma stem-like cells. Methods Cell Biol. 2022, 170, 1–19. [Google Scholar] [CrossRef] [PubMed]
  99. Safa, A.R.; Saadatzadeh, M.R.; Cohen-Gadol, A.A.; Pollok, K.E.; Bijangi-Vishehsaraei, K. Glioblastoma stem cells (GSCs) epigenetic plasticity and interconversion between differentiated non-GSCs and GSCs. Genes Dis. 2015, 2, 152–163. [Google Scholar] [CrossRef] [Green Version]
  100. De Bacco, F.; Casanova, E.; Medico, E.; Pellegatta, S.; Orzan, F.; Albano, R.; Luraghi, P.; Reato, G.; D’Ambrosio, A.; Porrati, P.; et al. The MET oncogene is a functional marker of a glioblastoma stem cell subtype. Cancer Res. 2012, 72, 4537–4550. [Google Scholar] [CrossRef] [Green Version]
  101. Marziali, G.; Signore, M.; Buccarelli, M.; Grande, S.; Palma, A.; Biffoni, M.; Rosi, A.; D’Alessandris, Q.G.; Martini, M.; Larocca, L.M.; et al. Metabolic/Proteomic Signature Defines Two Glioblastoma Subtypes With Different Clinical Outcome. Sci. Rep. 2016, 6, 21557. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Wang, L.; Babikir, H.; Muller, S.; Yagnik, G.; Shamardani, K.; Catalan, F.; Kohanbash, G.; Alvarado, B.; Di Lullo, E.; Kriegstein, A.; et al. The Phenotypes of Proliferating Glioblastoma Cells Reside on a Single Axis of Variation. Cancer Discov. 2019, 9, 1708–1719. [Google Scholar] [CrossRef] [Green Version]
  103. Park, N.I.; Guilhamon, P.; Desai, K.; McAdam, R.F.; Langille, E.; O’Connor, M.; Lan, X.; Whetstone, H.; Coutinho, F.J.; Vanner, R.J.; et al. ASCL1 Reorganizes Chromatin to Direct Neuronal Fate and Suppress Tumorigenicity of Glioblastoma Stem Cells. Cell Stem Cell 2017, 21, 411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Varn, F.S.; Johnson, K.C.; Martinek, J.; Huse, J.T.; Nasrallah, M.P.; Wesseling, P.; Cooper, L.A.D.; Malta, T.M.; Wade, T.E.; Sabedot, T.S.; et al. Glioma progression is shaped by genetic evolution and microenvironment interactions. Cell 2022, 185, 2184–2199.e2116. [Google Scholar] [CrossRef]
  105. Snuderl, M.; Fazlollahi, L.; Le, L.P.; Nitta, M.; Zhelyazkova, B.H.; Davidson, C.J.; Akhavanfard, S.; Cahill, D.P.; Aldape, K.D.; Betensky, R.A.; et al. Mosaic amplification of multiple receptor tyrosine kinase genes in glioblastoma. Cancer Cell 2011, 20, 810–817. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Piccirillo, S.G.M.; Colman, S.; Potter, N.E.; van Delft, F.W.; Lillis, S.; Carnicer, M.J.; Kearney, L.; Watts, C.; Greaves, M. Genetic and functional diversity of propagating cells in glioblastoma. Stem Cell Rep. 2015, 4, 7–15. [Google Scholar] [CrossRef] [Green Version]
  107. Sottoriva, A.; Spiteri, I.; Piccirillo, S.G.; Touloumis, A.; Collins, V.P.; Marioni, J.C.; Curtis, C.; Watts, C.; Tavare, S. Intratumor heterogeneity in human glioblastoma reflects cancer evolutionary dynamics. Proc. Natl. Acad. Sci. USA 2013, 110, 4009–4014. [Google Scholar] [CrossRef] [Green Version]
  108. Johnson, B.E.; Mazor, T.; Hong, C.; Barnes, M.; Aihara, K.; McLean, C.Y.; Fouse, S.D.; Yamamoto, S.; Ueda, H.; Tatsuno, K.; et al. Mutational analysis reveals the origin and therapy-driven evolution of recurrent glioma. Science 2014, 343, 189–193. [Google Scholar] [CrossRef] [Green Version]
  109. Lan, X.; Jorg, D.J.; Cavalli, F.M.G.; Richards, L.M.; Nguyen, L.V.; Vanner, R.J.; Guilhamon, P.; Lee, L.; Kushida, M.M.; Pellacani, D.; et al. Fate mapping of human glioblastoma reveals an invariant stem cell hierarchy. Nature 2017, 549, 227–232. [Google Scholar] [CrossRef] [Green Version]
  110. Gallo, M.; Coutinho, F.J.; Vanner, R.J.; Gayden, T.; Mack, S.C.; Murison, A.; Remke, M.; Li, R.; Takayama, N.; Desai, K.; et al. MLL5 Orchestrates a Cancer Self-Renewal State by Repressing the Histone Variant H3.3 and Globally Reorganizing Chromatin. Cancer Cell 2015, 28, 715–729. [Google Scholar] [CrossRef] [Green Version]
  111. Hensler, M.; Kasikova, L.; Fiser, K.; Rakova, J.; Skapa, P.; Laco, J.; Lanickova, T.; Pecen, L.; Truxova, I.; Vosahlikova, S.; et al. M2-like macrophages dictate clinically relevant immunosuppression in metastatic ovarian cancer. J. Immunother. Cancer 2020, 8, e000979. [Google Scholar] [CrossRef]
  112. Arlauckas, S.P.; Garren, S.B.; Garris, C.S.; Kohler, R.H.; Oh, J.; Pittet, M.J.; Weissleder, R. Arg1 expression defines immunosuppressive subsets of tumor-associated macrophages. Theranostics 2018, 8, 5842–5854. [Google Scholar] [CrossRef] [PubMed]
  113. Wu, L.; Wu, W.; Zhang, J.; Zhao, Z.; Li, L.; Zhu, M.; Wu, M.; Wu, F.; Zhou, F.; Du, Y.; et al. Natural Coevolution of Tumor and Immunoenvironment in Glioblastoma. Cancer Discov. 2022, 12, 2820–2837. [Google Scholar] [CrossRef]
  114. Spehalski, E.I.; Lee, J.A.; Peters, C.; Tofilon, P.; Camphausen, K. The Quiescent Metabolic Phenotype of Glioma Stem Cells. J. Proteom. Bioinform. 2019, 12, 96–103. [Google Scholar] [CrossRef] [Green Version]
  115. Gunther, H.S.; Schmidt, N.O.; Phillips, H.S.; Kemming, D.; Kharbanda, S.; Soriano, R.; Modrusan, Z.; Meissner, H.; Westphal, M.; Lamszus, K. Glioblastoma-derived stem cell-enriched cultures form distinct subgroups according to molecular and phenotypic criteria. Oncogene 2008, 27, 2897–2909. [Google Scholar] [CrossRef] [PubMed]
  116. Schulte, A.; Gunther, H.S.; Phillips, H.S.; Kemming, D.; Martens, T.; Kharbanda, S.; Soriano, R.H.; Modrusan, Z.; Zapf, S.; Westphal, M.; et al. A distinct subset of glioma cell lines with stem cell-like properties reflects the transcriptional phenotype of glioblastomas and overexpresses CXCR4 as therapeutic target. Glia 2011, 59, 590–602. [Google Scholar] [CrossRef] [PubMed]
  117. Oizel, K.; Chauvin, C.; Oliver, L.; Gratas, C.; Geraldo, F.; Jarry, U.; Scotet, E.; Rabe, M.; Alves-Guerra, M.C.; Teusan, R.; et al. Efficient Mitochondrial Glutamine Targeting Prevails Over Glioblastoma Metabolic Plasticity. Clin. Cancer Res. 2017, 23, 6292–6304. [Google Scholar] [CrossRef] [Green Version]
  118. Ye, F.; Zhang, Y.; Liu, Y.; Yamada, K.; Tso, J.L.; Menjivar, J.C.; Tian, J.Y.; Yong, W.H.; Schaue, D.; Mischel, P.S.; et al. Protective properties of radio-chemoresistant glioblastoma stem cell clones are associated with metabolic adaptation to reduced glucose dependence. PLoS ONE 2013, 8, e80397. [Google Scholar] [CrossRef] [Green Version]
  119. Garnier, D.; Renoult, O.; Alves-Guerra, M.C.; Paris, F.; Pecqueur, C. Glioblastoma Stem-Like Cells, Metabolic Strategy to Kill a Challenging Target. Front. Oncol. 2019, 9, 118. [Google Scholar] [CrossRef]
  120. Badr, C.E.; Silver, D.J.; Siebzehnrubl, F.A.; Deleyrolle, L.P. Metabolic heterogeneity and adaptability in brain tumors. Cell Mol. Life Sci. 2020, 77, 5101–5119. [Google Scholar] [CrossRef]
  121. Virtuoso, A.; Giovannoni, R.; De Luca, C.; Gargano, F.; Cerasuolo, M.; Maggio, N.; Lavitrano, M.; Papa, M. The Glioblastoma Microenvironment: Morphology, Metabolism, and Molecular Signature of Glial Dynamics to Discover Metabolic Rewiring Sequence. Int. J. Mol. Sci. 2021, 22, 3301. [Google Scholar] [CrossRef] [PubMed]
  122. Polyzos, A.A.; Lee, D.Y.; Datta, R.; Hauser, M.; Budworth, H.; Holt, A.; Mihalik, S.; Goldschmidt, P.; Frankel, K.; Trego, K.; et al. Metabolic Reprogramming in Astrocytes Distinguishes Region-Specific Neuronal Susceptibility in Huntington Mice. Cell Metab. 2019, 29, 1258–1273.e1211. [Google Scholar] [CrossRef] [PubMed]
  123. De Luca, C.; Colangelo, A.M.; Virtuoso, A.; Alberghina, L.; Papa, M. Neurons, Glia, Extracellular Matrix and Neurovascular Unit: A Systems Biology Approach to the Complexity of Synaptic Plasticity in Health and Disease. Int. J. Mol. Sci. 2020, 21, 1539. [Google Scholar] [CrossRef] [Green Version]
  124. Dong, Z.; Cui, H. Epigenetic modulation of metabolism in glioblastoma. Semin. Cancer Biol. 2019, 57, 45–51. [Google Scholar] [CrossRef]
  125. El Khayari, A.; Bouchmaa, N.; Taib, B.; Wei, Z.; Zeng, A.; El Fatimy, R. Metabolic Rewiring in Glioblastoma Cancer: EGFR, IDH and Beyond. Front. Oncol. 2022, 12, 901951. [Google Scholar] [CrossRef] [PubMed]
  126. Quinones, A.; Le, A. The Multifaceted Glioblastoma: From Genomic Alterations to Metabolic Adaptations. Adv. Exp. Med. Biol. 2021, 1311, 59–76. [Google Scholar] [CrossRef] [PubMed]
  127. Flavahan, W.A.; Wu, Q.; Hitomi, M.; Rahim, N.; Kim, Y.; Sloan, A.E.; Weil, R.J.; Nakano, I.; Sarkaria, J.N.; Stringer, B.W.; et al. Brain tumor initiating cells adapt to restricted nutrition through preferential glucose uptake. Nat. Neurosci. 2013, 16, 1373–1382. [Google Scholar] [CrossRef] [PubMed]
  128. Bhowmick, R.; Subramanian, A.; Sarkar, R.R. Exploring the differences in metabolic behavior of astrocyte and glioblastoma: A flux balance analysis approach. Syst. Synth. Biol. 2015, 9, 159–177. [Google Scholar] [CrossRef] [Green Version]
  129. Lin, H.; Muramatsu, R.; Maedera, N.; Tsunematsu, H.; Hamaguchi, M.; Koyama, Y.; Kuroda, M.; Ono, K.; Sawada, M.; Yamashita, T. Extracellular Lactate Dehydrogenase A Release From Damaged Neurons Drives Central Nervous System Angiogenesis. EBioMedicine 2018, 27, 71–85. [Google Scholar] [CrossRef] [Green Version]
  130. Brand, A.; Singer, K.; Koehl, G.E.; Kolitzus, M.; Schoenhammer, G.; Thiel, A.; Matos, C.; Bruss, C.; Klobuch, S.; Peter, K.; et al. LDHA-Associated Lactic Acid Production Blunts Tumor Immunosurveillance by T and NK Cells. Cell Metab. 2016, 24, 657–671. [Google Scholar] [CrossRef] [Green Version]
  131. Toyonaga, T.; Yamaguchi, S.; Hirata, K.; Kobayashi, K.; Manabe, O.; Watanabe, S.; Terasaka, S.; Kobayashi, H.; Hattori, N.; Shiga, T.; et al. Hypoxic glucose metabolism in glioblastoma as a potential prognostic factor. Eur. J. Nucl. Med. Mol. Imaging 2017, 44, 611–619. [Google Scholar] [CrossRef] [Green Version]
  132. Zhang, Y.; Yu, G.; Chu, H.; Wang, X.; Xiong, L.; Cai, G.; Liu, R.; Gao, H.; Tao, B.; Li, W.; et al. Macrophage-Associated PGK1 Phosphorylation Promotes Aerobic Glycolysis and Tumorigenesis. Mol. Cell 2018, 71, 201–215.e207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Babic, I.; Anderson, E.S.; Tanaka, K.; Guo, D.; Masui, K.; Li, B.; Zhu, S.; Gu, Y.; Villa, G.R.; Akhavan, D.; et al. EGFR mutation-induced alternative splicing of Max contributes to growth of glycolytic tumors in brain cancer. Cell Metab. 2013, 17, 1000–1008. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Wolf, A.; Agnihotri, S.; Micallef, J.; Mukherjee, J.; Sabha, N.; Cairns, R.; Hawkins, C.; Guha, A. Hexokinase 2 is a key mediator of aerobic glycolysis and promotes tumor growth in human glioblastoma multiforme. J. Exp. Med. 2011, 208, 313–326. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Guo, D.; Tong, Y.; Jiang, X.; Meng, Y.; Jiang, H.; Du, L.; Wu, Q.; Li, S.; Luo, S.; Li, M.; et al. Aerobic glycolysis promotes tumor immune evasion by hexokinase2-mediated phosphorylation of IkappaBalpha. Cell Metab. 2022, 34, 1312–1324.e1316. [Google Scholar] [CrossRef] [PubMed]
  136. Maddocks, O.D.; Vousden, K.H. Metabolic regulation by p53. J. Mol. Med. 2011, 89, 237–245. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Schwartzenberg-Bar-Yoseph, F.; Armoni, M.; Karnieli, E. The tumor suppressor p53 down-regulates glucose transporters GLUT1 and GLUT4 gene expression. Cancer Res. 2004, 64, 2627–2633. [Google Scholar] [CrossRef] [Green Version]
  138. Struys, E.A. 2-Hydroxyglutarate is not a metabolite; D-2-hydroxyglutarate and L-2-hydroxyglutarate are! Proc. Natl. Acad. Sci. USA 2013, 110, E4939. [Google Scholar] [CrossRef] [Green Version]
  139. Newsholme, P.; Lima, M.M.; Procopio, J.; Pithon-Curi, T.C.; Doi, S.Q.; Bazotte, R.B.; Curi, R. Glutamine and glutamate as vital metabolites. Braz. J. Med. Biol. Res. 2003, 36, 153–163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Nicklin, P.; Bergman, P.; Zhang, B.; Triantafellow, E.; Wang, H.; Nyfeler, B.; Yang, H.; Hild, M.; Kung, C.; Wilson, C.; et al. Bidirectional transport of amino acids regulates mTOR and autophagy. Cell 2009, 136, 521–534. [Google Scholar] [CrossRef] [Green Version]
  141. Marin-Valencia, I.; Yang, C.; Mashimo, T.; Cho, S.; Baek, H.; Yang, X.L.; Rajagopalan, K.N.; Maddie, M.; Vemireddy, V.; Zhao, Z.; et al. Analysis of tumor metabolism reveals mitochondrial glucose oxidation in genetically diverse human glioblastomas in the mouse brain in vivo. Cell Metab. 2012, 15, 827–837. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Rosati, A.; Poliani, P.L.; Todeschini, A.; Cominelli, M.; Medicina, D.; Cenzato, M.; Simoncini, E.L.; Magrini, S.M.; Buglione, M.; Grisanti, S.; et al. Glutamine synthetase expression as a valuable marker of epilepsy and longer survival in newly diagnosed glioblastoma multiforme. Neuro Oncol. 2013, 15, 618–625. [Google Scholar] [CrossRef] [PubMed]
  143. Oppermann, H.; Ding, Y.; Sharma, J.; Berndt Paetz, M.; Meixensberger, J.; Gaunitz, F.; Birkemeyer, C. Metabolic response of glioblastoma cells associated with glucose withdrawal and pyruvate substitution as revealed by GC-MS. Nutr. Metab. 2016, 13, 70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. DeBerardinis, R.J.; Mancuso, A.; Daikhin, E.; Nissim, I.; Yudkoff, M.; Wehrli, S.; Thompson, C.B. Beyond aerobic glycolysis: Transformed cells can engage in glutamine metabolism that exceeds the requirement for protein and nucleotide synthesis. Proc. Natl. Acad. Sci. USA 2007, 104, 19345–19350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Phang, J.M. Proline Metabolism in Cell Regulation and Cancer Biology: Recent Advances and Hypotheses. AntiOxid. Redox Signal 2019, 30, 635–649. [Google Scholar] [CrossRef] [Green Version]
  146. Panosyan, E.H.; Lin, H.J.; Koster, J.; Lasky, J.L., 3rd. In search of druggable targets for GBM amino acid metabolism. BMC Cancer 2017, 17, 162. [Google Scholar] [CrossRef] [Green Version]
  147. Rocha, C.R.; Kajitani, G.S.; Quinet, A.; Fortunato, R.S.; Menck, C.F. NRF2 and glutathione are key resistance mediators to temozolomide in glioma and melanoma cells. Oncotarget 2016, 7, 48081–48092. [Google Scholar] [CrossRef] [Green Version]
  148. Lv, H.; Zhen, C.; Liu, J.; Yang, P.; Hu, L.; Shang, P. Unraveling the Potential Role of Glutathione in Multiple Forms of Cell Death in Cancer Therapy. Oxid. Med. Cell Longev. 2019, 2019, 3150145. [Google Scholar] [CrossRef] [Green Version]
  149. Chinopoulos, C.; Seyfried, T.N. Mitochondrial Substrate-Level Phosphorylation as Energy Source for Glioblastoma: Review and Hypothesis. ASN Neuro 2018, 10, 1759091418818261. [Google Scholar] [CrossRef] [Green Version]
  150. Fahy, E.; Subramaniam, S.; Murphy, R.C.; Nishijima, M.; Raetz, C.R.; Shimizu, T.; Spener, F.; van Meer, G.; Wakelam, M.J.; Dennis, E.A. Update of the LIPID MAPS comprehensive classification system for lipids. J. Lipid Res. 2009, 50, S9–S14. [Google Scholar] [CrossRef] [Green Version]
  151. Taib, B.; Aboussalah, A.M.; Moniruzzaman, M.; Chen, S.; Haughey, N.J.; Kim, S.F.; Ahima, R.S. Lipid accumulation and oxidation in glioblastoma multiforme. Sci. Rep. 2019, 9, 19593. [Google Scholar] [CrossRef] [Green Version]
  152. Ledwozyw, A.; Lutnicki, K. Phospholipids and fatty acids in human brain tumors. Acta Physiol. Hung 1992, 79, 381–387. [Google Scholar] [PubMed]
  153. Guo, D.; Bell, E.H.; Chakravarti, A. Lipid metabolism emerges as a promising target for malignant glioma therapy. CNS Oncol. 2013, 2, 289–299. [Google Scholar] [CrossRef]
  154. Mashimo, T.; Pichumani, K.; Vemireddy, V.; Hatanpaa, K.J.; Singh, D.K.; Sirasanagandla, S.; Nannepaga, S.; Piccirillo, S.G.; Kovacs, Z.; Foong, C.; et al. Acetate is a bioenergetic substrate for human glioblastoma and brain metastases. Cell 2014, 159, 1603–1614. [Google Scholar] [CrossRef] [Green Version]
  155. Park, J.K.; Coffey, N.J.; Limoges, A.; Le, A. The Heterogeneity of Lipid Metabolism in Cancer. Adv. Exp. Med. Biol. 2021, 1311, 39–56. [Google Scholar] [CrossRef]
  156. Guo, D.; Prins, R.M.; Dang, J.; Kuga, D.; Iwanami, A.; Soto, H.; Lin, K.Y.; Huang, T.T.; Akhavan, D.; Hock, M.B.; et al. EGFR signaling through an Akt-SREBP-1-dependent, rapamycin-resistant pathway sensitizes glioblastomas to antilipogenic therapy. Sci. Signal. 2009, 2, ra82. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Duman, C.; Yaqubi, K.; Hoffmann, A.; Acikgoz, A.A.; Korshunov, A.; Bendszus, M.; Herold-Mende, C.; Liu, H.K.; Alfonso, J. Acyl-CoA-Binding Protein Drives Glioblastoma Tumorigenesis by Sustaining Fatty Acid Oxidation. Cell Metab. 2019, 30, 274–289.e275. [Google Scholar] [CrossRef] [PubMed]
  158. Huang, R.; Dong, R.; Wang, N.; He, Y.; Zhu, P.; Wang, C.; Lan, B.; Gao, Y.; Sun, L. Adaptive Changes Allow Targeting of Ferroptosis for Glioma Treatment. Cell Mol. Neurobiol. 2022, 42, 2055–2074. [Google Scholar] [CrossRef] [PubMed]
  159. Cheng, X.; Geng, F.; Guo, D. DGAT1 protects tumor from lipotoxicity, emerging as a promising metabolic target for cancer therapy. Mol. Cell Oncol. 2020, 7, 1805257. [Google Scholar] [CrossRef] [PubMed]
  160. Lautenberger, J.A.; Schulz, R.A.; Garon, C.F.; Tsichlis, P.N.; Papas, T.S. Molecular cloning of avian myelocytomatosis virus (MC29) transforming sequences. Proc. Natl. Acad. Sci. USA 1981, 78, 1518–1522. [Google Scholar] [CrossRef] [Green Version]
  161. 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] [PubMed]
  162. Dalla-Favera, R.; Wong-Staal, F.; Gallo, R.C. Onc gene amplification in promyelocytic leukaemia cell line HL-60 and primary leukaemic cells of the same patient. Nature 1982, 299, 61–63. [Google Scholar] [CrossRef]
  163. Crews, S.; Barth, R.; Hood, L.; Prehn, J.; Calame, K. Mouse c-myc oncogene is located on chromosome 15 and translocated to chromosome 12 in plasmacytomas. Science 1982, 218, 1319–1321. [Google Scholar] [CrossRef] [PubMed]
  164. Steffen, D. Proviruses are adjacent to c-myc in some murine leukemia virus-induced lymphomas. Proc. Natl. Acad. Sci. USA 1984, 81, 2097–2101. [Google Scholar] [CrossRef] [Green Version]
  165. 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] [PubMed]
  166. 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]
  167. Baluapuri, A.; Wolf, E.; Eilers, M. Target gene-independent functions of MYC oncoproteins. Nat. Rev. Mol. Cell Biol. 2020, 21, 255–267. [Google Scholar] [CrossRef]
  168. 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]
  169. 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]
  170. 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]
  171. Cole, M.D. Myc meets its Max. Cell 1991, 65, 715–716. [Google Scholar] [CrossRef] [PubMed]
  172. 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] [PubMed]
  173. Grandori, C.; Eisenman, R.N. Myc target genes. Trends Biochem. Sci. 1997, 22, 177–181. [Google Scholar] [CrossRef] [PubMed]
  174. 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]
  175. Yang, G.; Hurlin, P.J. MNT and Emerging Concepts of MNT-MYC Antagonism. Genes 2017, 8, 83. [Google Scholar] [CrossRef]
  176. Link, J.M.; Ota, S.; Zhou, Z.Q.; Daniel, C.J.; Sears, R.C.; Hurlin, P.J. A critical role for Mnt in Myc-driven T-cell proliferation and oncogenesis. Proc. Natl. Acad. Sci. USA 2012, 109, 19685–19690. [Google Scholar] [CrossRef] [Green Version]
  177. Nair, S.K.; Burley, S.K. Structural aspects of interactions within the Myc/Max/Mad network. Curr. Top Microbiol. Immunol. 2006, 302, 123–143. [Google Scholar] [CrossRef]
  178. Cowling, V.H.; Cole, M.D. Mechanism of transcriptional activation by the Myc oncoproteins. Semin. Cancer Biol. 2006, 16, 242–252. [Google Scholar] [CrossRef]
  179. Sears, R.; Nuckolls, F.; Haura, E.; Taya, Y.; Tamai, K.; Nevins, J.R. Multiple Ras-dependent phosphorylation pathways regulate Myc protein stability. Genes Dev. 2000, 14, 2501–2514. [Google Scholar] [CrossRef] [Green Version]
  180. Jiang, J.; Wang, J.; Yue, M.; Cai, X.; Wang, T.; Wu, C.; Su, H.; Wang, Y.; Han, M.; Zhang, Y.; et al. Direct Phosphorylation and Stabilization of MYC by Aurora B Kinase Promote T-cell Leukemogenesis. Cancer Cell 2020, 37, 200–215.e205. [Google Scholar] [CrossRef]
  181. Kim, S.Y.; Herbst, A.; Tworkowski, K.A.; Salghetti, S.E.; Tansey, W.P. Skp2 regulates Myc protein stability and activity. Mol. Cell 2003, 11, 1177–1188. [Google Scholar] [CrossRef]
  182. von der Lehr, N.; Johansson, S.; Wu, S.; Bahram, F.; Castell, A.; Cetinkaya, C.; Hydbring, P.; Weidung, I.; Nakayama, K.; Nakayama, K.I.; et al. The F-box protein Skp2 participates in c-Myc proteosomal degradation and acts as a cofactor for c-Myc-regulated transcription. Mol. Cell 2003, 11, 1189–1200. [Google Scholar] [CrossRef] [PubMed]
  183. Adhikary, S.; Marinoni, F.; Hock, A.; Hulleman, E.; Popov, N.; Beier, R.; Bernard, S.; Quarto, M.; Capra, M.; Goettig, S.; et al. The ubiquitin ligase HectH9 regulates transcriptional activation by Myc and is essential for tumor cell proliferation. Cell 2005, 123, 409–421. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Patel, J.H.; Du, Y.; Ard, P.G.; Phillips, C.; Carella, B.; Chen, C.J.; Rakowski, C.; Chatterjee, C.; Lieberman, P.M.; Lane, W.S.; et al. The c-MYC oncoprotein is a substrate of the acetyltransferases hGCN5/PCAF and TIP60. Mol. Cell Biol. 2004, 24, 10826–10834. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Faiola, F.; Liu, X.; Lo, S.; Pan, S.; Zhang, K.; Lymar, E.; Farina, A.; Martinez, E. Dual regulation of c-Myc by p300 via acetylation-dependent control of Myc protein turnover and coactivation of Myc-induced transcription. Mol. Cell Biol. 2005, 25, 10220–10234. [Google Scholar] [CrossRef] [Green Version]
  186. Zhang, K.; Faiola, F.; Martinez, E. Six lysine residues on c-Myc are direct substrates for acetylation by p300. Biochem. Biophys. Res. Commun. 2005, 336, 274–280. [Google Scholar] [CrossRef]
  187. Blanc, R.S.; Richard, S. Arginine Methylation: The Coming of Age. Mol. Cell 2017, 65, 8–24. [Google Scholar] [CrossRef] [Green Version]
  188. Tikhanovich, I.; Zhao, J.; Bridges, B.; Kumer, S.; Roberts, B.; Weinman, S.A. Arginine methylation regulates c-Myc-dependent transcription by altering promoter recruitment of the acetyltransferase p300. J. Biol. Chem. 2017, 292, 13333–13344. [Google Scholar] [CrossRef] [Green Version]
  189. Favia, A.; Salvatori, L.; Nanni, S.; Iwamoto-Stohl, L.K.; Valente, S.; Mai, A.; Scagnoli, F.; Fontanella, R.A.; Totta, P.; Nasi, S.; et al. The Protein Arginine Methyltransferases 1 and 5 affect Myc properties in glioblastoma stem cells. Sci. Rep. 2019, 9, 15925. [Google Scholar] [CrossRef] [Green Version]
  190. 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] [PubMed] [Green Version]
  191. 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]
  192. 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] [PubMed] [Green Version]
  193. Lorenzin, F.; Benary, U.; Baluapuri, A.; Walz, S.; Jung, L.A.; von Eyss, B.; Kisker, C.; Wolf, J.; Eilers, M.; Wolf, E. Different promoter affinities account for specificity in MYC-dependent gene regulation. eLife 2016, 5, e15161. [Google Scholar] [CrossRef]
  194. 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] [PubMed]
  195. 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]
  196. Psathas, J.N.; Thomas-Tikhonenko, A. MYC and the art of microRNA maintenance. Cold Spring Harb. Perspect. Med. 2014, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. 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] [PubMed] [Green Version]
  198. 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] [PubMed]
  199. 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] [PubMed]
  200. Zindy, F.; Eischen, C.M.; Randle, D.H.; Kamijo, T.; Cleveland, J.L.; Sherr, C.J.; Roussel, M.F. Myc signaling via the ARF tumor suppressor regulates p53-dependent apoptosis and immortalization. Genes Dev. 1998, 12, 2424–2433. [Google Scholar] [CrossRef] [Green Version]
  201. 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]
  202. McMahon, S.B. MYC and the control of apoptosis. Cold Spring Harb. Perspect. Med. 2014, 4, a014407. [Google Scholar] [CrossRef] [Green Version]
  203. Hermeking, H.; Eick, D. Mediation of c-Myc-induced apoptosis by p53. Science 1994, 265, 2091–2093. [Google Scholar] [CrossRef]
  204. Herold, S.; Wanzel, M.; Beuger, V.; Frohme, C.; Beul, D.; Hillukkala, T.; Syvaoja, J.; Saluz, H.P.; Haenel, F.; Eilers, M. Negative regulation of the mammalian UV response by Myc through association with Miz-1. Mol. Cell 2002, 10, 509–521. [Google Scholar] [CrossRef] [PubMed]
  205. Gartel, A.L.; Ye, X.; Goufman, E.; Shianov, P.; Hay, N.; Najmabadi, F.; Tyner, A.L. Myc represses the p21(WAF1/CIP1) promoter and interacts with Sp1/Sp3. Proc. Natl. Acad. Sci. USA 2001, 98, 4510–4515. [Google Scholar] [CrossRef] [Green Version]
  206. Eischen, C.M.; Packham, G.; Nip, J.; Fee, B.E.; Hiebert, S.W.; Zambetti, G.P.; Cleveland, J.L. Bcl-2 is an apoptotic target suppressed by both c-Myc and E2F-1. Oncogene 2001, 20, 6983–6993. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Maclean, K.H.; Keller, U.B.; Rodriguez-Galindo, C.; Nilsson, J.A.; Cleveland, J.L. c-Myc augments gamma irradiation-induced apoptosis by suppressing Bcl-XL. Mol. Cell Biol. 2003, 23, 7256–7270. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Dansen, T.B.; Whitfield, J.; Rostker, F.; Brown-Swigart, L.; Evan, G.I. Specific requirement for Bax, not Bak, in Myc-induced apoptosis and tumor suppression in vivo. J. Biol. Chem. 2006, 281, 10890–10895. [Google Scholar] [CrossRef] [Green Version]
  209. Eischen, C.M.; Roussel, M.F.; Korsmeyer, S.J.; Cleveland, J.L. Bax loss impairs Myc-induced apoptosis and circumvents the selection of p53 mutations during Myc-mediated lymphomagenesis. Mol. Cell Biol. 2001, 21, 7653–7662. [Google Scholar] [CrossRef] [Green Version]
  210. Iaccarino, I.; Hancock, D.; Evan, G.; Downward, J. c-Myc induces cytochrome c release in Rat1 fibroblasts by increasing outer mitochondrial membrane permeability in a Bid-dependent manner. Cell Death Differ. 2003, 10, 599–608. [Google Scholar] [CrossRef]
  211. Cao, X.; Bennett, R.L.; May, W.S. c-Myc and caspase-2 are involved in activating Bax during cytotoxic drug-induced apoptosis. J. Biol. Chem. 2008, 283, 14490–14496. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Zhuang, D.; Mannava, S.; Grachtchouk, V.; Tang, W.H.; Patil, S.; Wawrzyniak, J.A.; Berman, A.E.; Giordano, T.J.; Prochownik, E.V.; Soengas, M.S.; et al. C-MYC overexpression is required for continuous suppression of oncogene-induced senescence in melanoma cells. Oncogene 2008, 27, 6623–6634. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Campaner, S.; Doni, M.; Hydbring, P.; Verrecchia, A.; Bianchi, L.; Sardella, D.; Schleker, T.; Perna, D.; Tronnersjo, S.; Murga, M.; et al. Cdk2 suppresses cellular senescence induced by the c-myc oncogene. Nat. Cell Biol. 2010, 12, 54–59. [Google Scholar] [CrossRef]
  214. Katanasaka, Y.; Kodera, Y.; Kitamura, Y.; Morimoto, T.; Tamura, T.; Koizumi, F. Epidermal growth factor receptor variant type III markedly accelerates angiogenesis and tumor growth via inducing c-myc mediated angiopoietin-like 4 expression in malignant glioma. Mol. Cancer 2013, 12, 31. [Google Scholar] [CrossRef] [Green Version]
  215. Lee, D.H.; Kim, G.W.; Yoo, J.; Lee, S.W.; Jeon, Y.H.; Kim, S.Y.; Kang, H.G.; Kim, D.H.; Chun, K.H.; Choi, J.; et al. Histone demethylase KDM4C controls tumorigenesis of glioblastoma by epigenetically regulating p53 and c-Myc. Cell Death Dis. 2021, 12, 89. [Google Scholar] [CrossRef]
  216. Ke, X.X.; Zhang, R.; Zhong, X.; Zhang, L.; Cui, H. Deficiency of G9a Inhibits Cell Proliferation and Activates Autophagy via Transcriptionally Regulating c-Myc Expression in Glioblastoma. Front. Cell Dev. Biol. 2020, 8, 593964. [Google Scholar] [CrossRef] [PubMed]
  217. Hou, J.; Liu, Y.; Huang, P.; Wang, Y.; Pei, D.; Tan, R.; Zhang, Y.; Cui, H. RANBP10 promotes glioblastoma progression by regulating the FBXW7/c-Myc pathway. Cell Death Dis. 2021, 12, 967. [Google Scholar] [CrossRef]
  218. Wang, F.; Zhang, J.; Ke, X.; Peng, W.; Zhao, G.; Peng, S.; Xu, J.; Xu, B.; Cui, H. WDR5-Myc axis promotes the progression of glioblastoma and neuroblastoma by transcriptional activating CARM1. Biochem. Biophys. Res. Commun. 2020, 523, 699–706. [Google Scholar] [CrossRef]
  219. Liu, N.; Wang, Z.; Liu, D.; Xie, P. HOXC13-AS-miR-122-5p-SATB1-C-Myc feedback loop promotes migration, invasion and EMT process in glioma. Onco Targets Ther. 2019, 12, 7165–7173. [Google Scholar] [CrossRef] [Green Version]
  220. Huang, L.; Hu, C.; Chao, H.; Wang, R.; Lu, H.; Li, H.; Chen, H. miR-29c regulates resistance to paclitaxel in nasopharyngeal cancer by targeting ITGB1. Exp. Cell Res. 2019, 378, 1–10. [Google Scholar] [CrossRef]
  221. Pyko, I.V.; Nakada, M.; Sabit, H.; Teng, L.; Furuyama, N.; Hayashi, Y.; Kawakami, K.; Minamoto, T.; Fedulau, A.S.; Hamada, J. Glycogen synthase kinase 3beta inhibition sensitizes human glioblastoma cells to temozolomide by affecting O6-methylguanine DNA methyltransferase promoter methylation via c-Myc signaling. Carcinogenesis 2013, 34, 2206–2217. [Google Scholar] [CrossRef] [Green Version]
  222. Wang, J.; Li, Y.; Xu, B.; Dong, J.; Zhao, H.; Zhao, D.; Wu, Y. ALYREF Drives Cancer Cell Proliferation Through an ALYREF-MYC Positive Feedback Loop in Glioblastoma. Onco Targets Ther. 2021, 14, 145–155. [Google Scholar] [CrossRef] [PubMed]
  223. Chen, X.; Xu, H.; Yuan, P.; Fang, F.; Huss, M.; Vega, V.B.; Wong, E.; Orlov, Y.L.; Zhang, W.; Jiang, J.; et al. Integration of external signaling pathways with the core transcriptional network in embryonic stem cells. Cell 2008, 133, 1106–1117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. 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]
  225. Zhang, G.; Dong, Z.; Prager, B.C.; Kim, L.J.; Wu, Q.; Gimple, R.C.; Wang, X.; Bao, S.; Hamerlik, P.; Rich, J.N. Chromatin remodeler HELLS maintains glioma stem cells through E2F3 and MYC. JCI Insight 2019, 4, e126140. [Google Scholar] [CrossRef] [Green Version]
  226. Dixit, D.; Prager, B.C.; Gimple, R.C.; Poh, H.X.; Wang, Y.; Wu, Q.; Qiu, Z.; Kidwell, R.L.; Kim, L.J.Y.; Xie, Q.; et al. The RNA m6A Reader YTHDF2 Maintains Oncogene Expression and Is a Targetable Dependency in Glioblastoma Stem Cells. Cancer Discov. 2021, 11, 480–499. [Google Scholar] [CrossRef] [PubMed]
  227. Huang, H.; Yu, X.; Han, X.; Hao, J.; Zhao, J.; Bebek, G.; Bao, S.; Prayson, R.A.; Khalil, A.M.; Jankowsky, E.; et al. Piwil1 Regulates Glioma Stem Cell Maintenance and Glioblastoma Progression. Cell Rep. 2021, 34, 108522. [Google Scholar] [CrossRef] [PubMed]
  228. Fang, X.; Zhou, W.; Wu, Q.; Huang, Z.; Shi, Y.; Yang, K.; Chen, C.; Xie, Q.; Mack, S.C.; Wang, X.; et al. Deubiquitinase USP13 maintains glioblastoma stem cells by antagonizing FBXL14-mediated Myc ubiquitination. J. Exp. Med. 2017, 214, 245–267. [Google Scholar] [CrossRef] [Green Version]
  229. Fukasawa, K.; Kadota, T.; Horie, T.; Tokumura, K.; Terada, R.; Kitaguchi, Y.; Park, G.; Ochiai, S.; Iwahashi, S.; Okayama, Y.; et al. CDK8 maintains stemness and tumorigenicity of glioma stem cells by regulating the c-MYC pathway. Oncogene 2021, 40, 2803–2815. [Google Scholar] [CrossRef] [PubMed]
  230. Li, D.; Zhang, Q.; Li, L.; Chen, K.; Yang, J.; Dixit, D.; Gimple, R.C.; Ci, S.; Lu, C.; Hu, L.; et al. beta2-Microglobulin Maintains Glioblastoma Stem Cells and Induces M2-like Polarization of Tumor-Associated Macrophages. Cancer Res. 2022, 82, 3321–3334. [Google Scholar] [CrossRef] [PubMed]
  231. Chen, H.; Liu, H.; Qing, G. Targeting oncogenic Myc as a strategy for cancer treatment. Signal. Transduct. Target. Ther. 2018, 3, 5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Llombart, V.; Mansour, M.R. Therapeutic targeting of “undruggable” MYC. EBioMedicine 2022, 75, 103756. [Google Scholar] [CrossRef] [PubMed]
  233. Le Rhun, E.; von Achenbach, C.; Lohmann, B.; Silginer, M.; Schneider, H.; Meetze, K.; Szabo, E.; Weller, M. Profound, durable and MGMT-independent sensitivity of glioblastoma cells to cyclin-dependent kinase inhibition. Int. J. Cancer 2019, 145, 242–253. [Google Scholar] [CrossRef]
  234. Devaiah, B.N.; Mu, J.; Akman, B.; Uppal, S.; Weissman, J.D.; Cheng, D.; Baranello, L.; Nie, Z.; Levens, D.; Singer, D.S. MYC protein stability is negatively regulated by BRD4. Proc. Natl. Acad. Sci. USA 2020, 117, 13457–13467. [Google Scholar] [CrossRef] [PubMed]
  235. 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]
  236. Kato, F.; Fiorentino, F.P.; Alibes, A.; Perucho, M.; Sanchez-Cespedes, M.; Kohno, T.; Yokota, J. MYCL is a target of a BET bromodomain inhibitor, JQ1, on growth suppression efficacy in small cell lung cancer cells. Oncotarget 2016, 7, 77378–77388. [Google Scholar] [CrossRef] [Green Version]
  237. Zhou, S.; Zhang, S.; Wang, L.; Huang, S.; Yuan, Y.; Yang, J.; Wang, H.; Li, X.; Wang, P.; Zhou, L.; et al. BET protein inhibitor JQ1 downregulates chromatin accessibility and suppresses metastasis of gastric cancer via inactivating RUNX2/NID1 signaling. Oncogenesis 2020, 9, 33. [Google Scholar] [CrossRef] [Green Version]
  238. Falchook, G.; Rosen, S.; LoRusso, P.; Watts, J.; Gupta, S.; Coombs, C.C.; Talpaz, M.; Kurzrock, R.; Mita, M.; Cassaday, R.; et al. Development of 2 Bromodomain and Extraterminal Inhibitors With Distinct Pharmacokinetic and Pharmacodynamic Profiles for the Treatment of Advanced Malignancies. Clin. Cancer Res. 2020, 26, 1247–1257. [Google Scholar] [CrossRef] [Green Version]
  239. Leal, A.S.; Liu, P.; Krieger-Burke, T.; Ruggeri, B.; Liby, K.T. The Bromodomain Inhibitor, INCB057643, Targets Both Cancer Cells and the Tumor Microenvironment in Two Preclinical Models of Pancreatic Cancer. Cancers 2020, 13, 96. [Google Scholar] [CrossRef]
  240. Cole, M.D.; Nikiforov, M.A. Transcriptional activation by the Myc oncoprotein. Curr. Top. Microbiol. Immunol. 2006, 302, 33–50. [Google Scholar] [CrossRef]
  241. Wall, M.; Poortinga, G.; Hannan, K.M.; Pearson, R.B.; Hannan, R.D.; McArthur, G.A. Translational control of c-MYC by rapamycin promotes terminal myeloid differentiation. Blood 2008, 112, 2305–2317. [Google Scholar] [CrossRef] [Green Version]
  242. Cianfanelli, V.; Fuoco, C.; Lorente, M.; Salazar, M.; Quondamatteo, F.; Gherardini, P.F.; De Zio, D.; Nazio, F.; Antonioli, M.; D’Orazio, M.; et al. AMBRA1 links autophagy to cell proliferation and tumorigenesis by promoting c-Myc dephosphorylation and degradation. Nat. Cell Biol. 2015, 17, 20–30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Delpuech, O.; Griffiths, B.; East, P.; Essafi, A.; Lam, E.W.; Burgering, B.; Downward, J.; Schulze, A. Induction of Mxi1-SR alpha by FOXO3a contributes to repression of Myc-dependent gene expression. Mol. Cell Biol. 2007, 27, 4917–4930. [Google Scholar] [CrossRef] [Green Version]
  244. Jensen, K.S.; Binderup, T.; Jensen, K.T.; Therkelsen, I.; Borup, R.; Nilsson, E.; Multhaupt, H.; Bouchard, C.; Quistorff, B.; Kjaer, A.; et al. FoxO3A promotes metabolic adaptation to hypoxia by antagonizing Myc function. EMBO J. 2011, 30, 4554–4570. [Google Scholar] [CrossRef] [PubMed]
  245. Okuyama, H.; Endo, H.; Akashika, T.; Kato, K.; Inoue, M. Downregulation of c-MYC protein levels contributes to cancer cell survival under dual deficiency of oxygen and glucose. Cancer Res. 2010, 70, 10213–10223. [Google Scholar] [CrossRef] [Green Version]
  246. Gordan, J.D.; Thompson, C.B.; Simon, M.C. HIF and c-Myc: Sibling rivals for control of cancer cell metabolism and proliferation. Cancer Cell 2007, 12, 108–113. [Google Scholar] [CrossRef] [Green Version]
  247. 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] [PubMed] [Green Version]
  248. 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]
  249. 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] [PubMed] [Green Version]
  250. Sabo, A.; Kress, T.R.; Pelizzola, M.; de Pretis, S.; Gorski, M.M.; Tesi, A.; Morelli, M.J.; Bora, P.; Doni, M.; Verrecchia, A.; et al. Selective transcriptional regulation by Myc in cellular growth control and lymphomagenesis. Nature 2014, 511, 488–492. [Google Scholar] [CrossRef] [Green Version]
  251. Kim, J.W.; Zeller, K.I.; Wang, Y.; Jegga, A.G.; Aronow, B.J.; O’Donnell, K.A.; Dang, C.V. Evaluation of myc E-box phylogenetic footprints in glycolytic genes by chromatin immunoprecipitation assays. Mol. Cell Biol. 2004, 24, 5923–5936. [Google Scholar] [CrossRef] [Green Version]
  252. 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]
  253. Venkateswaran, N.; Lafita-Navarro, M.C.; Hao, Y.H.; Kilgore, J.A.; Perez-Castro, L.; Braverman, J.; Borenstein-Auerbach, N.; Kim, M.; Lesner, N.P.; Mishra, P.; et al. MYC promotes tryptophan uptake and metabolism by the kynurenine pathway in colon cancer. Genes Dev. 2019, 33, 1236–1251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  254. David, C.J.; Chen, M.; Assanah, M.; Canoll, P.; Manley, J.L. HnRNP proteins controlled by c-Myc deregulate pyruvate kinase mRNA splicing in cancer. Nature 2010, 463, 364–368. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Kaadige, M.R.; Elgort, M.G.; Ayer, D.E. Coordination of glucose and glutamine utilization by an expanded Myc network. Transcription 2010, 1, 36–40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Wang, R.; Dillon, C.P.; Shi, L.Z.; Milasta, S.; Carter, R.; Finkelstein, D.; McCormick, L.L.; Fitzgerald, P.; Chi, H.; Munger, J.; et al. The transcription factor Myc controls metabolic reprogramming upon T lymphocyte activation. Immunity 2011, 35, 871–882. [Google Scholar] [CrossRef] [Green Version]
  257. Morrish, F.; Isern, N.; Sadilek, M.; Jeffrey, M.; Hockenbery, D.M. c-Myc activates multiple metabolic networks to generate substrates for cell-cycle entry. Oncogene 2009, 28, 2485–2491. [Google Scholar] [CrossRef] [Green Version]
  258. Locasale, J.W. Serine, glycine and one-carbon units: Cancer metabolism in full circle. Nat. Rev. Cancer 2013, 13, 572–583. [Google Scholar] [CrossRef] [Green Version]
  259. Agarwal, S.; Chakravarthi, B.; Behring, M.; Kim, H.G.; Chandrashekar, D.S.; Gupta, N.; Bajpai, P.; Elkholy, A.; Balasubramanya, S.A.H.; Hardy, C.; et al. PAICS, a Purine Nucleotide Metabolic Enzyme, is Involved in Tumor Growth and the Metastasis of Colorectal Cancer. Cancers 2020, 12, 772. [Google Scholar] [CrossRef] [Green Version]
  260. Liu, Y.C.; Li, F.; Handler, J.; Huang, C.R.; Xiang, Y.; Neretti, N.; Sedivy, J.M.; Zeller, K.I.; Dang, C.V. Global regulation of nucleotide biosynthetic genes by c-Myc. PLoS ONE 2008, 3, e2722. [Google Scholar] [CrossRef]
  261. Cunningham, J.T.; Moreno, M.V.; Lodi, A.; Ronen, S.M.; Ruggero, D. Protein and nucleotide biosynthesis are coupled by a single rate-limiting enzyme, PRPS2, to drive cancer. Cell 2014, 157, 1088–1103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  262. Liu, T.; Han, Y.; Yu, C.; Ji, Y.; Wang, C.; Chen, X.; Wang, X.; Shen, J.; Zhang, Y.; Lang, J.Y. MYC predetermines the sensitivity of gastrointestinal cancer to antifolate drugs through regulating TYMS transcription. EBioMedicine 2019, 48, 289–300. [Google Scholar] [CrossRef] [Green Version]
  263. Yue, M.; Jiang, J.; Gao, P.; Liu, H.; Qing, G. Oncogenic MYC Activates a Feedforward Regulatory Loop Promoting Essential Amino Acid Metabolism and Tumorigenesis. Cell Rep. 2017, 21, 3819–3832. [Google Scholar] [CrossRef] [Green Version]
  264. Hattori, A.; Tsunoda, M.; Konuma, T.; Kobayashi, M.; Nagy, T.; Glushka, J.; Tayyari, F.; McSkimming, D.; Kannan, N.; Tojo, A.; et al. Cancer progression by reprogrammed BCAA metabolism in myeloid leukaemia. Nature 2017, 545, 500–504. [Google Scholar] [CrossRef] [Green Version]
  265. Cervenka, I.; Agudelo, L.Z.; Ruas, J.L. Kynurenines: Tryptophan’s metabolites in exercise, inflammation, and mental health. Science 2017, 357. [Google Scholar] [CrossRef] [Green Version]
  266. Gouw, A.M.; Margulis, K.; Liu, N.S.; Raman, S.J.; Mancuso, A.; Toal, G.G.; Tong, L.; Mosley, A.; Hsieh, A.L.; Sullivan, D.K.; et al. The MYC Oncogene Cooperates with Sterol-Regulated Element-Binding Protein to Regulate Lipogenesis Essential for Neoplastic Growth. Cell Metab. 2019, 30, 556–572.e555. [Google Scholar] [CrossRef]
  267. Priolo, C.; Pyne, S.; Rose, J.; Regan, E.R.; Zadra, G.; Photopoulos, C.; Cacciatore, S.; Schultz, D.; Scaglia, N.; McDunn, J.; et al. AKT1 and MYC induce distinctive metabolic fingerprints in human prostate cancer. Cancer Res. 2014, 74, 7198–7204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  268. Morrish, F.; Noonan, J.; Perez-Olsen, C.; Gafken, P.R.; Fitzgibbon, M.; Kelleher, J.; VanGilst, M.; Hockenbery, D. Myc-dependent mitochondrial generation of acetyl-CoA contributes to fatty acid biosynthesis and histone acetylation during cell cycle entry. J. Biol. Chem. 2010, 285, 36267–36274. [Google Scholar] [CrossRef] [Green Version]
  269. 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] [PubMed] [Green Version]
  270. Carroll, P.A.; Diolaiti, D.; McFerrin, L.; Gu, H.; Djukovic, D.; Du, J.; Cheng, P.F.; Anderson, S.; Ulrich, M.; Hurley, J.B.; et al. Deregulated Myc requires MondoA/Mlx for metabolic reprogramming and tumorigenesis. Cancer Cell 2015, 27, 271–285. [Google Scholar] [CrossRef] [Green Version]
  271. Zhong, C.; Fan, L.; Yao, F.; Shi, J.; Fang, W.; Zhao, H. HMGCR is necessary for the tumorigenecity of esophageal squamous cell carcinoma and is regulated by Myc. Tumour Biol. 2014, 35, 4123–4129. [Google Scholar] [CrossRef] [PubMed]
  272. Cao, Z.; Fan-Minogue, H.; Bellovin, D.I.; Yevtodiyenko, A.; Arzeno, J.; Yang, Q.; Gambhir, S.S.; Felsher, D.W. MYC phosphorylation, activation, and tumorigenic potential in hepatocellular carcinoma are regulated by HMG-CoA reductase. Cancer Res. 2011, 71, 2286–2297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  273. Camarda, R.; Zhou, A.Y.; Kohnz, R.A.; Balakrishnan, S.; Mahieu, C.; Anderton, B.; Eyob, H.; Kajimura, S.; Tward, A.; Krings, G.; et al. Inhibition of fatty acid oxidation as a therapy for MYC-overexpressing triple-negative breast cancer. Nat. Med. 2016, 22, 427–432. [Google Scholar] [CrossRef]
  274. Casciano, J.C.; Perry, C.; Cohen-Nowak, A.J.; Miller, K.D.; Vande Voorde, J.; Zhang, Q.; Chalmers, S.; Sandison, M.E.; Liu, Q.; Hedley, A.; et al. MYC regulates fatty acid metabolism through a multigenic program in claudin-low triple negative breast cancer. Br. J. Cancer 2020, 122, 868–884. [Google Scholar] [CrossRef] [Green Version]
  275. Mossmann, D.; Park, S.; Hall, M.N. mTOR signalling and cellular metabolism are mutual determinants in cancer. Nat. Rev. Cancer 2018, 18, 744–757. [Google Scholar] [CrossRef]
  276. DeHaven, J.E.; Robinson, K.A.; Nelson, B.A.; Buse, M.G. A novel variant of glutamine: Fructose-6-phosphate amidotransferase-1 (GFAT1) mRNA is selectively expressed in striated muscle. Diabetes 2001, 50, 2419–2424. [Google Scholar] [CrossRef]
  277. Liu, B.; Huang, Z.B.; Chen, X.; See, Y.X.; Chen, Z.K.; Yao, H.K. Mammalian Target of Rapamycin 2 (MTOR2) and C-MYC Modulate Glucosamine-6-Phosphate Synthesis in Glioblastoma (GBM) Cells Through Glutamine: Fructose-6-Phosphate Aminotransferase 1 (GFAT1). Cell Mol. Neurobiol. 2019, 39, 415–434. [Google Scholar] [CrossRef] [PubMed]
  278. Masui, K.; Tanaka, K.; Akhavan, D.; Babic, I.; Gini, B.; Matsutani, T.; Iwanami, A.; Liu, F.; Villa, G.R.; Gu, Y.; et al. mTOR complex 2 controls glycolytic metabolism in glioblastoma through FoxO acetylation and upregulation of c-Myc. Cell Metab. 2013, 18, 726–739. [Google Scholar] [CrossRef] [Green Version]
  279. Nguyen, T.T.T.; Shang, E.; Shu, C.; Kim, S.; Mela, A.; Humala, N.; Mahajan, A.; Yang, H.W.; Akman, H.O.; Quinzii, C.M.; et al. Aurora kinase A inhibition reverses the Warburg effect and elicits unique metabolic vulnerabilities in glioblastoma. Nat. Commun. 2021, 12, 5203. [Google Scholar] [CrossRef]
  280. Yang, W.; Xia, Y.; Hawke, D.; Li, X.; Liang, J.; Xing, D.; Aldape, K.; Hunter, T.; Alfred Yung, W.K.; Lu, Z. PKM2 phosphorylates histone H3 and promotes gene transcription and tumorigenesis. Cell 2012, 150, 685–696. [Google Scholar] [CrossRef] [Green Version]
  281. Lee, J.H.; Shao, F.; Ling, J.; Lu, S.; Liu, R.; Du, L.; Chung, J.W.; Koh, S.S.; Leem, S.H.; Shao, J.; et al. Phosphofructokinase 1 Platelet Isoform Promotes beta-Catenin Transactivation for Tumor Development. Front. Oncol. 2020, 10, 211. [Google Scholar] [CrossRef]
  282. Haynes, H.R.; Scott, H.L.; Killick-Cole, C.L.; Shaw, G.; Brend, T.; Hares, K.M.; Redondo, J.; Kemp, K.C.; Ballesteros, L.S.; Herman, A.; et al. shRNA-mediated PPARalpha knockdown in human glioma stem cells reduces in vitro proliferation and inhibits orthotopic xenograft tumour growth. J. Pathol. 2019, 247, 422–434. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  283. Li, J.; Liu, Q.; Liu, Z.; Xia, Q.; Zhang, Z.; Zhang, R.; Gao, T.; Gu, G.; Wang, Y.; Wang, D.; et al. KPNA2 promotes metabolic reprogramming in glioblastomas by regulation of c-myc. J. Exp. Clin. Cancer Res. 2018, 37, 194. [Google Scholar] [CrossRef] [PubMed]
  284. Rothhammer, V.; Quintana, F.J. The aryl hydrocarbon receptor: An environmental sensor integrating immune responses in health and disease. Nat. Rev. Immunol. 2019, 19, 184–197. [Google Scholar] [CrossRef]
  285. Denison, M.S.; Nagy, S.R. Activation of the aryl hydrocarbon receptor by structurally diverse exogenous and endogenous chemicals. Annu. Rev. Pharmacol. Toxicol. 2003, 43, 309–334. [Google Scholar] [CrossRef] [PubMed]
  286. Lafita-Navarro, M.C.; Perez-Castro, L.; Zacharias, L.G.; Barnes, S.; DeBerardinis, R.J.; Conacci-Sorrell, M. The transcription factors aryl hydrocarbon receptor and MYC cooperate in the regulation of cellular metabolism. J. Biol. Chem. 2020, 295, 12398–12407. [Google Scholar] [CrossRef]
  287. Mair, R.; Wright, A.J.; Ros, S.; Hu, D.E.; Booth, T.; Kreis, F.; Rao, J.; Watts, C.; Brindle, K.M. Metabolic Imaging Detects Low Levels of Glycolytic Activity That Vary with Levels of c-Myc Expression in Patient-Derived Xenograft Models of Glioblastoma. Cancer Res. 2018, 78, 5408–5418. [Google Scholar] [CrossRef] [Green Version]
  288. Nguyen, T.T.T.; Zhang, Y.; Shang, E.; Shu, C.; Torrini, C.; Zhao, J.; Bianchetti, E.; Mela, A.; Humala, N.; Mahajan, A.; et al. HDAC inhibitors elicit metabolic reprogramming by targeting super-enhancers in glioblastoma models. J. Clin. Investig. 2020, 130, 3699–3716. [Google Scholar] [CrossRef]
  289. Szeliga, M.; Albrecht, J. Opposing roles of glutaminase isoforms in determining glioblastoma cell phenotype. Neurochem. Int. 2015, 88, 6–9. [Google Scholar] [CrossRef]
  290. Yachi, K.; Tsuda, M.; Kohsaka, S.; Wang, L.; Oda, Y.; Tanikawa, S.; Ohba, Y.; Tanaka, S. miR-23a promotes invasion of glioblastoma via HOXD10-regulated glial-mesenchymal transition. Signal. Transduct. Target. Ther. 2018, 3, 33. [Google Scholar] [CrossRef] [Green Version]
  291. Xin, R.; Qu, D.; Su, S.; Zhao, B.; Chen, D. Downregulation of miR-23b by transcription factor c-Myc alleviates ischemic brain injury by upregulating Nrf2. Int. J. Biol. Sci. 2021, 17, 3659–3671. [Google Scholar] [CrossRef]
  292. Chang, T.C.; Yu, D.; Lee, Y.S.; Wentzel, E.A.; Arking, D.E.; West, K.M.; Dang, C.V.; Thomas-Tikhonenko, A.; Mendell, J.T. Widespread microRNA repression by Myc contributes to tumorigenesis. Nat. Genet. 2008, 40, 43–50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  293. Winter-Vann, A.M.; Casey, P.J. Post-prenylation-processing enzymes as new targets in oncogenesis. Nat. Rev. Cancer 2005, 5, 405–412. [Google Scholar] [CrossRef] [PubMed]
  294. Wang, X.; Huang, Z.; Wu, Q.; Prager, B.C.; Mack, S.C.; Yang, K.; Kim, L.J.Y.; Gimple, R.C.; Shi, Y.; Lai, S.; et al. MYC-Regulated Mevalonate Metabolism Maintains Brain Tumor-Initiating Cells. Cancer Res. 2017, 77, 4947–4960. [Google Scholar] [CrossRef] [Green Version]
  295. Wang, X.; Yang, K.; Xie, Q.; Wu, Q.; Mack, S.C.; Shi, Y.; Kim, L.J.Y.; Prager, B.C.; Flavahan, W.A.; Liu, X.; et al. Purine synthesis promotes maintenance of brain tumor initiating cells in glioma. Nat. Neurosci. 2017, 20, 661–673. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  296. Rosinska, S.; Gavard, J. Tumor Vessels Fuel the Fire in Glioblastoma. Int. J. Mol. Sci. 2021, 22, 6514. [Google Scholar] [CrossRef] [PubMed]
  297. Domenech, M.; Hernandez, A.; Plaja, A.; Martinez-Balibrea, E.; Balana, C. Hypoxia: The Cornerstone of Glioblastoma. Int. J. Mol. Sci. 2021, 22, 12608. [Google Scholar] [CrossRef] [PubMed]
  298. Lakka, S.S.; Rao, J.S. Antiangiogenic therapy in brain tumors. Expert. Rev. Neurother. 2008, 8, 1457–1473. [Google Scholar] [CrossRef] [PubMed]
  299. Carmeliet, P.; Jain, R.K. Principles and mechanisms of vessel normalization for cancer and other angiogenic diseases. Nat. Rev. Drug Discov. 2011, 10, 417–427. [Google Scholar] [CrossRef]
  300. De Bock, K.; Cauwenberghs, S.; Carmeliet, P. Vessel abnormalization: Another hallmark of cancer? Molecular mechanisms and therapeutic implications. Curr. Opin. Genet. Dev. 2011, 21, 73–79. [Google Scholar] [CrossRef]
  301. Nagy, J.A.; Chang, S.H.; Shih, S.C.; Dvorak, A.M.; Dvorak, H.F. Heterogeneity of the tumor vasculature. Semin. Thromb. Hemost. 2010, 36, 321–331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  302. Lee, J.W.; Bae, S.H.; Jeong, J.W.; Kim, S.H.; Kim, K.W. Hypoxia-inducible factor (HIF-1)alpha: Its protein stability and biological functions. Exp. Mol. Med. 2004, 36, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  303. Liu, Z.Z.; Tian, Y.F.; Wu, H.; Ouyang, S.Y.; Kuang, W.L. LncRNA H19 promotes glioma angiogenesis through miR-138/HIF-1alpha/VEGF axis. Neoplasma 2020, 67, 111–118. [Google Scholar] [CrossRef] [PubMed]
  304. Balandeh, E.; Mohammadshafie, K.; Mahmoudi, Y.; Hossein Pourhanifeh, M.; Rajabi, A.; Bahabadi, Z.R.; Mohammadi, A.H.; Rahimian, N.; Hamblin, M.R.; Mirzaei, H. Roles of Non-coding RNAs and Angiogenesis in Glioblastoma. Front. Cell Dev. Biol. 2021, 9, 716462. [Google Scholar] [CrossRef]
  305. Xie, T.X.; Xia, Z.; Zhang, N.; Gong, W.; Huang, S. Constitutive NF-kappaB activity regulates the expression of VEGF and IL-8 and tumor angiogenesis of human glioblastoma. Oncol. Rep. 2010, 23, 725–732. [Google Scholar]
  306. Yu, M.O.; Park, K.J.; Park, D.H.; Chung, Y.G.; Chi, S.G.; Kang, S.H. Reactive oxygen species production has a critical role in hypoxia-induced Stat3 activation and angiogenesis in human glioblastoma. J. Neurooncol. 2015, 125, 55–63. [Google Scholar] [CrossRef]
  307. Vallee, A.; Guillevin, R.; Vallee, J.N. Vasculogenesis and angiogenesis initiation under normoxic conditions through Wnt/beta-catenin pathway in gliomas. Rev. Neurosci. 2018, 29, 71–91. [Google Scholar] [CrossRef]
  308. O’Donnell, K.A.; Wentzel, E.A.; Zeller, K.I.; Dang, C.V.; Mendell, J.T. c-Myc-regulated microRNAs modulate E2F1 expression. Nature 2005, 435, 839–843. [Google Scholar] [CrossRef]
  309. Dews, M.; Fox, J.L.; Hultine, S.; Sundaram, P.; Wang, W.; Liu, Y.Y.; Furth, E.; Enders, G.H.; El-Deiry, W.; Schelter, J.M.; et al. The myc-miR-17~92 axis blunts TGFbeta signaling and production of multiple TGFbeta-dependent antiangiogenic factors. Cancer Res. 2010, 70, 8233–8246. [Google Scholar] [CrossRef] [Green Version]
  310. Chen, J.; Liu, G.; Wang, X.; Hong, H.; Li, T.; Li, L.; Wang, H.; Xie, J.; Li, B.; Li, T.; et al. Glioblastoma stem cell-specific histamine secretion drives pro-angiogenic tumor microenvironment remodeling. Cell Stem Cell 2022, 29, 1531–1546.e1537. [Google Scholar] [CrossRef]
  311. Abdel-Wahab, A.F.; Mahmoud, W.; Al-Harizy, R.M. Targeting glucose metabolism to suppress cancer progression: Prospective of anti-glycolytic cancer therapy. Pharmacol. Res. 2019, 150, 104511. [Google Scholar] [CrossRef]
  312. Mohanti, B.K.; Rath, G.K.; Anantha, N.; Kannan, V.; Das, B.S.; Chandramouli, B.A.; Banerjee, A.K.; Das, S.; Jena, A.; Ravichandran, R.; et al. Improving cancer radiotherapy with 2-deoxy-D-glucose: Phase I/II clinical trials on human cerebral gliomas. Int. J. Radiat. Oncol. Biol. Phys. 1996, 35, 103–111. [Google Scholar] [CrossRef]
  313. Coleman, M.C.; Asbury, C.R.; Daniels, D.; Du, J.; Aykin-Burns, N.; Smith, B.J.; Li, L.; Spitz, D.R.; Cullen, J.J. 2-deoxy-D-glucose causes cytotoxicity, oxidative stress, and radiosensitization in pancreatic cancer. Free Radic. Biol. Med. 2008, 44, 322–331. [Google Scholar] [CrossRef]
  314. Liapi, E.; Geschwind, J.F.; Vali, M.; Khwaja, A.A.; Prieto-Ventura, V.; Buijs, M.; Vossen, J.A.; Ganapathy-Kanniappan, S.; Wahl, R.L. Assessment of tumoricidal efficacy and response to treatment with 18F-FDG PET/CT after intraarterial infusion with the antiglycolytic agent 3-bromopyruvate in the VX2 model of liver tumor. J. Nucl. Med. 2011, 52, 225–230. [Google Scholar] [CrossRef] [Green Version]
  315. Chapiro, J.; Sur, S.; Savic, L.J.; Ganapathy-Kanniappan, S.; Reyes, J.; Duran, R.; Thiruganasambandam, S.C.; Moats, C.R.; Lin, M.; Luo, W.; et al. Systemic delivery of microencapsulated 3-bromopyruvate for the therapy of pancreatic cancer. Clin. Cancer Res. 2014, 20, 6406–6417. [Google Scholar] [CrossRef] [Green Version]
  316. Dunbar, E.M.; Coats, B.S.; Shroads, A.L.; Langaee, T.; Lew, A.; Forder, J.R.; Shuster, J.J.; Wagner, D.A.; Stacpoole, P.W. Phase 1 trial of dichloroacetate (DCA) in adults with recurrent malignant brain tumors. Investig. New Drugs 2014, 32, 452–464. [Google Scholar] [CrossRef] [Green Version]
  317. Shen, Y.C.; Ou, D.L.; Hsu, C.; Lin, K.L.; Chang, C.Y.; Lin, C.Y.; Liu, S.H.; Cheng, A.L. Activating oxidative phosphorylation by a pyruvate dehydrogenase kinase inhibitor overcomes sorafenib resistance of hepatocellular carcinoma. Br. J. Cancer 2013, 108, 72–81. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  318. Chu, Q.S.; Sangha, R.; Spratlin, J.; Vos, L.J.; Mackey, J.R.; McEwan, A.J.; Venner, P.; Michelakis, E.D. A phase I open-labeled, single-arm, dose-escalation, study of dichloroacetate (DCA) in patients with advanced solid tumors. Investig. New Drugs 2015, 33, 603–610. [Google Scholar] [CrossRef] [PubMed]
  319. Le, A.; Cooper, C.R.; Gouw, A.M.; Dinavahi, R.; Maitra, A.; Deck, L.M.; Royer, R.E.; Vander Jagt, D.L.; Semenza, G.L.; Dang, C.V. Inhibition of lactate dehydrogenase A induces oxidative stress and inhibits tumor progression. Proc. Natl. Acad. Sci. USA 2010, 107, 2037–2042. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  320. Zhou, M.; Zhao, Y.; Ding, Y.; Liu, H.; Liu, Z.; Fodstad, O.; Riker, A.I.; Kamarajugadda, S.; Lu, J.; Owen, L.B.; et al. Warburg effect in chemosensitivity: Targeting lactate dehydrogenase-A re-sensitizes taxol-resistant cancer cells to taxol. Mol. Cancer 2010, 9, 33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  321. Lin, S.T.; Tu, S.H.; Yang, P.S.; Hsu, S.P.; Lee, W.H.; Ho, C.T.; Wu, C.H.; Lai, Y.H.; Chen, M.Y.; Chen, L.C. Apple Polyphenol Phloretin Inhibits Colorectal Cancer Cell Growth via Inhibition of the Type 2 Glucose Transporter and Activation of p53-Mediated Signaling. J. Agric. Food Chem. 2016, 64, 6826–6837. [Google Scholar] [CrossRef] [PubMed]
  322. Chan, D.A.; Sutphin, P.D.; Nguyen, P.; Turcotte, S.; Lai, E.W.; Banh, A.; Reynolds, G.E.; Chi, J.T.; Wu, J.; Solow-Cordero, D.E.; et al. Targeting GLUT1 and the Warburg effect in renal cell carcinoma by chemical synthetic lethality. Sci. Transl. Med. 2011, 3, 94ra70. [Google Scholar] [CrossRef] [Green Version]
  323. Dalva-Aydemir, S.; Bajpai, R.; Martinez, M.; Adekola, K.U.; Kandela, I.; Wei, C.; Singhal, S.; Koblinski, J.E.; Raje, N.S.; Rosen, S.T.; et al. Targeting the metabolic plasticity of multiple myeloma with FDA-approved ritonavir and metformin. Clin. Cancer Res. 2015, 21, 1161–1171. [Google Scholar] [CrossRef] [Green Version]
  324. Yu, W.; Yang, X.; Zhang, Q.; Sun, L.; Yuan, S.; Xin, Y. Targeting GLS1 to cancer therapy through glutamine metabolism. Clin. Transl. Oncol. 2021, 23, 2253–2268. [Google Scholar] [CrossRef] [PubMed]
  325. Song, M.; Kim, S.H.; Im, C.Y.; Hwang, H.J. Recent Development of Small Molecule Glutaminase Inhibitors. Curr. Top. Med. Chem. 2018, 18, 432–443. [Google Scholar] [CrossRef]
  326. Matre, P.; Velez, J.; Jacamo, R.; Qi, Y.; Su, X.; Cai, T.; Chan, S.M.; Lodi, A.; Sweeney, S.R.; Ma, H.; et al. Inhibiting glutaminase in acute myeloid leukemia: Metabolic dependency of selected AML subtypes. Oncotarget 2016, 7, 79722–79735. [Google Scholar] [CrossRef] [Green Version]
  327. Tateishi, K.; Iafrate, A.J.; Ho, Q.; Curry, W.T.; Batchelor, T.T.; Flaherty, K.T.; Onozato, M.L.; Lelic, N.; Sundaram, S.; Cahill, D.P.; et al. Myc-Driven Glycolysis Is a Therapeutic Target in Glioblastoma. Clin. Cancer Res. 2016, 22, 4452–4465. [Google Scholar] [CrossRef] [Green Version]
  328. Hasmann, M.; Schemainda, I. FK866, a highly specific noncompetitive inhibitor of nicotinamide phosphoribosyltransferase, represents a novel mechanism for induction of tumor cell apoptosis. Cancer Res. 2003, 63, 7436–7442. [Google Scholar] [PubMed]
  329. Watson, M.; Roulston, A.; Belec, L.; Billot, X.; Marcellus, R.; Bedard, D.; Bernier, C.; Branchaud, S.; Chan, H.; Dairi, K.; et al. The small molecule GMX1778 is a potent inhibitor of NAD+ biosynthesis: Strategy for enhanced therapy in nicotinic acid phosphoribosyltransferase 1-deficient tumors. Mol. Cell Biol. 2009, 29, 5872–5888. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  330. 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]
Figure 1. The complex structure of the Myc oncoprotein. Structures of distinct Myc domains, relative post-translationally modified residues, and interacting regions with protein partners.
Figure 1. The complex structure of the Myc oncoprotein. Structures of distinct Myc domains, relative post-translationally modified residues, and interacting regions with protein partners.
Ijms 24 04217 g001
Figure 2. Myc regulation of gene transcription in healthy and cancer cells. Schematic representation of abundancy-dependent Myc binding at target promoters. (a) In healthy cells, Myc/Max and coregulators preferentially bind to high-affinity E-boxes (h-a-E-box). By inducing both the opening of chromatin through histone acetylation and RNAPII pause and release via PTEFb-dependent phosphorylation of Ser2 on RNAPII, Myc/Max heterodimers tune the transcription, splicing, and RNA capping of genes pivotal to cellular homeostasis. (b) In overexpressing Myc cancer cells, Myc/Max heterodimers also invade low-affinity binding sites (l-a-E-box) and push transcription, leading to the execution of aberrant gene expression programs.
Figure 2. Myc regulation of gene transcription in healthy and cancer cells. Schematic representation of abundancy-dependent Myc binding at target promoters. (a) In healthy cells, Myc/Max and coregulators preferentially bind to high-affinity E-boxes (h-a-E-box). By inducing both the opening of chromatin through histone acetylation and RNAPII pause and release via PTEFb-dependent phosphorylation of Ser2 on RNAPII, Myc/Max heterodimers tune the transcription, splicing, and RNA capping of genes pivotal to cellular homeostasis. (b) In overexpressing Myc cancer cells, Myc/Max heterodimers also invade low-affinity binding sites (l-a-E-box) and push transcription, leading to the execution of aberrant gene expression programs.
Ijms 24 04217 g002
Figure 3. Myc at the intersection between metabolism and gene expression in glioblastoma. The image depicts the major metabolic pathways that regulate and are regulated by Myc in GBM. Basically, Myc controls all the metabolic pathways in GBM, whereas glycolysis remains the main controller of Myc activity in this type of brain tumor. Glu—glucose; Gln—glutamine.
Figure 3. Myc at the intersection between metabolism and gene expression in glioblastoma. The image depicts the major metabolic pathways that regulate and are regulated by Myc in GBM. Basically, Myc controls all the metabolic pathways in GBM, whereas glycolysis remains the main controller of Myc activity in this type of brain tumor. Glu—glucose; Gln—glutamine.
Ijms 24 04217 g003
Table 1. Clinical trials related to Myc overexpressing brain tumours.
Table 1. Clinical trials related to Myc overexpressing brain tumours.
Trial N°StatusStudy TitleDrug/TreatmentConditionSponsorCollaborators
NCT03434262Active, not recruitingSJDAWN: St. Jude’s Children Research Hospital Phase-I study evaluating molecularly driven doublet therapies for children and young adults with recurrent brain tumorsRibociclib/trametinib,
ribociclib/sonidegib,
ribociclib/gemcitabine
Brain tumors, including medulloblstoma, ependymoma, oligodendroglioma, glioblastoma (IDHwt and IDHmut), xanthoastrocytoma, neuroblastoma, medulloepithelioma, and embryonal tumorsSt. Jude’s Children Research HospitalNovartis Pharmaceuticals
NCT03224104CompletedStudy of TG02 in elderly newly diagnosed or adult relapsed patients with anaplastic astrocytoma or glioblastomaTG02,
TMZ,
radiation
IDH1R132H-non mutated and MGMT-promoter-unmethylated anaplastic astrocytoma or GBM; IDH1R132H-non mutated and MGMT-promoter-methylated anaplastic astrocytoma or GBMEuropean Organization for Research and Treatment of Cancer (EORTC)Tragara Pharmaceuticals
NCT02711137TerminatedOpen-label safety and tolerability study of INCB057643 in subjects with advanced malignanciesINCB057643,
gemcitabine,
paclitaxel,
rucaparib,
abiraterone,
azacitidine,
ruxolitinib
Solid tumors, including brain tumors and lymphomaIncyte CorporationIncyte Corporation
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

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. https://doi.org/10.3390/ijms24044217

AMA Style

Cencioni C, Scagnoli F, Spallotta F, Nasi S, Illi B. The “Superoncogene” Myc at the Crossroad between Metabolism and Gene Expression in Glioblastoma Multiforme. International Journal of Molecular Sciences. 2023; 24(4):4217. https://doi.org/10.3390/ijms24044217

Chicago/Turabian Style

Cencioni, Chiara, Fiorella Scagnoli, Francesco Spallotta, Sergio Nasi, and Barbara Illi. 2023. "The “Superoncogene” Myc at the Crossroad between Metabolism and Gene Expression in Glioblastoma Multiforme" International Journal of Molecular Sciences 24, no. 4: 4217. https://doi.org/10.3390/ijms24044217

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

Article Metrics

Back to TopTop