Next Article in Journal
Low Repair Capacity of DNA Double-Strand Breaks Induced by Laser-Driven Ultrashort Electron Beams in Cancer Cells
Next Article in Special Issue
The Contribution of Autophagy and LncRNAs to MYC-Driven Gene Regulatory Networks in Cancers
Previous Article in Journal
Versatile Roles of Aquaporins in Plant Growth and Development
Previous Article in Special Issue
Pivotal Role of Fatty Acid Synthase in c-MYC Driven Hepatocarcinogenesis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Molecular ‘Myc-anisms’ behind Myc-Driven Tumorigenesis and the Relevant Myc-Directed Therapeutics

Department of Pathology, University of Michigan, Ann Arbor, MI 48109, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2020, 21(24), 9486; https://doi.org/10.3390/ijms21249486
Submission received: 16 November 2020 / Revised: 9 December 2020 / Accepted: 11 December 2020 / Published: 13 December 2020
(This article belongs to the Special Issue Myc-Driven Tumorigenesis and Cell Death)

Abstract

:
MYC, a well-studied proto-oncogene that is overexpressed in >20% of tumors across all cancers, is classically known as “undruggable” due to its crucial roles in cell processes and its lack of a drug binding pocket. Four decades of research and creativity led to the discovery of a myriad of indirect (and now some direct!) therapeutic strategies targeting Myc. This review explores the various mechanisms in which Myc promotes cancer and highlights five key therapeutic approaches to disrupt Myc, including transcription, Myc-Max dimerization, protein stability, cell cycle regulation, and metabolism, in order to develop more specific Myc-directed therapies.

1. Myc’s Role as a Transcription Factor

Think about any key cellular process, and the Myc family most likely has a role in it: proliferation, metabolism, differentiation, and apoptosis. The Myc transcription factor family consists of c-Myc, N-Myc, and L-Myc. Discovery of c-Myc led to finding N-Myc, primarily expressed during development or in neuroblastoma, and L-Myc, expressed in lung tissue and small-cell lung cancer [1,2]. Although the family shares stretches of homologous regions and some related targets, N-Myc and L-Myc are less characterized. c-Myc (hereon referred to as Myc), is the most well-studied family member as it is an influential protooncogenic transcription factor that binds to about 15% of genes [3,4,5]. In order to regulate gene expression, Myc recruits or interacts with many different cofactors, including histone acetyl transferases (CBP, p300, GCN5/TRAPP), P-TEFb and polymerases, and chromatin remodelers (BRD4, the SWI/SNF complex, SIRT1) [5,6]. It is important to note that Myc can also repress gene expression by binding to the promoter region and interacting with MIZ1 and SP1 to displace co-activators or by recruiting DNA methyltransferases [5]. Furthermore, there are two structural components of the MYC gene that are essential to drive its role as a transcription factor: the E box and the basic helix-loop-helix leucine zipper domain.
The canonical Myc E Box DNA binding motif (5′-CACGTG-3′) is one of the most frequent regulatory motifs in the human genome [7]. Although Myc is not the only transcription factor that can occupy this motif, elevated levels of Myc will replace the other bound transcription factors [8], demonstrating how Myc can influence the transcription of many genes and diverse processes in proliferating cells. A Myc core signature of 50 common target genes across four human cancer cell types and human embryonic stem cells revealed Myc’s influence in RNA processing and ribosome biogenesis [9]. Other diverse functions of Myc target genes include cell cycle regulation, metabolism, cell adhesion, and signal transduction [5,10,11].
However, Myc does not exclusively bind to the E-box to modulate transcription. In repression of gene transcription, cofactors recruit Myc to the promoters lacking the E-box and interfere with active transcription factors [12,13,14]. Furthermore, Myc can amplify transcriptional signals by accumulating at the promoters of active genes, even in those with low-affinity E-box-like sequences [15,16]. There is still debate of whether Myc drives global amplification of transcription [15,16] or if global amplification is an indirect consequence of Myc’s selective regulation of gene targets [17,18,19].
In addition to the E box binding motif, the basic helix-loop-helix leucine zipper (bHLHZip) domain is crucial for Myc’s activity. To take on its role as a transcription factor, Myc must heterodimerize with Myc-associated factor X (Max); Myc is incapable of homodimerizing and is inactive as a monomer. Max binds to Myc at the bHLHZip domain [20,21], and this heterodimerization is required to bind to the E box consensus sequence and activate transcription [22,23]. However, overexpression of Max leads to transcriptional repression as the Max homodimers antagonize Myc/Max heterodimers [22,24]. Mad, a transcriptional repressor, can also reduce Myc-driven transcription by dimerizing with Max [5].

2. Dysregulation of Myc Leads to Cancer

Normally, Myc expression is tightly controlled at each molecular level (transcriptionally, post-transcriptionally, translationally, and post-translationally via protein stability, and via protein interactions), and has a short half-life of 20–30 min [25,26,27,28,29]. Given that there are many levels of regulation, as a consequence, there are many opportunities for which control of MYC can go awry. For instance, point mutations, chromosomal translocations, and gene amplification, or other factors that activate transcription or stabilize Myc, have been found in a wide range of cancers, which are further described by Meyer and Penn and Kalkat et al. [30,31]. This oncogenic activation, which leads to sustained levels of Myc, contributes to tumorigenesis and evasion of tumor-suppressive checkpoints leading to uncontrolled cell growth. MYC expressing tumors thus become addicted to and depend on the oncogene, as shown in cancer models with conditional activation of MYC [32]. On the contrary, inactivation of MYC leads to tumor regression in transgenic mouse models, displaying Myc’s vital role in tumor initiation and maintenance [33,34,35].
MYC amplification is found in 21% of patients across 33 different cancers [36], particularly breast cancer, lung squamous cell carcinoma, uterine carcinoma, esophageal carcinoma, and ovarian cancer [25] (Figure 1). The highest rates of amplification are seen in high-grade serous ovarian cancer wherein greater than 50% of tumors harbor this genomic alteration. MYC translocation affects several hematological malignancies, including multiple myeloma, Burkitt’s lymphoma, diffuse large cell lymphoma, and T-cell acute leukemia [37]. Alternatively, some tumors that do not display MYC amplification show extreme phosphorylation levels which aid in Myc stability [38,39,40,41].
With Myc’s prominent role across many cancers, the idea of Myc as a clinical target is too good to be true. Although targeted inhibition of MYC via siRNA reduces tumor burden in mice with very few toxicities despite Myc’s influence on global transcription [33,35,42,43], global MYC knockout is embryonic lethal in mice. Thus, cautious measures in observing side effects of disrupting Myc need to be addressed [44]. The less expected problem is that direct inhibition of Myc is not possible with current therapeutic approaches—Myc lacks both enzymatic activity and an active site for a small molecule to disrupt protein-protein interactions [45]. Myc’s primary nuclear localization further escalates the problem. Nonetheless, scientific discoveries led to creative ways to downregulate Myc. This review focuses on how Myc’s oncogenic activation leads to tumorigenesis through initiating transcription, increasing stability, and influencing cell cycle and metabolism, coupled with descriptions of the indirect inhibitors of Myc that target each mechanism (Figure 2). The molecular changes in which MYC becomes an oncogene (mutations, translocations, and amplification) is beyond the scope of this review [30,31,32].

3. Disrupting Myc Stability to Inhibit Its Actions as a Transcription Factor

In cancer, Myc’s aberrant function as a transcription factor leads to increased cell proliferation, cell differentiation, cell adhesion, and angiogenesis [10]. Here we will focus on inhibiting transcription, disrupting Myc/Max dimerization, and enhancing protein degradation as strategies to disrupt Myc gene and protein stability and therefore Myc-driven tumorigenesis.

3.1. Myc Drives Aberrant Transcription

As discussed, MYC amplification is common among many cancer types. This amplification of MYC results in increased binding of Myc to promoters and enhancers of active genes, which magnifies the transcriptional signal [15,16] and as a consequence, increases global transcription. During transcription, Myc recruits the transcriptional pause-release complex P-TEFb (a heterodimer of cyclin-dependent kinase 9 (CDK9) and cyclin T1, T2, or K) [6,16,46]. P-TEFb leads to activation of transcriptional elongation by phosphorylating RNA polymerase II (Pol II) via CDK9, stimulating pause release [47,48,49]. Furthermore, BRD4, part of the bromodomain and extra-terminal motif (BET) protein family, also recruits P-TEFb to promoters to initiate transcription elongation [50]. The overlapping roles of BET proteins and Myc in recruiting P-TEFb suggests BET proteins or CDK9 as therapeutic targets. First, BET proteins are known to regulate MYC transcription [51]. A recent study demonstrated in normal cells that BRD4 has even more control over Myc by binding and phosphorylating Threonine 58 on Myc, leading to degradation [52]. However, Myc is also capable of regulating BRD4′s histone acetyl transferase activity [52]. Additional studies are needed to better understand how this circular balance may be affected in cancer. CDK9 is a potential target as it is part of P-TEFb, is necessary for proliferation and maintenance of MYC-overexpressing hepatocellular carcinoma [53], and is required for maintenance of gene silencing in several cancer cell lines [54]. Another tumorigenic feature of Myc is looping to tumor-specific super-enhancers (sites defined by multiple enhancers abnormally bound by a plethora of transcription factors, such as BRD4 and CDK9) [55]. Therefore, inhibiting MYC transcription indirectly via BET inhibitors or affecting transcription of Myc target genes by inhibiting CDK9 are promising strategies that have shown efficacy in Myc-driven cancers (Figure 3).

Targeting MYC Transcription—BET Inhibitors, BRD4 Degraders, CDK9 Inhibitors

The BET proteins, BRD2, BRD3, BRD4, and testis-specific BRDT, are epigenetic readers and histone acetyl transferases that activate transcription via binding to specific acetylated lysine residues on histone tails. The bound BET proteins regulate chromatin remodeling via H3K122 acetylation and act as scaffolds to form transcription complexes by recruiting transcriptional activators such as P-TEFb [50,56,57]. Furthermore, BRD4 influences mitotic progression by binding selectively to transcriptional start sites of M/G1 genes [58]. Oncogenes, such as MYC, have a transcriptional dependency on BRD4 and recent findings suggest additional non-transcriptional functions of BRD4 in cancer [59]. Bromodomains have a mostly hydrophobic pocket with aromatic rings and is an ideal size for protein–protein interactions, making bromodomains attractive and obtainable therapeutic targets, unlike Myc [56]. BET protein inhibitors compete for access to the bromodomain and upon binding, disrupt chromatin remodeling and prohibit expression of target genes, including MYC. Filippakopoulos et al. and Nicodeme et al. independently designed some of the first bromodomain inhibitors, known as JQ1 and iBET respectively, that are highly specific towards the BET protein family [60,61]. Initial studies showed efficacy of JQ1 downregulating both MYC expression and Myc’s transcriptome genome-wide in Myc-addicted hematological malignancies [62,63,64], and solid cancers [65,66,67,68]. iBET’s proof-of-concept in preventing BET proteins from binding to acetylated histones was demonstrated in an inflammation context [61], although a follow-up study exhibited that iBET was capable of downregulating MYC expression, but to a lesser extent than JQ1 [51]. It is important to note that both JQ1 and iBET lack specificity for a particular BET protein family member, which limits their therapeutic availability [69,70,71]. Therefore, these BET inhibitors serve best as tools to improve our understanding of targeting bromodomains and the effects on MYC. The discovery of JQ1 and iBET inspired the development of additional BET inhibitors, with 10 inhibitors being assessed in clinical trials, including MK-8628/OTX015. Phase Ib trials included six solid tumors such as NUT midline carcinoma (NMC), which harbors an oncogenic form of BRD4, known as BRD4-NUT. The trial (NCT02259114) completed with a recommended dose for Phase II studies, although the NMC patients that initially responded, relapsed several months after treatment [72]. BET inhibitors as a whole currently appear to have limited therapeutic response and dose-limiting toxicities. More preclinical research will increase the biological knowledge on mechanisms of action and resistance of BET inhibitors.
In addition to BET inhibitors, there are also BET degraders that utilize a concept designed in 2000: PROteolysis TArgeting Chimeric (PROTAC) [73]. PROTAC protein degraders link the protein of interest to an E3 ligase in order to ubiquitinate the protein of interest for degradation. This approach has been adapted to a variety of targets, including the androgen receptor, estrogen receptor, BCL2, CDK9, and BET proteins to name a few [74]. PROTAC technology has entered clinical trials, including Arvinas’s ARV-110 for patients with metastatic castration-resistant prostate cancer (NCT03888612) which has shown efficacy and a promising safety profile in Phase I [75]. The first PROTAC BET degraders, including MZ1, a BRD4-specific degrader, were designed in 2015 and demonstrated increased apoptotic response compared to nonspecific BET inhibitors, but a modest decrease in MYC expression [76,77].
It appears the antitumor efficacy of both BET inhibitors and BET degraders is most likely due to global transcription downregulation, rather than downregulation in MYC transcripts specifically [78]. Devaiah et al. recently discovered crucial molecular differences in Myc stability between BET inhibitors and BET degraders. Since endogenous BRD4 destabilizes Myc, treatment with a BRD4 degrader, such as MZ1, enhances Myc stability, but treatment with a BET inhibitor, such as JQ1, does not affect BRD4′s phosphorylation of Myc and therefore Myc’s half-life is unaffected while MYC transcription is downregulated [52]. Several PROTAC BRD4 degraders demonstrate robust decreases of MYC expression throughout 3–24 h [79,80], though there are no current clinical trials on BET/BRD4 degraders. Perhaps later timepoints and investigation of phophorylated-S62-c-Myc expression will aid in understanding long-term effects of BET degraders on Myc stability. For further reading, detailed reviews on bromodomains and their inhibitors are cited [56,81,82,83,84].
CDK9 is another potential therapeutic target, given its kinase activity in the P-TEFb complex which releases paused RNA Pol II to initiate transcription. CDK9 inhibitors demonstrate efficacy in downregulating MYC transcripts and Myc stability across hepatocellular carcinoma [53], mixed-lineage leukemia [85], diffuse large B-cell lymphoma [86], acute myeloid leukemia [87], and pancreatic cancer [88]. Although preclinical studies have shown efficacy in targeting CDK9, the sequence similarity to other cyclin-dependent kinases made specificity difficult. However, several groups succeeded in creating CDK9-specific inhibitors and PROTAC degraders [38,54,89]. A recent study demonstrated that CDK9-specific inhibitor, MC180295, downregulates MYC and leads to reactivation of epigenetically silenced tumor suppressor genes [54]. Thus, downregulation of MYC is not due to off-target effects of nonspecific CDK inhibition. Initial nonselective CDK inhibitors did not succeed in clinical trials, most likely due to toxicities from off-target effects. These trials included patients of many cancer types and were not selective to MYC-amplified patients [90]. However, CDK9-specific inhibitors, such as BAY 1143572 (NCT01938638), are beginning to enter clinical trials in patients with advanced cancer and will evaluate MYC expression as a biomarker [91,92].
Additionally, combining CDK9 and BET inhibitors synergistically improves anti-proliferative activity in several cancers, with no hematological toxicity or weight loss shown in vivo [93,94,95]. Of the same note, BET inhibitors were also efficacious when paired with additional inhibitors, such as PI3K, ERK, or BCL2 inhibitors [81]. Readers are referred to reviews further discussing targetable Myc cofactors that aid in tumorigenesis [13], such as G quadraplex stabilizers [96,97]. In all, BET and CDK9 inhibitors vastly affect transcription and as a result, downregulate MYC expression indirectly; improving their specificity is expected to increase their therapeutic benefit.

3.2. Myc/Max Dimerization

Another way of affecting Myc transcriptional stability is by preventing Myc from interacting with DNA. Myc must dimerize with Max in order to drive gene expression, though a recent structural study demonstrates that against previous belief, Myc is stabile in the absence of binding DNA [45]. Although, Max heterodimerization with Myc is required for Myc’s oncogenic activity [98]. Therefore, inhibiting Myc and Max dimerization prevents Myc from initiating gene transcription. There are two immediate challenges: (a) targeting the bHLHZip domain is nonspecific to Myc/Max and therefore could present off-target effects and (b) there are no apparent pockets for which a small molecule can bind [99,100]. Despite this, there has been success in disrupting the Myc/Max interaction with several mini-proteins or molecules, including Omomyc, 10058-FA, 10074-G5, KJ-Pyr-9, MYCMI-6, and KI-MS2-008 (Figure 4).

Disrupting Myc/Max Dimerization

The most well-known, and perhaps the first, Myc/Max dimerization inhibitor is Omomyc, a dominant negative mutant of Myc’s bHLHZip domain with 4 amino acid mutations in the leucine zipper that prevents Myc/Max heterodimerization [101]. Omomyc was a laboratory tool developed to bind and inhibit Myc. Over the past two decades, research produced a better understanding of how the molecular tool functions: Omomyc reduces the amount of Myc that can bind to promoters by either heterodimerizing with Myc in the cytoplasm, heterodimerizing with Max, or homodimerizing. Recent data show Omomyc preferentially binds to Max or homodimerizes [102]. The Omomyc homodimers or Max/Omomyc heterodimers are transcriptionally inactive complexes that bind specifically to E-box sequences and displace Myc/Max heterodimers resulting in decreased Myc-driven transcription [102,103,104,105]. Importantly, Omomyc is specific towards Myc’s function and does not suppress gene expression of other E-box-binding transcription factors [103].
Omomyc has shown efficacy in several tumor studies when it is conditionally or transiently expressed in the cell or linked with a cell penetrating Phylomer [43,106,107,108]. However, Omomyc is indeed capable of penetrating cells, including non-small cell lung cancer, neuroblastoma, glioblastoma, and melanoma cell lines, due to its basic region [103]. Until recently, in vivo proof-of-concept was lacking. Beaulieu et al. show Omomyc downregulates Myc target gene expression and prevents tumor progression in lung adenocarcinoma in vivo models via intranasal administration (2.37 mg/kg) over four weeks [103]. Similarly, in a lung adenocarcinoma xenograft model, paclitaxel combined with Omomyc administered intravenously diminished tumor growth over 30 days [103]. Both models showed no significant changes nor toxicities in blood counts or pathology reports of all major organs. In non-tumor-bearing mice, Demma et al. show Omomyc injected intravenously (5.22 mg/kg) primarily distributes to the liver and kidneys and has a short half-life in plasma [102]. Although this study used a higher dosage of Omomyc than the cancer study, toxicities of Omomyc in normal cells must be considered in future preclinical studies. Dr. Soucek, who created and studied Omomyc over the past 20 years, created the company Peptomyc to develop and sponsor Omomyc-derived clinical candidates; the first clinical trial is anticipated to start in 2021.
While Omomyc is capable of disrupting Myc/Max dimerization and preventing Myc from interacting with DNA, other Myc/Max inhibitors are typically characterized by one of those two actions. In 2002, Berg et al. demonstrated the proof-of-concept of using combinatorial chemical libraries to find small molecule inhibitors of protein–protein interactions, including Myc/Max [109]. Shortly after, Yin et al. identified specific Myc/Max inhibitors from a combinatorial library including 10058-FA and 10074-G5, that result in G0/G1 cell cycle arrest and apoptosis in vitro [110]. However, in vivo studies show both 10058-FA and 10074-G5 are rapidly metabolized and lack anti-tumor activity [111,112]. Therefore, these compounds best serve as molecular tools and a starting point for new compound development. More recent Myc/Max inhibitors include KJ-Pyr-9 and MYCMI-6. KJ-Pyr-9 has sufficient pharmacokinetic properties to penetrate tissue and prevent tumor growth, but cannot reduce existing tumors; its inability to decrease tumor size may be due to residual Myc activity as an effect of incomplete Myc inhibition [113]. MYCMI-6 was identified in 2018 as a promising Myc/Max-specific inhibitor that halts Myc-driven transcription, induces apoptosis, and reduced tumor proliferation in vivo [114]. Interestingly, these different Myc/Max inhibitors all initiate different biological effects.
Recently, Han et al. identified novel Myc-binding inhibitors, MYCi361 and MYCi975, that appear to act through disrupting Myc/Max dimers and increasing Threonine (T)58 phosphorylation of Myc, which leads to Myc degradation. MYCi-induced degradation could be a result of changes in Myc confirmation as it interacts with MYCi; it is important to note that not all Myc/Max inhibitors lead to Myc degradation [115]. Treating cells with proteasome inhibitor MG132 or exposing non-phosphorylatable Myc (T58A mutant) cells to MYCi361 rescues or prevents the MYCi361-induced Myc degradation [116]. In vivo studies utilized a Myc-driven prostate cancer mouse model, MycCaP, in which tumors were significantly decreased upon MYCi361 treatment. Additional studies are required to determine its efficacy in other cancers. Given that MYCi treatment modified the tumor microenvironment through increased expression of PD-L1, Han et al. demonstrated synergistic effects with MYCi361 and anti-PD1 in the MycCaP model, despite the model’s documented resistance to anti-PD1 therapy [116]. MYCi975 performs similarly to MYCi361 but has a higher therapeutic index and better tolerability in vivo of up to ten time the anti-tumor efficacious dose. There is promise for future studies on MYCi975 due to its inhibition of cancer cell growth and reduction of Myc target gene expression in vitro and decreased tumor growth in vivo with high tolerability. These compounds represent a new class of directly targeting Myc and inhibiting Myc/Max dimers, which led to Myc degradation.
Agents that inhibit Myc/Max from binding DNA have also been pursued, although they lack in vivo data and specificity towards Myc/Max [117]. One approach to prevent Myc binding to DNA is by targeting one of the many cofactors that recruits Myc to its target genes. WDR5 is an adapter protein that interacts with histone methyltransferase and serves as a scaffold for chromatin; it recruits Myc to chromatin and the Myc-WDR5 interaction is required for Myc-driven tumorigenesis [118]. Thomas et al. recently discovered that WDR5 stabilizes the Myc/Max interaction with DNA and a mutant Myc that cannot bind to WDR5 leads to tumor regression in a Burkitt lymphoma in vivo model [118]. However, the mutant Myc was capable of binding to chromatin, suggesting that targeting WDR5 does not affect Myc’s ability to interact with DNA. Given the antitumor effect and the druggable pockets within WDR5, it is a viable anti-Myc contender to pursue; additional recent advances with WDR5 are described in the Metabolism section of this review.
Alternative approaches that are not widely explored include stabilizing Max. In 2019, Struntz et al. discovered KI-MS2-008, which stabilizes Max homodimers while decreasing both Myc binding at promoters and Myc protein levels [119]. KI-MS2-008 proved efficacious in T cell acute lymphoblastic leukemia and hepatocellular carcinoma in vivo models with a reduction in tumor burden and no toxicities in liver or kidney [119]. Further studies are needed to determine the mechanism of action to optimize for in vivo use, but for now, KI-MS2-008 serves as an instrument to investigate the importance of Max dimerization in cancer.
Exploring the Myc/Max interaction has been a popular avenue for disrupting Myc-driven transcription. Sammak et al.’s high resolution crystal structure of the Myc and Max heterodimer in the absence of DNA will aid in development of future Myc-targeting therapeutics [45]. Pursing additional compound libraries, such as Carabet and colleagues’ computational screen to discover inhibitors of Myc-max in silico, can further broaden our understanding of inhibiting Myc/Max dimers [99]. Future Myc/Max dimerization inhibitors must overcome challenges faced by current therapeutics such as fast metabolism, poor penetrability, and nonspecific targets.

3.3. Myc Protein Stability

The short half-life of Myc is evidence for Myc’s highly controlled turnover. Myc’s stability is regulated by phosphorylation on serine 62 (S62) and threonine 58 (T58) by several proteins through the Raf-MEK-ERK kinase and phosphatidylinositol-3 kinase (PI3K)-Akt pathways [28,120]. First, extracellular signal-regulated kinase (ERK), CDK1, or growth signals stabilize Myc by phosphorylating S62. Glycogen synthase kinase 3 (GSK3) is recruited to phosphorylate T58, which is required for Myc degradation. In brief, Pin1 isomerizes proline 63 on Myc, in which protein phosphatase 2A (PP2A), a serine/threonine phosphatase, can now dephosphorylate Myc at S62 [121]. The unstable Myc, with only T58 phosphorylation remaining, becomes ubiquitinated by Fbw7 and is sent for degradation [122]. Again, these many levels of regulation provide multiple opportunities for cancer hijacking. In cancers that lack MYC amplification, there are increases in the stabilizing pS62-Myc and decreases in the degrading pT58-Myc, therefore promoting Myc’s stability and activity [38,39,40,41]. Studies show that mutating T58 to alanine, a non-phosphorylatable residue, results in stable Myc expression and tumorigenic properties, suggesting Myc stability has a role in transformation [123,124,125].
In cancer, we see faulty regulation of these proteins that modify the phosphorylation on Myc and promote stabilization. We will describe three scenarios—activation of PI3K/AKT signaling (which inhibits GSK3B), overexpression of Pin1, and suppression of PP2A activity—that stabilize Myc. PI3K, PTEN, and upstream components of the PI3K/AKT pathway are commonly mutated in cancer to promote pathway activation [126]. Activated AKT phosphorylates (and therefore inhibits) GSK3, which in turn enhances Myc stability [127,128] as GSK3 cannot phosphorylate T58-Myc. This is just one example of how an upstream signaling pathway (MAPK, Wnt, Notch) can quickly trickle down to promoting cancer through Myc.
Peptidyl-prolyl cis-trans isomerase NIMA-interacting 1 (Pin1), an isomerase that specifically recognizes the serine/threonine-proline motif, is overexpressed in several cancers including pancreatic, breast, and prostate and its expression correlates with poor clinical outcomes [129,130,131]. Furthermore, Pin1 promotes several hallmarks of cancer through inactivating 26 tumor suppressors and activating 56 oncogenes [132,133]. By catalyzing the cis/trans conformational change of the target protein, such as Myc, isomerases like Pin1 gain control of the target protein’s stability, activity, and localization [134]. As mentioned, Myc’s stability is regulated through phosphorylation on S62 and T58 and these sites are recognized by trans-specific phosphatases; therefore, Pin1 can stabilize Myc in the cis-confirmation and prevent degradation [135]. On the contrary, Pin1 can revert Myc back to the trans-confirmation after phosphorylation of T58, which allows PP2A to remove phosphorylation from S62 to promote Myc degradation [28]. However, another consideration is that PP2A is commonly inactivated in cancers (described below), and so even if Pin1 reverts Myc back to the trans-confirmation, Myc would unlikely get degraded in the absence of PP2A activity. More research is needed to better understand how T58 phosphorylation affects Pin1 activity and S62 dephosphorylation. Additionally, Pin1 can promote self-ubiquitination of Fbw7, the E3 ubiquitin ligase that ultimately degrades Myc [136]. Pin1’s influence on Myc’s transcriptional activity and stability potentiates tumorigenesis and is a potential therapeutic target for MYC-overexpressing cells [132,135,136,137].
Lastly, PP2A is a ubiquitously expressed tumor suppressor that accounts for a majority of the phosphatase activity in cells and dephosphorylates a range of substrates such as Akt, p53, β-catenin, and Myc [138]. The holoenzyme can contain a variety of different scaffold (A) and regulatory (B) subunits with a common catalytic (C) subunit, with multiple isoforms for each subunit [139]. Inactivation of PP2A through PP2A inhibitor okadaic acid results in tumorigenesis and cellular transformation [140]. PP2A is commonly inactivated in cancer, including lung, colon, breast, skin, cervix, and ovarian [139]. This PP2A inactivation occurs through phosphorylation, somatic mutation, or increased expression of endogenous inhibitors such as SET and CIP2A [141,142,143,144]. In the case of Myc, PP2A inactivation prevents dephosphorylation of S62, therefore stabilizing Myc and promoting transformation [28,121]. In sum, inhibition of GSK3 through PI3K, overexpression of Pin1, and inactivation of PP2A promote stability of Myc (Figure 5). Although there are many opportunities to increase Myc stability, many of these proteins are also potential therapeutic targets to promote Myc’s degradation.

Enhancing Degradation of Myc

Given the various levels regulating Myc degradation, numerous compounds have been developed to enhance Myc degradation through inhibition of PI3K or Pin1 and re-activation of PP2A. Becker and collogues demonstrated efficacy in combining a PI3K inhibitor with a microtubule destabilizer in high-Myc expressing cells. First, they eloquently demonstrated that unphosphorylated S62-Myc binds to mitotic tubules and is protected from degradation [145]. Given this interaction, treatment with a microtubule destabilizer, vincristine, drastically reduced Myc protein and P493-6 B-cell lymphoma cells with ectopic Myc expression were more sensitive to colony forming unit inhibition than Myc low-expressing cell lines. Since PI3K/AKT inhibits GSK3B activity and therefore stabilizes Myc, Becker and collogues investigated the addition of PI3K inhibitor idelalisib following the G2-M arrest induced by vincristine. Treating first with vincristine followed by idelalisib led to higher cell death and decreased clonogenic growth than either compound alone across 16 Burkitt lymphoma and DLBCL cell lines [145]. Furthermore, this combination lead to reduction of Myc and tumor viability in two lymphoma in vivo models, in which the compounds as single agents were not effective. These results suggest a novel avenue of disrupting Myc stability via microtubule destabilizers followed by PI3K inhibition to further decrease Myc protein levels. Another targetable signaling pathway that influences Myc degradation is the MEK/ERK pathway. As mentioned, ERK maintains S62 phosphorylation of Myc, which promotes Myc’s stability [120]. Therefore, inhibition of the MEK/ERK pathway through MEK inhibitor U0126 reduced Myc expression and growth in rhabdomyosarcoma cell lines [146]. Furthermore, inhibition of the MEK/ERK pathway or the consequent decrease in Myc expression, a known driver of radioresistance, sensitizes cancer cells to radiation therapy [147,148].
Aside from Pin1′s influence over Myc’s stability, there are several other mechanisms in which Pin1 can promote tumorigenesis such as sustaining proliferative signaling and downregulating tumor suppressors [132]. More than ten Pin1 inhibitors have been developed that demonstrate anticancer activity, including sensitizing various cancer cells to chemotherapy [132]. We will discuss two Pin1 inhibitors—All-trans retinoic acid (ATRA) and KPT-6566, that have more favorable specificity and safety profiles than other Pin1 inhibitors. ATRA is clinically used for acute promyelocytic leukemia (APL), although its drug target was unknown. Through a mechanism-based high throughput screen, Wei and collogues discovered ATRA directly binds and degrades Pin1 [149]. ATRA was capable of decreasing Pin1 and tumor growth in APL mouse models and APL human patients’ bone marrow, along with in vivo models of triple negative breast cancer [149] and acute myeloid leukemia [150]; both cancers overexpress Pin1. However, ATRA has a short half-life of 45 minutes and moderate anti-cancer activity. Yang and collogues developed an improved, controlled-release formulation of ATRA (ATRA-PLLA microparticles) that demonstrated selectivity for Pin1 inhibition and improved anti-cancer efficacy in xenografts of hepatocellular carcinoma, a cancer that is enhanced by Pin1 [151]. Several other liposomal ATRA delivery methods have been developed and performed well in clinical trials for APL patients [152], although it appears trials for solid tumors utilizing the improved ATRA formulation are lacking. Additionally, these studies did not specifically investigate the effects of ATRA and Myc. Several older studies across small cell lung cancer, breast cancer, and colon cancer demonstrated treatment with ATRA decreased Myc expression at the gene or protein level [153,154,155]. Selective Pin1 inhibitor KPT-6566, which was also identified through a mechanism-based screen, sets Pin1 for degradation. When KPT-6566 binds to the catalytic site of Pin1, reactive oxygen species are produced and DNA damage occurs, leading to cell death particularly in Pin1-overexpressing cancer cells [156]. There are no data on KPT-6566 decreasing tumor volume in vivo, but in mice injected with MDA-MB-231 cells, KPT-6566 daily treatment reduced metastatic spread and showed no toxicities in vital organs [156]. Again, these studies did not investigate the effects of Pin1 inhibition on Myc. More development is necessary to improve efficacy and drug-likeness of Pin1 inhibitors, especially in the context of Myc-driven cancers.
Compounds that target PP2A, which is the main phosphatase the regulates Myc stability, have shown promise in promoting Myc degradation and cell death. There are several methods published on indirectly activating PP2A as an anti-cancer treatment, such as antagonizing the endogenous PP2A inhibitors SET (via OP449 [41,157] or FTY720 [158,159]), and CIP2A (via bortezomib, erlotinib, or celastrol) or disrupting PP2A post translational modifications [143]. SET-inhibitor OP449 increased PP2A activity dose-dependently and OP449-treated leukemia xenografts had a two-fold reduction of tumor burden [157]. In breast cancer, OP449 decreased both phosphorylation levels of S62-Myc and Myc transcriptional activity across several cell lines in vitro. OP449 additionally induced apoptosis while reducing tumor volume and increasing PP2A activity in vivo [41]. In terms of disrupting CIP2A, the described inhibitors were primarily discovered as a proteasome inhibitor (bortezomib), EGFR kinase inhibitor (erlotinib), or anti-cancer (celastrol), but indirectly or independently reduce CIP2A expression or activity [143,160]. Small molecule activators of PP2A (SMAPs) have also emerged as a new class of validated compounds that re-activate PP2A through binding to the A scaffolding subunit of PP2A. As PP2A reactivates, S62-Myc becomes dephosphorylated and Myc is sent for degradation. Recently, SMAPs demonstrated efficacy through binding to PP2A in in vivo models of Burkitt’s lymphoma, non-small cell lung cancer, and triple-negative breast cancer—all Myc-driven cancers, representing Myc amplification, post-translational stabilization, and overexpression [161]. SMAPs also display efficacy in prostate and pancreatic cancer models [162,163]. Dr. Narla, one of the developer of SMAPs, serves as Chief Scientific Officer for Rappta Therapeutics to further develop these anti-cancer molecules that reactive PP2A [164,165]. In all, targeting Myc’s protein stability may help reduce toxicity that is expected with a complete loss of Myc.

4. Taking Advantage of MYC Overexpression to Initiate Synthetic Dosage Lethality in the Context of Cell Cycle

Since transcription factors pose as difficult drug targets, leveraging synthetic lethality offers an alternative approach of antitumoral therapy. Synthetic lethality occurs when a mutation or inhibition of two specific genes leads to cell death, but a mutation or inhibition of just one gene does not affect viability [166]. Synthetic dosage lethality is when manipulation of expression levels leads to cell death; for example, overexpression of gene A and presence of gene B is viable, but the combination of gene A overexpression and loss or lower expression of gene B results in cell death. Therefore, synthetic lethality, or more specifically synthetic dosage lethality, can be advantageous in cancer as the tumors already have mutations or oncogenic addiction, such as overexpression of MYC. A synthetic lethal approach affects the mutated tumor cells and spares the normal cells.
Synthetic lethal targets are identified in an unbiased, high-throughput fashion through RNA interference (RNAi) or CRISPR screens on isogenic cells—cells that differ by a mutation in a single gene. Although the idea sounds swift, identifying clinically relevant synthetic lethal interactions have proven difficult due to validation of lethal mutants by recovery, condition-dependent interactions, and rarity [166]. However, PARP inhibitors successfully demonstrated this concept clinically when given to cancer patients with BRCA mutations, such as in breast and ovarian cancer [167,168].
Understanding the biological results of MYC overexpression will help identify second-site targets that lead to synthetic lethality. Reports show that MYC-overexpressing cancer cells have increased sensitivity to apoptosis in response to cytotoxic drugs or radiation [169]. However, the opposite appears to be true in melanoma, in which lower MYC expression improves susceptibility to chemotherapy and radiation due to reactive oxygen species production and mismatch repair protein inhibition [170,171]. As Myc is a master regulator of cell proliferation and metabolism, genes affiliated with these processes offer a promising avenue to identify synthetic lethal targets.
Cells overexpressing MYC have more mitotic abnormalities, such as altered spindle morphology and mitotic timing [172]. During mitotic stress, Myc worsens mitotic dysfunction and enhances apoptosis, which explains the many cell cycle proteins as targets for synthetic lethality. In normal conditions, advancing through the cell cycle phases of G1, S, G2, and M requires four heterodimers of cyclin-dependent serine/threonine kinases (CDK) and cyclins: CDK1, 2, 4, 6, and cyclins A, B, E, D, all of which are Myc target genes [173]. Cyclin B1 binds to CDK1 at the G2-M transition, activating the complex to promote mitosis. CDK1 is the only essential CDK required for cell cycle progression and it is rarely dysregulated in cancer [174,175]. Inhibiting CDK1 typically results in a G2 arrest, but in MYC-overexpressing cells, CDK1 inhibition leads to apoptosis [176,177]. Cyclins and CDKs just scratch the surface of proteins involved in the cell cycle.
Myc also induces expression of Aurora A kinase, which reciprocally stabilizes Myc in addition to its role in cell cycle [178]. Aurora kinases A and B direct cell cycle progression through G2-M. Aurora A aids in centrosome function, spindle assembly, and mitotic entry while Aurora kinase B is the catalytic component within the chromosomal passenger protein complex (CPPC) to control chromosomal condensation and cytokinesis [179,180]. Aurora A and B kinases are overexpressed in breast and colon cancers, along with sarcoma, esophageal, and stomach cancers [181]. Myc is known to upregulate Aurora kinase A and B expression in B-cell lymphomas, which is necessary to maintain the lymphoma [182]. Similarly, overexpressing MYC in medulloblastoma cell lines with low MYC expression led to an associated increase in Aurora B expression [183]. Overall, cell cycle proteins such as Aurora kinases or CDK1 in MYC-overexpressing cells are potential therapeutic targets as inhibition leads to synthetic dosage lethality (Figure 6).

Targeting Cell Cycle Proteins

CDK1: Purvalanol is a potent CDK1 inhibitor that has selectivity for CDK1 over CDK2 at a low concentration of 4 nM [184]. Goga et al. discovered inhibiting CDK1 pharmacologically with purvalanol or genetically using a cell line with temperature-sensitive Cdk1 allele results in apoptosis in MYC-overexpressing cells [176]. This synthetic lethal interaction led to decreased tumor growth in MYC-expressing lymphoma and hepatoblastoma in vivo models. However, purvalanol is not suitable clinically as it is poorly soluble; new variations are needed to pursue CDK1 inhibitors as a clinical candidate. Additionally, Goga et al. also demonstrated promise for targeting survivin, an endogenous inhibitor of apoptosis and known CDK1 target. CDK1 inhibition via purvalnol degraded survivin, and depleting survivin independently resulted in similar results to CDK1 inhibition in that MYC-overexpressing cells were more sensitive to survivin degradation via peptide inhibitors [176]. Given that loss of p53 influences decreased apoptosis in MYC-overexpressing cells [185,186], the authors explored effects of p53 status on the efficacy of the CDK1 inhibitor. Through the use of wildtype and p53−/− MYC overexpressing mouse embryonic fibroblasts, Goga et al. determined p53 status is independent of purvalanol-induced apoptosis [176]. This is advantageous as many cancers have p53 deficiencies.
A second study from the same group reported that triple-negative breast cancer (TNBC) with elevated MYC expression displayed efficacy with CDK1 inhibition (purvalanol, dinaciclib, or siRNA), compared to lines with low MYC expression [187]. Although third-generation CDK inhibitor dinaciclib inhibits CDK1, CDK2, CDK5, and CDK9, efficacy in TNBC cell lines was shown to be specific to CDK1 inhibition by knocking down Cdk1 via siRNA. Furthermore, dinaciclib, which has improved pharmacokinetic and pharmacodynamic properties than previous CDK inhibitors, decreased tumor volume by about 50% in TNBC xenograft mice [187]. Several clinical trials testing dinaciclib in mainly hematologic malignancies recently completed and are pending results. In 2015 studies, 11% of relapsed multiple myeloma patients partially responded and 54% of lymphocytic leukemia patients partially responded [188,189]. In a randomized phase II in TNBC patients, dinaciclib failed to outperform capecitabine, the standard of care [190]. These studies did not consider MYC expression but offer promising results to continue dinaciclib in clinical research. An active Phase I clinical trial, NCT01676753, is assessing MYC overexpression with dinaciclib + pembrolizumab efficacy in advanced breast cancer [191].
Aurora Kinases: Given the similarities and genetic overlap between Aurora kinases A and B, many aurora kinase inhibitors are nonselective, other than Alisertib (MLN8237), which is specific to Aurora A and Barasertib (AZD1152), which is specific to Aurora B. VX-680 and AZD1152 have been studied specifically in MYC-overexpressing cancers preclinically. Yang et al. demonstrated proof of concept of synthetic lethality with non-specific aurora kinase inhibitor VX-680 [192]. Pulse treatment of VX-680 in Myc-driven models of lymphoma resulted in a 3-fold increase in survival. Yang and collogues propose synthetic lethality is a result of failed spindle checkpoint due to inhibition of Aurora B (resulting in a compromised CPPC) and MYC overexpression leads to polyploidy; the combination of the two proceeds to apoptosis. Yang and Goga both acknowledged defects in CPPC, through either survivin or Aurora B inhibition, which led to synthetic lethality in MYC-overexpressing cells. This suggests inhibition of other CPPC components as additional avenues to explore [176,192]. Furthermore, similar to Goga et al.’s findings, Yang et al. reported that the synthetic lethal interaction between Aurora kinase inhibitor and Myc is also independent of p53.
Lastly, another p53-deficient cancer, small cell lung cancer (SCLC), is also susceptible to Aurora kinase-Myc synthetic lethality. Helfirch et al. demonstrated MYC amplification is a good biomarker for predicting in vitro and in vivo growth inhibition of SCLC upon AZD1152 treatment [193]. Similarly, MYC-overexpressing medulloblastoma cells treated with AZD1152 were more sensitive to apoptosis than the low-expressing parent cell line, as was true for medulloblastoma cells with endogenous MYC overexpression [183]. Medulloblastoma xenograft models treated with AZD1152 had decreased tumor growth and prolonged survival.
Several Aurora kinase inhibitors have been tested clinically, including MK-0457 (or VX-680), AZD1152, PHA-739358, and MLN8237 [180]. Clinical trials that included MYC expression as a biomarker appear to be limited to Aurora A inhibitor Alisertib [194,195]. Recent in vivo studies of Alisterib in MYC-overexpressing lymphoma xenografts demonstrated synthetic lethality by caspase-independent cell death and complete tumor regression when paired with chemotherapy cyclophosphamide [196].
In vivo reports of CDK1 and Aurora kinase inhibitors support the inclusion of MYC overexpression as a recruitment factor in clinical trials or at least warrant further investigation of MYC a biomarker for these inhibitors. This stratification may result in an improved clinical outcome. Additional potential synthetic lethal interactions in Myc-driven cancers are described by Cermelli et al. [197].

5. Myc Drives Metabolism through Its Target Genes

Cellular proliferation is closely related to metabolism. In cancer, metabolic reprogramming, such as prompt ATP synthesis, increased anabolism of macromolecules, and redox homeostasis, support the rapidly proliferating cancer cells [198]. Understandably, Myc, the master regulator of growth, also aids in metabolic reprogramming. Several Myc target genes are involved in metabolic pathways, including glucose transporter GLUT1 (SLC2A1), glutaminase (GLS), hexokinase 2 (HK2), phosphofructokinase (PFKM), enolase 1 (ENO1), peroxisome proliferator-activated receptor gamma coactivator 1-beta (PGC-1β), nuclear respiratory factor 1 (NRF1), and inosine monophosphate dehydrogenase (IMPDH1/2) [199,200]. Through Myc’s many target genes, it can regulate aerobic glycolysis (the Warburg effect), mitochondria, and ribosome biogenesis, and metabolism of nucleotides, amino acids, and lipids, which contributes to its oncogenic function [198,199].
Cancer cells have increased levels of guanosine triphosphate (GTP), an energy source and signaling molecule [201]. Both GTP and the rate-limiting enzyme for GTP synthesis, inosine monophosphate dehydrogenase (IMPDH), are Myc target genes [202]. MYC and IMPDH expression significantly correlated and IMPDH overexpression has been observed in several cancers, including glioblastoma, leukemia, colorectal cancer, and small cell lung cancer [203,204,205,206,207]. Manipulating IMPDH expression in glioblastoma cells results in the same change in MYC expression [207]. Furthermore, Myc activates GTP synthesis and it has been shown in small cell lung cancer (SCLC), that IMPDH is depended upon by naïve and chemoresistant high-MYC SCLC cells [208]. In addition, IMPDH links Myc’s role in nucleotide biosynthesis and ribosome biogenesis as IMPDH-dependent GTP synthesis is needed for Pol I synthesis of pre-ribosomal RNA [208].
Glutaminase, another target gene of Myc, converts the abundant glutamine into glutamate. Myc-driven cancers depend on glutamine metabolism rather than glucose, especially when deprived of oxygen [209,210,211]. Furthermore, glutamine-depletion-induced apoptosis is dependent on Myc activity [212], and on the other hand, Myc-induced renal adenocarcinoma depends on glutaminase [213]. Shen et al. explored the concept of glutaminase dependency in the context of ovarian cancer, in which >45% of patients’ tumors overexpress MYC. Elevated MYC expression correlated with glutaminase in immortalized cell lines and primary cultures, and overexpression of MYC and GLS were associated with chemoresistance and worse disease outcome [214].
We will discuss inhibiting the proteins of metabolic Myc target genes, IMPDH and glutaminase, along with epigenetic cofactor WDR5 that recruits Myc to drive ribosome biogenesis (Figure 7). Details of Myc’s role in cancer metabolism are beyond the scope of this review and are summarized in several articles [17,198,199,215]. Additional potential Myc-driven metabolic targets are described by Dong et al. [198].

Targeting Metabolism Through Myc Target Genes and Cofactors

Myc’s effects on metabolism are highlighted by the negative impact of various metabolic inhibitors in Myc-driven cancer models. First, there are two clinically available IMPDH inhibitors: mycophenolic acid (MPA) and mizoribine. Both are clinically used as an immunosuppressants to prevent organ transplant rejection, but additional research is necessary to explore their anticancer properties. MPA appears to primarily serve as an anticancer tool; it has been preclinically tested in several cancers, but has dose-limiting toxicity due to gastro-intestinal side effects [216,217,218]. However, mizoribine has higher tolerability [219]. Studies connected the anticancer efficacy of IMPDH inhibitors to Myc; it appears that Myc is needed for the antiangiogenic properties of MPA [218]. Huang et al. show Myc-driven SCLC and hepatoblastoma demonstrate sensitivity to mizoribine and MPA, confirming a dependence on IMPDH. Importantly, mizoribine was capable of decreasing tumor growth in immunocompetent mice, despite its immunosuppressive properties [208]. This study warrants further testing and development of these clinically-available IMPDH inhibitors in Myc-driven cancer models.
A second approach is targeting glutaminase in MYC-overexpressing cells. The concept of pharmacologically inhibiting glutaminase has been discussed since 1975 [220], but there were concerns of targeting a major metabolic component. Allosteric glutaminase inhibitor Bis-2-(5-phenylacetamido-1,2,4-thiadiazol-2-yl)ethyl sulfide (BPTES) has been studied extensively. Treatment with BPTES increases reactive oxygen species production and hinders cell bioenergetics, leading to cell death [211]. In an in vivo renal adenocarcinoma model, BPTES reduced tumor growth by 32% [213]. However, modifications must be made to BPTES to improve its therapeutic potential as it has a moderate potency and poor solubility [221]. The newest glutaminase inhibitor, CB-839, is a BPTES derivative and is currently in Phase II clinical trials for several cancers including colorectal cancer, acute myeloid leukemia, and triple negative breast cancer. Shen et al. demonstrated preclinically that ovarian tumor xenografts treated with CB-839 resulted in increased sensitivity to PARP inhibitor olaparib as glutaminase inhibition led to replicative stress [214]. A Phase Ib/II study investigating CB-839 in combination with PARP inhibitor talazoparib for solid tumors is currently recruiting patients (NCT03875313) and a Phase I study combining CB-839 with PARP inhibitor niraparib in platinum resistance BRCA-wildtype ovarian cancer has posted (NCT03944902).
As mentioned, Myc’s role in metabolism primarily stems from expression of its target genes that are involved in metabolic pathways. The previously described Myc metabolic therapeutics inhibit Myc’s target genes, but not Myc’s activity. Thomas et al. discovered that epigenetic cofactor WDR5 recruits Myc to chromatin to promote expression of genes involved in biomass accumulation [222]. An inducible exon swap system in a Burkitt lymphoma cell line was created to study the interaction between Myc and WDR5 by implementing a mutant Myc that could not interact with WDR5. Inhibition of WDR5 prevented Myc’s function as a transcription factor by disrupting gene binding, which decreased transcription of translational machinery, including ribosome protein subunits and nucleolar RNAs. When the exon swap system was assessed in vivo, switching to the WDR5-interaction-defective Myc resulted in apoptosis, decreased tumor volume, and improved survival [222]. Thomas and collogues additionally reported that targeting the “WIN” site in WDR5 may also be a valuable target to displace Myc from chromatin. WDR5 inhibitors are currently being synthesized to further study the anticancer effects of disrupting the Myc-WDR5 interaction [223]. Targeting an epigenetic cofactor that aids in Myc’s activity will prevent Myc’s target gene from being transcribed; this route may be more beneficial than inhibiting already-transcribed genes under Myc’s control. This study opens an avenue outside the context of metabolism to explore other targetable Myc-interacting cofactors to prevent Myc binding to chromatin.

6. Conclusions

This review summarizes the main mechanisms by which c-Myc promotes tumorigenesis and the different therapeutic approaches that directly/indirectly target Myc. Broadly, we described inhibitors that prevent Myc’s actions as a transcription factor through altering Myc stability (transcription, dimerization, degradation), inducing synthetic lethality via cell cycle targets, and inhibiting Myc target genes involved in metabolism. There are other methods of disrupting Myc’s activity not listed here, including inhibiting Myc-recruited cofactors or epigenetic mechanisms [13,25,122]. Additionally, Myc’s robust control on microRNA expression is another area of interest [224,225,226], given Myc’s ability to activate oncogenic miRNAs and repress tumor suppressive miRNAs [225,227]. Modulating expression of the miRNAs to promote anticancer effects as a therapeutic option is being explored [228,229]. When clinically testing these therapeutics in Myc-driven cancers, it is important to consider patients’ MYC expression in the case of stratifying patients and identifying clinically relevant subgroup results. Furthermore, although this review was limited to well-studied c-Myc, the Myc family of c-Myc, N-Myc and L-Myc, can be functionally redundant, and therefore inhibition of the Myc family rather than one specific Myc may be required [230]. All in all, new discoveries improved our understanding of the “myc-anisms” behind Myc-driven cancers and enhanced the potential of targeting the “undruggable” Myc.

Author Contributions

Writing—original draft preparation, J.M.; writing—review and editing, A.D.; visualization, J.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

bHLHZipbasic helix-loop-helix leucine zipper
MaxMyc-associated factor X
Pol IIPolymerase II
BETBromodomain and Extra-Terminal motif
NMCNUT midline carcinoma
PROTACPROteolysis TArgeting Chimeric
CdkCyclin-dependent kinase
S62Serine 62
T58Threonine 58
GSK3Glycogen synthase kinase 3
ERKExtracellular signal-regulated kinase
PI3KPhosphoinositol 3-kinase
PP2AProtein phosphatase 2a
Pin1Peptidyl-prolyl cis-trans isomerase NIMA-interacting 1
ATRAAll-trans retinoic acid
APLAcute promyelocytic leukemia
RNAiRNA interference
CRISPRClustered Regularly Interspaced Short Palindromic Repeats
PARPPoly(ADP-ribose) polymerase
CPPCChromosomal Passenger Protein Complex
TNBCTriple Negative Breast Cancer
GTPGuanosine triphosphate
IMPDHInosine monophosphate dehydrogenase
SCLCSmall cell lung cancer
GlnGlutamine
GluGlutamate
MPAMycophenolic acid
BPTESBis-2-(5-phenylacetamido-1,2,4-thiadiazol-2-yl)ethyl sulfide

References

  1. Nau, M.M.; Brooks, B.J.; Battey, J.; Sausville, E.; Gazdar, A.F.; 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] [PubMed]
  2. Beltran, H. The N-myc Oncogene: Maximizing its Targets, Regulation, and Therapeutic Potential. Mol. Cancer Res. 2014, 12, 815–822. [Google Scholar] [CrossRef] [Green Version]
  3. Fernandez, P.C. Genomic targets of the human c-Myc protein. Genes Dev. 2003, 17, 1115–1129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Li, Z.; Van Calcar, S.; Qu, C.; Cavenee, W.K.; Zhang, M.Q.; Ren, B. A global transcriptional regulatory role for c-Myc in Burkitt’s lymphoma cells. Proc. Natl. Acad. Sci. USA 2003, 100, 8164–8169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Lüscher, B.; Vervoorts, J. Regulation of gene transcription by the oncoprotein MYC. Gene 2012, 494, 145–160. [Google Scholar] [CrossRef]
  6. Kanazawa, S.; Soucek, L.; Evan, G.; Okamoto, T.; Peterlin, B.M. c-Myc recruits P-TEFb for transcription, cellular proliferation and apoptosis. Oncogene 2003, 22, 5707–5711. [Google Scholar] [CrossRef] [Green Version]
  7. Xie, X.; Lu, J.; Kulbokas, E.J.; Golub, T.R.; Mootha, V.; Lindblad-Toh, K.; Lander, E.S.; Kellis, M. Systematic discovery of regulatory motifs in human promoters and 3′ UTRs by comparison of several mammals. Nature 2005, 434, 338–345. [Google Scholar] [CrossRef]
  8. Dang, C.V. MYC on the Path to Cancer. Cell 2012, 149, 22–35. [Google Scholar] [CrossRef] [Green Version]
  9. Ji, H.; Wu, G.; Zhan, X.; Nolan, A.; Koh, C.; De Marzo, A.; Doan, H.M.; Fan, J.; Cheadle, C.; Fallahi, M.; et al. Cell-Type Independent MYC Target Genes Reveal a Primordial Signature Involved in Biomass Accumulation. PLoS ONE 2011, 6, e26057. [Google Scholar] [CrossRef]
  10. Dang, C.V.; O’Donnell, K.A.; Zeller, K.I.; Nguyen, T.; Osthus, R.C.; Li, F. The c-Myc target gene network. Semin. Cancer Biol. 2006, 16, 253–264. [Google Scholar] [CrossRef]
  11. Zeller, K.I.; Zhao, X.; Lee, C.W.H.; Chiu, K.P.; Yao, F.; Yustein, J.T.; Ooi, H.S.; Orlov, Y.L.; Shahab, A.; Yong, H.C.; et al. Global mapping of c-Myc binding sites and target gene networks in human B cells. Proc. Natl. Acad. Sci. USA 2006, 103, 17834–17839. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Yap, C.-S.; Peterson, A.L.; Castellani, G.; Sedivy, J.M.; Neretti, N. Kinetic profiling of the c-Myc transcriptome and bioinformatic analysis of repressed gene promoters. Cell Cycle 2011, 10, 2184–2196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Caforio, M.; Sorino, C.; Iacovelli, S.; Fanciulli, M.; Locatelli, F.; Folgiero, V. Recent advances in searching c-Myc transcriptional cofactors during tumorigenesis. J. Exp. Clin. Cancer Res. 2018, 37, 239. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Herkert, B.; Eilers, M. Transcriptional Repression: The Dark Side of Myc. Genes Cancer 2010, 1, 580–586. [Google Scholar] [CrossRef] [Green Version]
  15. Nie, Z.; Hu, G.; Wei, G.; Cui, K.; Yamane, A.; Resch, W.; Wang, R.; Green, D.R.; Tessarollo, L.; Casellas, R.; et al. c-Myc Is a Universal Amplifier of Expressed Genes in Lymphocytes and Embryonic Stem Cells. Cell 2012, 151, 68–79. [Google Scholar] [CrossRef] [Green Version]
  16. Lin, C.Y.; Lovén, 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]
  17. Dang, C.V. MYC, Metabolism, Cell Growth, and Tumorigenesis. Cold Spring Harb. Perspect. Med. 2013, 3, a014217. [Google Scholar] [CrossRef]
  18. Sabò, 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]
  19. 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]
  20. Blackwood, E.; Eisenman, R. Max: A helix-loop-helix zipper protein that forms a sequence-specific DNA-binding complex with Myc. Science 1991, 251, 1211–1217. [Google Scholar] [CrossRef]
  21. Grinberg, A.V.; Hu, C.-D.; Kerppola, T.K. Visualization of Myc/Max/Mad Family Dimers and the Competition for Dimerization in Living Cells. Mol. Cell. Biol. 2004, 24, 4294–4308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Amati, B.; Dalton, S.; Brooks, M.W.; Littlewood, T.D.; Evan, G.I.; Land, H. Transcriptional activation by the human c-Myc oncoprotein in yeast requires interaction with Max. Nature 1992, 359, 423–426. [Google Scholar] [CrossRef] [PubMed]
  23. Kato, G.J.; Lee, W.M.; Chen, L.L.; Dang, C.V. Max: Functional domains and interaction with c-Myc. Genes Dev. 1992, 6, 81–92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Kretzner, L.; Blackwood, E.M.; Eisenman, R.N. Myc and Max proteins possess distinct transcriptional activities. Nature 1992, 359, 426–429. [Google Scholar] [CrossRef] [PubMed]
  25. Vita, M.; Henriksson, M. The Myc oncoprotein as a therapeutic target for human cancer. Semin. Cancer Biol. 2006, 16, 318–330. [Google Scholar] [CrossRef] [PubMed]
  26. Vervoorts, J.; Lüscher-Firzlaff, J.; Lüscher, B. The Ins and Outs of MYC Regulation by Posttranslational Mechanisms. J. Biol. Chem. 2006, 281, 34725–34729. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Ramsay, G.; Evan, G.I.; Bishop, J.M. The protein encoded by the human proto-oncogene c-myc. Proc. Natl. Acad. Sci. USA 1984, 81, 7742–7746. [Google Scholar] [CrossRef] [Green Version]
  28. Yeh, E.; Cunningham, M.; Arnold, H.; Chasse, D.; Monteith, T.; Ivaldi, G.; Hahn, W.C.; Stukenberg, P.T.; Shenolikar, S.; Uchida, T.; et al. A signalling pathway controlling c-Myc degradation that impacts oncogenic transformation of human cells. Nat. Cell Biol. 2004, 6, 308–318. [Google Scholar] [CrossRef]
  29. Salghetti, S.E. Destruction of Myc by ubiquitin-mediated proteolysis: Cancer-associated and transforming mutations stabilize Myc. EMBO J. 1999, 18, 717–726. [Google Scholar] [CrossRef] [Green Version]
  30. Meyer, N.; Penn, L.Z. Reflecting on 25 years with MYC. Nat. Rev. Cancer 2008, 8, 976–990. [Google Scholar] [CrossRef]
  31. Kalkat, M.; De Melo, J.; Hickman, K.; Lourenco, C.; Redel, C.; Resetca, D.; Tamachi, A.; Tu, W.; Penn, L. MYC Deregulation in Primary Human Cancers. Genes 2017, 8, 151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Gabay, M.; Li, Y.; Felsher, D.W. MYC Activation Is a Hallmark of Cancer Initiation and Maintenance. Cold Spring Harb. Perspect. Med. 2014, 4, a014241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Jain, M. Sustained Loss of a Neoplastic Phenotype by Brief Inactivation of MYC. Science 2002, 297, 102–104. [Google Scholar] [CrossRef] [PubMed]
  34. Pelengaris, S.; Littlewood, T.; Khan, M.; Elia, G.; Evan, G. Reversible Activation of c-Myc in Skin: Induction of a Complex Neoplastic Phenotype by a Single Oncogenic Lesion. Mol. Cell 1999, 3, 565–577. [Google Scholar] [CrossRef]
  35. Felsher, D.W.; Bishop, J.M. Reversible Tumorigenesis by MYC in Hematopoietic Lineages. Mol. Cell 1999, 4, 199–207. [Google Scholar] [CrossRef]
  36. Schaub, F.X.; Dhankani, V.; Berger, A.C.; Trivedi, M.; Richardson, A.B.; Shaw, R.; Zhao, W.; Zhang, X.; Ventura, A.; Liu, Y.; et al. Pan-cancer Alterations of the MYC Oncogene and Its Proximal Network across the Cancer Genome Atlas. Cell Syst. 2018, 6, 282–300.e2. [Google Scholar] [CrossRef] [Green Version]
  37. Boxer, L.M.; Dang, C.V. Translocations involving c-myc and c-myc function. Oncogene 2001, 20, 5595–5610. [Google Scholar] [CrossRef] [Green Version]
  38. Malempati, S.; Tibbitts, D.; Cunningham, M.; Akkari, Y.; Olson, S.; Fan, G.; Sears, R.C. Aberrant stabilization of c-Myc protein in some lymphoblastic leukemias. Leukemia 2006, 20, 1572–1581. [Google Scholar] [CrossRef] [Green Version]
  39. Zhang, X.; Farrell, A.S.; Daniel, C.J.; Arnold, H.; Scanlan, C.; Laraway, B.J.; Janghorban, M.; Lum, L.; Chen, D.; Troxell, M.; et al. Mechanistic insight into Myc stabilization in breast cancer involving aberrant Axin1 expression. Proc. Natl. Acad. Sci. USA 2012, 109, 2790–2795. [Google Scholar] [CrossRef] [Green Version]
  40. Farrell, A.S.; Allen-Petersen, B.; Daniel, C.J.; Wang, X.; Wang, Z.; Rodriguez, S.; Impey, S.; Oddo, J.; Vitek, M.P.; Lopez, C.; et al. Targeting Inhibitors of the Tumor Suppressor PP2A for the Treatment of Pancreatic Cancer. Mol. Cancer Res. 2014, 12, 924–939. [Google Scholar] [CrossRef] [Green Version]
  41. Janghorban, M.; Farrell, A.S.; Allen-Petersen, B.L.; Pelz, C.; Daniel, C.J.; Oddo, J.; Langer, E.M.; Christensen, D.J.; Sears, R.C. Targeting c-MYC by antagonizing PP2A inhibitors in breast cancer. Proc. Natl. Acad. Sci. USA 2014, 111, 9157–9162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Soucek, L.; Whitfield, J.R.; Sodir, N.M.; Masso-Valles, D.; Serrano, E.; Karnezis, A.N.; Swigart, L.B.; Evan, G.I. Inhibition of Myc family proteins eradicates KRas-driven lung cancer in mice. Genes Dev. 2013, 27, 504–513. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Annibali, D.; Whitfield, J.R.; Favuzzi, E.; Jauset, T.; Serrano, E.; Cuartas, I.; Redondo-Campos, S.; Folch, G.; Gonzàlez-Juncà, A.; Sodir, N.M.; et al. Myc inhibition is effective against glioma and reveals a role for Myc in proficient mitosis. Nat. Commun. 2014, 5, 4632. [Google Scholar] [CrossRef] [PubMed]
  44. Davis, A.C.; Wims, M.; Spotts, G.D.; Hann, S.R.; Bradley, A. A null c-myc mutation causes lethality before 10.5 days of gestation in homozygotes and reduced fertility in heterozygous female mice. Genes Dev. 1993, 7, 671–682. [Google Scholar] [CrossRef] [Green Version]
  45. Sammak, S.; Hamdani, N.; Gorrec, F.; Allen, M.D.; Freund, S.M.V.; Bycroft, M.; Zinzalla, G. Crystal Structures and Nuclear Magnetic Resonance Studies of the Apo Form of the c-MYC:MAX bHLHZip Complex Reveal a Helical Basic Region in the Absence of DNA. Biochemistry 2019, 58, 3144–3154. [Google Scholar] [CrossRef] [Green Version]
  46. 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]
  47. Marshall, N.F.; Price, D.H. Control of formation of two distinct classes of RNA polymerase II elongation complexes. Mol. Cell. Biol. 1992, 12, 2078–2090. [Google Scholar] [CrossRef] [Green Version]
  48. Marshall, N.F.; Peng, J.; Xie, Z.; Price, D.H. Control of RNA Polymerase II Elongation Potential by a Novel Carboxyl-terminal Domain Kinase. J. Biol. Chem. 1996, 271, 27176–27183. [Google Scholar] [CrossRef] [Green Version]
  49. Fu, T.-J.; Peng, J.; Lee, G.; Price, D.H.; Flores, O. Cyclin K Functions as a CDK9 Regulatory Subunit and Participates in RNA Polymerase II Transcription. J. Biol. Chem. 1999, 274, 34527–34530. [Google Scholar] [CrossRef] [Green Version]
  50. Yang, Z.; Yik, J.H.N.; Chen, R.; He, N.; Jang, M.K.; Ozato, K.; Zhou, Q. Recruitment of P-TEFb for Stimulation of Transcriptional Elongation by the Bromodomain Protein Brd4. Mol. Cell 2005, 19, 535–545. [Google Scholar] [CrossRef]
  51. 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]
  52. 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, 201919507. [Google Scholar] [CrossRef] [PubMed]
  53. Huang, C.-H.; Lujambio, A.; Zuber, J.; Tschaharganeh, D.F.; Doran, M.G.; Evans, M.J.; Kitzing, T.; Zhu, N.; de Stanchina, E.; Sawyers, C.L.; et al. CDK9-mediated transcription elongation is required for MYC addiction in hepatocellular carcinoma. Genes Dev. 2014, 28, 1800–1814. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Zhang, H.; Pandey, S.; Travers, M.; Sun, H.; Morton, G.; Madzo, J.; Chung, W.; Khowsathit, J.; Perez-Leal, O.; Barrero, C.A.; et al. Targeting CDK9 Reactivates Epigenetically Silenced Genes in Cancer. Cell 2018, 175, 1244–1258.e26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Schuijers, J.; Manteiga, J.C.; Weintraub, A.S.; Day, D.S.; Zamudio, A.V.; Hnisz, D.; Lee, T.I.; Young, R.A. Transcriptional Dysregulation of MYC Reveals Common Enhancer-Docking Mechanism. Cell Rep. 2018, 23, 349–360. [Google Scholar] [CrossRef] [Green Version]
  56. Muller, S.; Filippakopoulos, P.; Knapp, S. Bromodomains as therapeutic targets. Expert Rev. Mol. Med. 2011, 13, e29. [Google Scholar] [CrossRef] [Green Version]
  57. Devaiah, B.N.; Case-Borden, C.; Gegonne, A.; Hsu, C.H.; Chen, Q.; Meerzaman, D.; Dey, A.; Ozato, K.; Singer, D.S. BRD4 is a histone acetyltransferase that evicts nucleosomes from chromatin. Nat. Struct. Mol. Biol. 2016, 23, 540–548. [Google Scholar] [CrossRef] [Green Version]
  58. Dey, A.; Nishiyama, A.; Karpova, T.; McNally, J.; Ozato, K. Brd4 Marks Select Genes on Mitotic Chromatin and Directs Postmitotic Transcription. Mol. Biol. Cell 2009, 20, 4899–4909. [Google Scholar] [CrossRef] [Green Version]
  59. Donati, B.; Lorenzini, E.; Ciarrocchi, A. BRD4 and Cancer: Going beyond transcriptional regulation. Mol. Cancer 2018, 17, 164. [Google Scholar] [CrossRef]
  60. Filippakopoulos, P.; Qi, J.; Picaud, S.; Shen, Y.; Smith, W.B.; Fedorov, O.; Morse, E.M.; Keates, T.; Hickman, T.T.; Felletar, I.; et al. Selective inhibition of BET bromodomains. Nature 2010, 468, 1067–1073. [Google Scholar] [CrossRef] [Green Version]
  61. Nicodeme, E.; Jeffrey, K.L.; Schaefer, U.; Beinke, S.; Dewell, S.; Chung, C.; Chandwani, R.; Marazzi, I.; Wilson, P.; Coste, H.; et al. Suppression of inflammation by a synthetic histone mimic. Nature 2010, 468, 1119–1123. [Google Scholar] [CrossRef] [PubMed]
  62. Ott, C.J.; Kopp, N.; Bird, L.; Paranal, R.M.; Qi, J.; Bowman, T.; Rodig, S.J.; Kung, A.L.; Bradner, J.E.; Weinstock, D.M. BET bromodomain inhibition targets both c-Myc and IL7R in high-risk acute lymphoblastic leukemia. Blood 2012, 120, 2843–2852. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Mertz, J.A.; Conery, A.R.; Bryant, B.M.; Sandy, P.; Balasubramanian, S.; Mele, D.A.; Bergeron, L.; Sims, R.J. Targeting MYC dependence in cancer by inhibiting BET bromodomains. Proc. Natl. Acad. Sci. USA 2011, 108, 16669–16674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Chapuy, B.; McKeown, M.R.; Lin, C.Y.; Monti, S.; Roemer, M.G.M.; Qi, J.; Rahl, P.B.; Sun, H.H.; Yeda, K.T.; Doench, J.G.; et al. Discovery and Characterization of Super-Enhancer-Associated Dependencies in Diffuse Large B Cell Lymphoma. Cancer Cell 2013, 24, 777–790. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Shimamura, T.; Chen, Z.; Soucheray, M.; Carretero, J.; Kikuchi, E.; Tchaicha, J.H.; Gao, Y.; Cheng, K.A.; Cohoon, T.J.; Qi, J.; et al. Efficacy of BET Bromodomain Inhibition in Kras-Mutant Non-Small Cell Lung Cancer. Clin. Cancer Res. 2013, 19, 6183–6192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Sakaguchi, T.; Yoshino, H.; Sugita, S.; Miyamoto, K.; Yonemori, M.; Osako, Y.; Meguro-Horike, M.; Horike, S.-I.; Nakagawa, M.; Enokida, H. Bromodomain protein BRD4 inhibitor JQ1 regulates potential prognostic molecules in advanced renal cell carcinoma. Oncotarget 2018, 9, 23003–23017. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Qin, Z.; Wang, T.; Su, S.; Shen, L.; Zhu, G.; Liu, Q.; Zhang, L.; Liu, K.; Zhang, Y.; Zhou, Z.; et al. BRD4 promotes gastric cancer progression and metastasis through acetylation-dependent stabilization of Snail. Cancer Res. 2019, 79, 4869–4881. [Google Scholar] [CrossRef] [PubMed]
  68. Zhu, X.; Enomoto, K.; Zhao, L.; Zhu, Y.J.; Willingham, M.C.; Meltzer, P.; Qi, J.; Cheng, S. Bromodomain and Extraterminal Protein Inhibitor JQ1 Suppresses Thyroid Tumor Growth in a Mouse Model. Clin. Cancer Res. 2017, 23, 430–440. [Google Scholar] [CrossRef] [Green Version]
  69. Alqahtani, A.; Choucair, K.; Ashraf, M.; Hammouda, D.M.; Alloghbi, A.; Khan, T.; Senzer, N.; Nemunaitis, J. Bromodomain and extra-terminal motif inhibitors: A review of preclinical and clinical advances in cancer therapy. Future Sci. OA 2019, 5, FSO372. [Google Scholar] [CrossRef] [Green Version]
  70. Andrieu, G.; Belkina, A.C.; Denis, G.V. Clinical trials for BET inhibitors run ahead of the science. Drug Discov. Today Technol. 2016, 19, 45–50. [Google Scholar] [CrossRef] [Green Version]
  71. Xu, Y.; Vakoc, C.R. Targeting Cancer Cells with BET Bromodomain Inhibitors. Cold Spring Harb. Perspect. Med. 2017, 7, a026674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Stathis, A.; Zucca, E.; Bekradda, M.; Gomez-Roca, C.; Delord, J.-P.; de La Motte Rouge, T.; Uro-Coste, E.; de Braud, F.; Pelosi, G.; French, C.A. Clinical Response of Carcinomas Harboring the BRD4–NUT Oncoprotein to the Targeted Bromodomain Inhibitor OTX015/MK-8628. Cancer Discov. 2016, 6, 492–500. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Sakamoto, K.M.; Kim, K.B.; Kumagai, A.; Mercurio, F.; Crews, C.M.; Deshaies, R.J. Protacs: Chimeric molecules that target proteins to the Skp1-Cullin-F box complex for ubiquitination and degradation. Proc. Natl. Acad. Sci. USA 2001, 98, 8554–8559. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Sun, X.; Gao, H.; Yang, Y.; He, M.; Wu, Y.; Song, Y.; Tong, Y.; Rao, Y. PROTACs: Great opportunities for academia and industry. Signal Transduct. Target. Ther. 2019, 4, 64. [Google Scholar] [CrossRef] [Green Version]
  75. Petrylak, D.P.; Gao, X.; Vogelzang, N.J.; Garfield, M.H.; Taylor, I.; Moore, M.D.; Peck, R.A.; Burris, H.A. First-in-human phase I study of ARV-110, an androgen receptor (AR) PROTAC degrader in patients (pts) with metastatic castrate-resistant prostate cancer (mCRPC) following enzalutamide (ENZ) and/or abiraterone (ABI). J. Clin. Oncol. 2020, 38. [Google Scholar] [CrossRef]
  76. Winter, G.E.; Buckley, D.L.; Paulk, J.; Roberts, J.M.; Souza, A.; Dhe-Paganon, S.; Bradner, J.E. Phthalimide conjugation as a strategy for in vivo target protein degradation. Science 2015, 348, 1376–1381. [Google Scholar] [CrossRef] [Green Version]
  77. Zengerle, M.; Chan, K.-H.; Ciulli, A. Selective Small Molecule Induced Degradation of the BET Bromodomain Protein BRD4. ACS Chem. Biol. 2015, 10, 1770–1777. [Google Scholar] [CrossRef] [Green Version]
  78. Yang, C.-Y.; Qin, C.; Bai, L.; Wang, S. Small-molecule PROTAC degraders of the Bromodomain and Extra Terminal (BET) proteins—A review. Drug Discov. Today Technol. 2019, 31, 43–51. [Google Scholar] [CrossRef]
  79. Zhang, F.; Wu, Z.; Chen, P.; Zhang, J.; Wang, T.; Zhou, J.; Zhang, H. Discovery of a new class of PROTAC BRD4 degraders based on a dihydroquinazolinone derivative and lenalidomide/pomalidomide. Bioorg. Med. Chem. 2020, 28, 115228. [Google Scholar] [CrossRef]
  80. Shi, C.; Zhang, H.; Wang, P.; Wang, K.; Xu, D.; Wang, H.; Yin, L.; Zhang, S.; Zhang, Y. PROTAC induced-BET protein degradation exhibits potent anti-osteosarcoma activity by triggering apoptosis. Cell Death Dis. 2019, 10, 815. [Google Scholar] [CrossRef] [Green Version]
  81. Stathis, A.; Bertoni, F. BET Proteins as Targets for Anticancer Treatment. Cancer Discov. 2018, 8, 24–36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Morgado-Pascual, J.L.; Rayego-Mateos, S.; Tejedor, L.; Suarez-Alvarez, B.; Ruiz-Ortega, M. Bromodomain and Extraterminal Proteins as Novel Epigenetic Targets for Renal Diseases. Front. Pharmacol. 2019, 10, 1315. [Google Scholar] [CrossRef] [PubMed]
  83. Shu, S.; Polyak, K. BET Bromodomain Proteins as Cancer Therapeutic Targets. Cold Spring Harb. Symp. Quant. Biol. 2016, 81, 123–129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Cochran, A.G.; Conery, A.R.; Sims, R.J. Bromodomains: A new target class for drug development. Nat. Rev. Drug Discov. 2019, 18, 609–628. [Google Scholar] [CrossRef] [PubMed]
  85. Garcia-Cuellar, M.-P.; Füller, E.; Mäthner, E.; Breitinger, C.; Hetzner, K.; Zeitlmann, L.; Borkhardt, A.; Slany, R.K. Efficacy of cyclin-dependent-kinase 9 inhibitors in a murine model of mixed-lineage leukemia. Leukemia 2014, 28, 1427–1435. [Google Scholar] [CrossRef] [PubMed]
  86. Hashiguchi, T.; Bruss, N.; Best, S.; Lam, V.; Danilova, O.; Paiva, C.J.; Wolf, J.; Gilbert, E.W.; Okada, C.Y.; Kaur, P.; et al. Cyclin-Dependent Kinase-9 Is a Therapeutic Target in MYC-Expressing Diffuse Large B-Cell Lymphoma. Mol. Cancer Ther. 2019, 18, 1520–1532. [Google Scholar] [CrossRef]
  87. Boffo, S.; Damato, A.; Alfano, L.; Giordano, A. CDK9 inhibitors in acute myeloid leukemia. J. Exp. Clin. Cancer Res. 2018, 37, 36. [Google Scholar] [CrossRef] [Green Version]
  88. Blake, D.R.; Vaseva, A.V.; Hodge, R.G.; Kline, M.P.; Gilbert, T.S.K.; Tyagi, V.; Huang, D.; Whiten, G.C.; Larson, J.E.; Wang, X.; et al. Application of a MYC degradation screen identifies sensitivity to CDK9 inhibitors in KRAS-mutant pancreatic cancer. Sci. Signal. 2019, 12, eaav7259. [Google Scholar] [CrossRef]
  89. Olson, C.M.; Jiang, B.; Erb, M.A.; Liang, Y.; Doctor, Z.M.; Zhang, Z.; Zhang, T.; Kwiatkowski, N.; Boukhali, M.; Green, J.L.; et al. Pharmacological perturbation of CDK9 using selective CDK9 inhibition or degradation. Nat. Chem. Biol. 2018, 14, 163–170. [Google Scholar] [CrossRef]
  90. Asghar, U.; Witkiewicz, A.K.; Turner, N.C.; Knudsen, E.S. The history and future of targeting cyclin-dependent kinases in cancer therapy. Nat. Rev. Drug Discov. 2015, 14, 130–146. [Google Scholar] [CrossRef] [Green Version]
  91. Lücking, U.; Scholz, A.; Lienau, P.; Siemeister, G.; Kosemund, D.; Bohlmann, R.; Briem, H.; Terebesi, I.; Meyer, K.; Prelle, K.; et al. Identification of Atuveciclib (BAY 1143572), the First Highly Selective, Clinical PTEFb/CDK9 Inhibitor for the Treatment of Cancer. ChemMedChem 2017, 12, 1776–1793. [Google Scholar] [CrossRef] [PubMed]
  92. Cidado, J.; Boiko, S.; Proia, T.; Ferguson, D.; Criscione, S.W.; San Martin, M.; Pop-Damkov, P.; Su, N.; Roamio Franklin, V.N.; Chilamakuri, C.S.R.; et al. AZD4573 Is a Highly Selective CDK9 Inhibitor That Suppresses MCL-1 and Induces Apoptosis in Hematologic Cancer Cells. Clin. Cancer Res. 2020, 26, 922–934. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Moreno, N.; Holsten, T.; Mertins, J.; Zhogbi, A.; Johann, P.; Kool, M.; Meisterernst, M.; Kerl, K. Combined BRD4 and CDK9 inhibition as a new therapeutic approach in malignant rhabdoid tumors. Oncotarget 2017, 8, 84986–84995. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. McCalmont, H.; Li, K.L.; Jones, L.; Toubia, J.; Bray, S.C.; Casolari, D.A.; Mayoh, C.; Samaraweera, S.E.; Lewis, I.D.; Prinjha, R.K.; et al. Efficacy of combined CDK9/BET inhibition in preclinical models of MLL-rearranged acute leukemia. Blood Adv. 2020, 4, 296–300. [Google Scholar] [CrossRef] [PubMed]
  95. Gerlach, D.; Tontsch-Grunt, U.; Baum, A.; Popow, J.; Scharn, D.; Hofmann, M.H.; Engelhardt, H.; Kaya, O.; Beck, J.; Schweifer, N.; et al. The novel BET bromodomain inhibitor BI 894999 represses super-enhancer-associated transcription and synergizes with CDK9 inhibition in AML. Oncogene 2018, 37, 2687–2701. [Google Scholar] [CrossRef] [PubMed]
  96. Carvalho, J.; Mergny, J.-L.; Salgado, G.F.; Queiroz, J.A.; Cruz, C. G-quadruplex, Friend or Foe: The Role of the G-quartet in Anticancer Strategies. Trends Mol. Med. 2020, 26, 848–861. [Google Scholar] [CrossRef]
  97. Brooks, T.A.; Hurley, L.H. Targeting MYC Expression through G-Quadruplexes. Genes Cancer 2010, 1, 641–649. [Google Scholar] [CrossRef]
  98. Amati, B.; Brooks, M.W.; Naomi, L.; Littlewood, T.D.; Evan, G.I.; Land, H. Oncogenic Activity of the c-Myc Protein Requires Dimierzation with Max. Cell 1993, 72, 233–245. [Google Scholar] [CrossRef]
  99. Carabet, L.A.; Lallous, N.; Leblanc, E.; Ban, F.; Morin, H.; Lawn, S.; Ghaidi, F.; Lee, J.; Mills, I.G.; Gleave, M.E.; et al. Computer-aided drug discovery of Myc-Max inhibitors as potential therapeutics for prostate cancer. Eur. J. Med. Chem. 2018, 160, 108–119. [Google Scholar] [CrossRef] [Green Version]
  100. Posternak, V.; Cole, M.D. Strategically targeting MYC in cancer. F1000Research 2016, 5, 408. [Google Scholar] [CrossRef]
  101. Soucek, L.; Helmer-Citterich, M.; Sacco, A.; Jucker, R.; Cesareni, G.; Nasi, S. Design and properties of a Myc derivative that efficiently homodimerizes. Oncogene 1998, 17, 2463–2472. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Demma, M.J.; Mapelli, C.; Sun, A.; Bodea, S.; Ruprecht, B.; Javaid, S.; Wiswell, D.; Muise, E.; Chen, S.; Zelina, J.; et al. Omomyc Reveals New Mechanisms to Inhibit the MYC Oncogene. Mol. Cell. Biol. 2019, 39, e00248-19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Beaulieu, M.-E.; Jauset, T.; Massó-Vallés, D.; Martínez-Martín, S.; Rahl, P.; Maltais, L.; Zacarias-Fluck, M.F.; Casacuberta-Serra, S.; Serrano del Pozo, E.; Fiore, C.; et al. Intrinsic cell-penetrating activity propels Omomyc from proof of concept to viable anti-MYC therapy. Sci. Transl. Med. 2019, 11, eaar5012. [Google Scholar] [CrossRef] [PubMed]
  104. Massó-Vallés, D.; Soucek, L. Blocking Myc to Treat Cancer: Reflecting on Two Decades of Omomyc. Cells 2020, 9, 883. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Jung, L.A.; Gebhardt, A.; Koelmel, W.; Ade, C.P.; Walz, S.; Kuper, J.; von Eyss, B.; Letschert, S.; Redel, C.; d’Artista, L.; et al. OmoMYC blunts promoter invasion by oncogenic MYC to inhibit gene expression characteristic of MYC-dependent tumors. Oncogene 2017, 36, 1911–1924. [Google Scholar] [CrossRef] [Green Version]
  106. Soucek, L.; Whitfield, J.; Martins, C.P.; Finch, A.J.; Murphy, D.J.; Sodir, N.M.; Karnezis, A.N.; Swigart, L.B.; Nasi, S.; Evan, G.I. Modelling Myc inhibition as a cancer therapy. Nature 2008, 455, 679–683. [Google Scholar] [CrossRef] [Green Version]
  107. 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]
  108. Wang, E.; Sorolla, A.; Cunningham, P.T.; Bogdawa, H.M.; Beck, S.; Golden, E.; Dewhurst, R.E.; Florez, L.; Cruickshank, M.N.; Hoffmann, K.; et al. Tumor penetrating peptides inhibiting MYC as a potent targeted therapeutic strategy for triple-negative breast cancers. Oncogene 2019, 38, 140–150. [Google Scholar] [CrossRef]
  109. Berg, T.; Cohen, S.B.; Desharnais, J.; Sonderegger, C.; Maslyar, D.J.; Goldberg, J.; Boger, D.L.; Vogt, P.K. Small-molecule antagonists of Myc/Max dimerization inhibit Myc-induced transformation of chicken embryo fibroblasts. Proc. Natl. Acad. Sci. USA 2002, 99, 3830–3835. [Google Scholar] [CrossRef] [Green Version]
  110. Yin, X.; Giap, C.; Lazo, J.S.; Prochownik, E.V. Low molecular weight inhibitors of Myc–Max interaction and function. Oncogene 2003, 22, 6151–6159. [Google Scholar] [CrossRef] [Green Version]
  111. Guo, J.; Parise, R.A.; Joseph, E.; Egorin, M.J.; Lazo, J.S.; Prochownik, E.V.; Eiseman, J.L. Efficacy, pharmacokinetics, tisssue distribution, and metabolism of the Myc–Max disruptor, 10058-F4 [Z,E]-5-[4-ethylbenzylidine]-2-thioxothiazolidin-4-one, in mice. Cancer Chemother. Pharmacol. 2009, 63, 615–625. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Clausen, D.M.; Guo, J.; Parise, R.A.; Beumer, J.H.; Egorin, M.J.; Lazo, J.S.; Prochownik, E.V.; Eiseman, J.L. In Vitro Cytotoxicity and In Vivo Efficacy, Pharmacokinetics, and Metabolism of 10074-G5, a Novel Small-Molecule Inhibitor of c-Myc/Max Dimerization. J. Pharmacol. Exp. Ther. 2010, 335, 715–727. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Hart, J.R.; Garner, A.L.; Yu, J.; Ito, Y.; Sun, M.; Ueno, L.; Rhee, J.-K.; Baksh, M.M.; Stefan, E.; Hartl, M.; et al. Inhibitor of MYC identified in a Krohnke pyridine library. Proc. Natl. Acad. Sci. USA 2014, 111, 12556–12561. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Castell, A.; Yan, Q.; Fawkner, K.; Hydbring, P.; Zhang, F.; Verschut, V.; Franco, M.; Zakaria, S.M.; Bazzar, W.; Goodwin, J.; et al. A selective high affinity MYC-binding compound inhibits MYC:MAX interaction and MYC-dependent tumor cell proliferation. Sci. Rep. 2018, 8, 10064. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Truica, M.I.; Burns, M.C.; Han, H.; Abdulkadir, S.A. Turning up the heat on MYC: Progress in small molecule inhibitors. Cancer Res. 2020. [Google Scholar] [CrossRef] [PubMed]
  116. Han, H.; Jain, A.D.; Truica, M.I.; Izquierdo-Ferrer, J.; Anker, J.F.; Lysy, B.; Sagar, V.; Luan, Y.; Chalmers, Z.R.; Unno, K.; et al. Small-Molecule MYC Inhibitors Suppress Tumor Growth and Enhance Immunotherapy. Cancer Cell 2019, 36, 483–497.e15. [Google Scholar] [CrossRef] [PubMed]
  117. Carabet, L.; Rennie, P.; Cherkasov, A. Therapeutic Inhibition of Myc in Cancer. Structural Bases and Computer-Aided Drug Discovery Approaches. Int. J. Mol. Sci. 2018, 20, 120. [Google Scholar] [CrossRef] [Green Version]
  118. Thomas, L.R.; Wang, Q.; Grieb, B.C.; Phan, J.; Foshage, A.M.; Sun, Q.; Olejniczak, E.T.; Clark, T.; Dey, S.; Lorey, S.; et al. Interaction with WDR5 Promotes Target Gene Recognition and Tumorigenesis by MYC. Mol. Cell 2015, 58, 440–452. [Google Scholar] [CrossRef] [Green Version]
  119. Struntz, N.B.; Chen, A.; Deutzmann, A.; Wilson, R.M.; Stefan, E.; Evans, H.L.; Ramirez, M.A.; Liang, T.; Caballero, F.; Wildschut, M.H.E.; et al. Stabilization of the Max Homodimer with a Small Molecule Attenuates Myc-Driven Transcription. Cell Chem. Biol. 2019, 26, 711–723.e14. [Google Scholar] [CrossRef]
  120. 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]
  121. Arnold, H.K.; Sears, R.C. Protein Phosphatase 2A Regulatory Subunit B56 Associates with c-Myc and Negatively Regulates c-Myc Accumulation. Mol. Cell. Biol. 2006, 26, 13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Allen-Petersen, B.L.; Sears, R.C. Mission Possible: Advances in MYC Therapeutic Targeting in Cancer. BioDrugs 2019. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Henriksson, M.; Bakardjiev, A.; Klein, G.; Lüscher, B. Phosphorylation Sites Mapping in the N-terminal Domain of C-Myc Modulate Its Transforming Potential. Oncogene 1993, 8, 3199–3209. [Google Scholar] [PubMed]
  124. Pulverer, B.J.; Fisher, C.; Vousden, K.; Littlewood, T.; Evan, G.; Woodgett, J.R. Site-specific Modulation of c-Myc Cotransformation by Residues Phosphorylated in Vivo. Oncogene 1994, 9, 59–70. [Google Scholar]
  125. Wang, X.; Cunningham, M.; Zhang, X.; Tokarz, S.; Laraway, B.; Troxell, M.; Sears, R.C. Phosphorylation Regulates c-Myc’s Oncogenic Activity in the Mammary Gland. Cancer Res. 2011, 71, 925–936. [Google Scholar] [CrossRef] [Green Version]
  126. Yang, J.; Nie, J.; Ma, X.; Wei, Y.; Peng, Y.; Wei, X. Targeting PI3K in cancer: Mechanisms and advances in clinical trials. Mol. Cancer 2019, 18, 26. [Google Scholar] [CrossRef] [Green Version]
  127. Hermida, M.A.; Dinesh Kumar, J.; Leslie, N.R. GSK3 and its interactions with the PI3K/AKT/mTOR signalling network. Adv. Biol. Regul. 2017, 65, 5–15. [Google Scholar] [CrossRef]
  128. Gregory, M.A.; Qi, Y.; Hann, S.R. Phosphorylation by Glycogen Synthase Kinase-3 Controls c-Myc Proteolysis and Subnuclear Localization. J. Biol. Chem. 2003, 278, 51606–51612. [Google Scholar] [CrossRef] [Green Version]
  129. Liang, C.; Shi, S.; Liu, M.; Qin, Y.; Meng, Q.; Hua, J.; Ji, S.; Zhang, Y.; Yang, J.; Xu, J.; et al. PIN1 Maintains Redox Balance via the c-Myc/NRF2 Axis to Counteract Kras-Induced Mitochondrial Respiratory Injury in Pancreatic Cancer Cells. Cancer Res. 2019, 79, 133–145. [Google Scholar] [CrossRef] [Green Version]
  130. Wulf, G.M. Pin1 is overexpressed in breast cancer and cooperates with Ras signaling in increasing the transcriptional activity of c-Jun towards cyclin D1. EMBO J. 2001, 20, 3459–3472. [Google Scholar] [CrossRef] [Green Version]
  131. Ayala, G.; Wang, D.; Wulf, G.; Frolov, A.; Li, R.; Sowadski, J.; Wheeler, T.M.; Lu, K.P. The Prolyl Isomerase Pin1 Is a Novel Prognostic Marker in Human Prostate Cancer. Cancer Res. 2003, 63, 6244–6251. [Google Scholar] [PubMed]
  132. Chen, Y.; Wu, Y.; Yang, H.; Li, X.; Jie, M.; Hu, C.; Wu, Y.; Yang, S.; Yang, Y. Prolyl isomerase Pin1: A promoter of cancer and a target for therapy. Cell Death Dis. 2018, 9, 883. [Google Scholar] [CrossRef] [PubMed]
  133. Yu, J.H.; Im, C.Y.; Min, S.-H. Function of PIN1 in Cancer Development and Its Inhibitors as Cancer Therapeutics. Front. Cell Dev. Biol. 2020, 8, 120. [Google Scholar] [CrossRef] [PubMed]
  134. Cohn, G.M.; Liefwalker, D.F.; Langer, E.M.; Sears, R.C. PIN1 Provides Dynamic Control of MYC in Response to Extrinsic Signals. Front. Cell Dev. Biol. 2020, 8, 224. [Google Scholar] [CrossRef] [PubMed]
  135. Farrell, A.S.; Pelz, C.; Wang, X.; Daniel, C.J.; Wang, Z.; Su, Y.; Janghorban, M.; Zhang, X.; Morgan, C.; Impey, S.; et al. Pin1 regulates the dynamics of c-Myc DNA binding to facilitate target gene regulation and oncogenesis. Mol. Cell. Biol. 2013, 33, 2930–2949. [Google Scholar] [CrossRef] [Green Version]
  136. Min, S.-H.; Lau, A.W.; Lee, T.H.; Inuzuka, H.; Wei, S.; Huang, P.; Shaik, S.; Lee, D.Y.; Finn, G.; Balastik, M.; et al. Negative Regulation of the Stability and Tumor Suppressor Function of Fbw7 by the Pin1 Prolyl Isomerase. Mol. Cell 2012, 46, 771–783. [Google Scholar] [CrossRef] [Green Version]
  137. D’Artista, L.; Bisso, A.; Piontini, A.; Doni, M.; Verrecchia, A.; Kress, T.R.; Morelli, M.J.; Del Sal, G.; Amati, B.; Campaner, S. Pin1 is required for sustained B cell proliferation upon oncogenic activation of Myc. Oncotarget 2016, 7, 21786–21798. [Google Scholar] [CrossRef]
  138. Sablina, A.A.; Hector, M.; Colpaert, N.; Hahn, W.C. Identification of PP2A Complexes and Pathways Involved in Cell Transformation. Cancer Res. 2010, 70, 10474–10484. [Google Scholar] [CrossRef] [Green Version]
  139. Seshacharyulu, P.; Pandey, P.; Datta, K.; Batra, S.K. Phosphatase: PP2A structural importance, regulation and its aberrant expression in cancer. Cancer Lett. 2013, 335, 9–18. [Google Scholar] [CrossRef] [Green Version]
  140. Ruvolo, P.P. The broken “Off” switch in cancer signaling: PP2A as a regulator of tumorigenesis, drug resistance, and immune surveillance. BBA Clin. 2016, 6, 87–99. [Google Scholar] [CrossRef] [Green Version]
  141. Velmurugan, B.K.; Lee, C.-H.; Chiang, S.-L.; Hua, C.-H.; Chen, M.-C.; Lin, S.-H.; Yeh, K.-T.; Ko, Y.-C. PP2A deactivation is a common event in oral cancer and reactivation by FTY720 shows promising therapeutic potential. J. Cell. Physiol. 2018, 233, 1300–1311. [Google Scholar] [CrossRef]
  142. Neviani, P.; Santhanam, R.; Trotta, R.; Notari, M.; Blaser, B.W.; Liu, S.; Mao, H.; Chang, J.S.; Galietta, A.; Uttam, A.; et al. The tumor suppressor PP2A is functionally inactivated in blast crisis CML through the inhibitory activity of the BCR/ABL-regulated SET protein. Cancer Cell 2005, 8, 355–368. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. O’Connor, C.M.; Perl, A.; Leonard, D.; Sangodkar, J.; Narla, G. Therapeutic targeting of PP2A. Int. J. Biochem. Cell Biol. 2018, 96, 182–193. [Google Scholar] [CrossRef]
  144. Junttila, M.R.; Puustinen, P.; Niemelä, M.; Ahola, R.; Arnold, H.; Böttzauw, T.; Ala-aho, R.; Nielsen, C.; Ivaska, J.; Taya, Y.; et al. CIP2A Inhibits PP2A in Human Malignancies. Cell 2007, 130, 51–62. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Becker, S.; Kiecke, C.; Sch, E.; Griesinger, C.; Koch, R.; Rydzynska, Z.; Chapuy, B.; Kube, D.; Venkataramani, V.; Bohnenberger, H.; et al. Destruction of a Microtubule-Bound MYC Reservoir during Mitosis Contributes to Vincristine’s Anticancer Activity. Mol. Cancer Res. 2020, 18, 859–872. [Google Scholar] [PubMed] [Green Version]
  146. Marampon, F.; Ciccarelli, C.; Zani, B.M. Down-regulation of c-Myc following MEK/ERK inhibition halts the expression of malignant phenotype in rhabdomyosarcoma and in non muscle-derived human tumors. Mol. Cancer 2006, 5, 31. [Google Scholar] [CrossRef] [Green Version]
  147. Ciccarelli, C.; Di Rocco, A.; Gravina, G.L.; Mauro, A.; Festuccia, C.; Del Fattore, A.; Berardinelli, P.; De Felice, F.; Musio, D.; Bouché, M.; et al. Disruption of MEK/ERK/c-Myc signaling radiosensitizes prostate cancer cells in vitro and in vivo. J. Cancer Res. Clin. Oncol. 2018, 144, 1685–1699. [Google Scholar] [CrossRef]
  148. Gravina, G.L.; Festuccia, C.; Popov, V.M.; Rocco, A.D.; Colapietro, A.; Sanita, P.; Monache, S.D.; Musio, D.; Felice, F.D.; Cesare, E.D.; et al. c-Myc Sustains Transformed Phenotype and Promotes Radioresistance of Embryonal Rhabdomyosarcoma Cell Lines. Radiat. Res. 2016, 185, 411–422. [Google Scholar] [CrossRef]
  149. Wei, S.; Kozono, S.; Kats, L.; Nechama, M.; Li, W.; Guarnerio, J.; Luo, M.; You, M.-H.; Yao, Y.; Kondo, A.; et al. Active Pin1 is a key target of all-trans retinoic acid in acute promyelocytic leukemia and breast cancer. Nat. Med. 2015, 21, 457–466. [Google Scholar] [CrossRef]
  150. Lian, X.; Lin, Y.-M.; Kozono, S.; Herbert, M.K.; Li, X.; Yuan, X.; Guo, J.; Guo, Y.; Tang, M.; Lin, J.; et al. Pin1 inhibition exerts potent activity against acute myeloid leukemia through blocking multiple cancer-driving pathways. J. Hematol. Oncol. 2018, 11, 73. [Google Scholar] [CrossRef] [Green Version]
  151. Yang, D.; Luo, W.; Wang, J.; Zheng, M.; Liao, X.-H.; Zhang, N.; Lu, W.; Wang, L.; Chen, A.-Z.; Wu, W.-G.; et al. A novel controlled release formulation of the Pin1 inhibitor ATRA to improve liver cancer therapy by simultaneously blocking multiple cancer pathways. J. Controlled Release 2018, 269, 405–422. [Google Scholar] [CrossRef] [PubMed]
  152. Giuli, M.V.; Hanieh, P.N.; Giuliani, E.; Rinaldi, F.; Marianecci, C.; Screpanti, I.; Checquolo, S.; Carafa, M. Current Trends in ATRA Delivery for Cancer Therapy. Pharmaceutics 2020, 12, 707. [Google Scholar] [CrossRef] [PubMed]
  153. Kalemkerian, P.; Jasti, K.; Celano, P.; Nelkin, D.; Mabry, M. All-trans-retinoic acid alters myc gene expression and inhibits in vitro progression in small cell lung cancer. Cell Growth Differ. 1994, 5, 55–60. [Google Scholar] [PubMed]
  154. Saunders, D.E.; Christensen, C.; Wappler, N.L.; Schultz, J.F.; Lawrence, W.D.; Malviya, V.K.; Malone, J.M.; Deppe, G. Inhibition of c-myc in breast and ovarian carcinoma cells by 1,25-dihydroxyvitamin D3, retinoic acid and dexamethasone. Anticancer Drugs 1993, 4, 201–208. [Google Scholar] [CrossRef]
  155. Stopera, S.A.; Bird, R.P. Effects of all-trans retinoic acid as a potential chemopreventive agent on the formation of azoxymethane-induced aberrant crypt foci: Differential expression of c-myc and c-fos mrna and protein. Int. J. Cancer 1993, 53, 798–803. [Google Scholar] [CrossRef]
  156. Campaner, E.; Rustighi, A.; Zannini, A.; Cristiani, A.; Piazza, S.; Ciani, Y.; Kalid, O.; Golan, G.; Baloglu, E.; Shacham, S.; et al. A covalent PIN1 inhibitor selectively targets cancer cells by a dual mechanism of action. Nat. Commun. 2017, 8, 15772. [Google Scholar] [CrossRef]
  157. Agarwal, A.; MacKenzie, R.J.; Pippa, R.; Eide, C.A.; Oddo, J.; Tyner, J.W.; Sears, R.; Vitek, M.P.; Odero, M.D.; Christensen, D.J.; et al. Antagonism of SET Using OP449 Enhances the Efficacy of Tyrosine Kinase Inhibitors and Overcomes Drug Resistance in Myeloid Leukemia. Clin. Cancer Res. 2014, 20, 2092–2103. [Google Scholar] [CrossRef] [Green Version]
  158. Mandala, S. Alteration of Lymphocyte Trafficking by Sphingosine-1-Phosphate Receptor Agonists. Science 2002, 296, 346–349. [Google Scholar] [CrossRef]
  159. Brinkmann, V.; Davis, M.D.; Heise, C.E.; Albert, R.; Cottens, S.; Hof, R.; Bruns, C.; Prieschl, E.; Baumruker, T.; Hiestand, P.; et al. The Immune Modulator FTY720 Targets Sphingosine 1-Phosphate Receptors. J. Biol. Chem. 2002, 277, 21453–21457. [Google Scholar] [CrossRef] [Green Version]
  160. Chen, K.-F.; Liu, C.-Y.; Lin, Y.-C.; Yu, H.-C.; Liu, T.-H.; Hou, D.-R.; Chen, P.-J.; Cheng, A.-L. CIP2A mediates effects of bortezomib on phospho-Akt and apoptosis in hepatocellular carcinoma cells. Oncogene 2010, 29, 6257–6266. [Google Scholar] [CrossRef] [Green Version]
  161. Farrington, C.C.; Yuan, E.; Mazhar, S.; Izadmehr, S.; Hurst, L.; Allen-Petersen, B.L.; Janghorban, M.; Chung, E.; Wolczanski, G.; Galsky, M.; et al. Protein phosphatase 2A activation as a therapeutic strategy for managing MYC-driven cancers. J. Biol. Chem. 2020, 295, 757–770. [Google Scholar] [CrossRef] [PubMed]
  162. McClinch, K.; Avelar, R.A.; Callejas, D.; Izadmehr, S.; Wiredja, D.; Perl, A.; Sangodkar, J.; Kastrinsky, D.B.; Schlatzer, D.; Cooper, M.; et al. Small-Molecule Activators of Protein Phosphatase 2A for the Treatment of Castration-Resistant Prostate Cancer. Cancer Res. 2018, 78, 2065–2080. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Allen-Petersen, B.L.; Risom, T.; Feng, Z.; Wang, Z.; Jenny, Z.P.; Thoma, M.C.; Pelz, K.R.; Morton, J.P.; Sansom, O.J.; Lopez, C.D.; et al. Activation of PP2A and Inhibition of mTOR Synergistically Reduce MYC Signaling and Decrease Tumor Growth in Pancreatic Ductal Adenocarcinoma. Cancer Res. 2019, 79, 209–219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Sangodkar, J.; Perl, A.; Tohme, R.; Kiselar, J.; Kastrinsky, D.B.; Zaware, N.; Izadmehr, S.; Mazhar, S.; Wiredja, D.D.; O’Connor, C.M.; et al. Activation of tumor suppressor protein PP2A inhibits KRAS-driven tumor growth. J. Clin. Investig. 2017, 127, 2081–2090. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Leonard, D.; Huang, W.; Izadmehr, S.; O’Connor, C.M.; Wiredja, D.D.; Wang, Z.; Zaware, N.; Chen, Y.; Schlatzer, D.M.; Kiselar, J.; et al. Selective PP2A Enhancement through Biased Heterotrimer Stabilization. Cell 2020, 181, 688–701.e16. [Google Scholar] [CrossRef] [PubMed]
  166. O’Neil, N.J.; Bailey, M.L.; Hieter, P. Synthetic lethality and cancer. Nat. Rev. Genet. 2017, 18, 613–623. [Google Scholar] [CrossRef]
  167. Bryant, H.E.; Schultz, N.; Thomas, H.D.; Parker, K.M.; Flower, D.; Lopez, E.; Kyle, S.; Meuth, M.; Curtin, N.J.; Helleday, T. Specific killing of BRCA2-deficient tumours with inhibitors of poly(ADP-ribose) polymerase. Nature 2005, 434, 913–917. [Google Scholar] [CrossRef]
  168. Farmer, H.; McCabe, N.; Lord, C.J.; Tutt, A.N.J.; Johnson, D.A.; Richardson, T.B.; Santarosa, M.; Dillon, K.J.; Hickson, I.; Knights, C.; et al. Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy. Nature 2005, 434, 917–921. [Google Scholar] [CrossRef]
  169. Iba, T.; Kigawa, J.; Kanamori, Y.; Itamochi, H.; Oishi, T.; Simada, M.; Uegaki, K.; Naniwa, J.; Terakawa, N. Expression of the c-myc gene as a predictor of chemotherapy response and a prognostic factor in patients with ovarian cancer. Cancer Sci. 2004, 95, 418–423. [Google Scholar] [CrossRef] [Green Version]
  170. Biroccio, A.; Benassi, B.; Amodei, S.; Gabellini, C.; Del Bufalo, D.; Zupi, G. C-Myc Down-Regulation Increases Susceptibility to Cisplatin through Reactive Oxygen Species-Mediated Apoptosis in M14 Human Melanoma Cells. Mol. Pharmacol. 2001, 60, 174–182. [Google Scholar] [CrossRef]
  171. Bucci, B. Myc Down-Regulation Sensitizes Melanoma Cells to Radiotherapy by Inhibiting MLH1 and MSH2 Mismatch Repair Proteins. Clin. Cancer Res. 2005, 11, 2756–2767. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Littler, S.; Sloss, O.; Geary, B.; Pierce, A.; Whetton, A.D.; Taylor, S.S. Oncogenic MYC amplifies mitotic perturbations. Open Biol. 2019, 9, 190136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Bretones, G.; Delgado, M.D.; León, J. Myc and cell cycle control. Biochim. Biophys. Acta BBA Gene Regul. Mech. 2015, 1849, 506–516. [Google Scholar] [CrossRef] [PubMed]
  174. Santamaría, D.; Barrière, C.; Cerqueira, A.; Hunt, S.; Tardy, C.; Newton, K.; Cáceres, J.F.; Dubus, P.; Malumbres, M.; Barbacid, M. Cdk1 is sufficient to drive the mammalian cell cycle. Nature 2007, 448, 811–815. [Google Scholar] [CrossRef]
  175. Otto, T.; Sicinski, P. Cell cycle proteins as promising targets in cancer therapy. Nat. Rev. Cancer 2017, 17, 93–115. [Google Scholar] [CrossRef] [Green Version]
  176. Goga, A.; Yang, D.; Tward, A.D.; Morgan, D.O.; Bishop, J.M. Inhibition of CDK1 as a potential therapy for tumors over-expressing MYC. Nat. Med. 2007, 13, 820–827. [Google Scholar] [CrossRef]
  177. Kang, J.; Sergio, C.M.; Sutherland, R.L.; Musgrove, E.A. Targeting cyclin-dependent kinase 1 (CDK1) but not CDK4/6 or CDK2 is selectively lethal to MYC-dependent human breast cancer cells. BMC Cancer 2014, 14, 32. [Google Scholar] [CrossRef]
  178. Dauch, D.; Rudalska, R.; Cossa, G.; Nault, J.-C.; Kang, T.-W.; Wuestefeld, T.; Hohmeyer, A.; Imbeaud, S.; Yevsa, T.; Hoenicke, L.; et al. A MYC–aurora kinase A protein complex represents an actionable drug target in p53-altered liver cancer. Nat. Med. 2016, 22, 12. [Google Scholar] [CrossRef]
  179. Fu, J.; Bian, M.; Jiang, Q.; Zhang, C. Roles of Aurora Kinases in Mitosis and Tumorigenesis. Mol. Cancer Res. 2007, 5, 1–10. [Google Scholar] [CrossRef] [Green Version]
  180. Gautschi, O.; Heighway, J.; Mack, P.C.; Purnell, P.R.; Lara, P.N.; Gandara, D.R. Aurora Kinases as Anticancer Drug Targets. Clin. Cancer Res. 2008, 14, 1639–1648. [Google Scholar] [CrossRef] [Green Version]
  181. Cerami, E.; Gao, J.; Dogrusoz, U.; Gross, B.E.; Sumer, S.O.; Aksoy, B.A.; Jacobsen, A.; Byrne, C.J.; Heuer, M.L.; Larsson, E.; et al. The cBio Cancer Genomics Portal: An Open Platform for Exploring Multidimensional Cancer Genomics Data: Figure 1. Cancer Discov. 2012, 2, 401–404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Den Hollander, J.; Rimpi, S.; Doherty, J.R.; Rudelius, M.; Buck, A.; Hoellein, A.; Kremer, M.; Graf, N.; Scheerer, M.; Hall, M.A.; et al. Aurora kinases A and B are up-regulated by Myc and are essential for maintenance of the malignant state. Blood 2010, 116, 1498–1505. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Diaz, R.J.; Golbourn, B.; Faria, C.; Picard, D.; Shih, D.; Raynaud, D.; Leadly, M.; MacKenzie, D.; Bryant, M.; Bebenek, M.; et al. Mechanism of action and therapeutic efficacy of Aurora kinase B inhibition in MYC overexpressing medulloblastoma. Oncotarget 2015, 6, 3359–3374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Gray, N.S. Exploiting Chemical Libraries, Structure, and Genomics in the Search for Kinase Inhibitors. Science 1998, 281, 533–538. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. 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] [PubMed] [Green Version]
  186. Eischen, C.M.; Weber, J.D.; Roussel, M.F.; Sherr, C.J.; Cleveland, J.L. Disruption of the ARF-Mdm2-p53 tumor suppressor pathway in Myc-induced lymphomagenesis. Genes Dev. 1999, 13, 2658–2669. [Google Scholar] [CrossRef] [Green Version]
  187. Horiuchi, D.; Kusdra, L.; Huskey, N.E.; Chandriani, S.; Lenburg, M.E.; Gonzalez-Angulo, A.M.; Creasman, K.J.; Bazarov, A.V.; Smyth, J.W.; Davis, S.E.; et al. MYC pathway activation in triple-negative breast cancer is synthetic lethal with CDK inhibition. J. Exp. Med. 2012, 209, 679–696. [Google Scholar] [CrossRef] [Green Version]
  188. Kumar, S.K.; LaPlant, B.; Chng, W.J.; Zonder, J.; Callander, N.; Fonseca, R.; Fruth, B.; Roy, V.; Erlichman, C.; Stewart, A.K. Dinaciclib, a novel CDK inhibitor, demonstrates encouraging single-agent activity in patients with relapsed multiple myeloma. Blood 2015, 125, 443–448. [Google Scholar] [CrossRef]
  189. Flynn, J.; Jones, J.; Johnson, A.J.; Andritsos, L.; Maddocks, K.; Jaglowski, S.; Hessler, J.; Grever, M.R.; Im, E.; Zhou, H.; et al. Dinaciclib is a novel cyclin-dependent kinase inhibitor with significant clinical activity in relapsed and refractory chronic lymphocytic leukemia. Leukemia 2015, 29, 1524–1529. [Google Scholar] [CrossRef] [Green Version]
  190. Mita, M.M.; Joy, A.A.; Mita, A.; Sankhala, K.; Jou, Y.-M.; Zhang, D.; Statkevich, P.; Zhu, Y.; Yao, S.-L.; Small, K.; et al. Randomized Phase II Trial of the Cyclin-Dependent Kinase Inhibitor Dinaciclib (MK-7965) Versus Capecitabine in Patients With Advanced Breast Cancer. Clin. Breast Cancer 2014, 14, 169–176. [Google Scholar] [CrossRef]
  191. US National Library of Medicine. Available online: https://www.clinicaltrials.gov/ct2/show/NCT01676753 (accessed on 24 September 2020).
  192. Yang, D.; Liu, H.; Goga, A.; Kim, S.; Yuneva, M.; Bishop, J.M. Therapeutic potential of a synthetic lethal interaction between the MYC proto-oncogene and inhibition of aurora-B kinase. Proc. Natl. Acad. Sci. USA 2010, 107, 13836–13841. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Helfrich, B.A.; Kim, J.; Gao, D.; Chan, D.C.; Zhang, Z.; Tan, A.-C.; Bunn, P.A. Barasertib (AZD1152), a Small Molecule Aurora B Inhibitor, Inhibits the Growth of SCLC Cell Lines In Vitro and In Vivo. Mol. Cancer Ther. 2016, 15, 2314–2322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. US National Library of Medicine. NCT02293005. Available online: https://clinicaltrials.gov/ct2/show/NCT02293005?term=aurora+kinase%2C+myc&draw=2&rank=2 (accessed on 29 September 2020).
  195. US National Library of Medicine. NCT01897012. Available online: https://clinicaltrials.gov/ct2/show/NCT01897012?term=aurora+kinase%2C+myc&draw=2&rank=5 (accessed on 29 September 2020).
  196. Park, S.I.; Lin, C.P.; Ren, N.; Angus, S.P.; Dittmer, D.P.; Foote, M.; Parton, T.; Bhatt, A.P.; Fedoriw, Y.D.; Roth, D.P.; et al. Inhibition of Aurora A Kinase in Combination with Chemotherapy Induces Synthetic Lethality and Overcomes Chemoresistance in Myc-Overexpressing Lymphoma. Target. Oncol. 2019. [Google Scholar] [CrossRef] [PubMed]
  197. Cermelli, S.; Jang, I.S.; Bernard, B.; Grandori, C. Synthetic Lethal Screens as a Means to Understand and Treat MYC-Driven Cancers. Cold Spring Harb. Perspect. Med. 2014, 4, a014209. [Google Scholar] [CrossRef] [Green Version]
  198. Dong, Y.; Tu, R.; Liu, H.; Qing, G. Regulation of cancer cell metabolism: Oncogenic MYC in the driver’s seat. Signal Transduct. Target. Ther. 2020, 5, 124. [Google Scholar] [CrossRef]
  199. Miller, D.M.; Thomas, S.D.; Islam, A.; Muench, D.; Sedoris, K. c-Myc and Cancer Metabolism. Clin. Cancer Res. 2012, 18, 5546–5553. [Google Scholar] [CrossRef] [Green Version]
  200. Kim, J.; Lee, J.; Iyer, V.R. Global Identification of Myc Target Genes Reveals Its Direct Role in Mitochondrial Biogenesis and Its E-Box Usage In Vivo. PLoS ONE 2008, 3, e1798. [Google Scholar] [CrossRef] [Green Version]
  201. Weber, G.; Nakamura, H.; Natsumeda, Y.; Szekeres, T.; Nagai, M. Regulation of GTP biosynthesis. Adv. Enzyme Regul. 1992, 32, 57–69. [Google Scholar] [CrossRef]
  202. Liu, Y.-C.; Li, F.; Handler, J.; Huang, C.R.L.; 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]
  203. Kofuji, S.; Hirayama, A.; Eberhardt, A.O.; Kawaguchi, R.; Sugiura, Y.; Sampetrean, O.; Ikeda, Y.; Warren, M.; Sakamoto, N.; Kitahara, S.; et al. IMP dehydrogenase-2 drives aberrant nucleolar activity and promotes tumorigenesis in glioblastoma. Nat. Cell Biol. 2019, 21, 1003–1014. [Google Scholar] [CrossRef]
  204. Huang, F.; Ni, M.; Chalishazar, M.D.; Huffman, K.E.; Kim, J.; Cai, L.; Shi, X.; Cai, F.; Zacharias, L.G.; Ireland, A.S.; et al. Inosine Monophosphate Dehydrogenase Dependence in a Subset of Small Cell Lung Cancers. Cell Metab. 2018, 28, 369–382.e5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Nagai, M.; Natsumeda, Y.; Konno, Y.; Hoffman, R.; Irino, S.; Weber, G. Selective up-regulation of type II inosine 5′-monophosphate dehydrogenase messenger RNA expression in human leukemias. Cancer Res. 1991, 51, 3886–3890. [Google Scholar] [PubMed]
  206. He, Y.; Mou, Z.; Li, W.; Liu, B.; Fu, T.; Zhao, S.; Xiang, D.; Wu, Y. Identification of IMPDH2 as a tumor-associated antigen in colorectal cancer using immunoproteomics analysis. Int. J. Colorectal Dis. 2009, 24, 1271–1279. [Google Scholar] [CrossRef] [PubMed]
  207. Kofuji, S.; Sasaki, A.T. GTP metabolic reprogramming by IMPDH2: Unlocking cancer cells’ fuelling mechanism. J. Biochem. 2020, 168, 319–328. [Google Scholar] [CrossRef]
  208. Huang, F.; Huffman, K.; Wang, Z.; Wang, X.; Li, K.; Cai, F.; Yang, C.; Cai, L.; Shih, T.S.; Zacharias, L.G.; et al. Guanosine triphosphate links MYC-dependent metabolic and ribosome programs in small cell lung cancer. J. Clin. Investig. 2020. [Google Scholar] [CrossRef]
  209. Wise, D.R.; DeBerardinis, R.J.; Mancuso, A.; Sayed, N.; Zhang, X.-Y.; Pfeiffer, H.K.; Nissim, I.; Daikhin, E.; Yudkoff, M.; McMahon, S.B.; et al. Myc regulates a transcriptional program that stimulates mitochondrial glutaminolysis and leads to glutamine addiction. Proc. Natl. Acad. Sci. USA 2008, 105, 18782–18787. [Google Scholar] [CrossRef] [Green Version]
  210. 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]
  211. Le, A.; Lane, A.N.; Hamaker, M.; Bose, S.; Gouw, A.; Barbi, J.; Tsukamoto, T.; Rojas, C.J.; Slusher, B.S.; Zhang, H.; et al. Glucose-Independent Glutamine Metabolism via TCA Cycling for Proliferation and Survival in B Cells. Cell Metab. 2012, 15, 110–121. [Google Scholar] [CrossRef] [Green Version]
  212. Yuneva, M.; Zamboni, N.; Oefner, P.; Sachidanandam, R.; Lazebnik, Y. Deficiency in glutamine but not glucose induces MYC-dependent apoptosis in human cells. J. Cell Biol. 2007, 178, 93–105. [Google Scholar] [CrossRef] [Green Version]
  213. Shroff, E.H.; Eberlin, L.S.; Dang, V.M.; Gouw, A.M.; Gabay, M.; Adam, S.J.; Bellovin, D.I.; Tran, P.T.; Philbrick, W.M.; Garcia-Ocana, A.; et al. MYC oncogene overexpression drives renal cell carcinoma in a mouse model through glutamine metabolism. Proc. Natl. Acad. Sci. USA 2015, 112, 6539–6544. [Google Scholar] [CrossRef] [Green Version]
  214. Shen, Y.-A.; Hong, J.; Asaka, R.; Asaka, S.; Hsu, F.-C.; Suryo Rahmanto, Y.; Jung, J.-G.; Chen, Y.-W.; Yen, T.-T.; Tomaszewski, A.; et al. Inhibition of the MYC-regulated glutaminase metabolic axis is an effective synthetic lethal approach for treating chemoresistant cancers. Cancer Res. 2020, 80, 4514–4526. [Google Scholar] [CrossRef] [PubMed]
  215. Wahlström, T.; Arsenian Henriksson, M. Impact of MYC in regulation of tumor cell metabolism. Biochim. Biophys. Acta BBA Gene Regul. Mech. 2015, 1849, 563–569. [Google Scholar] [CrossRef] [PubMed]
  216. Naffouje, R.; Grover, P.; Yu, H.; Sendilnathan, A.; Wolfe, K.; Majd, N.; Smith, E.P.; Takeuchi, K.; Senda, T.; Kofuji, S.; et al. Anti-Tumor Potential of IMP Dehydrogenase Inhibitors: A Century-Long Story. Cancers 2019, 11, 1346. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Klangjorhor, J.; Chaiyawat, P.; Teeyakasem, P.; Sirikaew, N.; Phanphaisarn, A.; Settakorn, J.; Lirdprapamongkol, K.; Yama, S.; Svasti, J.; Pruksakorn, D. Mycophenolic acid is a drug with the potential to be repurposed for suppressing tumor growth and metastasis in osteosarcoma treatment. Int. J. Cancer 2020, 146, 3397–3409. [Google Scholar] [CrossRef]
  218. Domhan, S.; Muschal, S.; Schwager, C.; Morath, C.; Wirkner, U.; Ansorge, W.; Maercker, C.; Zeier, M.; Huber, P.E.; Abdollahi, A. Molecular mechanisms of the antiangiogenic and antitumor effects of mycophenolic acid. Mol. Cancer Ther. 2008, 7, 1656–1668. [Google Scholar] [CrossRef] [Green Version]
  219. Akiyama, T.; Okazaki, H.; Takahashi, K.; Hasegawa, A.; Tanabe, K.; Uchida, K.; Takahara, S.; Toma, H. Mizoribine in Combination Therapy with Tacrolimus For Living Donor Renal Transplantation: Analysis of a Nationwide Study in Japan. Transplant. Proc. 2005, 37, 843–845. [Google Scholar] [CrossRef]
  220. Shapiro, R.; Ckark, V.; Curthoys, N. Inactivation of rat renal phosphate-dependent glutaminase with 6-diazo-5-oxo-L-norleucine. Evidence for interaction at the glutamine binding site. J. Biol. Chem. 1979, 254, 2835–2838. [Google Scholar]
  221. Shukla, K.; Ferraris, D.V.; Thomas, A.G.; Stathis, M.; Duvall, B.; Delahanty, G.; Alt, J.; Rais, R.; Rojas, C.; Gao, P.; et al. Design, Synthesis, and Pharmacological Evaluation of Bis-2-(5-phenylacetamido-1,2,4-thiadiazol-2-yl)ethyl Sulfide 3 (BPTES) Analogs as Glutaminase Inhibitors. J. Med. Chem. 2012, 55, 10551–10563. [Google Scholar] [CrossRef] [Green Version]
  222. Thomas, L.R.; Adams, C.M.; Wang, J.; Weissmiller, A.M.; Creighton, J.; Lorey, S.L.; Liu, Q.; Fesik, S.W.; Eischen, C.M.; Tansey, W.P. Interaction of the oncoprotein transcription factor MYC with its chromatin cofactor WDR5 is essential for tumor maintenance. Proc. Natl. Acad. Sci. USA 2019, 116, 25260–25268. [Google Scholar] [CrossRef] [Green Version]
  223. Chacón, S.; Wang, F.; Thomas, L.R.; Phan, J.; Zhao, B.; Olejniczak, E.T.; Macdonald, J.D.; Shaw, J.G.; Schlund, C.; Payne, W.; et al. Discovery of WD Repeat-Containing Protein 5 (WDR5)–MYC Inhibitors Using Fragment-Based Methods and Structure-Based Design. J. Med. Chem. 2020, 63, 4315–4333. [Google Scholar] [CrossRef]
  224. Psathas, J.N.; Thomas-Tikhonenko, A. MYC and the Art of MicroRNA Maintenance. Cold Spring Harbor Perspect. Med. 2014, 4, a014175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. 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]
  226. Bui, T.V.; Mendell, J.T. Myc: Maestro of MicroRNAs. Genes Cancer 2010, 1, 568–575. [Google Scholar] [CrossRef] [PubMed]
  227. 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]
  228. Frenzel, A.; Lovén, J.; Henriksson, M.A. Targeting MYC-Regulated miRNAs to Combat Cancer. Genes Cancer 2010, 1, 660–667. [Google Scholar] [CrossRef]
  229. Dhanasekaran, R.; Gabay-Ryan, M.; Baylot, V.; Lai, I.; Mosley, A.; Huang, X.; Zabludoff, S.; Li, J.; Kaimal, V.; Karmali, P.; et al. Anti-miR-17 therapy delays tumorigenesis in MYC-driven hepatocellular carcinoma (HCC). Oncotarget 2018, 9, 5517–5528. [Google Scholar] [CrossRef] [Green Version]
  230. Malynn, B.A.; de Alboran, I.M.; Davidson, L.; DePinho, R.A.; Alt, F.W. N-myc can functionally replace c-myc in murine development, cellular growth, and differentiation. Genes Dev. 2000, 14, 1390–1399. [Google Scholar]
Figure 1. MYC amplification across cancers. Percentage represents number of patients with MYC amplification for that cancer type. Red bars represent cancers in which >10% of patients harbor MYC mutations. Data from The Cancer Genome Atlas Pan Cancer 2018 Dataset, cancer.gov/TCGA.
Figure 1. MYC amplification across cancers. Percentage represents number of patients with MYC amplification for that cancer type. Red bars represent cancers in which >10% of patients harbor MYC mutations. Data from The Cancer Genome Atlas Pan Cancer 2018 Dataset, cancer.gov/TCGA.
Ijms 21 09486 g001
Figure 2. Schematic presenting the various cellular processes to target through inhibition or reactivation in the nucleus (left) or cytoplasm (right) upon Myc-induced tumorigenesis. All will be described in detail in this review.
Figure 2. Schematic presenting the various cellular processes to target through inhibition or reactivation in the nucleus (left) or cytoplasm (right) upon Myc-induced tumorigenesis. All will be described in detail in this review.
Ijms 21 09486 g002
Figure 3. Upon MYC amplification in cancer, Myc recruits additional transcriptional cofactors to drive transcription: (A) BRD4 binds to acetylated lysines on histone tails and recruits P-TEFb (which includes CDK9), that phosphorylates the carboxy terminal domain of RNA Pol II. Myc can also individually recruit P-TEFb. (B) Treatment with bromodomain and extra-terminal motif (BET)/BRD4 inhibitors prevents BRD4 from binding to histone tails and treatment with CDK9 inhibitors disrupts CDK9′s kinase activity. Thus, both result in failure of activating transcription of MYC or Myc target genes.
Figure 3. Upon MYC amplification in cancer, Myc recruits additional transcriptional cofactors to drive transcription: (A) BRD4 binds to acetylated lysines on histone tails and recruits P-TEFb (which includes CDK9), that phosphorylates the carboxy terminal domain of RNA Pol II. Myc can also individually recruit P-TEFb. (B) Treatment with bromodomain and extra-terminal motif (BET)/BRD4 inhibitors prevents BRD4 from binding to histone tails and treatment with CDK9 inhibitors disrupts CDK9′s kinase activity. Thus, both result in failure of activating transcription of MYC or Myc target genes.
Ijms 21 09486 g003
Figure 4. Heterodimerization with Max is required for Myc’s oncogenic activity: (A) Upon heterodimerization, Myc/Max binds to the E-box and initiates transcription. This is a normal cellular process, but in cancer, Myc amplification further increases Myc activity; (B) Treating with Omomyc, a dimerization inhibitor that preferentially binds to Max or homodimerizes, displaces Myc at E-boxes and decreases Myc transcription. Other discussed Myc/Max inhibitors either disrupt Myc/Max dimers or block Myc’s interaction with DNA, but not both.
Figure 4. Heterodimerization with Max is required for Myc’s oncogenic activity: (A) Upon heterodimerization, Myc/Max binds to the E-box and initiates transcription. This is a normal cellular process, but in cancer, Myc amplification further increases Myc activity; (B) Treating with Omomyc, a dimerization inhibitor that preferentially binds to Max or homodimerizes, displaces Myc at E-boxes and decreases Myc transcription. Other discussed Myc/Max inhibitors either disrupt Myc/Max dimers or block Myc’s interaction with DNA, but not both.
Ijms 21 09486 g004
Figure 5. Disrupting Myc stability: (A) In cancer, PI3K signaling inactivates GSK3, preventing phosphorylation of T58 Myc. Pin1 overexpression keeps Myc in the cis-confirmation, preventing PP2A trans-specific enzyme from binding to Myc. Furthermore, PP2A is inactivated in several cancers, and therefore S62 remains phosphorylated. All of this leads to high Myc stability. (B) Inhibition of PI3K allows for GSK3 to phosphorylate T58 on Myc, which is required for degradation. Pin1 inhibitors and PP2A activators allow for PP2A to recognize and remove the phosphorylation of S62, leading to low stability and Myc’s degradation.
Figure 5. Disrupting Myc stability: (A) In cancer, PI3K signaling inactivates GSK3, preventing phosphorylation of T58 Myc. Pin1 overexpression keeps Myc in the cis-confirmation, preventing PP2A trans-specific enzyme from binding to Myc. Furthermore, PP2A is inactivated in several cancers, and therefore S62 remains phosphorylated. All of this leads to high Myc stability. (B) Inhibition of PI3K allows for GSK3 to phosphorylate T58 on Myc, which is required for degradation. Pin1 inhibitors and PP2A activators allow for PP2A to recognize and remove the phosphorylation of S62, leading to low stability and Myc’s degradation.
Ijms 21 09486 g005
Figure 6. Taking advantage of oncogenic addiction in cancer with synthetic dosage lethality. (A) Myc produces target genes Aurora kinases A and B and CDK1, which help stabilize Myc and promote the cell cycle. MYC is commonly amplified in cancer and cells remain viable. (B) In MYC-overexpressing cancers, inhibiting CDK1 or Aurora kinases leads to cell cycle arrest and synthetic dosage lethality.
Figure 6. Taking advantage of oncogenic addiction in cancer with synthetic dosage lethality. (A) Myc produces target genes Aurora kinases A and B and CDK1, which help stabilize Myc and promote the cell cycle. MYC is commonly amplified in cancer and cells remain viable. (B) In MYC-overexpressing cancers, inhibiting CDK1 or Aurora kinases leads to cell cycle arrest and synthetic dosage lethality.
Ijms 21 09486 g006
Figure 7. Myc influences metabolism through its target genes: (A) In cancer, MYC overexpression correlates with IMPDH expression and transcribes other target genes, including GTP and glutaminase (GLN). IMPDH catalyzes GDP to GTP and glutaminase converts glutamine (Gln) to glutamate (Glu), a major energy source in cancer. Lastly, epigenetic co-factor WDR5 recruits Myc to chromatin to express genes involved in biomass accumulation. (B) Although Myc still has control over target genes IMPDH, GTP, and GLN, the function of these proteins can be inhibited. Cancer’s energy supply can be depleted by inhibiting IMPDH, which prevents GTP production, or by inhibiting glutaminase, which will limit the pool of glutamate. Finally, inhibiting WDR5 will prevent Myc’s target gene expression of biomass related genes.
Figure 7. Myc influences metabolism through its target genes: (A) In cancer, MYC overexpression correlates with IMPDH expression and transcribes other target genes, including GTP and glutaminase (GLN). IMPDH catalyzes GDP to GTP and glutaminase converts glutamine (Gln) to glutamate (Glu), a major energy source in cancer. Lastly, epigenetic co-factor WDR5 recruits Myc to chromatin to express genes involved in biomass accumulation. (B) Although Myc still has control over target genes IMPDH, GTP, and GLN, the function of these proteins can be inhibited. Cancer’s energy supply can be depleted by inhibiting IMPDH, which prevents GTP production, or by inhibiting glutaminase, which will limit the pool of glutamate. Finally, inhibiting WDR5 will prevent Myc’s target gene expression of biomass related genes.
Ijms 21 09486 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

McAnulty, J.; DiFeo, A. The Molecular ‘Myc-anisms’ behind Myc-Driven Tumorigenesis and the Relevant Myc-Directed Therapeutics. Int. J. Mol. Sci. 2020, 21, 9486. https://doi.org/10.3390/ijms21249486

AMA Style

McAnulty J, DiFeo A. The Molecular ‘Myc-anisms’ behind Myc-Driven Tumorigenesis and the Relevant Myc-Directed Therapeutics. International Journal of Molecular Sciences. 2020; 21(24):9486. https://doi.org/10.3390/ijms21249486

Chicago/Turabian Style

McAnulty, Jessica, and Analisa DiFeo. 2020. "The Molecular ‘Myc-anisms’ behind Myc-Driven Tumorigenesis and the Relevant Myc-Directed Therapeutics" International Journal of Molecular Sciences 21, no. 24: 9486. https://doi.org/10.3390/ijms21249486

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