Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2016 Dec 12.
Published in final edited form as: RNA Dis. 2016 May 30;3(3):e1330. doi: 10.14800/rd.1330

MicroRNA-based screens for synthetic lethal interactions with c-Myc

Youjun Li 1,2, Yahui Zhu 1, Edward V Prochownik 3
PMCID: PMC5152767  NIHMSID: NIHMS791534  PMID: 27975083

Abstract

microRNAs (miRs) are small, non-coding RNAs, which play crucial roles in the development and progression of human cancer. Given that miRs are stable, easy to synthetize and readily introduced into cells, they have been viewed as having potential therapeutic benefit in cancer. c-Myc (Myc) is one of the most commonly deregulated oncogenic transcription factors and has important roles in the pathogenesis of cancer, thus making it an important, albeit elusive therapeutic target. Here we review the miRs that have been identified as being both positive and negative targets for Myc and how these participate in the complex phenotypes that arise as a result of Myc-driven transformation. We also discussseveral recent reports of Myc-synthetic lethal interactions with miRs.These highlight the importance and complexity of miRs in Myc-mediated biological functions and the opportunities for Myc-driven human cancer therapies.

Keywords: microRNAs, c-Myc, synthetic lethality, cancer

Introduction

microRNAs (miRs) are a small class of endogenous non-coding RNA molecules that are first transcribed into primary miR (pri-miR) products by RNA polymerase II. The pri-miRs are then cleaved into precursor miR (pre-miR) by the type III RNase Drosha. Following their transport from the nucleus to the cytoplasm by Exportin-5/Ran-GTP, pre-miRs are processed to double-stranded RNAs by Dicer, another type III RNase. After the double-stranded RNAs are unwound, individual mature miRs, which are 19-21 nt in length, are incorporated into an RNA-induced silencing complex (RISC) and bind to the 3’-untranslated region (3’UTR) of target mRNAs. These then affect gene expression either by inhibiting translation or by promoting target degradation[1-3]. It is noteworthy that in most cases, the complementarity of miRs to their target sequences is less than 100%. As a result, individual miRs can bind imperfectly to multiple transcripts with slightly different target sequences and affinities and thereby affect multiple cellular phenotypes[1-3].

Although many aspects of miR biosynthesis and function are understood, there remain a number of questions concerning the precise means by which miRs affect gene expression.There are several ways by which this is accomplished depending on whether their match to 3’UTR target gene sequences are perfect or imperfect. These include induction of target mRNA degradation, translation-coupled protein degradation, inhibition of translation initiation, inhibition of translation elongation and termination of translation[1-3]. Some miRs can also inhibit mRNA processing by interfering with splicing [1-3].The imprecise binding of miRs can make the absolute prediction of true physiologic target sites challenging. Indeed, target mRNAs with otherwise identical miR binding sites can possess widely differing degrees of miR binding in an actual cellular context and even those targets with similar degrees of miR binding can show large variations in response. miRBase, a data base for miRs, contains the sequences of more than 2000 human pri-miR genes and 2588 mature human miR sequences; however, the functions for many of these remain to be defined. miRs are involved in various normal biological processes, such as development, cell growth and metabolic programming[3-5]. Loss of control of miR expression can contribute to the development of many diseases, including cancer[3-6]. Because miRs are very stable, easy to synthetize and easily introduced into cells, they have been viewed as having therapeutic potential for a wide range of diseases including cancer[6-7].

Differental miR expression patterns between tumors and adjacent normal tissues provide useful information for tumor classification, prognosis and management[3, 5-7]. Specific miRs may display tumor suppressor or oncogenic function. For example, the enforced expression of let-7, miR-26 and miR-34a can inhibit malignant transformation and tumorigenesis whereas enforcing expression ofmiR-21 promotes B cell lymphomagenesis[3, 5-10]. Altered miR expression in cancers canresult from both genetic and epigenetic changes. An example of the former is seen with deletion of 13q14, which inactivates the tumor suppressors miR-15a and miR-16 in chronic lymphocytic leukemia. An example of the latter occurs with hypermethylation of CpG islands in the promoters of miR-127and miR-124, which leads to loss of their tumor suppressorfunction in human bladder cancer and hepatocellular carcinoma, respectively[3,5-10].

c-Myc and miRs

c-Myc (Myc), a helix-loop-helix leucine zipper (HLH-ZIP) oncogenic transcription factor, is de-regulated and/or over-expressed in a large fractions of human malignancies[11-14]. Myc dimerizes with Max, another HLH-ZIP protein, and binds to E-box sequences to activate transcription of target genes, including somemiRs[11-14]. Myc also acts as a transcriptional repressor by interacting with and suppressing other transcription factors and/or by modulating chromatin status [11-14].Through a myriad of downstream targets, Myc supervises a variety of biological processes, including cell proliferation, survival, metabolism and transformation[11-19]. Over-expressing Myc invarious mouse models of cancer causes tumors of a variety of types whereas homozygous deletion of Myc is associated with developmental abnormalities and embryonic lethality [13,17-19]. Homozygous deletion of Myc in rat fibroblasts cells prolongs doubling time, indicating that it plays a central role in regulating cell proliferation[11-14]. The basis for this involves a profound impairment of mitochondrial structure and function, which lead to chronic ATP depletion and the up-regulation of AMP-activated protein kinase, which futilely attempts to restore a normal energy balance by suppressing energy-consuming and anabolic processessuch as protein translation and proliferation[20].

Myc is one of the most frequently altered oncogenes in human cancers[15]. Its over-expression occurs as a consequence of DNA amplification, chromosomal translocation, or protein stabilization either by direct point mutations in the Myc protein itself or, more commonly by activation of pathways that promote Myc's phosphorylation-dependent stabilization[11-15]. Myc has been shown to be necessary for cancer progression and maintenance and its up-regulation is often correlated with poor clinical outcomes and aggressive cancer phenotypes in human neoplasms[21-25].

Myc transcriptionally regulates the expression of mRNAs, long non-coding RNAs and miRs[12, 26-35]. The latter in particular represent an important subset of Myc's transcriptional targets and significantly contribute tothe complexity of its function and biological readoutby virtue of their post-transcriptional effects on mRNA stability and translation[12, 25-35]. Myc can both activate and suppress miR expression, thereby regulating a variety of functions pertaining to cell growth, apoptosis, metabolic re-programming, tumorigenesis, angiogenesis and metastasis[25-37]. Myc may also affect the maturation of some pri-miRs. In the case of let-7, this occurs indirectly as a result of Myc's up-regulation of the RNA binding proteins Lin-28 and Lin-28B, which then negatively regulates let-7 [38]. Myc and miRs can also reciprocally regulate one another's expression. miRs have been shown to regulate a variety of Myc-induced phenotypes[29, 30,34,39].

Myc-upregulated miRs

The first reported effect of Myc on miR expression was its up-regulation of the polycistronic miR-17-92 cluster[26, 27]. Myc induces transcription of pri-17-92 via its direct binding to an E-box within the first intron of thepri-17-92 gene. In humans, the miR-17-92 cluster is located in an intron of the MIR17HG gene located on chromosome 13q31.3. The primary pri-17-92 transcript is processed into seven different mature miRs: miR-17, miR-18a, miR-19a, miR-19b, miR-20a and miR-92a which are widely expressed in different tissues and are essential for many developmental and pathogenic processes [26, 27]. miR-17-92 deletion in mice results in perinatal death due to a combination of severe pulmonary hypoplasia and ventricular septal defects. In humans, homozygous germ-line deletion of MIR17HG significantly reduces mature miR-17-92 levels and is associated with a syndrome characterized by microcephaly, short stature, and digital defects[40,41]. miR-17-92 acts as an oncogene “complex” that regulates multiple cellular processes which promote malignant transformation, increased survival, rapid proliferation, and angiogenesis[28-30,42,43]. miR-17-92 was reported to be involved in some human epithelial and hematopoietic malignancies. In the latter case, these include neoplasms as diverse as diffuse large B-cell lymphoma, Burkitt's lymphoma, mantle cell lymphoma, and chronic lymphocytic leukemia[28-30,42,43]. A direct role for miR-17-92 in lymphomagenesis has been demonstrated by virtue of the ability of Eμ-drivenmiR-17-92 in transgenic mice to promote the development of B-cell malignancies and massive splenomegaly with a high degree of penetrance[30, 42]. miR-17-92 appears to promote tumorigenesis by antagonizing tumor-suppressing mechanisms, notably apoptosis and senescence, through the activities of different miR components encoded by this cluster[30, 42].

miR-9 expression is also activated by Myc as well as by Myc's close relative N-Myc, both of which bind to a miR-9 promoter E-box. miR-9 is up-regulated in breast cancer cells and directly targets E-cadherin so as to increase cell motility and invasiveness[44-47].

Interestingly, miR-378 was reported to be a Myc-activated miR that can cooperate with activated Ras or HER2/neu to promote cellular transformation by directly inhibiting the anti-proliferative BTG family member, TOB2, which transcriptionally represses the proto-oncogene cyclin D1[48]. Additional Myc-upregulated miRs and some of their known functions are summarized in Table 1.

Table 1.

Myc-upregulated miRs

No. miR Location Targets Function Reference(s)
1 miR-9 1q22 E-cadherin, E2F1 Proliferation, EMT, survival, metastasis [40-47]
2 miR-17 13q31.3 E2F1,TGFBR2, CHEK2, TSP1,CTGF Migration, tumor growth [27,49-54]
3 miR-18a 13q31.3 PIAS3, TSP1,CTGF Proliferation, apoptosis [49-55]
4 miR-19a 13q31.3 PTEN, BIM, TSP1,CTGF EMT, angiogenesis [49-54,56]
5 miR-19b 13q31.3 MXD1, TSP1,CTGF Angiogenesis, cell cycle [49-54,56]
6 miR-20a 13q31.3 E2F1, PTEN, TSP1,CTGF Proliferation, metastasis, cell cycle [49-54,57]
7 miR-20b Xq26.2 PTEN, NCOA3, CAPRIN2 Lymphoma progression [49-57]
8 miR-25 7q22.1 BIM, DR4 Proliferation, apoptosis [58-59]
9 miR-92a 13q31.3 P57, PTEN Proliferation, cell cycle [49-57]
10 miR-93 7q22.1 P21 Proliferation, apoptosis [60-63]
11 miR-106a Xq26.2 PTEN, Hcyp19A1,IRS-2 Growth, metastasis, [63,64]
12 miR-106b 7q22.1 P21, PTEN Tumorigenesis [58,65]
13 miR-221 Xp11.3 SOCS3 Migration, invasion [66-67]
14 miR-378a 5q32 TOB2 Cell growth [30,48]

Myc-downregulated miRs

Although protein-coding gene activation tends to occur more commonly than does repression in response to Myc activation, the reverse seems to be true for miRs[12,26-37,68]. For example, Burkitt's lymphoma is a highly aggressive B-cell neoplasm, which originates from germinal center B cells and harbors chromosomal translocations involving Myc and immunoglobulin promoter-enhancer loci[69]. Through the analysis of human and mouse B cell lymphoma models, Chang et al. found Myc to be responsible for the repression of multiple miRs, including members of the let-7 and miR-30 familiesas well as, miR-15a/16, miR-22, miR-26a/b, miR-29a/b/c, miR-34a, miR-99a/b, mir-146a, miR-150 and miR-195. They found that Myc directly bound to these miR promoters, thereby attesting to their direct suppression, and that enforced re-expression of the repressed miRs diminished the tumorigenic potential of lymphoma cells. This suggested that widespread miR repression by Myc contributes to tumorigenesis and can be reversed by normalizing the expression of at least some miRs[31].

In γ-irradiation-induced lymphomas, miR-15a, miR-22, miR-23b, miR-125b, miR-26a/b, miR-29a/b and several let-7 family members have been reported to be repressed in response to Myc activation[69]. Restoration of miR-26a expression in the cells attenuated cycle progression and proliferation by inhibiting the Polycomb complex protein EZH2[70]. EZH2 is a histone-lysine N-methyltransferase enzyme, which catalyzes the methylation of histone H3 and thereby serves to epigenetically repress transcription. Presumably, EZH2 down-regulates genes that would otherwise prevent transformation and/or proliferation in response to Myc de-regulation.

Along similar lines, Zhang et al. found that miR-29 is repressed by Myc through a co-repressor complex comprised of EZH2 and histone deacetylase 3 (HDAC3). They showed that dual inhibition of HDAC3 and EZH2 disrupted theMyc-EZH2-miR-29 axis, leading to restoration of miR-29 expression, down-regulation of miR-29-targeted genes, and lymphoma suppression[33,35,71].

miR-28 was also found to be significantly repressed by Myc in Burkitt's lymphoma. Restoring miR-28 expression impaired cell proliferation and clonogenicity by inhibiting spindle checkpoint proteins including MAD2L1[72].

In human P-493 B lymphoma and PC3 prostate cells, Myc transcriptionally represses miR-23a/b, which can directly target mitochondrial glutaminase. This increases glutaminase mRNA and protein levels, which promotes glutamine catabolism. Glutaminase catalyzes the conversion of glutamine to glutamate, which in turn in converted into α-ketoglutarate, a key TCA cycle substrate that, is also needed for nucleotide and amino acid biosynthesis and maintenance of redox homeostasis[16]. Increasing the supply of α-ketoglutarate represents a form of metabolic re-programming that may serve to provide substrates to mitochondria other than those derived from glucose which are increasingly used in support of the Warburg effect [12,16,20].

Inhibition of miR-34a by Myc is also essential for tumorigenesis and survival. Repression of miR-34a was shown to be required for IL-6-induced epithelial-to-mesenchymal transition (EMT) and invasion. In miR-34a-deficient mice, colitis-associated intestinal tumors induced by azoxymethane/dextran sodium sulfate showed up-regulation of the EMT facilitators p-STAT3, IL-6R and SNAIL and progressed to invasive carcinomas [73].

Myc is pathologically activated in and essential for promoting human hepatocellular carcinoma (HCC)[18, 23-25]. In these tumors, Myc represses miR-122 expression directly by binding its promoter. Over-expression of miR-122 in miR-122−/− hepatocytes was shown to reduce c-Myc protein levels, while its depletion in miR-122+/+ hepatocytes led to an increase. miR-122 indirectly inhibited Myc transcription by targeting Tfdp2 and E2f1 in HCC, both of which bind to the Myc promoter and increase transcription. These suggest a double-negative feedback loop between miR-122 and Myc[74].

Similarly, Myc was also found to bind directly to conserved regions in the promoters of miR-148a and miR-363 and repress their expression. miR-148a-5p directly binds Myc's 3’UTR and inhibits its expression whereas miR-363-3p destabilizes Myc indirectly by inhibiting expression of USP28, a Myc deubiquitinase. Together, these two miRs provide for a novel and elegant form of positive feedback control by which Myc indirectly promotes the stabilization of both its own mRNA and protein.The importance of these interactions is underscored by the fact that inhibition of either miR-148a-5p or miR-363-3p induces hepatocarcinogenesis, whereas activation of them has the opposite effects[25].

miR-129-5p is another miR that is transcriptionally repressed by Myc. miR-129-5p levels negatively correlate with the clinical stage of human HCC and thus with survival. Restoring miR-129-5p expression suppresses diethylnitros- amine (DEN)-induced hepatocarcinogenesis in mice by targeting and down-regulating the activity of pyruvate dehydrogenase kinase 4 (PDK4)[75]. PDK4 is a critical negative regulator of pyruvate dehydrogenase (PDH), a rate-limiting enzyme that connects glycolysis and oxidative phosphorylation by converting pyruvate to acetyl coenzyme A. PDH's inhibition via PDK4-mediated phosphorylation should allow for the accumulation of glycolytic substrates, and their diversion into the anabolic pathways needed to support tumor growth. PDK4 inhibition via miR-129-5p normalizes the flow of pyruvate into the TCA cycle and an overall inhibition of the Warburg effect by depriving tumor cells of these critical anabolic substrates. This, along with the changes in glutamine flow discussed above, represents an additional Myc-regulated pathway that is responsible for the metabolic re-programming so commonly associated with cancer cells[75]. We summarize all the above Myc-suppressed miRs as well as additional ones in Table 2.

Table 2.

Myc-downregulated miRs

No. miR Location Targets Function Reference(s)
1 let7 family member 9q22.32
22q13.31
19q13.41
Myc,Bcl-2,Bcl-Xl,CCND2,HMGA2, SLC5A5 Growth, migration,invasion, EMT [68,69,76]
2 miR-15a 13q14.2 CCND1, Bcl-2,Bcl-xL, YAP1,FOXP3 Autophagy, proliferation [30,77]
3 miR-16 13q14.2 CCND1,CCND2, HSP70 Apoptosis, growth, proliferation [30,76,78]
4 miR-23a 19p13.13 GLS, ABCF1 Metabolism, Chemoresistance [16,79-81]
5 miR-23b 9q22.32 GLS, Smads,HMGB2 Metabolism, Chemoresistance [16,79,81]
6 miR-24 9q22.32 OCT4, Smads, DHFR Development, cancer progression [79,80]
7 miR-26a 3p22.2 EZH2,Lin28B,IL6, HMGA1,EphA2 Proliferation, tumorigenesis, [70, 82]
8 miR-27a 19p13.13 FOXO1,Apaf-1,DPD Differentiation, metastasis [30,80]
9 miR-27b 9q22.32 ST14, Smads Differentiation,apoptosis [79,80]
10 miR-28 3q28 Nrf2, MPL, CCND1, HOXB3, NM23-H1 Proliferation, migration [72]
11 miR-29a/b 7q32.3 CDK6,IGF1RLAMC1, MCL1,AKT2 Growth,migration [71,76]
12 miR-30a 6q13 IGF1R, HP1gamma, SOX4, Drp-1 Migration, invasion [83]
13 miR-34a 1p36.22 IL6R,HNF4G,cyclin E1 Proliferation, migration, invasion [30,73]
14 miR-122 18q21.31 IL-1α,DLX4,OCLN Fibrosis, proliferation, metabolism [74]
15 miR-129-2 11p11.2 PDK4, SOX4,BDKRB2 Migration,chemoresistance [75]
16 miR-148a 7p15.2 Myc,Wnt1, Dnmt1, IKKβ, LDLR, PKM2 Differentiation, cell cycle [25,30]
17 miR-185 22q11.21 DNMT1,Myc,SOCS3,VEGFA,E2F6,HIF2a Proliferation [84]
18 miR-363 Xq26.2 USP28,S1PR1,E2F3, MBP-1,Mcl-1 Cell cycle,cisplatin-induced apoptosis [25,30]
19 miR-449a 5q11.2 N-Myc,E2F3,BCL2,CDC25A,NOTCH1 Survival, cisplatin-induced cytotoxicity, invasion, proliferation [85]

Regulation of Myc expression by miRs

Myc expression is tightly controlled, both at the transcriptional and post-translational level and miRs play critical roles in these processes[11-15,28-31]. Some of the mechanisms underlying these varied forms of control have already been discussed above.

The first reported Myc-targeting miR was let-7a, which down-regulates Myc and reverts Myc-induced growth in Burkitt's lymphoma cells[86,87]. miR-145 was reported to be transcriptionally activated by p53 interacting with its response element in the miR-145 promoter. Myc is a direct target for miR-145. Enforced expression miR-145 silences Myc expression while anti-miR-145 enhances its expression. miR-145 silencing of Myc expression accounts at least in part for the miR-145-mediated tumor cell growth inhibition both in vitro and in vivo. miR-145 inhibitionis also able to reverse p53-mediated Myc repression[39,88].

miR-34a, a positive target of tumor suppressor p53, has been reported to function as a tumor suppressor that binds to Myc's 3’UTR and decreases its expression in prostate cancer and renal cell carcinoma[89]. In addition to this negative regulation of Myc, miR-34a was also found to target FoxM1, which, together with Myc, is involved in the activation of telomerase reverse transcriptase, which represents the critical step in immortalization [90, 91]. It has been shown that miR-34b/c's targeting of Myc requires mitogen-activated protein kinase-activated protein kinase 5 (MK5). MK5 activates miR-34b/c expression via phosphorylation of FoxO3a, thereby promoting nuclear localization of FoxO3a and enabling it to induce miR-34b/c expression. miR-34b/c binds to the 3'UTR of Myc and inhibits its expression. Myc in turn directly activates MK5 expression, forming a negative feedback loop that is dysregulated in colorectal tumorigenesis[92].

Interestingly miR-29a was reported to be a tumor suppressor that is induced by PRIMA-1Met, a small molecule with anti-tumor activity. miR-29a overexpression or exposure to PRIMA-1Met reduced multiple myeloma cell proliferation by targeting Myc. On the other hand, over-expression of Myc at least partially reverted the inhibitory effects caused by PRIMA-1Met or miR-29a overexpression suggesting that the miR-29a/Myc axis mediates anti-myeloma effects of PRIMA-1Met. Importantly, intratumoral delivery of miR-29a mimics induced regression of tumors in mouse xenograft model of MM and this effect synergized with PRIMA-1Met[93].

miR-33b is down-regulated in osteosarcoma tumors and cell lines and its overexpression significantly inhibits proliferation, migration, and invasion of and by osteosarcoma cells. Mechanically miR-33b negatively regulates Myc at the posttranscriptional level, via a specific target site within the 3'UTR. Over-expression of Myc impaired miR-33b-induced inhibition of proliferation and invasion in osteosarcoma cells. The expression of Myc was inversely correlated with miR-33b expression in osteosarcoma tumors and cells. Taken together, these findings suggest that miR-33b inhibits osteosarcoma cell migration and invasion and serves as a tumor suppressor by targeting Myc expression [94].

miR-130a has also been shown to be a Myc-targeting miR. In these studies, ribosomal protein L11 promoted miR-130a's targeting of the 3'-UTR of Myc mRNA thus repressing Myc Expressionin response to UV irradiation. miR-130a overexpression promoted Ago2 binding to Myc mRNA, significantly reduced the levels of both Myc mRNA and protein and inhibited cell proliferation. UV treatment markedly enhanced the binding of L11 to miR-130a, Myc mRNA and Ago2 in cells. Inhibiting miR-130a significantly suppressed UV-mediated Myc reduction. L11 was re-localized from the nucleolus to the cytoplasm where it associated with Myc mRNA upon UV treatment. These findings reveal a novel mechanism underlying Myc down-regulation in response to UV-mediated DNA damage[95]. Interestingly miR-24 inhibited Myc expression by recognizing seedless but highly complementary 3'-UTR sequences[96]. All known Myc-targeting miRs are listed in Table 3.

Table 3.

Myc-targeting miRs

No. miR Location Function Cancer type Reference(s)
1 let7 family
member
9q22.32,
19q13.41,
22q13.31
Growth, transformation, cisplatin resistance,
radio-sensitivity, invasion, Differentiation
Lymphoma, HCC,melanoma,ovarian
carcinoma, breast cancer, lung
cancer,medulloblastoma
[30,69,86,87]
2 miR-24 9q22.32 Development, cancer progression HCC,lymphoma [96]
3 miR-29a 7q32.3 Growth, migration, invasion Leukemia, lymphoma, lung cancer [93]
4 miR-33b 17p11.2 EMT,migration, invasion Melanoma,osteosarcoma [94]
5 miR-34a 1p36.22 Proliferation, migration, invasion Bladder cancer,sarcomas [30,89-91]
6 miR-34b/c 11q23.1 Metastasis, EMTgrowth, apoptosis,
chemo-resistance,self-renewal,
Colon cancer, breast
cancer,HCC,renalcancer,lung cancer
[30,92,97]
7 miR-125b 11q24.1 Tumor growth, proliferation Breast
cancer,pancreaticcancer,melanoma,HCC
[69]
8 miR-130a 11q12.1 Proliferation,migration,invasion, gefitinib
resistance
Breast cancer,HCC, lung cancer [95]
9 miR-132 17p13.3 Growth, migration, invasion Glioma,lung cancer [69]
10 miR-135a 3p21.1 Migration, invasion,metastasis Prostate cancer,HCC [98]
11 miR-145 5q32 Metastasis,invasion Bladder cancer,HCC,prostate cancer [39,88]
12 miR-148a 7p15.2 Cell cycle, Differentiation HCC, Squamous cell carcinoma [25]
13 miR-154 14q32.31 Growth, EMT Colorectal cancer,lung cancer, prostate
cancer.
[69]
14 miR-184 15q25.1 Proliferation, invasion Ovarian cancer, glioma,lung cancers [99,100]
15 miR-320b 1q42.1 Proliferation, invasion Colorectal cancer, [101]
16 miR-451 17q11.2 Radio-resistance, proliferation Lung, head and neck cancer [102-104]
17 miR-494 14q32.31 Growth, proliferation, cell cycle Gastric carcinoma, lymphoma, HCC,
ovarian cancer
[105-107]
18 miR-744 17p12 Proliferation HCC [108]

Myc-synthetic lethal interactions

In some murine models of Myc-dependent cancers, the suppression of Myc leads to tumor regression, thus indicating that continuous Myc expression is necessary to maintain tumor growth and/or viability[11-13,18]. Myc's role in the pathogenesis of cancer and the apparent “addiction” of some cancer to continuous Myc expression makes it an enticing, albeit elusive chemotherapeutic target[11-13,18,109-113]. Strategies aimed at inhibiting Myc itself including antisense oligodeoxy nucleotides and RNAi, small molecule disruptors of the Myc-Max heterodimer and inhibitors of chromatin modifying complexes with which Myc associates have been investigated as options for attacking Myc-driven cancers[11-13,18,109-120]. However these attempts to pharmacologically inhibit Myc have remained, to various degrees, ineffective and/or non-specific[109-120].

Because Myc is essential for the growth of some normal as well as transformed cells, it has sometimes been viewed as being a poor therapeutic target given that the protein is also seldom mutated in cancer and possesses no obviously “druggable” domains or enzymatic activity[11-13,109,110,113]. Thus it has been suggested that taking advantage of Myc's synthetic lethal interactions could be exploited as an effective therapeutic strategy in Myc-driven human cancers[109,110,112,121-137].

A synthetic lethal interaction is defined as a combination of two mutations that confers lethality to a cell without either individual mutation alone having such an effect[109,110,112,121-127]. A classical example of synthetic lethality is seen with the use of poly ADP-ribose polymerase (PARP) inhibitors to treat BRCA1- and BRCA2-deficient breast cancers. The inability of these cancers to repair damaged DNA in response to PARP inhibition occurs only in tumor cells that are defective in BRCA1/2-mediated DNA repair and not in otherwise normal cells[121-125]. Because tumor cells contain numerous genetic, epigenetic and metabolic alterations, a large number of heterogenous synthetic lethal genes have been identified whose inhibition confers synthetic lethality[109,110,112,121-127]. Among the molecules that have been used to identify synthetic lethal targets, both in vitro and in vivo, are low molecular weight drug-like compounds, siRNA/shRNAs, CRISPRs and miRs[109,110,112,121-127].

Previous siRNA/shRNA screens for Myc-synthetic lethal have identified regulators of the Myc network, components of RNA transcription initiation and elongation, proteins involved in sumoylation and ubiquitylation and kinases with roles in DNA repair and cell cycle checkpoints[109,110,112,128-137]. More recently, the spliceosome was discoveredto be a new target of oncogenic stress in Myc-driven cancers. BUD31, a core component of the spliceosome, was identified as a Myc synthetic lethal gene in human mammary epithelial cells. Other core spliceosome factors that interact with BUD31, such U2AF1 and SF3B1, were also identified as Myc synthetic lethals. Myc hyperactivation increases the total cellular burden of precursor mRNA synthesis, thus likely placing greater demands and stress on the pre-RNA processing function of spliceosomes. In contrast to normal cells, partial inhibition of the spliceosome in Myc-hyperactivated cells was found to be associated with global intron retention, widespread defects in pre-mRNA maturation and deregulation of many essential cell processes. Genetic or pharmacological inhibition of the spliceosome in vivo also impaired survival, tumorigenicity and metastatic proclivity of breast cancers with high Myc level but had a less pronounced effect in tumors with lower Myc levels. Collectively, these findings suggest that oncogenic Myc confers a collateral stress on splicing, and that components of the spliceosome may be viable synthetic lethal therapeutic targets for Myc-driven cancers[128,129].

SAE2, an endonuclease that plays an important role in DNA repair and that is necessary for the growth of cancers with high Myc levels, has also been shown to interact with de-regulated Myc in a synthetic lethal manner[130].

Inhibition of some cyclin-dependent kinases was also reported to be selectively toxic in breast cancers with high Myc levels [131-133]. For example, CDK9 is a component of the Myc-regulated RNA polymerase II-directed transcriptional elongation machinery and functions to phosphorylate the C-terminal domain of RNA polymerase II's largest subunit[133]. Myc-synthetic lethal genes are summarized in Table 4.

Table 4.

Summary of synthetic lethal targets of Myc oncogene

No. Genes Location Function Reference
1 AURKB 17p13.1 Cell division [134]
2 BUD31 7q22.1 RNA splicing [128,129]
3 CDK1 10q21.1 Cell cycle [131,132]
4 CDK9 9q34.1 Transcription elongation [133]
5 DR5 8p22-p21 Cell death [135]
6 EIF4F 4q23 Protein synthesis [136]
7 MAP3K13 3q27 Protein kinase [137]
8 miR-206 6p12.2 Tumor suppressor miR [137]
9 SAE2 18q11.2 Sumoylation/DNA repair [130]
10 SF3B1 2q33.1 RNA splicing [128,129]
11 U2AF1 21q22.3 RNA splicing [128,129]

A few reports have used miRs to identify synthetic lethal relationships in cancer cells that over-express Myc or other oncogenes. By functionally screening a miR library comprised of 1254 individual miR expression vector, we identified miR-206 as being able to impart a synthetic lethal phenotype in Myc over-expressing human cancer cells[137]. miR-206had similar growth inhibitory effects on both the in vitro and in vivo proliferation of breast cancer cells, but only when they expressed high levels of Myc as would be expected of a classic synthetic lethal relationship. This appeared to be dependent on miR-206's ability to target MAP3K13 as indicated by the ability to recapitulate similar phenotypes by suppressing MAPK3K13 through other means, including two different shRNAs. Bioinformatics-based analyses of gene expression profiles from the TCGA collection of breast cancers indicated that those tumors with the highest Myc levels tended to express the lowest levels of miR-206 and the highest levels of MAP3K13 while simultaneously being associated with more adverse clinical outcomes. The critical link between miR-206 and MAP3K13 in the development of Myc over-expressing human cancers suggested potential points of therapeutic intervention for this molecular sub-category.These results showed that MAP3K13 and perhaps downstream targets of MAP3K13 such as members of the JNK signaling pathway are Myc-synthetic lethal targets[137].

Future directions

The direct inhibition of certain“undruggable”oncoproteins, such as Myc remains an elusive goal in cancer therapy. As a way of circumventing this problem, synthetic lethal interactions could be exploited as an alternate therapeutic strategy, thus avoiding the inherent complexities of direct Myc inhibition. Given that miRs are easy to synthetize and can be readily introduced into cells, they possess significant potential for cancer therapy. Alternatively, even easier and more direct approaches could be envisioned in cases where the synthetic lethal genes identified by such screens encoded enzymes for which small molecule inhibitors could be generated. Such small molecules could then be used as substitutes for miRs or perhaps even in combination. miRs thus have potential not only as therapeutic entities but as tools to drive the ongoing discovery of new synthetic lethal interactions. Many approaches to identify synthetic lethal interactions have been successfully developed. Currently, most screens continue to be performed in vitro using cultured cells, primarily because of the relative ease and directness of high throughput as well as the potential for multiplexed-based screens. The latter are likely to grow in popularity, given the ease with which such screens can now be combined withdeep sequencing-based approaches to identify relevant miRs or shRNAs that have been selectively depleted from complex libraries. The potential for screening in vivo using tractable organisms such as zebrafish are also high and further promises to reduce the time needed to identify promising hits, characterize their mechanisms of action and proceed to clinical trials[138, 139].

Acknowledgements

This work was supported by grants from the National Natural Science Foundation of China (81472549) and by National Institutes of Health grants RO1 CA174713 and RO1 CA140624 to EVP.

Abbreviations

MiRs

microRNAs

Myc

c-Myc

pri-miR

primary miR

pre-miR

precursor miR

RISC

RNA-induced silencing complex

3’UTR

3’-untranslated region

HLH-ZIP

helix-loop-helix leucine zipper

EMT

epithelial-to-mesenchymal transition

HCC

hepatocellular carcinoma

DEN

diethylnitrosamine

PDK4

pyruvate dehydrogenase kinase 4

PDH

pyruvate dehydrogenase

MK5

mitogen-activated protein kinase-activated protein kinase 5

PARP

poly ADP-ribose polymerase

Footnotes

Conflicting interests

The authors have declared that no conflict of interests exist.

References

  • 1.Eulalio A, Huntzinger E, Izaurralde E. Getting to the root of miRNA-mediated gene silencing. Cell. 2008;132:9–14. doi: 10.1016/j.cell.2007.12.024. [DOI] [PubMed] [Google Scholar]
  • 2.Bartel DP. MicroRNAs: target recognition and regulatory functions. Cell. 2009;136:215–233. doi: 10.1016/j.cell.2009.01.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Adams BD, Kasinski AL, Slack FJ. Aberrant regulation and function of microRNAs in cancer. CurrBiol. 2014;24:R762–776. doi: 10.1016/j.cub.2014.06.043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Mendell JT, Olson EN. MicroRNAs in stress signaling and human disease. Cell. 2012;148:1172–1187. doi: 10.1016/j.cell.2012.02.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Calin GA, Croce CM. MicroRNA signatures in human cancers. Nat Rev Cancer. 2006;6:857–866. doi: 10.1038/nrc1997. [DOI] [PubMed] [Google Scholar]
  • 6.Iorio MV, Croce CM. MicroRNAs in cancer: small molecules with a huge impact. J ClinOncol. 2009;27:5848–5856. doi: 10.1200/JCO.2009.24.0317. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Kasinski AL, Slack FJ. Epigenetics and genetics. MicroRNAs en route to the clinic: progress in validating and targeting microRNAs for cancer therapy. Nat Rev Cancer. 2011;11:849–864. doi: 10.1038/nrc3166. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Lujambio A, Calin GA, Villanueva A, Ropero S, Sanchez-Cespedes M, Blanco D, et al. A microRNA DNA methylation signature for human cancer metastasis. ProcNatlAcadSci U S A. 2008;105:13556–13561. doi: 10.1073/pnas.0803055105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Datta J, Kutay H, Nasser MW, Nuovo GJ, Wang B, Majumder S, et al. Methylation mediated silencing of MicroRNA-1 gene and its role in hepatocellular carcinogenesis. Cancer Res. 2008;68:5049–5058. doi: 10.1158/0008-5472.CAN-07-6655. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 10.Medina PP, Nolde M, Slack FJ. OncomiR addiction in an in vivo model of microRNA-21-induced pre-B-cell lymphoma. Nature. 2010;467:86–90. doi: 10.1038/nature09284. [DOI] [PubMed] [Google Scholar]
  • 11.Meyer N, Penn LZ. Reflecting on 25 years with MYC. Nat Rev Cancer. 2008;8:976–990. doi: 10.1038/nrc2231. [DOI] [PubMed] [Google Scholar]
  • 12.Dang CV. MYC on the path to cancer. Cell. 2012;149:22–35. doi: 10.1016/j.cell.2012.03.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Laurenti E, Wilson A, Trumpp A. Myc's other life: stem cells and beyond. CurrOpin Cell Biol. 2009;21:844–854. doi: 10.1016/j.ceb.2009.09.006. [DOI] [PubMed] [Google Scholar]
  • 14.Mateyak MK, Obaya AJ, Adachi S, Sedivy JM. Phenotypes ofc-Myc-deficient rat fibroblasts isolated by targeted homologous recombination. Cell Growth Differ. 1997;8:1039–1048. [PubMed] [Google Scholar]
  • 15.Nesbit CE, Tersak JM, Prochownik EV. MYC oncogenes and human neoplastic disease. Oncogene. 1999;18:3004–3016. doi: 10.1038/sj.onc.1202746. [DOI] [PubMed] [Google Scholar]
  • 16.Gao P, Tchernyshyov I, Chang TC, Lee YS, Kita K, Ochi T, et al. c-Myc suppression of miR-23a/b enhances mitochondrial glutaminase expression and glutamine metabolism. Nature. 2009;458:762–765. doi: 10.1038/nature07823. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Secombe J, Li L, Carlos L, Eisenman RN. The Trithorax group protein Lid is a trimethyl histone H3K4 demethylase required for dMyc-induced cell growth. Genes Dev. 2007;21:537–551. doi: 10.1101/gad.1523007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Shachaf CM, Kopelman AM, Arvanitis C, Karlsson A, Beer S, Mandl S, et al. MYC inactivation uncovers pluripotent differentiation and tumour dormancy in hepatocellular cancer. Nature. 2004;431:1112–1117. doi: 10.1038/nature03043. [DOI] [PubMed] [Google Scholar]
  • 19.Gomez-Roman N, Grandori C, Eisenman RN, White RJ. Direct activation of RNA polymerase III transcription by c-Myc. Nature. 2003;421:290–294. doi: 10.1038/nature01327. [DOI] [PubMed] [Google Scholar]
  • 20.Wang H, Sharma L, Lu J, Finch P, Fletcher S, Prochownik EV. Structurally diverse c-Myc inhibitors share a common mechanism of action involving ATP depletion. Oncotarget. 2015;6:15857–15870. doi: 10.18632/oncotarget.4327. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Wang H, Mannava S, Grachtchouk V, Zhuang D, Soengas MS, Gudkov AV, et al. c-Myc depletion inhibits proliferation of human tumor cells at various stages of the cell cycle. Oncogene. 2008;27:1905–1915. doi: 10.1038/sj.onc.1210823. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Vita M, Henriksson M. The Myconcoprotein as a therapeutic target for human cancer. Semin Cancer Biol. 2006;16:318–330. doi: 10.1016/j.semcancer.2006.07.015. [DOI] [PubMed] [Google Scholar]
  • 23.Lin CP, Liu CR, Lee CN, Chan TS, Liu HE. Targeting c-Myc as a novel approach for hepatocellular carcinoma. World J Hepatol. 2010;2:16–20. doi: 10.4254/wjh.v2.i1.16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Chan KL, Guan XY, Ng IO. High-throughput tissue microarray analysis of c-myc activation in chronic liver diseases and hepatocellular carcinoma. Hum Pathol. 2004;35:1324–1331. doi: 10.1016/j.humpath.2004.06.012. [DOI] [PubMed] [Google Scholar]
  • 25.Han H, Sun D, Li W, Shen H, Zhu Y, Li C, et al. A c-Myc-MicroRNA functional feedback loop affects hepatocarcinogenesis. Hepatology. 2013;57:2378–2389. doi: 10.1002/hep.26302. [DOI] [PubMed] [Google Scholar]
  • 26.He L, Thomson JM, Hemann MT, Hernando-Monge E, Mu D, Goodson S, et al. A microRNA polycistron as a potential human oncogene. Nature. 2005;435:828–833. doi: 10.1038/nature03552. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.O'Donnell KA, Wentzel EA, Zeller KI, Dang CV, Mendell JT. c-Myc-regulated microRNAs modulate E2F1 expression. Nature. 2005;435:839–843. doi: 10.1038/nature03677. [DOI] [PubMed] [Google Scholar]
  • 28.Dews M, Homayouni A, Yu D, Murphy D, Sevignani C, Wentzel E, et al. Augmentation of tumor angiogenesis by a Myc-activated microRNA cluster. Nat Genet. 2006;38:1060–1065. doi: 10.1038/ng1855. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Bui TV, Mendell JT. Myc: Maestro of MicroRNAs. Genes Cancer. 2010;1:568–575. doi: 10.1177/1947601910377491. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Jackstadt R, Hermeking H. MicroRNAs as regulators and mediators of c-MYC function. BiochimBiophysActa. 2015;1849:544–553. doi: 10.1016/j.bbagrm.2014.04.003. [DOI] [PubMed] [Google Scholar]
  • 31.Chang TC, Yu D, Lee YS, Wentzel EA, Arking DE, West KM, et al. Widespread microRNA repression by Myc contributes to tumorigenesis. Nat Genet. 2008;40:43–50. doi: 10.1038/ng.2007.30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Thorgeirsson SS. The almighty MYC: orchestrating the micro-RNA universe to generate aggressive liver cancer. J Hepatol. 2011;55:486–487. doi: 10.1016/j.jhep.2011.01.042. [DOI] [PubMed] [Google Scholar]
  • 33.Wang GG, Konze KD, Tao J. Polycomb genes, miRNA, and their deregulation in B-cell malignancies. Blood. 2015;125:1217–1225. doi: 10.1182/blood-2014-10-606822. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Psathas JN, Thomas-Tikhonenko A. MYC and the art of microRNA maintenance. Cold Spring HarbPerspect Med. 2014;4:a014175. doi: 10.1101/cshperspect.a014175. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Tao J, Zhao X, Tao J. c-MYC-miRNA circuitry: a central regulator of aggressive B-cell malignancies. Cell Cycle2014. 13:191–198. doi: 10.4161/cc.27646. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Cairo S, Wang Y, de Reynies A, Duroure K, Dahan J, Redon MJ, et al. Stem cell-like micro-RNA signature driven by Myc in aggressive liver cancer. ProcNatlAcadSci U S A. 2010;107:20471–20476. doi: 10.1073/pnas.1009009107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Buendia MA, Bourre L, Cairo S. MyctargetmiRs and liver cancer: small molecules to get Myc sick. Gastroenterology. 2012;142:214–218. doi: 10.1053/j.gastro.2011.12.023. [DOI] [PubMed] [Google Scholar]
  • 38.Chang TC, Zeitels LR, Hwang HW, Chivukula RR, Wentzel EA, Dews M, et al. Lin-28B transactivation is necessary for Myc-mediated let-7 repression and proliferation. ProcNatlAcadSci U S A. 2009;106:3384–3389. doi: 10.1073/pnas.0808300106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Sachdeva M, Zhu S, Wu F, Wu H, Walia V, Kumar S, et al. p53 represses c-Myc through induction of the tumor suppressor miR-145. ProcNatlAcadSci U S A. 2009;106:3207–3212. doi: 10.1073/pnas.0808042106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Ventura A, Young AG, Winslow MM, Lintault L, Meissner A, Erkeland SJ, et al. Targeted deletion reveals essential and overlapping functions of the miR-17 through 92 family of miRNA clusters. Cell. 2008;132:875–886. doi: 10.1016/j.cell.2008.02.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.dePontual L, Yao E, Callier P, Faivre L, Drouin V, Cariou S, et al. Germline deletion of the miR-17 approximately 92 cluster causes skeletal and growth defects in humans. Nat Genet. 2011;43:1026–1030. doi: 10.1038/ng.915. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Sandhu SK, Fassan M, Volinia S, Lovat F, Balatti V, Pekarsky Y, et al. B-cell malignancies in microRNA Eμ-miR-17~92 transgenic mice. ProcNatlAcadSci USA. 2013;110:18208–18213. doi: 10.1073/pnas.1315365110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Lu Y, Thomson JM, Wong HY, Hammond SM, Hogan BL. Transgenic over-expression of the microRNA miR-17-92 cluster promotes proliferation and inhibits differentiation of lung epithelial progenitor cells. DevBiol. 2007;310:442–453. doi: 10.1016/j.ydbio.2007.08.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Ma L, Young J, Prabhala H, Pan E, Mestdagh P, Muth D, et al. miR-9, a MYC/MYCN-activated microRNA, regulates E-cadherin and cancer metastasis. Nat Cell Biol. 2010;12:247–256. doi: 10.1038/ncb2024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Zou Z, Chen J, Liu A, Zhou X, Song Q, Jia C, et al. mTORC2 promotes cell survival through c-Myc-dependent up-regulation of E2F1. J Cell Biol. 2015;211:105–122. doi: 10.1083/jcb.201411128. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Song Y, Li J, Zhu Y, Dai Y, Zeng T, Liu L, et al. MicroRNA-9 promotes tumor metastasis via repressing E-cadherin in esophageal squamous cell carcinoma. Oncotarget. 2014;5:11669–11680. doi: 10.18632/oncotarget.2581. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Liu M, Zhu H, Yang S, Wang Z, Bai J, Xu N. c-Myc suppressed E-cadherin through miR-9 at the post-transcriptional level. Cell BiolInt. 2013;37:197–202. doi: 10.1002/cbin.10039. [DOI] [PubMed] [Google Scholar]
  • 48.Feng M, Li Z, Aau M, Wong CH, Yang X, Yu Q. Myc/miR-378/TOB2/cyclin D1 functional module regulates oncogenic transformation. Oncogene. 2011;30:2242–2251. doi: 10.1038/onc.2010.602. [DOI] [PubMed] [Google Scholar]
  • 49.Mihailovich M, Bremang M, Spadotto V, Musiani D, Vitale E, Varano G, et al. miR-17-92fine-tunes MYC expression and function to ensure optimal B cell lymphoma growth. Nat Commun. 2015;6:8725. doi: 10.1038/ncomms9725. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Zhu H, Han C, Lu D, Wu T. miR-17-92cluster promotes cholangiocarcinoma growth: evidence for PTEN as downstream target and IL-6/Stat3 as upstream activator. Am J Pathol. 2014;184:2828–2839. doi: 10.1016/j.ajpath.2014.06.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Liu XS, Chopp M, Wang XL, Zhang L, Hozeska-Solgot A, Tang T, et al. MicroRNA-17-92 cluster mediates the proliferation and survival of neural progenitor cells after stroke. J BiolChem. 2013;288:12478–12488. doi: 10.1074/jbc.M112.449025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Yang S, Cho YJ, Jin L, Yuan G, Datta A, Buckhaults P, et al. An epigenetic auto-feedback loop regulates TGF-beta type II receptor expression and function in NSCLC. Oncotarget. 2015;6:33237–33252. doi: 10.18632/oncotarget.4893. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Dhar S, Kumar A, Rimando AM, Zhang X, Levenson AS. Resveratrol and pterostilbene epigenetically restorePTENexpression by targeting oncomiRs of the miR-17 family in prostate cancer. Oncotarget. 2015;6:27214–27226. doi: 10.18632/oncotarget.4877. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Battistella M, Romero M, Castro-Vega LJ, Gapihan G, Bouhidel F, Bagot M, et al. The High Expression of the microRNA 17-92 Cluster and its Paralogs, and the Downregulation of the Target Gene PTEN, Is Associated with Primary Cutaneous B-Cell Lymphoma Progression. J Invest Dermatol. 2015;135:1659–1667. doi: 10.1038/jid.2015.27. [DOI] [PubMed] [Google Scholar]
  • 55.Wu W, Takanashi M, Borjigin N, Ohno SI, Fujita K, Hoshino S, et al. MicroRNA-18a modulates STAT3 activity through negative regulation of PIAS3 during gastric adenocarcinogenesis. Br J Cancer. 2013;108:653–661. doi: 10.1038/bjc.2012.587. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Wu Q, Yang Z, An Y, Hu H, Yin J, Zhang P, et al. MiR-19a/b modulate the metastasis of gastric cancer cells by targeting the tumour suppressor MXD1. Cell Death Dis. 2014;5:e1144. doi: 10.1038/cddis.2014.110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Hu S, Liu L, Chang EB, Wang JY, Raufman JP. Butyrate inhibits pro-proliferative miR-92a by diminishing c-Myc-induced miR-17-92a cluster transcription in human colon cancer cells. Mol Cancer. 2015;14:180. doi: 10.1186/s12943-015-0450-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Kan T, Sato F, Ito T, Matsumura N, David S, Cheng Y, et al. The miR-106b-25 polycistron, activated by genomic amplification, functions as an oncogene by suppressing p21 and Bim. Gastroenterology. 2009;136:1689–1700. doi: 10.1053/j.gastro.2009.02.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Razumilava N, Bronk SF, Smoot RL, Fingas CD, Werneburg NW, Roberts LR, et al. miR-25 targets TNF-related apoptosis inducing ligand (TRAIL) death receptor-4 and promotes apoptosis resistance in cholangiocarcinoma. Hepatology. 2012;55:465–475. doi: 10.1002/hep.24698. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Wang Z, Liu M, Zhu H, Zhang W, He S, Hu C, et al. Suppression of p21 by c-Myc through members of miR-17 family at the post-transcriptional level. Int J Oncol. 2010;37:1315–1321. [PubMed] [Google Scholar]
  • 61.Tang Q, Zou Z, Zou C, Zhang Q, Huang R, Guan X, et al. MicroRNA-93 suppress colorectal cancer development via Wnt/beta-catenin pathway downregulating. TumourBiol. 2015;36:1701–1710. doi: 10.1007/s13277-014-2771-6. [DOI] [PubMed] [Google Scholar]
  • 62.Castellano L, Giamas G, Jacob J, Coombes RC, Lucchesi W, Thiruchelvam P, et al. The estrogen receptor-alpha-induced microRNA signature regulates itself and its transcriptional response. ProcNatlAcadSci U S A. 2009;106:15732–15737. doi: 10.1073/pnas.0906947106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Xie X, Liu HT, Mei J, Ding FB, Xiao HB, Hu FQ, et al. miR-106a promotes growth and metastasis of non-small cell lung cancer by targeting PTEN. Int J ClinExpPathol. 2015;8:3827–3834. [PMC free article] [PubMed] [Google Scholar]
  • 64.Kumar P, Luo Y, Tudela C, Alexander JM, Mendelson CR. The c-Myc-regulated microRNA-17~92 and miR-106a~363 clusters target hCYP19A1 and hGCM1 to inhibit human trophoblast differentiation. Mol Cell Biol. 2013;33:1782–1796. doi: 10.1128/MCB.01228-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Tan W, Li Y, Lim SG, Tan TM. miR-106b-25/miR-17-92 clusters: polycistrons with oncogenic roles in hepatocellular carcinoma. World J Gastroenterol. 2014;20:5962–5672. doi: 10.3748/wjg.v20.i20.5962. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Navarro A, Pairet S, Alvarez-Larran A, Pons A, Ferrer G, Longaron R, et al. miR-203 and miR-221 regulate SOCS1 and SOCS3 in essential thrombocythemia. Blood Cancer J. 2016;6:e406. doi: 10.1038/bcj.2016.10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Yeh CH, Jin L, Shen F, Balian G, Li X. miR-221 attenuates the osteogenic differentiation of human annulus fibrosus cells. Spine J. 2016:S1529–9430. doi: 10.1016/j.spinee.2016.03.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Mitxelena J, Apraiz A, Vallejo-Rodriguez J, Malumbres M, Zubiaga AM. E2F7 regulates transcription and maturation of multiple microRNAs to restrain cell proliferation. Nucleic Acids Res. 2016:gkw146. doi: 10.1093/nar/gkw146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Bueno MJ, Gomez de Cedron M, Gomez-Lopez G, Perez de Castro I, Di Lisio L, Montes-Moreno S, et al. Combinatorial effects of microRNAs to suppress the Myc oncogenic pathway. Blood. 2011;117:6255–6266. doi: 10.1182/blood-2010-10-315432. [DOI] [PubMed] [Google Scholar]
  • 70.Sander S, Bullinger L, Klapproth K, Fiedler K, Kestler HA, Barth TF, et al. MYC stimulates EZH2 expression by repression of its negative regulator miR-26a. Blood. 2008;112:4202–12. doi: 10.1182/blood-2008-03-147645. [DOI] [PubMed] [Google Scholar]
  • 71.Zhang X, Zhao X, Fiskus W, Lin J, Lwin T, Rao R, et al. Coordinated silencing ofMYC-mediated miR-29 by HDAC3 and EZH2 as a therapeutic target of histone modification in aggressive B-Cell lymphomas. Cancer Cell. 2012;22:506–523. doi: 10.1016/j.ccr.2012.09.003. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 72.Schneider C, Setty M, Holmes AB, Maute RL, Leslie CS, Mussolin L, et al. MicroRNA-28 controls cell proliferation and is down-regulated in B-cell lymphomas. ProcNatlAcadSci U S A. 2014;111:8185–8190. doi: 10.1073/pnas.1322466111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.RokavecM,Öner MG, Li H, Jackstadt R, Jiang L, Lodygin D, et al. IL-6R/STAT3/miR-34afeedbacklooppromotesEMT-mediatedcolor ectal cancerinvasion and metastasis. J Clin Invest. 2014;124:1853–1867. [Google Scholar]
  • 74.Wang B, Hsu SH, Wang X, Kutay H, Bid HK, Yu J, et al. Reciprocal regulation of microRNA-122 andc-Mycin hepatocellular cancer: role of E2F1 and transcription factor dimerization partner 2. Hepatology. 2014;59:555–566. doi: 10.1002/hep.26712. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Han H, Li W, Shen H, Zhang J, Zhu Y, Li Y. microRNA-129-5p, a c-Myc negative target, affects hepatocellular carcinoma progression by blocking the Warburg effect. J Mol Cell Biol. 2016:mjw010. doi: 10.1093/jmcb/mjw010. [DOI] [PubMed] [Google Scholar]
  • 76.Kawano M, Tanaka K, Itonaga I, Iwasaki T, Tsumura H. c-Myc Represses Tumor-Suppressive microRNAs, let-7a, miR-16 and miR-29b, and Induces Cyclin D2-Mediated Cell Proliferation in Ewing's Sarcoma Cell Line. PLoS One. 2015;10:e0138560. doi: 10.1371/journal.pone.0138560. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Liu Z, Cheng C, Luo X, Xia Q, Zhang Y, Long X, et al. CDK4 and miR-15a comprise an abnormal automodulatory feedback loop stimulating the pathogenesis and inducing chemotherapy resistance in nasopharyngeal carcinoma. BMC Cancer. 2016;16:238. doi: 10.1186/s12885-016-2277-2. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 78.Jiang Q, Zhang Y, Zhao M, Li Q, Chen R, Long X, et al. miR-16 induction after CDK4 knockdown is mediated by c-Myc suppression and inhibits cell growth as well as sensitizes nasopharyngeal carcinoma cells to chemotherapy. TumourBiol. 2015 doi: 10.1007/s13277-015-3966-1. doi:10.1007/s13277-015-3966-1. [DOI] [PubMed] [Google Scholar]
  • 79.Rogler CE, Levoci L, Ader T, Massimi A, Tchaikovskaya T, Norel R, et al. MicroRNA-23b cluster microRNAs regulate transforming growth factor-beta/bone morphogenetic protein signaling and liver stem cell differentiation by targeting Smads. Hepatology. 2009;50:575–584. doi: 10.1002/hep.22982. [DOI] [PubMed] [Google Scholar]
  • 80.Ma Y, Yao N, Liu G, Dong L, Liu Y, Zhang M, et al. Functional screen reveals essential roles of miR-27a/24 in differentiation of embryonic stem cells. EMBO J. 2015;34:361–378. doi: 10.15252/embj.201489957. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Chen Q, Zhang F, Wang Y, Liu Z, Sun A, Zen K, et al. The transcription factor c-Myc suppresses MiR-23b and MiR-27b transcription during fetal distress and increases the sensitivity of neurons to hypoxia-induced apoptosis. PLoS One. 2015;10:e0120217. doi: 10.1371/journal.pone.0120217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Huang H, Jiang X, Wang J, Li Y, Song CX, Chen P, et al. Identification of MLL-fusion/MYC dash, vertical miR-26 dash, vertical TET1 signaling circuit in MLL-rearranged leukemia. Cancer Lett. 2016;372:157–165. doi: 10.1016/j.canlet.2015.12.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Hand NJ, Master ZR, Eauclaire SF, Weinblatt DE, Matthews RP, Friedman JR. The microRNA-30 family is required for vertebrate hepatobiliary development. Gastroenterology. 2009;136:1081–1090. doi: 10.1053/j.gastro.2008.12.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Liao JM, Lu H. Autoregulatory suppression of c-Myc by miR-185-3p. J BiolChem. 2011;286:33901–33909. doi: 10.1074/jbc.M111.262030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Sanford EL, Choy KW, Donahoe PK, Tracy AA, Hila R, Loscertales M, et al. MiR-449a Affects Epithelial Proliferation during the Pseudoglandular and Canalicular Phases of Avian and Mammal Lung Development. PLoS One. 2016;11:e0149425. doi: 10.1371/journal.pone.0149425. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Sampson VB, Rong NH, Han J, Yang Q, Aris V, Soteropoulos P, et al. MicroRNA let-7a down-regulates MYC and reverts MYC-induced growth in Burkitt lymphoma cells. Cancer Res. 2007;67:9762–9770. doi: 10.1158/0008-5472.CAN-07-2462. [DOI] [PubMed] [Google Scholar]
  • 87.Yang X, Cai H, Liang Y, Chen L, Wang X, Si R, et al. Inhibition of c-Myc by let-7b mimic reverses mutidrug resistance in gastric cancer cells. Oncol Rep. 2015;33:1723–1730. doi: 10.3892/or.2015.3757. [DOI] [PubMed] [Google Scholar]
  • 88.Wang F, Xia J, Wang N, Zong H. miR-145 inhibits proliferation and invasion of esophageal squamous cell carcinoma in part by targeting c-Myc. Onkologie. 2013;36:754–758. doi: 10.1159/000356978. [DOI] [PubMed] [Google Scholar]
  • 89.Yamamura S, Saini S, Majid S, Hirata H, Ueno K, Deng G, et al. MicroRNA-34a modulates c-Myc transcriptional complexes to suppress malignancy in human prostate cancer cells. PLoS One. 2012;7:e29722. doi: 10.1371/journal.pone.0029722. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Xu X, Chen W, Miao R, Zhou Y, Wang Z, Zhang L, et al. miR-34a induces cellular senescence via modulation of telomerase activity in human hepatocellular carcinoma by targeting FoxM1/c-Myc pathway. Oncotarget. 2015;6:3988–4004. doi: 10.18632/oncotarget.2905. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Yamamura S, Saini S, Majid S, Hirata H, Ueno K, Chang I, et al. MicroRNA-34a suppresses malignant transformation by targeting c-Myc transcriptional complexes in human renal cell carcinoma. Carcinogenesis. 2012;33:294–300. doi: 10.1093/carcin/bgr286. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Kress TR, Cannell IG, Brenkman AB, Samans B, Gaestel M, Roepman P, et al. The MK5/PRAK kinase and Myc form a negative feedback loop that is disrupted during colorectal tumorigenesis. Mol Cell. 2011;41:445–457. doi: 10.1016/j.molcel.2011.01.023. [DOI] [PubMed] [Google Scholar]
  • 93.Saha MN, Abdi J, Yang Y, Chang H. miRNA-29a as a tumor suppressor mediates PRIMA-1Met-induced anti-myeloma activity by targeting c-Myc. Oncotarget. 2016;7:7149–7160. doi: 10.18632/oncotarget.6880. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Xu N, Li Z, Yu Z, Yan F, Liu Y, Lu X, et al. MicroRNA-33b suppresses migration and invasion by targeting c-Myc in osteosarcoma cells. PLoS One. 2014;9:e115300. doi: 10.1371/journal.pone.0115300. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 95.Li Y, Challagundla KB, Sun XX, Zhang Q, Dai MS. MicroRNA-130a associates with ribosomal protein L11 to suppress c-Myc expression in response to UV irradiation. Oncotarget. 2015;6:1101–1014. doi: 10.18632/oncotarget.2728. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Lal A, Navarro F, Maher CA, Maliszewski LE, Yan N, O'Day E, et al. miR-24 Inhibits cell proliferation by targeting E2F2, MYC, and other cell-cycle genes via binding to “seedless” 3′UTR microRNA recognition elements. Mol Cell. 2009;35:610–625. doi: 10.1016/j.molcel.2009.08.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Benassi B, Flavin R, Marchionni L, Zanata S, Pan Y, Chowdhury D, et al. MYC is activated by USP2a-mediated modulation of microRNAs in prostate cancer. Cancer Discov. 2012;2:236–247. doi: 10.1158/2159-8290.CD-11-0219. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Yamada Y, Hidaka H, Seki N, Yoshino H, Yamasaki T, Itesako T, et al. Tumor-suppressive microRNA-135a inhibits cancer cell proliferation by targeting the c-MYC oncogene in renal cell carcinoma. Cancer Sci. 2013;104:304–312. doi: 10.1111/cas.12072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Liu Z, Mai C, Yang H, Zhen Y, Yu X, Hua S, et al. Candidate tumour suppressor CCDC19 regulates miR-184 direct targeting of C-Myc thereby suppressing cell growth in non-small cell lung cancers. J Cell Mol Med. 2014;18:1667–1679. doi: 10.1111/jcmm.12317. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Zhen Y, Liu Z, Yang H, Yu X, Wu Q, Hua S, et al. Tumor suppressor PDCD4 modulates miR-184-mediated direct suppression of C-MYC and BCL2 blocking cell growth and survival in nasopharyngeal carcinoma. Cell Death Dis. 2013;4:e872. doi: 10.1038/cddis.2013.376. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Wang H, Cao F, Li X, Miao H, E J, Xing J, et al. miR-320b suppresses cell proliferation by targeting c-Myc in human colorectal cancer cells. BMC Cancer. 2015;15:748. doi: 10.1186/s12885-015-1728-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Wang H, Zhang G, Wu Z, Lu B, Yuan D, Li X, et al. MicoRNA-451 is a novel tumor suppressor via targeting c-myc in head and neck squamous cell carcinomas. J Cancer Res Ther. 2015;11:C216–221. doi: 10.4103/0973-1482.168189. [DOI] [PubMed] [Google Scholar]
  • 103.Chen D, Huang J, Zhang K, Pan B, Chen J, De W, et al. MicroRNA-451 induces epithelial-mesenchymal transition in docetaxel-resistant lung adenocarcinoma cells by targeting proto-oncogene c-Myc. Eur J Cancer. 2014;50:3050–3067. doi: 10.1016/j.ejca.2014.09.008. [DOI] [PubMed] [Google Scholar]
  • 104.Wang R, Chen DQ, Huang JY, Zhang K, Feng B, Pan BZ, et al. Acquisition of radioresistance in docetaxel-resistant human lung adenocarcinoma cells is linked with dysregulation of miR-451/c-Myc-survivin/rad-51 signaling. Oncotarget. 2014;5:6113–6129. doi: 10.18632/oncotarget.2176. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Lim L, Balakrishnan A, Huskey N, Jones KD, Jodari M, Ng R, et al. MicroRNA-494 within an oncogenic microRNA megacluster regulates G1/S transition in liver tumorigenesis through suppression of mutated in colorectal cancer. Hepatology. 2014;59:202–215. doi: 10.1002/hep.26662. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.He W, Li Y, Chen X, Lu L, Tang B, Wang Z, et al. miR-494 acts as an anti-oncogene in gastric carcinoma by targeting c-myc. J GastroenterolHepatol. 2014;29:1427–1434. doi: 10.1111/jgh.12558. [DOI] [PubMed] [Google Scholar]
  • 107.Yuan J, Wang K, Xi M. MiR-494 Inhibits Epithelial Ovarian Cancer Growth by Targeting c-Myc. Med SciMonit. 2016;22:617–624. doi: 10.12659/MSM.897288. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Lin F, Ding R, Zheng S, Xing D, Hong W, Zhou Z, et al. Decrease expression of microRNA-744 promotes cell proliferation by targeting c-Myc in human hepatocellular carcinoma. Cancer Cell Int. 2014;14:58. doi: 10.1186/1475-2867-14-58. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Prochownik EV, Vogt PK. Therapeutic Targeting of Myc. Genes Cancer. 2010;1:650–659. doi: 10.1177/1947601910377494. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Fletcher S, Prochownik EV. Small-molecule inhibitors of the Myconcoprotein. BiochimBiophysActa. 2015;1849:525–543. doi: 10.1016/j.bbagrm.2014.03.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Chrzan P, Skokowski J, Karmolinski A, Pawelczyk T. Amplification of c-myc gene and overexpression of c-Myc protein in breast cancer and adjacent non-neoplastic tissue. ClinBiochem. 2001;34:557–562. doi: 10.1016/s0009-9120(01)00260-0. [DOI] [PubMed] [Google Scholar]
  • 112.Horiuchi D, Kusdra L, Huskey NE, Chandriani S, Lenburg ME, Gonzalez-Angulo AM, et al. MYC pathway activation in triple-negative breast cancer is synthetic lethal with CDK inhibition. J Exp Med. 2012;209:679–696. doi: 10.1084/jem.20111512. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Hopkins AL, Groom CR. The druggable genome. Nat Rev Drug Discov. 2002;1:727–730. doi: 10.1038/nrd892. [DOI] [PubMed] [Google Scholar]
  • 114.Jiang H, Bower KE, Beuscher AEt, Zhou B, Bobkov AA, Olson AJ, et al. Stabilizers of the Max homodimer identified in virtual ligand screening inhibit Myc function. MolPharmacol. 2009;76:491–502. doi: 10.1124/mol.109.054858. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Holt JT, Redner RL, Nienhuis AW. An oligomer complementary to c-myc mRNA inhibits proliferation of HL-60 promyelocytic cells and induces differentiation. Mol Cell Biol. 1988;8:963–973. doi: 10.1128/mcb.8.2.963. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Prochownik EV, Kukowska J, Rodgers C. c-myc antisense transcripts accelerate differentiation and inhibit G1 progression in murine erythroleukemia cells. Mol Cell Biol. 1988;8:3683–3695. doi: 10.1128/mcb.8.9.3683. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Kimura S, Maekawa T, Hirakawa K, Murakami A, Abe T. Alterations of c-myc expression by antisense oligodeoxynucleotides enhance the induction of apoptosis in HL-60 cells. Cancer Res. 1995;55:1379–1384. [PubMed] [Google Scholar]
  • 118.Wang YH, Liu S, Zhang G, Zhou CQ, Zhu HX, Zhou XB, et al. Knockdown of c-Myc expression by RNAi inhibits MCF-7 breast tumor cells growth in vitro and in vivo. Breast Cancer Res. 2005;7:R220–228. doi: 10.1186/bcr975. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Sundberg TB, Ney GM, Subramanian C, Opipari AW, Glick GD. The immunomodulatory benzodiazepine Bz-423 inhibits B-cell proliferation by targeting c-myc protein for rapid and specific degradation. Cancer Res. 2006;66:1775–1782. doi: 10.1158/0008-5472.CAN-05-3476. [DOI] [PubMed] [Google Scholar]
  • 120.Pan C, Zhu D, Zhuo J, Li L, Wang D, Zhang CY. Role of Signal Regulatory Protein alpha in Arsenic Trioxide-induced Promyelocytic Leukemia Cell Apoptosis. Sci Rep. 2016;6:23710. doi: 10.1038/srep23710. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Reinhardt HC, Jiang H, Hemann MT, Yaffe MB. Exploiting synthetic lethal interactions for targeted cancer therapy. Cell Cycle. 2009;8:3112–3119. doi: 10.4161/cc.8.19.9626. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Cermelli S, Jang IS, Bernard B, Grandori C. Synthetic lethal screens as a means to understand and treat MYC-driven cancers. Cold Spring HarbPerspect Med. 2014;4(3):a014209. doi: 10.1101/cshperspect.a014209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Chan DA, Giaccia AJ. Harnessing synthetic lethal interactions in anticancer drug discovery. Nat Rev Drug Discov. 2011;10:351–364. doi: 10.1038/nrd3374. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.McLornan DP, List A, Mufti GJ. Applying synthetic lethality for the selective targeting of cancer. N Engl J Med. 2014;371:1725–1735. doi: 10.1056/NEJMra1407390. [DOI] [PubMed] [Google Scholar]
  • 125.Thompson JM, Nguyen QH, Singh M, Razorenova OV. Approaches to identifying synthetic lethal interactions in cancer. Yale J Biol Med. 2015;88:145–155. [PMC free article] [PubMed] [Google Scholar]
  • 126.Grandori C. A high-throughput siRNA screening platform to identify MYC-synthetic lethal genes as candidate therapeutic targets. Methods MolBiol. 2013;1012:187–200. doi: 10.1007/978-1-62703-429-6_12. [DOI] [PubMed] [Google Scholar]
  • 127.Toyoshima M, Howie HL, Imakura M, Walsh RM, Annis JE, Chang AN, et al. Functional genomics identifies therapeutic targets for MYC-driven cancer. ProcNatlAcadSci U S A. 2012;109:9545–9550. doi: 10.1073/pnas.1121119109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Stine ZE, Dang CV. Splicing and Dicing MYC-Mediated Synthetic Lethality. Cancer Cell. 2015;28:405–406. doi: 10.1016/j.ccell.2015.09.016. [DOI] [PubMed] [Google Scholar]
  • 129.Hsu TY, Simon LM, Neill NJ, Marcotte R, Sayad A, Bland CS, et al. The spliceosome is a therapeutic vulnerability in MYC-driven cancer. Nature. 2015;525:384–388. doi: 10.1038/nature14985. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Kessler JD, Kahle KT, Sun T, Meerbrey KL, Schlabach MR, Schmitt EM, et al. A SUMOylation-dependent transcriptional subprogram is required for Myc-driven tumorigenesis. Science. 2012;335:348–353. doi: 10.1126/science.1212728. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Goga A, Yang D, Tward AD, Morgan DO, Bishop JM. Inhibition of CDK1 as a potential therapy for tumors over-expressing MYC. Nat Med. 2007;13:820–827. doi: 10.1038/nm1606. [DOI] [PubMed] [Google Scholar]
  • 132.Kang J, Sergio CM, Sutherland RL, Musgrove EA. Targeting cyclin-dependent kinase 1 but not CDK4/6 or CDK2 is selectively lethal to MYC-dependent human breast cancer cells. BMC Cancer. 2014;14:32. doi: 10.1186/1471-2407-14-32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Huang CH, Lujambio A, Zuber J, Tschaharganeh DF, Doran MG, Evans MJ, et al. CDK9-mediated transcription elongation is required for MYC addiction in hepatocellular carcinoma. Genes Dev. 2014;28:1800–1814. doi: 10.1101/gad.244368.114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Yang D, Liu H, Goga A, Kim S, Yuneva M, Bishop JM. Therapeutic potential of a synthetic lethal interaction between the MYC proto-oncogene and inhibition of aurora-B kinase. ProcNatlAcadSci U S A. 2010;107:13836–13841. doi: 10.1073/pnas.1008366107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Wang Y, Engels IH, Knee DA, Nasoff M, Deveraux QL, Quon KC. Synthetic lethal targeting of MYC by activation of the DR5 death receptor pathway. Cancer Cell. 2004;5:501–512. doi: 10.1016/s1535-6108(04)00113-8. [DOI] [PubMed] [Google Scholar]
  • 136.Lin CJ, Nasr Z, Premsrirut PK, Porco JA, Jr., Hippo Y, Lowe SW, et al. Targeting synthetic lethal interactions between Myc and the eIF4F complex impedes tumorigenesis. Cell Rep. 2012;1:325–333. doi: 10.1016/j.celrep.2012.02.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Han H, Chen Y, Cheng L, Prochownik EV, Li Y. microRNA-206 impairs c-Myc-driven cancer in a synthetic lethal manner by directly inhibiting MAP3K13. Oncotarget. 2016 doi: 10.18632/oncotarget.7653. doi:10.18632/oncotarget.7653. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Bugel SM, Tanguay RL, Planchart A. Zebrafish: A marvel of high-throughput biology for 21st century toxicology. Curr Environ Health Rep. 2014;1:341–352. doi: 10.1007/s40572-014-0029-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Breyer MD, Look AT, Cifra A. From bench to patient: model systems in drug discovery. Dis Model Mech. 2015;8:1171–1174. doi: 10.1242/dmm.023036. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES