BBAGRM-00702; No. of pages: 10; 4C: 2, 5 Biochimica et Biophysica Acta xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Biochimica et Biophysica Acta

Review

2

MicroRNAs as regulators and mediators of c-MYC function☆

3Q1

Rene Jackstadt, Heiko Hermeking ⁎

4

Experimental and Molecular Pathology, Institute of Pathology, Ludwig-Maximilians-Universität München, D-80337 Munich, Germany

5

a r t i c l e

6 7 8 9 10

Article history: Received 5 December 2013 Received in revised form 27 March 2014 Accepted 4 April 2014 Available online xxxx

11 12 13 14 15

Keywords: MYC Oncogene miRNA Cancer

a b s t r a c t

R O

i n f o

O

1

E

D

P

In the past eight years microRNAs (miRNAs) have been widely implicated as components of tumor suppressive and oncogenic pathways. Also the proto-typic oncogene c-MYC has been connected to miRNAs. The c-MYC gene is activated in approximately half of all tumors, and its product, the c-MYC transcription factor, regulates numerous processes e.g. cell cycle progression, metabolism, epithelial–mesenchymal transition (EMT), metastasis, stemness, and angiogenesis, thereby facilitating tumor initiation and progression. c-MYC target-genes, which mediate these functions of c-MYC, represent a complex network of protein- and non-coding RNAs, including numerous miRNAs. For example, c-MYC directly regulates expression of the miR-17–92 cluster, miR-34a, miR15a/ 16-1 and miR-9. Moreover, the expression and activity of c-MYC itself are under the control of miRNAs. Here, we survey how these networks mediate and regulate c-MYC functions. In the future, miRNAs connected to c-MYC may be used for diagnostic and therapeutic approaches. This article is part of a Special Issue entitled: Myc proteins in cell biology and pathology. © 2014 Published by Elsevier B.V.

27

T

31 29 28 30

1. Introduction

33

The protein product of the proto-oncogene c-MYC (hereafter MYC) is a transcription factor that regulates a broad range of cellular processes, which may contribute to the initiation and progression of tumors [1,2]. In tumors MYC expression is often found to be activated constitutively via several mechanisms: e.g. chromosomal translocations [3], amplification [4], loss of auto-regulation [5], increased translation of MYC mRNA [5,6], stabilizing mutations [7,8], or activating mutations in components of upstream regulatory pathways, such as Wnt signaling [9–11]. In most cases these alterations lead to constitutive expression of the MYC gene, which is normally only expressed during certain phases of the cell cycle. Ectopic MYC is sufficient to induce cell cycle progression and proliferation. MYC has been shown to induce or repress the expression of hundreds of genes, which are thought to mediate its biological functions [12–14]. Cells have evolved a failsafe mechanism against aberrant proliferation caused by deregulation of MYC (and other mitogenic oncogenes), which involved activation of the transcription factor p53, which then mediates apoptosis and/or senescence to eliminate cells with deregulated MYC expression [15]. The activation of p53 by MYC may occur via activation of the ATM/ATR-pathway due to DNA damage

42 43 44 45 46 47 48 49 50 51

E

R

R

40 41

N C O

38 39

U

36 37

C

32

34 35

F

journal homepage: www.elsevier.com/locate/bbagrm

☆ This article is part of a Special Issue entitled: Myc proteins in cell biology and pathology. ⁎ Corresponding author at: Experimental and Molecular Pathology, Institute of Pathology, Ludwig-Maximilians-Universität München, Thalkirchner Strasse 36, D-80337 Munich, Germany. Tel.: +49 89 2180 73685; fax: +49 89 2180 73697. E-mail address: [email protected] (H. Hermeking).

caused by MYC-induced replication stress or by MYC-mediated upregulation of the ARF/MDM2 pathway [15,16]. MYC belongs to the basic-helix–loop–helix leucine-zipper (bHLH-LZ) transcription factor family. The C-terminus of MYC harbors a HLH-LZ motif, which mediates the heterodimerization with the bHLH-LZ transcription factor MAX. The basic regions of the MYC/MAX heterodimer mediate the DNA binding to the consensus E-Box sequence CA(C/T)GTG [17]. The N-terminus of MYC encompasses two evolutionary conserved regions, Myc-Box I and II, which act as transactivation domains. Upon DNA binding the MYC/MAX heterodimer recruits co-factors, which mediate multiple effects of MYC on gene expression in a context-dependent manner [18]. Several MYC-mediated chromatin modifications, which result in transcriptional silencing, promoter clearance [19] and transcriptional elongation [20] have been described as consequences of MYC occupancy at promoters. Besides with MYC, MAX may heterodimerize with MAD, MXI1 and MNT, which thereby antagonize MYC function [21,22]. Whereas direct E-Box binding of MYC/MAX heterodimers generally mediates the induction of MYC target gene transcription, repression by MYC is often mediated by binding of a MYC/MAX/MIZ1 complex to initiator (Inr) elements [23–26]. The Inr represents a 17 bp pyrimidine-rich motif, which mediates transcriptional initiation from TATA-less promoters [27]. miRNAs represent small non-coding RNAs, which bind to 3′-UTRs of target mRNAs and mediate translational repression or mRNA degradation [28]. Induction of miRNA-encoding genes represents an alternative mechanism by which MYC indirectly represses gene expression. Evidence described in this review suggests that this mode of repression by MYC is rather common. Genes which encode miRNAs are mostly transcribed by RNA polymerase II, resulting in a primary transcript miRNA (pri-miRNA) (Fig. 1A).

http://dx.doi.org/10.1016/j.bbagrm.2014.04.003 1874-9399/© 2014 Published by Elsevier B.V.

Please cite this article as: R. Jackstadt, H. Hermeking, MicroRNAs as regulators and mediators of c-MYC function, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbagrm.2014.04.003

16 17 18 19 20 21 22 23 24 25 26

52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80

104 105

F

O

R O

102 103

D

100 101

E

98 99

T

96 97

C

94 95

E

92 93

R

90 91

R

88 89

O

87

C

85 86

Besides regulating the expression of miRNA-encoding genes by directly binding to their promoters MYC regulates miRNA expression by alternative mechanisms. An example for this is represented by the transcriptional activation of the genes encoding the RNA binding proteins Lin28 and Lin28b by MYC [34] (Fig. 1B). Lin28/Lin28b interacts directly with let-7 pre-miRNA stem–loops and thereby blocks the processing of let-7 by DROSHA and DICER1 [35,36]. Furthermore, Lin28/Lin28b recruits the 3′ terminal uridylyl transferase TUT4 to pre-let-7, resulting in uridylation and subsequent decay of the pre-miRNA [34,37–39]. MYC may also enhance pri-miRNA processing by directly inducing the expression of DROSHA [40] (Fig. 1C). As evidenced by miRNA processing assays MYC promotes miRNA processing by up-regulation of DROSHA expression [40]. Alterations in the miRNA processing machinery may affect MYCinduced tumorigenesis. For example, DICER1 was characterized as a haploinsufficient tumor suppressor gene in a mouse model of softtissue sarcoma [41], suggesting that global loss of miRNA expression is pro-tumorigenic. In support of this conclusion, decreased expression of DICER1 correlates with poor prognosis in human lung cancer [42]. Surprisingly, a deletion of one DICER1 allele in B cells failed to promote B cell malignancy or accelerate MYC-induced B cell lymphomagenesis in mice [43]. Moreover, deletion of both DICER1 alleles in B cells of Eμ-myc mice significantly inhibited lymphomagenesis [43]. Therefore, the function of DICER1, and therefore general miRNA expression, seems to be either tumor suppressive or oncogenic in a context-dependent manner.

N

83 84

Subsequently, this transcript becomes processed by the double-stranded RNA (dsRNA)-specific RNase III endonuclease DROSHA into precursor miRNAs (pre-miRNAs). Then Exportin 5 (XPO5) binds these ~ 70 nucleotide-long molecules and shuttles them to the cytoplasm, where the mature miRNA is generated by the dsRNA-specific RNase III enzyme DICER1, followed by incorporation into the RNA-induced silencing complex (RISC). The guide strand of the duplex associates with an Argonaute (AGO) protein, while the opposite strand (miRNA*) is often degraded and therefore not detectable [29–31]. The association with target mRNAs occurs via ~7 nucleotide-long stretches, the so-called seed-sequence, present in the 5′-region of the miRNA. The complementary sequence is mostly located in the 3′-untranslated region (3′-UTR) of the respective target gene. Additional base pairings can occur through nucleotides in the central region and 3′-ends of the miRNA. As a consequence of the recruitment of the miRNA/Ago complex translation of the bound mRNA is inhibited and often secondary degradation of the mRNA is observed. This involves mechanisms reviewed in [32]. Since a relatively short seed region is important for target recognition, a single miRNA might regulate several target mRNAs [32]. The influence of miRNAs on gene expression is widespread, with N 60% of human protein-coding genes predicted to be subject to regulation by miRNAs [33]. The biological functions of miRNAs are highly dependent on cellular context, which may be due to the differential expression of their target mRNAs. Some miRNAs function either as oncomiRs or as tumor suppressive miRNAs, in celltype dependent manner.

U

81 82

R. Jackstadt, H. Hermeking / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

P

2

Fig. 1. Transcriptional and post-transcriptional influences of MYC on miRNA processing. (A) MYC binding to the promoter of miRNA genes, by recognition of E-boxes, regulates the transcription of pri-miRNAs transcripts. (B) MYC effect on miRNA processing (let-7), by transcriptional regulation of Lin28. (C) Inhibition of miRNA processing by transcriptional repression of DROSHA.

Please cite this article as: R. Jackstadt, H. Hermeking, MicroRNAs as regulators and mediators of c-MYC function, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbagrm.2014.04.003

106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130

R. Jackstadt, H. Hermeking / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

149 150 151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 186 187 188 189 190 191 192 193 194 195

F

O

147 148

R O

145 146

C

143 144

E

141 142

R

139 140

R

137 138

N C O

136

U

134 Q3 135

P

The first evidence, that cancer-related mechanisms, as differentiation, apoptosis and proliferation are regulated by miRNAs, was obtained in C. elegans and Drosophila [44,45]. Since these processes are also relevant during carcinogenesis, an involvement of miRNA deregulation in tumor initiation and progression seemed plausible. The first demonstration that miRNAs are indeed regulated directly by the oncogenic transcription factor MYC initiated an active field of research [46,47]. The first identified, MYC-regulated miRNAcluster, was miR17–92 (oncomiR-1). The miR-17–92 miRNA cluster, encodes miR-17, miR-18a, miR-19a, miR-20a, miR-19b-1, and miR92a-1 [48,49]. Two paralogs of the miR-17–92 cluster have been identified: the miR-106b-25 cluster is located in the 13th intron of the MCM7 gene, whereas the miR-106a-363 is encoded by its own mRNA [50]. Both miR-17–92 and miR-106b-25 derived miRNAs are expressed at high levels in several adult tissues and during embryogenesis, whereas miR-106a-363 members are commonly expressed at low levels [51,52]. The 15 miRNAs of the miR-17–92 family and its two paralogs can be subdivided into 4 groups with similar seed sequences, which show high conservation in vertebrates [50]. These 15 miRNAs are frequently activated in several solid tumors and B-cell lymphomas [49,53,54]. Furthermore, ectopic expression of the miR-17–92 cluster, under the control of an Eμ-enhancer and Ig heavy-chain promoter, leads to development of B-cell lymphomas in mice [49,55]. In addition, conditional knock-out of miR-17–92 in MYC-driven lymphomas leads to increased cell death and reduced tumorigenicity [56]. Furthermore, knock-out of single members of the miR-17–92 cluster revealed that the six miRNAs encoded by the cluster are not functionally equivalent during cancer development. Two members of the miR-19 seed subgroup were found to be required and sufficient to recapitulate the oncogenic activity of the full cluster [57]. Recently, the lab of Carlo Croce showed that ectopic expression of the miR-17–92 cluster is sufficient to cause B-cell lymphoma or leukemia in mice [58]. Additionally, ectopic miR-17–92 expression in a murine retinoblastoma mouse model in which tumors are initiated by deletion of the RB family members RB and p107, as well as in colorectal colonocytes and granule neuron progenitors, enhances tumorigenesis [59–61]. Taken together, these results show that the miR-17–92 cluster represents a bona fide oncogene and a mediator of MYC-induced tumorigenesis. MYC activates miR-17–92 expression by binding to an E-box in its first intron [48]. Furthermore, several MYC binding sites have been characterized in the vicinity of the transcription start sites (TSS) of miR-106a-363 and miR-106b-25. Recently, the direct regulation of the miR-106a-363 cluster by MYC during trophoblast differentiation was described [62]. Besides MYC, the mitogenic transcription factor E2F was shown to regulate the expression of miR-17–92 [63,64]. Since the miRNAs miR-17 and miR-20 directly target E2F1 via identical seedmatching sequences, E2F1 and these miRNAs constitute a negative feedback-loop [65], which may represent an auto-regulatory failsafe mechanism to restrict E2F1 expression and thereby control the balance between proliferation and apoptosis [48,63,64]. Additionally, MCM7, the host gene of miR-106b-25, is regulated by the E2F family member E2F1 and its expression levels correlate with miR-106b-25 expression in primary gastric tumors and normal mucosa [66,67]. Interestingly, MYC also regulates the expression of MCM7 and therefore also activates the miR-106b-25 cluster [65]. Members of the miR-17–92 cluster target many different mRNAs, which in turn mediate several MYC-associated cancer promoting functions e.g. cell proliferation, cell survival, angiogenesis, and metabolic reprogramming (see also Table 1). Another MYC-induced miRNA, namely miR-22, was recently shown to act as a very potent proto-oncogenic miRNA via genome-wide deregulation of the epigenetic state through inhibition of methylcytosine dioxygenase TET (ten-11 translocation) proteins [68]. Thereby, miR-22 influences chromatin remodeling and silences miR-200 genes which

196 197 Q4 198 199 200 201 202 203 204 205 206 207 208 209 210 211 212 Q5 213 214 215 216 217 218

3. c-MYC-repressed miRNAs

219

The first experimental evidence that MYC represses miRNAs was provided by Chang and co-workers [39], who showed that multiple miRNAs are negatively regulated by MYC. These miRNAs represent potent cell-cycle inhibitors or apoptosis promoters [79–81] (Table 1). Furthermore, MYC activates several oncogenic factors by repression of miRNAs, e.g. MYC-mediated repression of let-7 results in up-regulation of RAS or HMGA2 expression [39,82]. Another tumor promoting activity of MYC is to increase mitochondrial glutaminase (GLS) activity by repressing miR-23a and miR-23b, which both target the GLS transcript [83]. GLS is a key enzyme that converts glutamine to glutamate, which serves as a substrate in the TCA cycle for the production of ATP. In addition, MYC regulates the glutamine metabolism through several mechanisms, including the direct transcriptional activation of glutamine transporters [84]. Both glutamine and GLS are required for MYCmediated cancer cell survival and proliferation. Besides enhancing the supply of cells with glutamine MYC activation also enhances the levels of glucose, the second major carbon substrate needed by proliferative cancer cells, by activating aerobic glycolysis via inducing several enzymes involved, such as LDH-A [85]. A miRNA-mediated enhancement of glycolysis by MYC was recently reported: by inducing miR-19 family members, which down-regulate PTEN, MYC enhances PI3K (phosphoinositide 3kinase) activity [57], which strongly promotes glucose metabolism by increasing glucose transporter expression and enhancing glycolytic enzyme activity [86]. Therefore, MYC coordinates both glucose and glutamine metabolism through miRNA regulation. Furthermore, the miR15a/16-1 cluster is directly repressed by MYC [39]. The miR-15a/16-1 cluster is located in an intron of the DLEU2 gene, which is frequently deleted or down-regulated in human tumors [87,88]. In fact, miR-15a/16-1 represent the first miRNAs, which were shown to be the target of cancer-specific deletions [89]. Later, a knock-out of DLEU2 or the miR-15a/16-1 bearing intron in mice revealed that loss of miR-15a/16-1 expression is sufficient to cause chronic lymphatic leukemia (CLL) [90]. Therefore, DLEU2 is presumably the tumor suppressor gene located in the 13q14 region. Importantly, this mouse knock-out study provided the first genetic proof of a miRNA to represent a bona fide tumor suppressor gene. miR-15/16 targets multiple mRNAs, which encode oncogenic factors, such as BCL2 and cyclin D1, and thereby exerts its tumor suppressive function by inducing apoptosis or inhibiting cell cycle progression [91–94]. Interestingly, the DLEU2 gene does neither seem to encode a functional protein, nor a functional

220 221

D

132 133

promotes EMT and enhances stemness, breast cancer development and metastasis. In addition, miR-22 was characterized as a key regulator of the self-renewal machinery of the hematopoietic system [69]. miR-22 was found to reduce the global level of 5-hydroxymethylcytosine (5hmC) modifications in mouse hematopoietic stem cells (HSCs) and thereby triggers an increase in HSC self-renewal capability. The TET2 enzyme represents a critical target of miR-22 in this context. Furthermore, miR-22 was shown to repress the tumor-suppressor PTEN (phosphatase tensin and homolog) in prostate cancer and mouse models of cardiac hypertrophy [70]. However, further studies suggested, that miR-22 may also act as a tumor-suppressor, e.g. by targeting ERBB3 in lung cancer cells [70–72]. This is consistent with the notion that miRNAs often display a context-dependent function. miR-378 represents an additional, oncogenic MYC-induced miRNA [73]. Interestingly, miR-378 cooperates with activated RAS or HER2 to promote cellular transformation. miR-378 acts as oncogenic by targeting TOB2, and acts as tumor suppressive by repressing cyclin D1. As expected, the c-MYC homolog N-MYC also induces the miR-17–92 cluster and several other c-MYC-regulated miRNAs [74–77]. One of them is miR-9 (further discussed below), a repressor of the epithelial marker E-cadherin, which is often downregulated during cancer progression [78]. miR-130a and miR-214 represent further miRNAs, which are induced by both, c-MYC and N-MYC [74].

E

2. MYC-induced miRNAs

T

131

3

Please cite this article as: R. Jackstadt, H. Hermeking, MicroRNAs as regulators and mediators of c-MYC function, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbagrm.2014.04.003

222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 257 258 259

4 t1:1 t1:2 t1:3 t1:4 t1:5 Q2

R. Jackstadt, H. Hermeking / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

Table 1 Compilation of miRNAs induced/repressed by MYC or targeting c-MYC. The miRNAs, the respective host genes and the methods by which the regulation was characterized are summarized. The respective references are given in the last column. The putative transcription start site of the miRNA transcript or of the host gene was analyzed, in order to determine the direct regulation by MYC via DNA binding (ChIP). The down-stream miRNA targets and processes of these regulations are indicated. Lin28 is the direct MYC target gene. Its protein product influences the processing of the let-7 miRNAs.

t1:6

miRNA

t1:7 t1:8 t1:9 t1:10 t1:11 t1:12 t1:13 t1:14 t1:15

Target mRNAs

Process

Reference

MYC-induced miRNAs miR-9 C1ORF61 miR-17 C13ORF25 miR-18a C13ORF25 miR-19a C13ORF25 miR-19b-1 C13ORF25 miR-20a C13ORF25 miR-22 MGC14276 miR-25 MCM7

ChIP-on-chip ChIP ChIP ChIP ChIP ChIP – ChIP

E-cadherin, LIFR, CREB, NF1 STAT3, MAPK14, p21, p57, TGFBR2, PTEN, BIM, E2F1–3 SMAD2/4, CTGF, TSP-1, ISL1, TBX1 CTGF, TSP1, PTEN, BIM CTGF, TSP1, PTEN, BIM STAT3, MAPK14, p21, p57, TGFBR2, PTEN, BIM PTEN, TET, TIAM, MMP2, MMP9 ISL1, TBX1, PTEN, BIM, RECK, SMAD7, WWP2, FBXW7, EZH2, EP300

[78,185] [48,49] [48,49] [48,49] [48,49] [48,49] [39] [62]

t1:16 t1:17 t1:18

miR-92a-1 miR-93 miR-106b

C13ORF25 MCM7 MCM7

ChIP ChIP ChIP

ISL1, TBX1, PTEN, BIM, ITGA5, RECK STAT3, MAPK14, p21, p57, TGFBR2, PTEN, BIM, DAB2, LATS2, EP300 STAT3, MAPK14, p21, p57, TGFBR2, PTEN, BIM, EP300

t1:19 t1:20 t1:21 t1:22 t1:23 t1:24

ChIP ChIP –

ATG2B, DICER1, c-MET, HOXA5, SMAD4 PTEN, BCL2L2, TFAP2C, EZH2, LTF TOB2, MAPK1, SOX2, SUFU, FUS1

EMT, cell cycle proliferation, invasion, metastasis EMT, metastasis, stemness, angiogenesis, cell cycle

[34,39] [39,96]

t1:25 t1:26 t1:27 t1:28 t1:29

miR-130a – miR-214 – miR-378 PPARGC1B MYC-repressed miRNAs Lin28/let-7 – miR-15a/ DLEU2 16-1 miR-23a – miR-23b C9orf3 miR-26a CTDSP1 miR-29a/b – miR-34a –

EMT, metastasis Cell cycle, apoptosis Angiogenesis, apoptosis Survival, angiogenesis, metabolism, apoptosis Survival, angiogenesis, metabolism, apoptosis Cell cycle, apoptosis EMT, metastasis Apoptosis, heart development, reprogramming, EMT, metastasis Apoptosis, heart development, angiogenesis Cell cycle, apoptosis, angiogenesis, metastasis Apoptosis, heart development, reprogramming, EMT, metastasis Lung development, drug resistance, cell cycle Proliferation, invasion, migration, apoptosis Survival, proliferation angiogenesis

t1:30 t1:31

miR-148a miR-185

– C22orf25

t1:32 t1:33 t1:34

miR-363



t1:35 t1:36 t1:37 t1:38 t1:39 t1:40 t1:41 t1:42 t1:43 t1:44

miRNAs targeting c-MYC let-7 miR-24 miR-34b/c miR-148a miR-184 miR-185 miR-196b miR-449c miR-494

O

R O

P

ChIP, luc-reporter HMAG, RAS, TWIST, CDC25A, CCND2 ChIP, luc-reporter AP4, BCL2, BMI1, VEGF, CARD10, CDC27

[74] [74] [73]

Metabolism, EMT Metabolism, cell cycle, invasion, migration Cell cycle, invasion, metastasis Cell cycle Cell cycle, apoptosis, senescence, stemness, EMT, metastasis EMT, metastasis, apoptosis, proliferation Cell cycle, metastasis

[83] [83] [39,99] [39,101] [39]

USP28, PDPN

Invasion, metastasis

[123]

E

T

Host gene

Method of validation

– C9ORF3

C

ChIP ChIP; lucreporter ChIP

[48,49] [62] [62]

GLS, E-cadherin PTEN, GLS, SRC, PYK2 PTEN, EZH2, Lin28B, ZCCHC11, FGF9, MCL1, HGM1, IL6 CDK6, IGF1R CDK6, MET, CD44, LDHA, SIRT1, SNAIL, ZNF281, c-KIT, Notch1, BCL2, BIRC5, IL6R MYC, CCKBR, BCL2, WNT1, DNMT1, ROCK1, CDC25B DNMT1, ATR, MYC, RHOA, CDC42

D

– ChIP ChIP ChIP ChIP

Western blot, qPCR Western blot, qPCR Western blot, qPCR, luc-reporter GFP-reporter, Western Blot Western blot, qPCR, luc-reporter Western blot, qPCR, luc-reporter qPCR Western blot, qPCR, luc-reporter Western blot, qPCR, luc-reporter

E

− − C22orf25 HOXA9 CDC20B –

R

miRNA

F

Method of validation

R

Host gene

[123] [124]

SMS

Reference

+ − + + + + + + +

[125] [133,134] [132] [123] [127] [124] [126] [128] [100]

Abbreviations: SMS = seed-matching sequence, qPCR = quantitative polymerase chain reaction, luc = luciferase assay including MYC-binding site or SMS mutagenesis, GFP = green fluorescent protein, ChIP = chromatin immunoprecipitation.

260

RNA besides miR-15a/16-1. We could recently show that the MYCinduced bHLH-LZ transcription-factor AP4 directly represses DLEU2 expression [95]. Since AP4 also represents a target of miR-15a/16-1, AP4 and miR-15a/16-1 constitute a double-negative feedback loop, which controls EMT and metastasis (discussed below). Zhang et al. demonstrated the co-localization of MYC and HDAC3 at the two alternative promoters of the miR-15a/16-1-encoding DLEU2 gene [96]. In contrast to MYC, p53 induces expression of miR-15a/16-1 by several mechanisms. Initially, p53 was shown to enhance the post-transcriptional maturation of miR-15a/16-1 by directly interacting with DROSHA via the DEAD-box RNA helicase DDX5 and thereby influencing the processing from pri-miRNAs to premiRNAs [97]. Later, the DLEU2 gene was shown to be a transcriptional target of p53 in B-cells [97,98]. The same group could show that a repressive complex containing MYC, HDAC3, and EZH2 down-regulates the expression of the miR-29 family in mantle cell lymphoma and other MYC-associated lymphomas. Furthermore, MYC up-regulates EZH2 expression via repression of the EZH2-targeting miR-26a [99], and EZH2 induces MYC via inhibition of the MYC targeting miR-494 [100]. These regulations result in a

265 266 267 268 269 270 271 272 273 274 275 276 277 278 279

C

N

263 264

U

261 262

O

t1:45 t1:46

double-positive feedback loop (Fig. 2B). Additional studies in cholangiocarcinoma cells showed repression of miR-29b by MYC, NF-κB and Hedgehog [101]. Since replacement of miRNAs repressed by MYC in tumors seemed to be an attractive therapeutic option (discussed below), Kota et al. employed a MYC-driven mouse hepatocellular carcinoma model and an adenovirus associate virus (AAV) vector delivery system, which allowed the reintroduction of miR-26a and reversal of disease progression [102]. Notably, delivery of miR-26a into the liver led to regression of established liver tumors by induction of apoptosis in cancer cells but not in non-malignant hepatocytes. Among the p53-induced miRNAs, miR-34a displays the most pronounced induction by p53 [39,80,87,98,103]. Interestingly, miR-34a is directly repressed by MYC via binding in the vicinity of the miR-34a promoter, as shown by ChIP analyses in Epstein–Barr virus-immortalized human B-cells [39]. Therefore, the miR-34a regulation may serve as another platform for the antagonism between MYC and p53. In mammalians, the miR-34 family comprises at least six processed miRNAs that are encoded by three different genes: miR-34a is encoded by its own transcript, whereas miR-34b and miR-34c are encoded by a common

Please cite this article as: R. Jackstadt, H. Hermeking, MicroRNAs as regulators and mediators of c-MYC function, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbagrm.2014.04.003

280 281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299

R. Jackstadt, H. Hermeking / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

A Proliferation

B E2F

MYC

C Auto-regulation

Chromatin regulation MYC

5

miR-26a

MYC

miR-148

EZH2

USP28

miR-363

miR-29 miR-106b-25

miR-494

E

D p53 pathway MYC

SIRT1 miR-34b/c

MYC

MK5

FOXO3a

PTEN

p53

HDM2

Cell cycle

H Angiogenesis

MYC

MIZ1

miR-17 miR-106b

P

AP4

VEGF

E

D

p21

Epithelial-mesenchymal transition

C SNAIL

AP4

Vimentin E-cadherin

mesenchymal

ZNF281

miR-34a

R

E-cadherin miR-15a/16-1

miR-106a/b miR-20a; miR-93 miR-17

E

miR-9

T

MYC

miR-22 miR-19a; miR-106b miR-25; miR-93 miR-214

miR-15a miR-16-1 miR-34a

MYC

I

miR-23b miR-26a

F

MYC miR-34

miR-145

G

Mitogenic signaling

R O

ARF

F

Mitogenic signaling

O

miR-17-92

Vimentin E-cadherin

epithelial

R

p53

301 302 303 304 305 306 307 308 309 310 311 312 313 314 315

primary transcript, the miR-449 cluster, which encodes the highly conserved miR-449a, miR-449b and miR-449c, belongs to the miR-34 family, since their seed-sequences are highly related [104]. Therefore, the targets of miR-34 and miR-449 miRNA family members are presumably overlapping. However, the regulation of miR-449 differs from miR-34 since it is not induced by p53, but by E2F1 [105]. Since E2F1 is generally induced after MYC activation, miR-449 expression is presumably elevated by MYC. In mice, miR-34a is ubiquitously expressed with the highest expression in the brain, whereas miR-34b/c is mainly expressed in lung tissues [106]. By repression of miR-34a MYC may antagonize several functions ascribed to miR-34a, such as inhibition of cell cycle progression [103] and stemness [107], as well as induction of senescence [108] and mesenchymal–epithelial transition [109] (MET; discussed below in detail; Table 1). More recently, we could show that genetic ablation of miR-34a allows invasion of colitis-associated colon cancer by activation of an IL6R/STAT3/miR-34a feedback loop [110]. Furthermore,

U

300

N C O

Fig. 2. MYC involving miRNA networks. MYC participates in several miRNA networks, which regulate (A) proliferation, (B) chromatin modifications, (C) auto-regulation, (D) p53 pathway, (E and F) mitogenic signaling, (G) cell cycle and (H) angiogenesis. (I) Role of MYC and connected miRNAs in EMT. MYC positively regulates the indicated EMT inducing factors, thereby promoting a mesenchymal state allowing invasion and metastasis. The mesenchymal state is characterized e.g. by high expression of Vimentin and low expression of E-cadherin. In contrary p53 promotes the epithelial state characterized by high E-cadherin and low Vimentin expression. Factors acting as oncogenic or representing oncogenes are indicated in orange, whereas tumor suppressive factors or tumor suppressor genes are indicated in green.

loss of miR-34a/b/c cooperates with p53 loss in prostate cancer progression in mouse tumor models [111]. In addition, miR-34a/b/c deletion cooperates with deletion of one p53 allele in a lung cancer model [112]. By inhibiting these tumor-suppressive activities of miR-34 MYC may contribute to tumor progression. Besides being repressed by MYC, the miR-34a/b/c genes are down-regulated during tumor development and metastasis by promoter methylation [113–115]. Similar to the therapeutic application described above for miR-26, Sotillo and colleagues determined whether restoration of miR-34a levels in B-lymphoid cells with enhanced MYC expression would induce apoptosis [116]. Unexpectedly, delivery of miR-34a, which does not target p53 directly, severely decreased the steady-state p53 levels. This effect was preceded and mediated by down-regulation of MYC, which enhances p53 protein levels via the ARF–HDM2 pathway. As a result, in the presence of MYC, miR-34a inhibited p53-dependent bortezomib-induced apoptosis as efficiently as p53-specific knock-down. Conversely, inhibition of miR-34a

Please cite this article as: R. Jackstadt, H. Hermeking, MicroRNAs as regulators and mediators of c-MYC function, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbagrm.2014.04.003

316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331

343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361

4. MYC regulating miRNAs

363

Several regulatory mechanisms control MYC expression by influencing its transcription and translation, as well as mRNA and protein stability [1]. More recently, a number of miRNAs have been described that regulate MYC expression. Bioinformatics analyses predicted let-7a seed-matching sequences in the MYC 3′-UTR. Indeed, ectopic expression of let-7a down-regulated MYC on mRNA and protein level and reverted MYC-induced proliferation of Burkitt lymphoma cells [125]. Interestingly two different miRNAs, namely miR-196b and miR-184, both concomitantly regulate MYC and BCL2 to block cell proliferation and survival [126,127]. Interestingly, miR-184 is directly regulated by the tumor suppressor PDCD4, which inhibits cell proliferation, migration and invasion [127]. Ectopic miR-449c suppresses invasion and migration via direct repression of MYC [128]. Interestingly, miR-449 also targets the transcription factor E2F1 to achieve its tumor suppressive functions [104]. However, in non-small lung cancer (NSCLC) miR-449c is increased in tumor tissue compared to normal tissue, suggesting an oncogenic role of miR-449c in NSCLC. Therefore, miR-449 displays both tumor suppressive and oncogenic properties in a contextdependent manner. In addition to the previously described transcriptional repression of MYC by p53, the p53-mediated induction of miR-145, which targets MYC, provides an alternative mechanism for p53-mediated repression of MYC (Fig. 2D) [129]. Additionally, p53-induced miR-34a represses MYC [108,130,131]. Furthermore, miR-34b/c is involved in a negative feedback-loop with the MK5/PRAK kinase and FOXO3a to regulate MYC expression (Fig. 2E) [132]. MK5 phosphorylates FOXO3a, which leads to nuclear localization of FOXO3a and transcriptional activation of miR-34b/c in colorectal cancer (CRC) cells. Subsequently, miR-34b/c bind to the 3′-UTR of MYC and thereby block its translation. Interestingly, the expression of MK5 in CRC tumors correlates with increased differentiation and therefore enhanced survival of colorectal cancer patients.

373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394

E

R

R

O

371 372

C

369 370

N

368

U

366 367

C

362

364 365

5. Regulatory circuits controlled by MYC and miRNAs

404

miRNA-target circuits have been described to stabilize or fine tune biological processes by reinforcing transcriptional programs, constituting bistable switches or providing robustness [135]. Since MYC is an important factor during normal and tumor development, MYC/miRNA circuits and their deregulation are likely to contribute to the oncogenic functions of MYC [136]. An intensively studied feed-back loop involves the miR-17–92 family, E2Fs and MYC (Fig. 2A) [48,137]. The E2F transcription factors participate in the control of cell proliferation and apoptosis, and depending on their levels of expression they can act as oncogenes or tumor suppressors [138,139]. Positive feedback loops in the regulation of these factors may lead to bi-stability, which allows switch-like transitions in the concentrations of transcription factors in response to minor changes of extra-cellular signals [140]. Furthermore, post-translational inhibition of E2Fs by members of the miR-17–92 family and the induction of the miR-17–92 transcript by E2F1 and MYC represent negative feedback loops, which may serve to prevent overshooting inductions of the respective pathways [48,64]. Several additional complex feed-forward loops involving MYC are known: for example the MYC/PTEN/miR-106b, miR-93, miR-25, miR19a, miR-22, miR-26a, and miR-23b circuit (Fig. 2F). The mRNA encoded by the PTEN tumor suppressor gene is a target of several miRNAs, which are also regulated by MYC [67,141,142]. The regulation of the cell cycle inhibitor CDKN1A/p21 by MYC represents another coherent feedforward loop (Fig. 2G). Initially, MYC was shown to bind to MIZ1 and thereby repress p21 expression [24]. More recently, the bHLH-LZ transcription factor AP4, was shown to be directly induced by MYC and repress p21 [18]. The MYC-induced miR-17, miR-20a and miR-106b suppress p21 expression and therefore represent another mechanism of MYC-mediated suppression of p21 [47,67,143]. A third circuitry involves MYC/VEGF/miR-106b, miR-106a, miR-93, miR-34a, miR-20a, miR-17, miR-16 and miR-15a (Fig. 2H). All of these eight miRNAs are under the control of MYC and target VEGF [59,144–147]. Depending on the regulation, this loop can act as coherent or incoherent feedforward loop.

405

6. MYC and miRNA networks in EMT and metastasis

440

During cancer progression, metastasis represents the last and death causing step. The molecular mechanisms underlying the invasionmetastasis cascade are not fully understood. Although the precise role of MYC in this process is still elusive, multiple studies have shown that MYC controls and supports this complex multistep process at different stages [148]. Furthermore, elevated expression and amplification of MYC correlates with metastasis and poor survival in several tumor types [149,150]. In order to form metastases at distant sites tumor cells have to acquire migratory and invasive capacities, invade into the stroma surrounding the primary tumor, overcome the basement membrane, intravasate into blood vessels and disseminate via the circulation [151,152]. This is often achieved by undergoing an EMT, which is followed by an MET when cells have reached a metastatic niche in the target organ and resume proliferation and undergo partial differentiation to form micro- and macro-metastases [153,154]. The cellular EMT program presumably evolved to facilitate multiple developmental

441

F

341 342

O

339 340

R O

338

395 396

P

336 337

The Lieberman lab showed that miR-24, which is up-regulated during terminal differentiation of multiple lineages and inhibits the cell cycle can repress MYC expression via “seedless” 3′-UTR microRNA recognition elements [133]. Later the ribosomal protein L11 was shown to regulate the turnover of MYC mRNA [134]. A L11/AGO2/RISC/miR-24 complex binds to MYC mRNA 3′-UTR leading to mRNA degradation. Knockdown of L11 drastically increases the levels and stability of MYC mRNA. Knock-down of AGO2 abrogated the L11-mediated repression of MYC, whereas knockdown of L11 rescued miR-24-mediated mRNA decay.

D

334 335

using antagomirs sensitized lymphoma cells to apoptosis. Thus, in certain tumors with deregulated MYC expression, miR-34a confers drug resistance and could be considered as a therapeutic target [116]. However, reintroduction of miR-34a into neuroblastoma cells, which lacked the 1p36 region containing the miR-34a gene, or into cells, which express low levels of miR-34a, resulted in a pronounced growth inhibition, which was in part due to the direct regulation of N-MYC by miR-34a [117–119]. In addition, miR-34 prevented cancer initiation and progression in a K-RAS and p53-induced mouse model of lung adenocarcinoma [120]. Furthermore, tail vein injection of miR-34a oligonucleotide complexes with a lipid-based delivery agent led to reduced prostate cancer growth and metastasis in an orthotropic mouse model, by directly regulating the stemness marker CD44 [121]. Because of these promising results, treatment with miR-34 mimics entered a Phase I clinical trial in patients with advanced hepatocellular carcinoma (HCC) in April 2013 [122]. MYC plays an important role in normal development and tumorigenesis of the liver [46]. Besides the repression of miR-26 MYC may contribute to HCC through an additional miRNA-mediated feedback loop comprising of miR-148a, miR-363, and the ubiquitin-specific protease 28 (USP28) (Fig. 2C): MYC directly binds to conserved regions in the promoters of the host-genes of the two miRNAs and represses their expression. miR-148a directly targets and inhibits MYC, whereas miR-363 destabilizes MYC by directly targeting and inhibiting USP28. Inhibition of miR-148a or miR-363 induced hepatocellular tumorigenesis by promoting G1/S-phase transition, whereas activation of these miRNAs had the opposite effect [123]. In addition, another auto-regulatory mechanism involves miR-185, which directly represses MYC translation and is repressed by MYC [124]. Therefore, MYC activation is blunted by the MYC/miR-185 feedback loop.

T

332 333

R. Jackstadt, H. Hermeking / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

E

6

Please cite this article as: R. Jackstadt, H. Hermeking, MicroRNAs as regulators and mediators of c-MYC function, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbagrm.2014.04.003

397 398 399 400 401 402 403

406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439

442 443 444 445 446 447 448 449 450 451 452 453 454 455 456

R. Jackstadt, H. Hermeking / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 Q6 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522

F

534

525 526 527 528 529 530 531 532 533

535

8. Conclusions and outlook

580

Here we have summarized the current state of the literature describing the influence of MYC/miRNA circuits on diverse oncogenic processes, such as cell cycle regulation, apoptosis, metabolism, angiogenesis, EMT and metastasis. miRNAs add a functionally relevant layer of

581

O

Targeting of MYC-induced miRNAs or restoring the expression of MYC-repressed miRNAs seems to be an obvious strategy to combat MYC-driven cancers [16,176,177]. Several MYC-driven mouse models are suitable for pre-clinical studies of targeting MYC-regulated miRNAs (reviewed in [178]). The biggest obstacle when using miRNA mimetics, is their delivery to the tissue of interest. Enhanced stability and increased uptake of miRNAs into tumor cells are critical points that need further improvement. A detailed discussion of the recent developments in this area can be found in [122,179,180]. To use MYC as a therapeutic target several strategies are conceivable. For example, silencing of MYC by RNA interference, modulation of MYC protein stability, disruption of MYC/MAX interaction by small molecules and drug-mediated inhibition of enzymes down-stream target of MYC (discussed in detail in [176,181]). For the last approach miRNAs seem to be good tools. One possible strategy is to re-introduce MYC downregulated miRNAs, which may be regarded as tumor suppressive. This was successful in the case of miR-26a, miR-34a and let-7, since it caused regression of already established tumors or reduction of tumor formation [102,120,182,183]. The alternative strategy is to down-regulate MYC-induced, oncogenic miRNAs by using e.g. antagomirs. This was tested in neuroblastoma, which frequently harbors a N-MYC amplification. Silencing of miR-17–92 members induced cell cycle arrest and apopotosis, and caused inhibition of tumor growth of xenografted NMYC-amplified neuroblastoma cells [75,143]. Currently, several labs are developing for small-molecule drugs that target specific miRNAs (SMIRs) and modulate their activities. SMIRs have several disadvantages as they display poor specificity, undesirable miRNA-independent effects and require a complicated design when compared with oligonucleotide-based therapeutics. The most beneficial concept seems to be the combinatorial usage of different miRNA agents with conventional chemotherapy [122]. This led to positive results in a clinical Phase II study and further recruitment of participants [180]. Furthermore, multiplex strategies may be advantageous. The combined use of multiple miRNAs, which inhibit the same pathway, may allow to achieve a superior effect in inducing e.g. the apoptosis. One example, would be the combined delivery of miR 15a and miR 34a mimetics, which may reduce the expression of anti-apoptotic proteins such as MCL1 (by miR 15a), SIRT1 (by miR 34a) and BCL 2 (by miR 15a and miR 34a). Furthermore the combined restoration of miR15a/16-1 and miR-34a could induce a MET and thereby prevent metastasis. However, it should be mentioned that the ectopic expression of the tumor-suppressive miR-200 increased the number of metastasis in a xenograft mouse model, since the formation of micro-metastasis often involves an MET process [184]. This could be a hurdle for the use of MET-inducing miRNAs for tumor therapeutic purposes.

R O

476 477

7. Therapeutic targeting of c-MYC-regulated miRNAs

P

474 475

523 524

D

472 473

influence EMT-related processes, since SIRT1 interacts with the EMTTF ZEB1 to enhance prostate cancer invasion and metastasis [173]. In addition, TGF-β driven EMT of mammary epithelial cells involves upregulation of SIRT1, which may silence the miR-200 promoter through histone deacetylation [174]. Notably, SIRT1 and miR-200 seem to form a negative feedback loop, as miR-200 targets the 3′-UTR of SIRT1 [174]. However, the role of SIRT1 in EMT is still controversial, since a gain of mesenchymal markers by inhibition of SIRT1 was observed, in mammary epithelial. Furthermore, SIRT1 knock-down led to significantly reduced survival in a xenograft mouse model of metastasis formation [175].

E

470 471

T

468 469

C

466 467

E

464 465

R

463

R

461 462

N C O

459 460

processes, such as gastrulation and neural crest formation, that involve the migration of epithelial cells, which transiently acquire a mesenchymal state [155]. It was shown that MYC induces EMT through the GSK3β/SNAIL axis or via repression of the WNT inhibitors DKK1 and SFRP1 in mammary epithelial cells [156,157]. Furthermore, induction of SNAIL by TGF-β was abrogated by knock-down of MYC, indicating that MYC is central for this process [158]. Recent research on MYCregulated miRNA networks has shed light onto the regulation of the metastatic process [159]. It was shown that miR-9 is directly induced by MYC and N-MYC to promote metastasis in breast cancer [78,160, 161]. miR-9 achieved this by repressing the cell adhesion protein Ecadherin and thereby promoting migration and invasiveness of breast cancer cells. The repression of E-cadherin also stimulates WNT signaling, which induces the expression of the vascular endothelial growth factor (VEGF) and tumor angiogenesis. Furthermore, the miR-9 levels correlate with N-MYC amplification and metastatic spread in patients with neuroblastoma [78]. In addition, miR-9 can also promote metastasis in an E-cadherin-independent manner via the repression of leukemia inhibitory factor receptor (LIFR). LIFR can inhibit metastasis through the Hippo–YAP pathway and in human breast cancer the loss of LIFR is associated with poor prognosis [162]. Interestingly, miR-9, which is secreted by tumors, promotes endothelial cell migration and angiogenesis by activating the JAK–STAT pathway [163]. Therefore, MYC-induced miR-9 may also act in a paracrine fashion. Surprisingly, recent evidence suggests a role of MYC in suppressing metastasis formation and invasion, which may also involve the induction of the miR-17–92 family members miR-17/20 [164,165]. It is therefore likely that the observed functions of MYC during the metastatic process are dependent on the stage and tissue under investigation. We could recently show that MYC associates with [166] and induces the Krüppel-type transcription factor ZNF281 [167]. Interestingly, ZNF281 itself represents a miR-34 target [167] and MYC-mediated repression of miR-34 may therefore mediate the induction of ZNF281 by MYC. Furthermore, ectopic expression of ZNF281 induced EMT, migration and invasion which was associated with increased stemness, and enhanced Wnt pathway activity. ZNF281-induced EMT was dependent on the induction of SNAIL and also SNAIL-induced EMT required ZNF281 expression (Fig. 2I). miR-34 may represent an important link between these two EMT-TFs, which allows a feed-forward regulation. During tumor progression this regulation may be shifted towards the mesenchymal state by down-regulation of miR-34, e.g. by p53 mutation, miR-34a and miR-34b/c promoter methylation or by activation of MYC [87,109]. Previously, we could show that the MYC-induced transcription factor AP4 induces EMT and is required for lung-metastasis formation of xenotransplanted CRC cell lines [168]. Furthermore, elevated AP4 expression is associated with distant metastasis and poor survival of colorectal cancer patients. Therefore, AP4 represents a new EMT-TF. In addition, AP4 directly down-regulates p16 and p21 to suppress senescence and mediate transformation [169]. Moreover, we recently found that AP4 forms a negative feed-back with the p53-induced and MYCrepressed miRNA miR-15a/16-1 [95]. This loop is similar to the SNAIL/ miR-34 feedback-regulation. Similar to ectopic miR-34, miR-15a/16-1 also induces MET, albeit in an AP4-repression dependent manner [18, 95,170]. Furthermore, miR-15a/16-1 inhibited lung-metastasis formation of xenotransplanted CRC cell lines and its expression in CRC patient samples is inversely correlated with AP4 and poor survival of patients, which is generally associated with metastasis. Since, both MYC and AP4 directly repress miR-15a/16-1, the down-regulation of these antimetastatic miRNAs presumably represents an important step towards metastatic disease. MYC and SIRT1 form a positive feedback loop, which connects MYC to the metabolic state of cells, since SIRT1 is a NAD-dependent enzyme with deacetylating activity [171]. Interestingly, the MYC-repressed miR34a negatively regulates SIRT1 [172]. Besides, its impact on the regulation of chromatin modifications and states, this circuitry may also

U

457 458

7

Please cite this article as: R. Jackstadt, H. Hermeking, MicroRNAs as regulators and mediators of c-MYC function, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbagrm.2014.04.003

536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579

582 583 584

O

R

R

E

C

F

T

[1] C.V. Dang, MYC on the path to cancer, Cell 149 (1) (2012) 22–35. [2] N. Meyer, L.Z. Penn, Reflecting on 25 years with MYC, Nat. Rev. Cancer 8 (12) (2008) 976–990. [3] D.A. Thorley-Lawson, M.J. Allday, The curious case of the tumour virus: 50 years of Burkitt's lymphoma, Nat. Rev. Microbiol. 6 (12) (2008) 913–924. [4] C.E. Nesbit, J.M. Tersak, E.V. Prochownik, MYC oncogenes and human neoplastic disease, Oncogene 18 (19) (1999) 3004–3016. [5] K.B. Marcu, S.A. Bossone, A.J. Patel, Myc function and regulation, Annu. Rev. Biochem. 61 (1992) 809–860. [6] S.A. Chappell, et al., A mutation in the c-myc-IRES leads to enhanced internal ribosome entry in multiple myeloma: a novel mechanism of oncogene de-regulation, Oncogene 19 (38) (2000) 4437–4440. [7] T. Albert, et al., Ongoing mutations in the N-terminal domain of c-Myc affect transactivation in Burkitt's lymphoma cell lines, Oncogene 9 (3) (1994) 759–763. [8] S.E. Salghetti, S.Y. Kim, W.P. Tansey, Destruction of Myc by ubiquitin-mediated proteolysis: cancer-associated and transforming mutations stabilize Myc, EMBO J. 18 (3) (1999) 717–726. [9] T.C. He, et al., Identification of c-MYC as a target of the APC pathway, Science 281 (5382) (1998) 1509–1512. [10] M. van de Wetering, et al., The beta-catenin/TCF-4 complex imposes a crypt progenitor phenotype on colorectal cancer cells, Cell 111 (2) (2002) 241–250. [11] O.J. Sansom, et al., Loss of Apc in vivo immediately perturbs Wnt signaling, differentiation, and migration, Genes Dev. 18 (12) (2004) 1385–1390. [12] C.V. Dang, et al., The c-Myc target gene network, Semin. Cancer Biol. 16 (4) (2006) 253–264. [13] A. Menssen, H. Hermeking, Characterization of the c-MYC-regulated transcriptome by SAGE: identification and analysis of c-MYC target genes, Proc. Natl. Acad. Sci. U. S. A. 99 (9) (2002) 6274–6279. [14] R. Jackstadt, A. Menssen, H. Hermeking, Genome-wide analysis of c-MYC-regulated mRNAs and miRNAs, and c-MYC DNA binding by next-generation sequencing, Methods Mol. Biol. 1012 (2013) 145–185. [15] F. Zindy, et al., Myc signaling via the ARF tumor suppressor regulates p53dependent apoptosis and immortalization, Genes Dev. 12 (15) (1998) 2424–2433. [16] O. Vafa, et al., c-Myc can induce DNA damage, increase reactive oxygen species, and mitigate p53 function: a mechanism for oncogene-induced genetic instability, Mol. Cell 9 (5) (2002) 1031–1044. [17] M. Eilers, R.N. Eisenman, Myc's broad reach, Genes Dev. 22 (20) (2008) 2755–2766. [18] P. Jung, H. Hermeking, The c-MYC–AP4–p21 cascade, Cell Cycle 8 (7) (2009) 982–989. [19] V.H. Cowling, M.D. Cole, Mechanism of transcriptional activation by the Myc oncoproteins, Semin. Cancer Biol. 16 (4) (2006) 242–252. [20] P.B. Rahl, et al., c-Myc regulates transcriptional pause release, Cell 141 (3) (2010) 432–445. [21] S. Adhikary, M. Eilers, Transcriptional regulation and transformation by Myc proteins, Nat. Rev. Mol. Cell Biol. 6 (8) (2005) 635–645. [22] D.E. Ayer, Q.A. Lawrence, R.N. Eisenman, Mad-Max transcriptional repression is mediated by ternary complex formation with mammalian homologs of yeast repressor Sin3, Cell 80 (5) (1995) 767–776. [23] K. Peukert, et al., An alternative pathway for gene regulation by Myc, EMBO J. 16 (18) (1997) 5672–5686. [24] S. Herold, et al., Negative regulation of the mammalian UV response by Myc through association with Miz-1, Mol. Cell 10 (3) (2002) 509–521. [25] D.Y. Mao, et al., Analysis of Myc bound loci identified by CpG island arrays shows that Max is essential for Myc-dependent repression, Curr. Biol. 13 (10) (2003) 882–886. [26] P. Staller, et al., Repression of p15INK4b expression by Myc through association with Miz-1, Nat. Cell Biol. 3 (4) (2001) 392–399. [27] S.T. Smale, D. Baltimore, The “initiator” as a transcription control element, Cell 57 (1) (1989) 103–113. [28] D.P. Bartel, MicroRNAs: target recognition and regulatory functions, Cell 136 (2) (2009) 215–233. [29] S.M. Hammond, et al., Argonaute2, a link between genetic and biochemical analyses of RNAi, Science 293 (5532) (2001) 1146–1150. [30] S.M. Hammond, et al., An RNA-directed nuclease mediates post-transcriptional gene silencing in Drosophila cells, Nature 404 (6775) (2000) 293–296. [31] S.M. Elbashir, et al., Functional anatomy of siRNAs for mediating efficient RNAi in Drosophila melanogaster embryo lysate, EMBO J. 20 (23) (2001) 6877–6888.

C

597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664

N

592 593

U

591

O

References

589 590

R O

596

587 588

[32] M.R. Fabian, N. Sonenberg, W. Filipowicz, Regulation of mRNA translation and stability by microRNAs, Annu. Rev. Biochem. 79 (2010) 351–379. [33] R.C. Friedman, et al., Most mammalian mRNAs are conserved targets of microRNAs, Genome Res. 19 (1) (2009) 92–105. [34] T.C. Chang, et al., Lin-28B transactivation is necessary for Myc-mediated let-7 repression and proliferation, Proc. Natl. Acad. Sci. U. S. A. 106 (9) (2009) 3384–3389. [35] M.A. Newman, J.M. Thomson, S.M. Hammond, Lin-28 interaction with the Let-7 precursor loop mediates regulated microRNA processing, RNA 14 (8) (2008) 1539–1549. [36] A. Rybak, et al., A feedback loop comprising lin-28 and let-7 controls pre-let-7 maturation during neural stem-cell commitment, Nat. Cell Biol. 10 (8) (2008) 987–993. [37] J.P. Hagan, E. Piskounova, R.I. Gregory, Lin28 recruits the TUTase Zcchc11 to inhibit let-7 maturation in mouse embryonic stem cells, Nat. Struct. Mol. Biol. 16 (10) (2009) 1021–1025. [38] I. Heo, et al., TUT4 in concert with Lin28 suppresses microRNA biogenesis through pre-microRNA uridylation, Cell 138 (4) (2009) 696–708. [39] T.C. Chang, et al., Widespread microRNA repression by Myc contributes to tumorigenesis, Nat. Genet. 40 (1) (2008) 43–50. [40] X. Wang, et al., c-Myc modulates microRNA processing via the transcriptional regulation of Drosha, Sci. Rep. 3 (2013) 1942. [41] M.S. Kumar, et al., Dicer1 functions as a haploinsufficient tumor suppressor, Genes Dev. 23 (23) (2009) 2700–2704. [42] Y. Karube, et al., Reduced expression of Dicer associated with poor prognosis in lung cancer patients, Cancer Sci. 96 (2) (2005) 111–115. [43] M.P. Arrate, et al., MicroRNA biogenesis is required for Myc-induced B-cell lymphoma development and survival, Cancer Res. 70 (14) (2010) 6083–6092. [44] J. Brennecke, et al., Bantam encodes a developmentally regulated microRNA that controls cell proliferation and regulates the proapoptotic gene hid in Drosophila, Cell 113 (1) (2003) 25–36. [45] J. Brennecke, S.M. Cohen, Towards a complete description of the microRNA complement of animal genomes, Genome Biol. 4 (9) (2003) 228. [46] J.D. Yang, L.R. Roberts, Hepatocellular carcinoma: a global view, Nat. Rev. Gastroenterol. Hepatol. 7 (8) (2010) 448–458. [47] J.H. Gibcus, et al., MiR-17/106b seed family regulates p21 in Hodgkin's lymphoma, J. Pathol. 225 (4) (2011) 609–617. [48] K.A. O'Donnell, et al., c-Myc-regulated microRNAs modulate E2F1 expression, Nature 435 (7043) (2005) 839–843. [49] L. He, et al., A microRNA polycistron as a potential human oncogene, Nature 435 (7043) (2005) 828–833. [50] A. Tanzer, P.F. Stadler, Molecular evolution of a microRNA cluster, J. Mol. Biol. 339 (2) (2004) 327–335. [51] A. Ventura, et al., Targeted deletion reveals essential and overlapping functions of the miR-17 through 92 family of miRNA clusters, Cell 132 (5) (2008) 875–886. [52] H.B. Houbaviy, M.F. Murray, P.A. Sharp, Embryonic stem cell-specific microRNAs, Dev. Cell 5 (2) (2003) 351–358. [53] S. Volinia, et al., A microRNA expression signature of human solid tumors defines cancer gene targets, Proc. Natl. Acad. Sci. U. S. A. 103 (7) (2006) 2257–2261. [54] A. Ota, et al., Identification and characterization of a novel gene, C13orf25, as a target for 13q31–q32 amplification in malignant lymphoma, Cancer Res. 64 (9) (2004) 3087–3095. [55] C. Xiao, et al., Lymphoproliferative disease and autoimmunity in mice with increased miR-17–92 expression in lymphocytes, Nat. Immunol. 9 (4) (2008) 405–414. [56] P. Mu, et al., Genetic dissection of the miR-17~92 cluster of microRNAs in Mycinduced B-cell lymphomas, Genes Dev. 23 (24) (2009) 2806–2811. [57] V. Olive, et al., miR-19 is a key oncogenic component of mir-17–92, Genes Dev. 23 (24) (2009) 2839–2849. [58] S.K. Sandhu, et al., B-cell malignancies in microRNA Emu-miR-17~92 transgenic mice, Proc. Natl. Acad. Sci. U. S. A. 110 (45) (2013) 18208–18213. [59] M. Dews, et al., Augmentation of tumor angiogenesis by a Myc-activated microRNA cluster, Nat. Genet. 38 (9) (2006) 1060–1065. [60] T. Uziel, et al., The miR-17~92 cluster collaborates with the Sonic Hedgehog pathway in medulloblastoma, Proc. Natl. Acad. Sci. U. S. A. 106 (8) (2009) 2812–2817. [61] K. Conkrite, et al., miR-17~92 cooperates with RB pathway mutations to promote retinoblastoma, Genes Dev. 25 (16) (2011) 1734–1745. [62] P. Kumar, et al., The c-Myc-regulated microRNA-17~92 (miR-17~92) and miR106a~363 clusters target hCYP19A1 and hGCM1 to inhibit human trophoblast differentiation, Mol. Cell. Biol. 33 (9) (2013) 1782–1796. [63] K. Woods, J.M. Thomson, S.M. Hammond, Direct regulation of an oncogenic microRNA cluster by E2F transcription factors, J. Biol. Chem. 282 (4) (2007) 2130–2134. [64] Y. Sylvestre, et al., An E2F/miR-20a autoregulatory feedback loop, J. Biol. Chem. 282 (4) (2007) 2135–2143. [65] F. Petrocca, A. Vecchione, C.M. Croce, Emerging role of miR-106b-25/miR-17–92 clusters in the control of transforming growth factor beta signaling, Cancer Res. 68 (20) (2008) 8191–8194. [66] S. Ambs, et al., Genomic profiling of microRNA and messenger RNA reveals deregulated microRNA expression in prostate cancer, Cancer Res. 68 (15) (2008) 6162–6170. [67] F. Petrocca, et al., E2F1-regulated microRNAs impair TGFbeta-dependent cell-cycle arrest and apoptosis in gastric cancer, Cancer Cell 13 (3) (2008) 272–286. [68] S.J. Song, et al., MicroRNA-antagonism regulates breast cancer stemness and metastasis via TET-family-dependent chromatin remodeling, Cell 154 (2) (2013) 311–324. [69] S.J. Song, et al., The oncogenic microRNA miR-22 targets the TET2 tumor suppressor to promote hematopoietic stem cell self-renewal and transformation, Cell Stem Cell 13 (1) (2013) 87–101.

P

594 595

complexity to the MYC network of cellular regulation. In the future, more publicly available datasets of miRNA expression in cancer patient cohorts which allow to determine correlations with mutations, epigenetic changes and clinical data will become available. This and optimized technologies, as next-generation sequencing or HITS-CLIP will bring new insights into the role of MYC/miRNA regulations in tumors and their clinical relevance. Taken together, these possibilities will hopefully lead to the rapid translation of knowledge derived from analysis of the MYC/miRNA network into diagnostic and therapeutic applications. Further studies of MYC/miRNA regulations are necessary to achieve a complete understanding of MYC-driven tumorigenesis.

D

585 586

R. Jackstadt, H. Hermeking / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

E

8

Please cite this article as: R. Jackstadt, H. Hermeking, MicroRNAs as regulators and mediators of c-MYC function, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbagrm.2014.04.003

665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750

R. Jackstadt, H. Hermeking / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

N C O

R

R

E

C

D

P

R O

O

F

[106] G.T. Bommer, et al., p53-mediated activation of miRNA34 candidate tumorsuppressor genes, Curr. Biol. 17 (15) (2007) 1298–1307. [107] H. Siemens, et al., Repression of c-Kit by p53 is mediated by miR-34 and is associated with reduced chemoresistance, migration and stemness, Oncotarget 4 (9) (2013) 1399–1415. [108] N.R. Christoffersen, et al., p53-independent upregulation of miR-34a during oncogene-induced senescence represses MYC, Cell Death Differ. 17 (2) (2010) 236–245. [109] H. Siemens, et al., miR-34 and SNAIL form a double-negative feedback loop to regulate epithelial–mesenchymal transitions, Cell Cycle 10 (24) (2011) 4256–4271. [110] M. Rokavec, et al., IL-6R/STAT3/miR-34a feedback loop promotes EMT-mediated colorectal cancer invasion and metastasis, J. Clin. Invest. (2014), http://dx.doi. org/10.1172/JCI73531 ([Epub ahead of print]). [111] C.Y. Cheng, et al., miR-34 cooperates with p53 in suppression of prostate cancer by joint regulation of stem cell compartment, Cell. Reprogram. (2014), http://dx.doi. org/10.1016/j.celrep.2014.02.023 ([Epub ahead of print]). [112] N. Okada, et al., A positive feedback between p53 and miR-34 miRNAs mediates tumor suppression, Genes Dev. 28 (5) (2014) 438–450. [113] D. Lodygin, et al., Inactivation of miR-34a by aberrant CpG methylation in multiple types of cancer, Cell Cycle 7 (16) (2008) 2591–2600. [114] M. Vogt, et al., Frequent concomitant inactivation of miR-34a and miR-34b/c by CpG methylation in colorectal, pancreatic, mammary, ovarian, urothelial, and renal cell carcinomas and soft tissue sarcomas, Virchows Arch. 458 (3) (2011) 313–322. [115] H. Siemens, et al., Detection of miR-34a promoter methylation in combination with elevated expression of c-Met and beta-catenin predicts distant metastasis of colon cancer, Clin. Cancer Res. 19 (3) (2013) 710–720. [116] E. Sotillo, et al., Myc overexpression brings out unexpected antiapoptotic effects of miR-34a, Oncogene 30 (22) (2011) 2587–2594. [117] J.S. Wei, et al., The MYCN oncogene is a direct target of miR-34a, Oncogene 27 (39) (2008) 5204–5213. [118] C. Welch, Y. Chen, R.L. Stallings, MicroRNA-34a functions as a potential tumor suppressor by inducing apoptosis in neuroblastoma cells, Oncogene 26 (34) (2007) 5017–5022. [119] K.A. Cole, et al., A functional screen identifies miR-34a as a candidate neuroblastoma tumor suppressor gene, Mol. Cancer Res. 6 (5) (2008) 735–742. [120] A.L. Kasinski, F.J. Slack, miRNA-34 prevents cancer initiation and progression in a therapeutically resistant K-ras and p53-induced mouse model of lung adenocarcinoma, Cancer Res. 72 (21) (2012) 5576–5587. [121] C. Liu, et al., The microRNA miR-34a inhibits prostate cancer stem cells and metastasis by directly repressing CD44, Nat. Med. 17 (2) (2011) 211–215. [122] H. Ling, M. Fabbri, G.A. Calin, MicroRNAs and other non-coding RNAs as targets for anticancer drug development, Nat. Rev. Drug Discov. 12 (11) (2013) 847–865. [123] H. Han, et al., A c-Myc-microRNA functional feedback loop affects hepatocarcinogenesis, Hepatology 57 (6) (2013) 2378–2389. [124] J.M. Liao, H. Lu, Autoregulatory suppression of c-Myc by miR-185-3p, J. Biol. Chem. 286 (39) (2011) 33901–33909. [125] V.B. Sampson, et al., MicroRNA let-7a down-regulates MYC and reverts MYCinduced growth in Burkitt lymphoma cells, Cancer Res. 67 (20) (2007) 9762–9770. [126] W. Abe, et al., miR-196b targets c-myc and Bcl-2 expression, inhibits proliferation and induces apoptosis in endometriotic stromal cells, Hum. Reprod. 28 (3) (2013) 750–761. [127] Y. Zhen, 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. 4 (2013) e872. [128] L.J. Miao, et al., MiR-449c targets c-Myc and inhibits NSCLC cell progression, FEBS Lett. 587 (9) (2013) 1359–1365. [129] J.S. Ho, et al., p53-Dependent transcriptional repression of c-myc is required for G1 cell cycle arrest, Mol. Cell. Biol. 25 (17) (2005) 7423–7431. [130] S. Yamamura, et al., MicroRNA-34a suppresses malignant transformation by targeting c-Myc transcriptional complexes in human renal cell carcinoma, Carcinogenesis 33 (2) (2012) 294–300. [131] S. Yamamura, et al., MicroRNA-34a modulates c-Myc transcriptional complexes to suppress malignancy in human prostate cancer cells, PLoS One 7 (1) (2012) e29722. [132] T.R. Kress, et al., The MK5/PRAK kinase and Myc form a negative feedback loop that is disrupted during colorectal tumorigenesis, Mol. Cell 41 (4) (2011) 445–457. [133] A. Lal, 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 35 (5) (2009) 610–625. [134] K.B. Challagundla, et al., Ribosomal protein L11 recruits miR-24/miRISC to repress c-Myc expression in response to ribosomal stress, Mol. Cell. Biol. 31 (19) (2011) 4007–4021. [135] M.S. Ebert, P.A. Sharp, Roles for microRNAs in conferring robustness to biological processes, Cell 149 (3) (2012) 515–524. [136] M. El Baroudi, et al., A curated database of miRNA mediated feed-forward loops involving MYC as master regulator, PLoS One 6 (3) (2011) e14742. [137] H.A. Coller, J.J. Forman, A. Legesse-Miller, “Myc'ed messages”: myc induces transcription of E2F1 while inhibiting its translation via a microRNA polycistron, PLoS Genet. 3 (8) (2007) e146. [138] D.G. Johnson, The paradox of E2F1: oncogene and tumor suppressor gene, Mol. Carcinog. 27 (3) (2000) 151–157. [139] L. Soucek, G.I. Evan, The ups and downs of Myc biology, Curr. Opin. Genet. Dev. 20 (1) (2010) 91–95.

E

T

[70] X.D. Xu, et al., Attenuation of microRNA-22 derepressed PTEN to effectively protect rat cardiomyocytes from hypertrophy, J. Cell. Physiol. 227 (4) (2012) 1391–1398. [71] N. Bar, R. Dikstein, miR-22 forms a regulatory loop in PTEN/AKT pathway and modulates signaling kinetics, PLoS One 5 (5) (2010) e10859. [72] B. Ling, et al., Tumor suppressor miR-22 suppresses lung cancer cell progression through post-transcriptional regulation of ErbB3, J. Cancer Res. Clin. Oncol. 138 (8) (2012) 1355–1361. [73] M. Feng, et al., Myc/miR-378/TOB2/cyclin D1 functional module regulates oncogenic transformation, Oncogene 30 (19) (2011) 2242–2251. [74] P. Mestdagh, et al., MYCN/c-MYC-induced microRNAs repress coding gene networks associated with poor outcome in MYCN/c-MYC-activated tumors, Oncogene 29 (9) (2010) 1394–1404. [75] J. Loven, et al., MYCN-regulated microRNAs repress estrogen receptor-alpha (ESR1) expression and neuronal differentiation in human neuroblastoma, Proc. Natl. Acad. Sci. U. S. A. 107 (4) (2010) 1553–1558. [76] J.H. Schulte, et al., MYCN regulates oncogenic microRNAs in neuroblastoma, Int. J. Cancer 122 (3) (2008) 699–704. [77] P.A. Northcott, et al., The miR-17/92 polycistron is up-regulated in sonic hedgehogdriven medulloblastomas and induced by N-myc in sonic hedgehog-treated cerebellar neural precursors, Cancer Res. 69 (8) (2009) 3249–3255. [78] L. Ma, et al., miR-9, a MYC/MYCN-activated microRNA, regulates E-cadherin and cancer metastasis, Nat. Cell Biol. 12 (3) (2010) 247–256. [79] R.R. Chivukula, J.T. Mendell, Circular reasoning: microRNAs and cell-cycle control, Trends Biochem. Sci. 33 (10) (2008) 474–481. [80] H. Hermeking, p53 enters the microRNA world, Cancer Cell 12 (5) (2007) 414–418. [81] H. Hermeking, The miR-34 family in cancer and apoptosis, Cell Death Differ. 17 (2) (2010) 193–199. [82] Y.S. Lee, A. Dutta, The tumor suppressor microRNA let-7 represses the HMGA2 oncogene, Genes Dev. 21 (9) (2007) 1025–1030. [83] P. Gao, et al., c-Myc suppression of miR-23a/b enhances mitochondrial glutaminase expression and glutamine metabolism, Nature 458 (7239) (2009) 762–765. [84] D.R. Wise, et al., Myc regulates a transcriptional program that stimulates mitochondrial glutaminolysis and leads to glutamine addiction, Proc. Natl. Acad. Sci. U. S. A. 105 (48) (2008) 18782–18787. [85] M.G. Vander Heiden, L.C. Cantley, C.B. Thompson, Understanding the Warburg effect: the metabolic requirements of cell proliferation, Science 324 (5930) (2009) 1029–1033. [86] R.J. DeBerardinis, et al., The biology of cancer: metabolic reprogramming fuels cell growth and proliferation, Cell Metab. 7 (1) (2008) 11–20. [87] H. Hermeking, MicroRNAs in the p53 network: micromanagement of tumour suppression, Nat. Rev. Cancer 12 (9) (2012) 613–626. [88] R.I. Aqeilan, G.A. Calin, C.M. Croce, miR-15a and miR-16-1 in cancer: discovery, function and future perspectives, Cell Death Differ. 17 (2) (2010) 215–220. [89] G.A. Calin, et al., Frequent deletions and down-regulation of micro-RNA genes miR15 and miR16 at 13q14 in chronic lymphocytic leukemia, Proc. Natl. Acad. Sci. U. S. A. 99 (24) (2002) 15524–15529. [90] U. Klein, et al., The DLEU2/miR-15a/16-1 cluster controls B cell proliferation and its deletion leads to chronic lymphocytic leukemia, Cancer Cell 17 (1) (2010) 28–40. [91] D. Bonci, et al., The miR-15a–miR-16-1 cluster controls prostate cancer by targeting multiple oncogenic activities, Nat. Med. 14 (11) (2008) 1271–1277. [92] A. Cimmino, et al., miR-15 and miR-16 induce apoptosis by targeting BCL2, Proc. Natl. Acad. Sci. U. S. A. 102 (39) (2005) 13944–13949. [93] Q. Liu, et al., miR-16 family induces cell cycle arrest by regulating multiple cell cycle genes, Nucleic Acids Res. 36 (16) (2008) 5391–5404. [94] O.S. Rissland, S.J. Hong, D.P. Bartel, MicroRNA destabilization enables dynamic regulation of the miR-16 family in response to cell-cycle changes, Mol. Cell 43 (6) (2011) 993–1004. [95] L. Shi, et al., p53-induced miR-15a/16-1 and AP4 form a double-negative feedback loop to regulate epithelial–mesenchymal transition and metastasis in colorectal cancer, Cancer Res. 74 (2) (2014) 532–542. [96] X. Zhang, et al., Myc represses miR-15a/miR-16-1 expression through recruitment of HDAC3 in mantle cell and other non-Hodgkin B-cell lymphomas, Oncogene 31 (24) (2012) 3002–3008. [97] H.I. Suzuki, et al., Modulation of microRNA processing by p53, Nature 460 (7254) (2009) 529–533. [98] M. Fabbri, et al., Association of a microRNA/TP53 feedback circuitry with pathogenesis and outcome of B-cell chronic lymphocytic leukemia, JAMA 305 (1) (2011) 59–67. [99] S. Sander, et al., MYC stimulates EZH2 expression by repression of its negative regulator miR-26a, Blood 112 (10) (2008) 4202–4212. [100] X. Zhang, et al., Coordinated silencing of MYC-mediated miR-29 by HDAC3 and EZH2 as a therapeutic target of histone modification in aggressive B-cell lymphomas, Cancer Cell 22 (4) (2012) 506–523. [101] J.L. Mott, et al., Transcriptional suppression of mir-29b-1/mir-29a promoter by cMyc, hedgehog, and NF-kappaB, J. Cell. Biochem. 110 (5) (2010) 1155–1164. [102] J. Kota, et al., Therapeutic microRNA delivery suppresses tumorigenesis in a murine liver cancer model, Cell 137 (6) (2009) 1005–1017. [103] V. Tarasov, et al., Differential regulation of microRNAs by p53 revealed by massively parallel sequencing: miR-34a is a p53 target that induces apoptosis and G1arrest, Cell Cycle 6 (13) (2007) 1586–1593. [104] M. Lize, A. Klimke, M. Dobbelstein, MicroRNA-449 in cell fate determination, Cell Cycle 10 (17) (2011) 2874–2882. [105] M. Lize, S. Pilarski, M. Dobbelstein, E2F1-inducible microRNA 449a/b suppresses cell proliferation and promotes apoptosis, Cell Death Differ. 17 (3) (2010) 452–458.

U

751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805 806 807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836

9

Please cite this article as: R. Jackstadt, H. Hermeking, MicroRNAs as regulators and mediators of c-MYC function, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbagrm.2014.04.003

837 838 839 840 841 842 843 844 845 846 847 848 849 850 851 852 853 854 855 856 857 858 859 860 861 862 863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889 890 891 892 893 894 895 896 897 898 899 900 901 902 903 904 905 906 907 908 909 910 911 912 913 914 915 916 917 918 919 920 921 922

D

P

R O

O

F

[164] H. Liu, et al., MYC suppresses cancer metastasis by direct transcriptional silencing of alphav and beta3 integrin subunits, Nat. Cell Biol. 14 (6) (2012) 567–574. [165] Z. Yu, et al., MicroRNA 17/20 inhibits cellular invasion and tumor metastasis in breast cancer by heterotypic signaling, Proc. Natl. Acad. Sci. U. S. A. 107 (18) (2010) 8231–8236. [166] H.B. Koch, et al., Large-scale identification of c-MYC-associated proteins using a combined TAP/MudPIT approach, Cell Cycle 6 (2) (2007) 205–217. [167] S. Hahn, et al., SNAIL and miR-34a feed-forward regulation of ZNF281/ZBP99 promotes epithelial–mesenchymal transition, EMBO J. 32 (23) (2013) 3079–3095. [168] R. Jackstadt, et al., AP4 is a mediator of epithelial–mesenchymal transition and metastasis in colorectal cancer, J. Exp. Med. 210 (7) (2013) 1331–1350. [169] R. Jackstadt, P. Jung, H. Hermeking, AP4 directly downregulates p16 and p21 to suppress senescence and mediate transformation, Cell Death Dis. 4 (2013) e775. [170] P. Jung, et al., AP4 encodes a c-MYC-inducible repressor of p21, Proc. Natl. Acad. Sci. U. S. A. 105 (39) (2008) 15046–15051. [171] A. Menssen, et al., The c-MYC oncoprotein, the NAMPT enzyme, the SIRT1-inhibitor DBC1, and the SIRT1 deacetylase form a positive feedback loop, Proc. Natl. Acad. Sci. U. S. A. 109 (4) (2012) E187–E196. [172] M. Yamakuchi, M. Ferlito, C.J. Lowenstein, miR-34a repression of SIRT1 regulates apoptosis, Proc. Natl. Acad. Sci. U. S. A. 105 (36) (2008) 13421–13426. [173] V. Byles, et al., SIRT1 induces EMT by cooperating with EMT transcription factors and enhances prostate cancer cell migration and metastasis, Oncogene 31 (43) (2012) 4619–4629. [174] G. Eades, et al., miR-200a regulates SIRT1 expression and epithelial to mesenchymal transition (EMT)-like transformation in mammary epithelial cells, J. Biol. Chem. 286 (29) (2011) 25992–26002. [175] P. Simic, et al., SIRT1 suppresses the epithelial-to-mesenchymal transition in cancer metastasis and organ fibrosis, Cell. Reprogram. 3 (4) (2013) 1175–1186. [176] A. Frenzel, J. Loven, M.A. Henriksson, Targeting MYC-regulated miRNAs to combat cancer, Genes Cancer 1 (6) (2010) 660–667. [177] M. Uhlen, et al., A human protein atlas for normal and cancer tissues based on antibody proteomics, Mol. Cell. Proteomics 4 (12) (2005) 1920–1932. [178] J.P. Morton, O.J. Sansom, MYC-y mice: from tumour initiation to therapeutic targeting of endogenous MYC, Mol. Oncol. 7 (2) (2013) 248–258. [179] C.J. Cheng, W.M. Saltzman, F.J. Slack, Canonical and non-canonical barriers facing antimiR cancer therapeutics, Curr. Med. Chem. 20 (29) (2013) 3582–3593. [180] A.L. Kasinski, F.J. Slack, Epigenetics and genetics. MicroRNAs en route to the clinic: progress in validating and targeting microRNAs for cancer therapy, Nat. Rev. Cancer 11 (12) (2011) 849–864. [181] E.V. Prochownik, P.K. Vogt, Therapeutic targeting of Myc, Genes Cancer 1 (6) (2010) 650–659. [182] M.S. Kumar, et al., Suppression of non-small cell lung tumor development by the let-7 microRNA family, Proc. Natl. Acad. Sci. U. S. A. 105 (10) (2008) 3903–3908. [183] P. Trang, et al., Regression of murine lung tumors by the let-7 microRNA, Oncogene 29 (11) (2010) 1580–1587. [184] M. Korpal, et al., Direct targeting of Sec23a by miR-200s influences cancer cell secretome and promotes metastatic colonization, Nat. Med. 17 (9) (2011) 1101–1108. [185] S. Liu, et al., MicroRNA-9 up-regulates E-cadherin through inhibition of NFkappaB1–Snail1 pathway in melanoma, J. Pathol. 226 (1) (2012) 61–72

C

T

[140] B.D. Aguda, et al., MicroRNA regulation of a cancer network: consequences of the feedback loops involving miR-17–92, E2F, and Myc, Proc. Natl. Acad. Sci. U. S. A. 105 (50) (2008) 19678–19683. [141] J.W. Kim, S. Mori, J.R. Nevins, Myc-induced microRNAs integrate Myc-mediated cell proliferation and cell fate, Cancer Res. 70 (12) (2010) 4820–4828. [142] L. Poliseno, et al., Identification of the miR-106b~25 microRNA cluster as a protooncogenic PTEN-targeting intron that cooperates with its host gene MCM7 in transformation, Sci. Signal. 3 (117) (2010) ra29. [143] L. Fontana, et al., Antagomir-17-5p abolishes the growth of therapy-resistant neuroblastoma through p21 and BIM, PLoS One 3 (5) (2008) e2236. [144] C.Y. Sun, et al., miR-15a and miR-16 affect the angiogenesis of multiple myeloma by targeting VEGF, Carcinogenesis 34 (2) (2013) 426–435. [145] K.J. Yin, et al., Vascular endothelial cell-specific microRNA-15a inhibits angiogenesis in hindlimb ischemia, J. Biol. Chem. 287 (32) (2012) 27055–27064. [146] S. Ling, et al., MicroRNA-dependent cross-talk between VEGF and HIF1alpha in the diabetic retina, Cell. Signal. 25 (12) (2013) 2840–2847. [147] S. Wang, E.N. Olson, AngiomiRs—key regulators of angiogenesis, Curr. Opin. Genet. Dev. 19 (3) (2009) 205–211. [148] A. Wolfer, S. Ramaswamy, MYC and metastasis, Cancer Res. 71 (6) (2011) 2034–2037. [149] A. Wolfer, et al., MYC regulation of a “poor-prognosis” metastatic cancer cell state, Proc. Natl. Acad. Sci. U. S. A. 107 (8) (2010) 3698–3703. [150] L. Kozma, et al., Investigation of c-myc oncogene amplification in colorectal cancer, Cancer Lett. 81 (2) (1994) 165–169. [151] J.P. Thiery, et al., Epithelial–mesenchymal transitions in development and disease, Cell 139 (5) (2009) 871–890. [152] S. Vanharanta, J. Massague, Origins of metastatic traits, Cancer Cell 24 (4) (2013) 410–421. [153] W.L. Tam, R.A. Weinberg, The epigenetics of epithelial–mesenchymal plasticity in cancer, Nat. Med. 19 (11) (2013) 1438–1449. [154] T. Brabletz, To differentiate or not—routes towards metastasis, Nat. Rev. Cancer 12 (6) (2012) 425–436. [155] J. Yang, R.A. Weinberg, Epithelial–mesenchymal transition: at the crossroads of development and tumor metastasis, Dev. Cell 14 (6) (2008) 818–829. [156] K.B. Cho, et al., Overexpression of c-Myc induces epithelial mesenchymal transition in mammary epithelial cells, Cancer Lett. 293 (2) (2010) 230–239. [157] V.H. Cowling, et al., c-Myc transforms human mammary epithelial cells through repression of the Wnt inhibitors DKK1 and SFRP1, Mol. Cell. Biol. 27 (14) (2007) 5135–5146. [158] A.P. Smith, et al., A positive role for Myc in TGFbeta-induced Snail transcription and epithelial-to-mesenchymal transition, Oncogene 28 (3) (2009) 422–430. [159] M.A. Nieto, Epithelial plasticity: a common theme in embryonic and cancer cells, Science 342 (6159) (2013) 1234850. [160] M. Liu, et al., c-Myc suppressed E-cadherin through miR-9 at the posttranscriptional level, Cell Biol. Int. 37 (3) (2013) 197–202. [161] Y. Sun, et al., Expression profile of microRNAs in c-Myc induced mouse mammary tumors, Breast Cancer Res. Treat. 118 (1) (2009) 185–196. [162] D. Chen, et al., LIFR is a breast cancer metastasis suppressor upstream of the Hippo– YAP pathway and a prognostic marker, Nat. Med. 18 (10) (2012) 1511–1517. [163] G. Zhuang, et al., Tumour-secreted miR-9 promotes endothelial cell migration and angiogenesis by activating the JAK–STAT pathway, EMBO J. 31 (17) (2012) 3513–3523.

E

923 924 925 926 927 928 929 930 931 932 933 934 935 936 937 938 939 940 941 942 943 944 945 946 947 948 949 950 951 952 953 954 955 956 957 958 959 960 961 962 963 964 965 966 967 968 969 970 971 972

R. Jackstadt, H. Hermeking / Biochimica et Biophysica Acta xxx (2014) xxx–xxx

E

10

U

N

C

O

R

R

1023

Please cite this article as: R. Jackstadt, H. Hermeking, MicroRNAs as regulators and mediators of c-MYC function, Biochim. Biophys. Acta (2014), http://dx.doi.org/10.1016/j.bbagrm.2014.04.003

973 974 975 976 977 978 979 980 981 982 983 984 985 986 987 988 989 990 991 992 993 994 995 996 997 998 999 1000 1001 1002 1003 1004 1005 1006 1007 1008 1009 1010 1011 1012 1013 1014 1015 1016 1017 1018 1019 1020 1021 1022

MicroRNAs as regulators and mediators of c-MYC function.

In the past ten years microRNAs (miRNAs) have been widely implicated as components of tumor suppressive and oncogenic pathways. Also the proto-typic o...
926KB Sizes 2 Downloads 3 Views