+ All Categories
Home > Documents > Tumor suppressor microRNAs: A novel non-coding alliance against cancer

Tumor suppressor microRNAs: A novel non-coding alliance against cancer

Date post: 20-Dec-2016
Category:
Upload: yosef
View: 231 times
Download: 5 times
Share this document with a friend
13
1 2 Review 4 Tumor suppressor microRNAs: A novel non-coding alliance against 5 cancer 6 7 8 Giovanni Blandino a,Q1 , Francesco Fazi b , Sara Donzelli a , Merav Kedmi c , Aldema Sas-Chen c , Paola Muti d , 9 Sabrina Strano e , Yosef Yarden c 10 a Translational Oncogenomics Unit, Italian National Cancer Institute ‘Regina Elena’, Rome, Italy 11 b Department of Anatomical, Histological, Forensic & Orthopaedic Sciences, Section of Histology & Medical Embryology, Sapienza University of Rome, Rome, Italy 12 c Weizmann Institute of Science, Department of Biological Regulation, Rehovot, Israel 13 d Department of Oncology, Juravinski Cancer Center-McMaster University Hamilton, Ontario, Canada 14 e Molecular Chemoprevention Unit, Italian National Cancer Institute ‘Regina Elena’, Rome, Italy 15 17 article info 18 Article history: 19 Received 28 January 2014 20 Revised 14 March 2014 21 Accepted 17 March 2014 22 Available online xxxx 23 24 Edited by Shairaz Baksh and Wilhelm Just 25 26 abstract 27 Tumor initiation and progression are the outcomes of a stepwise accumulation of genetic altera- 28 tions. Among these, gene amplification and aberrant expression of oncogenic proteins, as well as 29 deletion or inactivation of tumor suppressor genes, represent hallmark steps. Mounting evidence 30 collected over the last few years has identified different populations of non-coding RNAs as major 31 players in tumor suppression in almost all cancer types. Elucidating the diverse molecular mecha- 32 nisms underlying the roles of non-coding RNAs in tumor progression might provide illuminating 33 insights, potentially able to assist improved diagnosis, better staging and effective treatments of 34 human cancers. Here we focus on several groups of tumor suppressor microRNAs, whose downreg- 35 ulation exerts a profound oncologic impact and might be harnessed for the benefit of cancer 36 patients. 37 Ó 2014 Published by Elsevier B.V. on behalf of the Federation of European Biochemical Societies. 38 39 Q3 40 1. Introduction Q4 41 In most epithelial tissues, cancer develops through separate and 42 interrelated steps of clonal expansion, genetic diversification, and 43 clonal selection. During cancer development, cancer cells acquire 44 diverse biological capabilities that are conferred by numerous ge- 45 netic and epigenetic modifications [1]. In recent years, different 46 high-throughput platforms have been used for the genomic, tran- 47 scriptomic, proteomic, and epigenetic analyses to search for new 48 biomarkers involved in cancer and to bring new insights into the 49 several aspects of cancer pathophysiology [1]. In addition to the 50 classical transcriptional cell regulators involved in cancer develop- 51 ment, a class of non-coding RNAs, termed microRNAs (miRNAs) 52 has emerged as critical regulators of gene expression acting pre- 53 dominantly at the post-transcriptional level. MiRNAs were first 54 identified through their ability to regulate developmental pro- 55 cesses, such as developmental timing and cell fate transitions [2]. 56 Subsequently, miRNAs have been studied in relation to cancer 57 development. A large number of miRNAs that map to specific re- 58 gions of the human genome have been shown to be frequently de- 59 leted or amplified in cancer [3]. Several lines of evidence indicate 60 that miRNAs might be differentially expressed in cancer cells, in 61 which they form unique expression patterns or signatures [4]. 62 Sevignani and colleagues reported a significant association be- 63 tween the chromosomal location of miRNAs and those of mouse 64 susceptibility loci that influence the development of solid tumors 65 [5]. Dysregulation of miRNAs in cancer can occur through both epi- 66 genetic changes, including aberrant DNA methylation and histone 67 modification [6], and genetic alterations. These two biological 68 mechanisms can affect the production of the primary RNAs, their 69 processing to the mature miRNA forms, and/or interactions with 70 mRNA targets [7]. 71 More recent studies indicate that mutations affecting proteins 72 involved in the processing and maturation of miRNA, such as TAR- 73 BP2, DICER1 and XPO5, can also lead to overall reductions in miR- 74 NA expression [8–10]. Consistent with these observations, miRNAs 75 are thought to act mainly as tumor suppressor genes, and their 76 deregulation is currently recognized as a common feature of hu- 77 man cancers. Later on, additional data indicated that the expres- 78 sion of miRNAs is mainly downregulated in tumor tissues, as 79 compared to corresponding healthy tissues, which supported the 80 role of miRNAs as primarily tumor suppressors [4,8,9,11,12]. In 81 the same vein, there is evidence that an extensive downregulation 82 of miRNAs is one of the first outcomes of the stimulation of signal- 83 ing cascades downstream to specific growth factor receptors http://dx.doi.org/10.1016/j.febslet.2014.03.033 0014-5793/Ó 2014 Published by Elsevier B.V. on behalf of the Federation of European Biochemical Societies. Corresponding author Q2 . E-mail address: [email protected] (G. Blandino). FEBS Letters xxx (2014) xxx–xxx journal homepage: www.FEBSLetters.org FEBS 36528 No. of Pages 13, Model 5G 27 March 2014 Please cite this article in press as: Blandino, G., et al. Tumor suppressor microRNAs: A novel non-coding alliance against cancer. FEBS Lett. (2014), http:// dx.doi.org/10.1016/j.febslet.2014.03.033
Transcript
Page 1: Tumor suppressor microRNAs: A novel non-coding alliance against cancer

1

2

4

5

6

7

8 Q1

9

1011121314

15

1 7

1819202122232425

2 6

Q3

40 Q4

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

Q2

FEBS Letters xxx (2014) xxx–xxx

FEBS 36528 No. of Pages 13, Model 5G

27 March 2014

journal homepage: www.FEBSLetters .org

Review

Tumor suppressor microRNAs: A novel non-coding alliance againstcancer

http://dx.doi.org/10.1016/j.febslet.2014.03.0330014-5793/� 2014 Published by Elsevier B.V. on behalf of the Federation of European Biochemical Societies.

⇑ Corresponding author.E-mail address: [email protected] (G. Blandino).

Please cite this article in press as: Blandino, G., et al. Tumor suppressor microRNAs: A novel non-coding alliance against cancer. FEBS Lett. (2014),dx.doi.org/10.1016/j.febslet.2014.03.033

Giovanni Blandino a,⇑, Francesco Fazi b, Sara Donzelli a, Merav Kedmi c, Aldema Sas-Chen c, Paola Muti d,Sabrina Strano e, Yosef Yarden c

a Translational Oncogenomics Unit, Italian National Cancer Institute ‘Regina Elena’, Rome, Italyb Department of Anatomical, Histological, Forensic & Orthopaedic Sciences, Section of Histology & Medical Embryology, Sapienza University of Rome, Rome, Italyc Weizmann Institute of Science, Department of Biological Regulation, Rehovot, Israeld Department of Oncology, Juravinski Cancer Center-McMaster University Hamilton, Ontario, Canadae Molecular Chemoprevention Unit, Italian National Cancer Institute ‘Regina Elena’, Rome, Italy

a r t i c l e i n f o

27282930313233

Article history:Received 28 January 2014Revised 14 March 2014Accepted 17 March 2014Available online xxxx

Edited by Shairaz Baksh and Wilhelm Just

3435363738

a b s t r a c t

Tumor initiation and progression are the outcomes of a stepwise accumulation of genetic altera-tions. Among these, gene amplification and aberrant expression of oncogenic proteins, as well asdeletion or inactivation of tumor suppressor genes, represent hallmark steps. Mounting evidencecollected over the last few years has identified different populations of non-coding RNAs as majorplayers in tumor suppression in almost all cancer types. Elucidating the diverse molecular mecha-nisms underlying the roles of non-coding RNAs in tumor progression might provide illuminatinginsights, potentially able to assist improved diagnosis, better staging and effective treatments ofhuman cancers. Here we focus on several groups of tumor suppressor microRNAs, whose downreg-ulation exerts a profound oncologic impact and might be harnessed for the benefit of cancerpatients.

� 2014 Published by Elsevier B.V. on behalf of the Federation of European Biochemical Societies.

39

60

61

62

63

64

65

66

67

68

69

70

71

72

73

74

75

76

77

78

79

80

1. Introduction

In most epithelial tissues, cancer develops through separate andinterrelated steps of clonal expansion, genetic diversification, andclonal selection. During cancer development, cancer cells acquirediverse biological capabilities that are conferred by numerous ge-netic and epigenetic modifications [1]. In recent years, differenthigh-throughput platforms have been used for the genomic, tran-scriptomic, proteomic, and epigenetic analyses to search for newbiomarkers involved in cancer and to bring new insights into theseveral aspects of cancer pathophysiology [1]. In addition to theclassical transcriptional cell regulators involved in cancer develop-ment, a class of non-coding RNAs, termed microRNAs (miRNAs)has emerged as critical regulators of gene expression acting pre-dominantly at the post-transcriptional level. MiRNAs were firstidentified through their ability to regulate developmental pro-cesses, such as developmental timing and cell fate transitions [2].Subsequently, miRNAs have been studied in relation to cancerdevelopment. A large number of miRNAs that map to specific re-gions of the human genome have been shown to be frequently de-leted or amplified in cancer [3]. Several lines of evidence indicate

81

82

83

that miRNAs might be differentially expressed in cancer cells, inwhich they form unique expression patterns or signatures [4].Sevignani and colleagues reported a significant association be-tween the chromosomal location of miRNAs and those of mousesusceptibility loci that influence the development of solid tumors[5]. Dysregulation of miRNAs in cancer can occur through both epi-genetic changes, including aberrant DNA methylation and histonemodification [6], and genetic alterations. These two biologicalmechanisms can affect the production of the primary RNAs, theirprocessing to the mature miRNA forms, and/or interactions withmRNA targets [7].

More recent studies indicate that mutations affecting proteinsinvolved in the processing and maturation of miRNA, such as TAR-BP2, DICER1 and XPO5, can also lead to overall reductions in miR-NA expression [8–10]. Consistent with these observations, miRNAsare thought to act mainly as tumor suppressor genes, and theirderegulation is currently recognized as a common feature of hu-man cancers. Later on, additional data indicated that the expres-sion of miRNAs is mainly downregulated in tumor tissues, ascompared to corresponding healthy tissues, which supported therole of miRNAs as primarily tumor suppressors [4,8,9,11,12]. Inthe same vein, there is evidence that an extensive downregulationof miRNAs is one of the first outcomes of the stimulation of signal-ing cascades downstream to specific growth factor receptors

http://

Page 2: Tumor suppressor microRNAs: A novel non-coding alliance against cancer

84

85

86

87

88

89

90

91

92

93

94

95

96

97

98

99

100

101

102

103

104

105

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

131

132

133

134

135

136

137

138

139

140

141

142

143

144

145

146

147

148

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

196

197

198

199

200

201

202

203

204

205

206

207

208

209

210

211

212

213

2 G. Blandino et al. / FEBS Letters xxx (2014) xxx–xxx

FEBS 36528 No. of Pages 13, Model 5G

27 March 2014

implicated in a number of human cancers, including breast cancer[13]. For example, EGF signaling rapidly and simultaneously in-duces an extensive downregulation of multiple miRNAs, reflectingcoordinated regulation at the level of miRNA synthesis, processingor degradation [13].

Along with the dominance of tumor suppressor microRNAs,several well-characterized oncogenic miRNAs have been reportedin tumors. An interplay between RNA-binding proteins and onco-genic miRNAs, which drive expression of proto-oncogenes or main-tenance of stem cell phenotypes, contributes to human cancer [14].One example relates to oncogenic receptors for growth factors,such as the EGF-receptor (EGFR/ErbB) family of receptor tyrosinekinases, the expression of which is regulated by several miRNAs[15]. In the same vein, signaling pathways are ideal candidatesfor miRNA-mediated regulation, owing to the sharp dose-sensitivenature of their effects. For instance, EGFR activation induces miR-7expression through a RAS-MYC pathway. In support of this, MYCbinds to and activates the miR-7 promoter and ectopic miR-7 pro-motes cell growth and tumor formation in lung cancer cells [16].Thus, in addition to the EGFR/ErbB family, oncogenic miRNAs(onco-miRs) affect the responsiveness of cells to signaling mole-cules, such as transforming growth factor-beta, WNT and Notch[17]. miRNAs control not only cellular proliferation and pro-grammed cell death, but also dissemination of tumor cells and col-onization of distant organs (metastasis). Indeed, some miRNAs areassociated with the invasive and metastatic phenotype of breastand other cancer cell lines or metastatic tumor tissues [18,19].

MiRNAs are also deregulated upon exposure to both metaboliccancer risk factors and exposures to carcinogenic substances[20,21]. Thus, miRNAs may represent at the same time both predic-tors and players of cancer development. A number of life-style fac-tors (e.g., diet rich in fats and refined carbohydrates) andpathological conditions (e.g., obesity), often related to inflamma-tion and cancer, result in deregulation of specific miRNAs [22–25]. In addition, there is evidence of an altered expression of miR-NAs in relation to the exposure to well-known carcinogenic sub-stances such as asbestos, formaldehyde and cigarette smoke inlung and hepatic tissue [26,27]. In regard to this evidence, onestudy examined the expression of 484 miRNAs in the lungs of ratsexposed to environmental cigarette smoke for 4 weeks. It wasfound that 126 miRNAs were down-regulated at least 2-fold and24 miRNAs were downregulated more than 3-fold [28].

In this review, we highlight the contribution of miRNA modula-tion, in particular prevalent downregulation of specific miRNAs, tocancer development. Due to space consideration, this review con-centrates on a selected group of tumor suppressor microRNAs.Table 1 lists some additional molecules within this category, whichwe do not discuss in the main text. They include miR-34, a p53 tar-get gene [29,30], miR-31, an inhibitor of metastasis [31], as well asmiR-205 [32], miR-375 [33], miR-203 [34–36], as well as miR-15a[37–39]. Because many downregulated miRNAs function as tumorsuppressors, better understanding of the biological mechanismsunderlying their modulation will likely enable new strategies forprevention, early detection and therapy of cancer.

1.1. MiR-10b3p: the early arm of the miR-10b locus

The so-called miRNA-10b locus is located on chromosome 2,within the cluster of the HOXD genes, in an intergenic region be-tween HOXD4 and HOXD8 [40]. Processing by Drosha and Dicertransforms the RNA product of the miRNA-10b locus into a 22-nucleotide RNA duplex that contains two distinct 50 phosphory-lated strands with 30 overhangs (Fig. 1). The functional strand ofthe duplex, referred to as the guide strand, is miR-10b5p whilethe other, passenger strand generates miR-10b3p. MiR-10b5pwas originally identified as a molecules down-regulated in primary

Please cite this article in press as: Blandino, G., et al. Tumor suppressor microRdx.doi.org/10.1016/j.febslet.2014.03.033

breast tumors, compared to normal breast tissues [41]. Similarly,downregulation of miR-10b5p by promoter hyper-methylationhas been reported in gastric tumors [42]. The Weinberg’s group la-ter reported that miR-10b5p acts as a metastasis-supporting miR-NA, due to its ability to favor migration and invasion of breastcancer cells [43,44]. In line with this, miR-10b5p targets theHOXD10 gene, a repressor of several modulators of cell migration[44]. The expression of miR-10b5p is tightly controlled by the tran-scription factor Twist, a well-established regulator of epithelial-to-mesenchymal transition (EMT). Increased expression of miR-10b5p was detected in the vasculature of breast IDC (invasive duc-tal carcinoma) grade III tumors, compared to lower expression inDCIS (ductal carcinoma in situ) [45]. The pleiotropic activity ofmiR-10b5p could also rely on its ability to target the expressionof diverse tumor suppressors, including TP53, HODX10, PAX6,NOTCH1 and FOXO3 (see Table 2) [46].

Biagioni et al. originally reported that the expression of miR-10b3p was down-regulated in breast tumors, relative to matchedperi-tumoral tissues [12]. This downregulation occurred, at leastin part, through the methylation of CpG islands located withinthe regulatory regions of the miR-10b locus [12]. Ectopic expres-sion of miR-10b3p inhibited proliferation of breast cancer cell linesand reduced the size of xenografted breast tumors. Three pivotalproteins involved in the control of cell proliferation, namelyBUB1, PLK1, and CCNA2, were shown to serve as targets of miR-10b3p. Accordingly, intratumoral injection of a mimic of miR-10b3p reduced the expression of BUB1, PLK1 and CCNA2 proteins[12]. The prognostic role of miR-10b3p and of its target was eval-uated in the MEATBRIC dataset. This analysis included 1286 breastcancers from 5 different subtypes: HER2+ (127 patients), basal-like(209 patients), luminal A (479 patients), luminal B (312 patients),and normal-like (151 patients), for which both mRNA and miRNAdata were available [12]. mRNAs and miRNAs were measured foreach tumoral and normal samples of the METABRIC dataset. Kap-lan-Meyer analysis revealed a significant association betweenlow expression levels of miR-10b3p and poor disease-specific sur-vival [12]. This association was not evidenced for the augmentedexpression of miR-10b5p. The combined application of the COSMICalgorithm [47] and miRanda predictions uncovered 15 targetmRNAs of miR-10b3p [12]. Among those targets three, BUB1,PLK1 and CCNA2 were confirmed, and additional cell cycle relatedtargets were also identified. Increased expression of BUB1, PLK1and CCNA2 was associated with poor survival (Table 2) [12] .

These findings have several implications to the roles played bymiR-10b in breast tumorigenesis. Presumably, the expression ofmiR-10b3p is altered in the early stages of mammary cell transfor-mation. This could lead to aberrant cell proliferation, mediated byincreased expression of the cell cycle related targets of miR-10b3p.In line with early alterations, downregulation of miR-10b3pexpression appears to occur independently from the subtype ofbreast cancer, suggesting that it might represent an event preced-ing specification of breast cancer subtypes. Interestingly, the regu-lation of the expression of the two strands derived from the miR-10b locus is controlled by the combination of epigenetic and tran-scriptional events. While downregulation of miR10b-3p occursthrough methylation of CpG islands, the transcription factor TWISTup-regulates the expression of miR-10b5p [44]. This might under-lie mechanistically the dual and opposite roles of the miR-10b lo-cus: Early in breast tumorigenesis miR-10b3p downregulationleads to aberrant cell proliferation, while TWIST-mediated tran-scriptional induction of miR-10b5p contributes, as a late step, toshape a metastatic phenotype.

Unlike downregulation of miR-10b3p, which occurs indepen-dently from the breast cancer subtype, up-regulation ofmiR-10b5p might be specifically selected in highly metastaticbreast tumors. Thus, waves of miR-10b3p and 5p targets might

NAs: A novel non-coding alliance against cancer. FEBS Lett. (2014), http://

Page 3: Tumor suppressor microRNAs: A novel non-coding alliance against cancer

Fig. 1. Biogenesis and targets of miRNAs encoded by the miR-10b locus. RNA polymerase II transcribes miRNA genes, generating long primary transcripts (pri miRNAs).Drosha-mediated cleavage of pri-miRNAs leads to the formation of a hairpin molecule, the pre-miRNA, that is exported to the cytoplasm by the exportin-5/RAN-GTP complex.In the cytoplasm, the pre-miRNA is cleaved by Dicer, to produce a �22-nt RNA duplex. In the case of the mir-10b locus, these comprise two distinct 50 phosphorylated strandswith 30 overhangs, miR-10b-5p and miR-10b-3p. The functional strand of the duplex is incorporated into a multi-protein complex, the RNA-induced silencing complex (RISC),which regulates protein expression. In particular, in breast cancer functional miR-10b-3p targets BUB1, PLK1 and CCNA2 mRNAs, inducing a decrease in protein expressionand a consequent inhibition of cancer cell proliferation.

Table 1Major tumor suppressor microRNAs that are not discussed in the main text of this review.Q9

microRNA Target Function References

miR-34 CDK4cylinE2c-Met

CDK6NotchHMGA2BCL-2SIRT1AXL

Induces cell cycle arrest and apoptosisA p53 target geneEpigenetically silences in some tumors

[29,30]

miR-205 E2F1LAMC1

Induces senescence and reduces cell proliferationDirectly transactivated by wild type p53

[32]

miR-375 PDK114-3-3fAEG-1IGF1-receptor

Suppresses cell growth through the AKT pathwayEpigenetically silenced by DNA methylation

[33]

miR-31 Integrin alpha5RhoAMMP16RadixinWAVE3

Inhibits motility and invasivenessInduces anoikis

Valastian et al. (2010)

miR-203 Caveolin 1LASP1

Reduction of proliferation and inhibition of migration [34–36]

miR-15a Cyclin D1WNT3ACryptoBCL2

Inhibition of ovarian, breast and lung cell proliferation and invasionInduces apoptosis

[37–39]

G. Blandino et al. / FEBS Letters xxx (2014) xxx–xxx 3

FEBS 36528 No. of Pages 13, Model 5G

27 March 2014

Please cite this article in press as: Blandino, G., et al. Tumor suppressor microRNAs: A novel non-coding alliance against cancer. FEBS Lett. (2014), http://dx.doi.org/10.1016/j.febslet.2014.03.033

Page 4: Tumor suppressor microRNAs: A novel non-coding alliance against cancer

214

215

216

217

218

219

220

221

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

260

261

262

263

264

265

266

267

268

269

270

271

272

273

274

275

276

277

278

279

280

281

282

283

284

285

286

287

288

289

290

291

292

293

294

295

296

297

298

299

300

301

302Q5

303

304

305

306

307

308

309

310

311

312

313

314

315

316

317

318

319

320

321

322

323

324

325

Table 2miR-10b locus target genes.

Tissue Target Function References

miR-10b-5pGlioma TP53

FOXO3CYLDPAX6PTCH1HOXD10NOTCH1

Proliferation EMT invasion angiogenesis [46]

Breastcancer

HOXD10 Cell migration and invasion [44]

miR-10b-3pBreast

cancerBUB1PLK1CCNA2

Inhibition of cancer cell proliferation [12]

4 G. Blandino et al. / FEBS Letters xxx (2014) xxx–xxx

FEBS 36528 No. of Pages 13, Model 5G

27 March 2014

tune the pro-proliferative and pro-metastatic activities of the aber-rantly activated miR-10b locus. Intriguingly, miR-10b3p (previ-ously named miR-10b⁄) could represent a prototype of miRNAsderived from passenger strands, which target specific mRNAs andexert biological activities as efficiently as those originated fromguide strands. While the role of miR-10b5p is relatively well doc-umented in different human cancers, that of miR-10b3p is poorlyinvestigated. This also indicates that the miR-10b locus, via down-regulation of the 3p strand or up-regulation of 5p plays a pivotalrole in breast cancer establishment or dissemination. Once, themolecular events responsible for aberrant activation of the miR-10b locus in tumors will be fully understood, they might holdpromise for novel therapeutic strategies. This might turn true alsofor passenger derived tumor suppressor miRNAs.

1.2. Let-7c acts as a tumor suppressor miRNA

The let-7 family is one of the most ancient and conservedgroups of miRNAs, showing high conservation across species fromCaenorhabditis elegans to mammals [48]. In humans, the let-7 fam-ily is comprised of ten members (let-7a, let-7b, let-7c, let-7d, let-7e, let-7f, let-7g, let-7i, miR-98 and miR202), which differ in theirnucleotide sequences. Some of the isoforms appear in multiplecopies in the genome, hence a number is added as a suffix to theirname (e.g., let-7a-1, let-7a-2 and let-7a-3) (reviewed in [49]).Interestingly, many of the let-7 miRNAs are located in fragile sitesand specific genomic regions that relate to cancer [50]. Thus, forexample, in the human genome, the cluster let-7a-1/let-7f-1/let-7d is included in a frequently deleted region of chromosome 9 (re-gion B at 9q22.3). In breast carcinomas, a region of LOH (loss ofheterozygosity) at 11q23 was shown to harbor the cluster miR-125b1/let-7a-2/miR-100, and the cluster miR-99a/let-7c/miR-125b-2 resides at 21p11.1, a region of frequent HD (homozygousdeletions) in lung cancers. Furthermore, let-7a-1, let-7f-1, let-7dand miR-202 are located close to class II homeotic genes in the hu-man HOX gene clusters [50].

The let-7 family plays crucial roles in cellular differentiationand in development, and it displays specific temporal and spatialexpression patterns during development of several species[48,51]. For example, in mammals, the level of let-7 increases dur-ing embryogenesis and during brain development [52]. In concor-dance with developmental roles for the let-7 family, members ofthis family are also involved in cancer, and as discussed below,many of them act as tumor suppressor miRNAs. Downregulationof most let-7 family miRNAs was demonstrated in several typesof human cancer, including lung, head and neck squamous cell car-cinoma, breast, melanoma, ovarian and prostate cancers. Interest-ingly, silencing all let-7 family members in ovarian cancer cell lines

Please cite this article in press as: Blandino, G., et al. Tumor suppressor microRdx.doi.org/10.1016/j.febslet.2014.03.033

increased cell survival, invasion and adhesion [53]. The targets oflet-7 are also conserved from worms to humans, and they includeknown oncogenic transcripts, such as MYC and RAS [54]. Interest-ingly, the viral homolog of KRAS contains a single nucleotide poly-morphism (SNP) in its potential let-7 binding site, which increasesthe risk for lung, oral and colorectal cancers [55–57]. Another can-cer related target of the let-7 family is the high-mobility group AT-hook 2 (HMGA2) oncoprotein, a chromatin-associated non-histoneprotein that affects transcription by modulating chromatin’s archi-tecture [58].

Let-7c, a member of the let-7 family, functions as a tumor sup-pressor in several types of cancer (Fig. 2). It regulates EMT [55] andtargets various oncogenes and cancer related genes [59], such as:N-RAS [54], c-MYC [60], HMGA2 [58], MMP11, PBX2, PBX3, TRIB2,ITGB3, TGFBR1 [61], BCL-XL and MAP4K3 (see Table 3). In supportof a role for let-7c as a tumor suppressor, its reduced expressionwas demonstrated in tumors and cultured cells from prostate,lung, colorectal and hepatocellular tumors [60,62–65]. Addition-ally, downregulation of let-7c was associated with poor prognosisin colorectal cancer, where let-7c expression was significantly low-er in patients presenting lymph node involvement or distantmetastases, compared to patients without any detectable metasta-sis [65]. In non-small cell lung carcinoma (NSCLC), low expressionof let-7c was associated with poor patient survival, as well as withtumor spread, venous invasion and advanced disease stages [64].Interestingly, let-7c is encoded by chromosomal locus 21q21,which shows loss of heterozygosity in lung cancer [66].

The maturation process of miRNAs of the let-7 family, includingthat of let-7c, is regulated by the RNA-binding proteins calledLin28 and Lin28B, which employ several different regulatorymechanisms [67]. Lin28 is expressed in the cytoplasm and blocksprocessing of let-7 by Dicer through direct binding to the terminalloop of pre-let-7 and by recruiting the terminal uridyl transferase(TUTase) Zcchc11/TUT4 to catalyze the oligouridine tail, thusmarking pre-let-7 for degradation [68,69]. Lin28B is found mainlyin the nucleolus, and upon binding to pri-let-7 it blocks the activityof the microprocessor complex through TUTase-independentmechanism (reviewed by [70]). Interestingly, miRNAs of the let-7family have been suggested to reciprocally downregulate theexpression of Lin28 and Lin28B and thus increase the levels of ma-ture let-7 family miRNAs [71,72]. Let-7c’s expression levels canalso be regulated by the peroxisome proliferator-activated recep-tor a (PPARa) in hepatocellular carcinoma cells [63] (see Table 4).

Manipulation of let-7c expression in cancer cell lines furtherestablished its role as a tumor suppressor. For example, overex-pression of let-7c decreased, whereas depletion of let-7c increasedcell proliferation and clonogenicity of prostate cancer cells in vitro.Let-7c also exerted an anti-proliferative effect in vivo, when testedby intratumoral injection, which significantly reduced tumor sizein xenografts of human prostate cancer cells [60]. According to re-cent studies, let-7c can also function as a metastasis suppressor. Inhighly metastatic colorectal adenocarcinoma cells, ectopic over-expression of let-7c led to reduced migration and invasionin vitro, and almost completely inhibited tumor growth and metas-tasis. When tested in weakly metastatic cells, let-7c inhibition re-sulted in increased cell motility and invasion. These effects of let-7c on motility were likely mediated by targeting of KRAS,MMP11 and PBX3 [65]. Additionally, ectopic let-7c expression inchemotherapy-resistant lung adenocarcinoma cells reduced thenumber of metastatic nodules in lung and liver, probably via inac-tivation of the AKT pathway [55].

In addition to the AKT pathway, let-7c controls several othercancer-related signaling pathways. Thus, let-7c plays an importantrole in the regulation of androgen receptor (AR) signaling by di-rectly targeting MYC, thereby controlling prostate cancer prolifer-ation. Accordingly, the expression of let-7c and AR are negatively

NAs: A novel non-coding alliance against cancer. FEBS Lett. (2014), http://

Page 5: Tumor suppressor microRNAs: A novel non-coding alliance against cancer

326

327

328

329

330

331

332

333

334

335

336

337

338

339

340

341

342

343

Fig. 2. Let-7c regulation and functions. The biogenesis of the let-7 family, including let-7c, is controlled by the RNA-binding proteins Lin28 and Lin28B. In particular, Lin28Binhibits Drosha-mediated formation of pre-let-7, as well as Dicer-mediated formation of let-7 duplexes. In turn let-7 family members downregulate expression of Lin28 andLin28B, thereby establishing a negative feedback loop. Let-7c is downregulated in many types of tumors and its overexpression has a negative effect on cancer cellsproliferation, metastasis and epithelial to mesenchymal transition.

Table 3miR-let7c target genes.

Tissue Target Function References

Lung cancer RASTRIB2

Inhibition of proliferation [54], Wang et al. (2012)

Bcl-xL Inhibition of chemo- or radioresistance and EMT [55]ITGB3MAP4K3

Inhibition of migration and invasion [64]

Acute myeloid leukemia PBX2 Inhibition of leukemogenesis [62]Uterine leiomyoma HMGA2 Inhibition of proliferation [58]Colorectal cancer MMP11

PBX3Inhibition of metastatization [65]

Prostate cancer MYC Suppression of androgen receptor via regulation of MYC [60]Kidney fibrosis TGF-beta receptor 1 Inhibition of fibrosis by suppression of a receptor for TGF-beta [61]

G. Blandino et al. / FEBS Letters xxx (2014) xxx–xxx 5

FEBS 36528 No. of Pages 13, Model 5G

27 March 2014

correlated in human prostate cancer, where AR, MYC and Lin28expression levels are high while let-7c expression level is low[60]. In acute leukaemia, let7-c targets PBX2, a homeodomain pro-tein, which upon aberrant expression enhances HoxA9-dependentleukemogenesis, and promotes granulocytic differentiation [62].Let-7c may also suppress cell growth and control cancer pathogen-esis by regulating the mitogen-activated protein kinase (MAPK)pathway. In lung adenocarcinoma, let-7c directly targets TRIB2,which in turn activates the downstream components, C/EBP-a

Please cite this article in press as: Blandino, G., et al. Tumor suppressor microRdx.doi.org/10.1016/j.febslet.2014.03.033

and a phosphorylated p38-MAPK [73]. Another pathway by whichlet-7c can regulate the MAPK pathway is through its role in con-trolling NRAS expression [54]. One of the key metastasis-drivingprocesses is EMT [74], which is affected by several transcriptionalswitches, including a let-7c regulated switch. Docetaxel-resistantlung adenocarcinoma cells are characterized by fibroblast-likemorphology and adhesion, which are typical of the mesenchymalphenotype. However, upon expression of the let-7c precursor,these cells gained an epithelial phenotype, which might contribute

NAs: A novel non-coding alliance against cancer. FEBS Lett. (2014), http://

Page 6: Tumor suppressor microRNAs: A novel non-coding alliance against cancer

344

345

346

347

348

349

350

351

352

353

354

355

356

357

358

359

360

361

362

363

364

365

366

367

368

369

370

371

372

373

374

375

376

377

378

379

380

381

382

383

384

385

386

387

388

389

390

391

392

393

394

395

396

397

398

399

400

401

402

403

404

405

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

440

441

442

443

Table 4miR-223 target genes.

Tissue Target Function References

Acute myeloid leukemia E2F1 Inhibition of proliferation [81]Hepato-carcinoma STMN1

IGF-1RABCB1

Inhibition of proliferation and of multidrug resistance [86,85], Yang et al. (2013)

Lewis lung carcinoma CDK2 Inhibition of proliferation and invasion [88]Colorectal cancer FOXO1 Inhibition of proliferation [89]Glioblastoma PAX6 Proliferation invasion Huang et al. (2013)Breast cancer Caprin-1 Inhibition of proliferation and invasion Gong et al. (2013)Breast and colon cancer STMN1 Inhibition of chemoresistance [87]Gastric cancer EPB41L3 Invasion and metastatization Li et al. (2011)Esophageal carcinoma ARTN Inhibition of invasion and metastatization Li et al (2011)

6 G. Blandino et al. / FEBS Letters xxx (2014) xxx–xxx

FEBS 36528 No. of Pages 13, Model 5G

27 March 2014

to the restoration of chemo- and radio-sensitivity. This effect of let-7c is probably achieved through direct targeting of BCL-XL [55].

In summary, let-7c plays a crucial role in cancer pathogenesisthrough targeting key cancer-related proteins and acting as a sup-pressor of both tumor growth and tumor spread. Thus, let-7c mightbe considered an attractive candidate for drug-induced manipula-tion in cancer therapy.

1.3. miR-223 in cell differentiation and in tumor suppression

MiR-223 was initially identified as a miRNA nearly exclusivelyexpressed in bone marrow, such that its functional role in the reg-ulation of the cell fate determination of human hematopoietic pro-genitors cells (HPCs) rapidly emerged [75,76]. Recently, Vian andcolleagues evidenced that in human CD34 + HPCs undergoing uni-lineage differentiation/maturation, miR-223 is up-regulated duringgranulopoiesis rather than during monocytopoiesis and is main-tained at low levels during erythropoiesis [77]. Interestingly,miR-223 overexpression in human CD34 + HPCs favors granulopoi-esis and impairs erythroid and monocytic/macrophagic differenti-ation [77].

The fine-tuning of miR-223 expression levels during hemato-poietic differentiation of HPCs, as well as of myeloid cell lines, intoerythroid, granulocytic and monocytic/macrophagic lineages is theresult of the coordinated recruitment and function of lineage-spe-cific transcription factors (TFs) on two different regulatory regionsof the miR-223 promoter [76,77]. For example, during granulocyticdifferentiation of human myeloid progenitors, the induction ofmiR-223 expression is transcriptionally regulated by competitivebinding of two transcription factors (TFs), Nuclear Factor I (NFI-A) and CCAAT/enhanced-binding protein alpha (C/EBPa), to theproximal miR-223 regulatory region. NFI-A maintains miR-223transcription at low levels, while its replacement by C/EBPa resultsin miR-223 induction and granulocytic differentiation [78]. NFI-Awas also identified as a target for miR-223 at transcriptional andtranslational level, thus establishing a feedback regulatory cir-cuitry in the control of granulopoiesis [76,79]. Interestingly, miR-223 expression is downregulated in different subtypes of acutemyeloid leukemia (AML), which represents the clonal expansionof hematopoietic precursors blocked at different stages of differen-tiation [80]. In particular, primary blasts carrying the t(8;21) gen-erating AML1/ETO, the most common acute myeloid leukemia-associated fusion protein, present very low expression levels ofmiR-223. In these cells the AML1-ETO oncoprotein, which recruitsan epigenetic silencing complex consisting of HDACs, DNMTs, andmethyl-CpG-binding proteins on the AML1 binding site of the miR-223 promoter, links the epigenetic silencing of a miRNA locus tothe differentiation block of this acute myeloid leukemia subtype[80]. Of note, the enforced expression of miR-223 in primaryAML blasts and in AML cell lines affects cell cycle progressionand enhances granulocyte differentiation [76,77,80,81].

Please cite this article in press as: Blandino, G., et al. Tumor suppressor microRdx.doi.org/10.1016/j.febslet.2014.03.033

The tumor suppressor function of miR-223 in acute myeloidleukemia is further supported by a recent study showing that theinduction of miR-223 inhibits the translation of the cell-cycle reg-ulator E2F1 and significantly down-regulates the proliferation rateof myeloid progenitors cells [81]. Of note, E2F1 transcriptionallyrepresses miR-223 gene in AML cells, suggesting that miR-223functions as a key regulator of myeloid cell proliferation is strictlylinked with E2F1 activity in a mutual negative feedback loop.

Exosomes or micro-vesicles are emerging as important media-tors of intercellular cross-talk for the regulation of a variety of bio-logical functions, such as cellular communication, proliferation anddifferentiation [82]. Micro-vesicle transfer represents a novelmechanism by which infiltrating mononuclear phagocytes maycontribute to cellular activation, survival, and immune function.MiR-223 was evidenced to be the most highly expressed miRNAin the macrophage-derived micro-vesicles that are able to inducecellular differentiation when added to naive monocytes, support-ing a functional amplification loop to enhance immune function(Fig. 3A) [83].

In addition to the exosome-mediated transfer of nucleic acids,tissue macrophages were also shown to be able to transfer miR-223 to hepatocarcinoma cells (HCCs) in a manner that requiredintercellular contact and gap junction communication (Fig. 3B)[84]. Functionally, the transfer of miR-223 from macrophages toHuH7 cells resulted in the inhibition of proliferation of these HCCscancer cells, thus highlighting intercellular transfer of miRNA fromimmune cells as a new, possible mechanism of defense againstneoplastic cell transformation or tumor growth [84]. The transferof miR-223 influences protein expression in HuH7 cells. Specifi-cally, miR-223 decreases the expression of Stathmin-1 (STMN1)and insulin-like growth factor-1 receptor (IGF-1R), which bothinfluence cellular proliferation and can support the growth of tu-mors [85,86]. STMN1 is a key microtubule-regulatory protein thatcontrols microtubule dynamics, cell proliferation and, in particular,the S-phase of the cell cycle. High-levels of STMN1 have been asso-ciated with increased histologic grading, shorter patient survivaltimes, and increased drug resistance in different tumors. STMN1is a protein that is usually present at low levels in healthy hepato-cytes but is expressed at high levels in hepatocarcinomas [86]. InHCC cell lines, a strong inverse relationship between STMN1’smRNA and miR-223 expression was observed and a substantialreduction in STMN1 protein was demonstrated upon restorationof miR-223 expression, resulting in a consistent inhibitory effecton cell viability [86].

Consistent with these lines of evidence, it was recently reportedthat modulation of miR-223/STMN-1 pathway represents anotherway by which mutant p53 increases cellular resistance to chemo-therapeutic drugs [87]. The induction of mutant p53R175H in breastand colon cancer cell lines decreases miR-223 expression. Mutantp53R175H, together with the transcriptional repressor ZEB-1, bindsto the miR-223 promoter and decreases miR-223 expression,

NAs: A novel non-coding alliance against cancer. FEBS Lett. (2014), http://

Page 7: Tumor suppressor microRNAs: A novel non-coding alliance against cancer

444

445

446

447

448

449

450

451

452

453

454

455

456

457

458

459

460

461

462

463

464

465

466

467

468

469

470

471

472

473

474

475

476

477

478

479

480

481

482

483

484

485

486

487

488

489

490

491

492

493

494

495

496

497

498

499

500

501

502

503

504

505

506

507

508

509

Fig. 3. Macrophage-mediated miR-223 transfer. (A) miR-223 is released by macrophages into micro-vesicles and it induces cellular differentiation of naive monocytes toincrease in immune functions. (B) Macrophages mediate miR-223 transfer to cancer cells by mean of cell contact and gap junction communication. The transferred miR-223 isfunctionally active, and through decreased expression of Stathmin-1 (STMN1) and insulin-like growth factor-1 receptor (IGF-1R), it induces growth arrest of cancer cells.

G. Blandino et al. / FEBS Letters xxx (2014) xxx–xxx 7

FEBS 36528 No. of Pages 13, Model 5G

27 March 2014

resulting in an up-regulation of STMN-1 and in an increased cellu-lar resistance to chemotherapeutic agents. On the contrary, ectopicexpression of miR-223 can lower the levels of STMN-1 and sensi-tize breast and colon cancer cell lines expressing mutant p53 totreatment with DNA-damaging drugs [87].

By targeting IGF-1R and the cyclin-dependent kinase 2 (CDK2),miR-223 functions as a potent tumor suppressor also in the Lewislung carcinoma (LLC) cell line [88]. Ectopic expression of miR-223suppresses proliferation, invasion and tumorigenicity of LLC cells,induces G2/M phase arrest and inhibits tumor growth in vivo, pro-viding the basis for novel therapies targeting IGF-1R in the treat-ment of NSCLC [88].

The inhibition of cancer cell proliferation by miR-223 was re-cently reported also in colorectal cancer [89]. The colorectal cancerHCT116 cells express very low levels of endogenous miR-223 andoverexpression of miR-223 in these cells reduces mRNA and pro-tein expression levels of FOXO1, whose abnormal expression oractivation can result in aberrant apoptotic pathways, proliferation,and cell cycle regulation. Interestingly, miR-223 overexpressionmainly increases the un-phosphorylated FOXO1 protein, its nucle-ar localization, as well as cyclin D1/p21/p27 at either mRNA or pro-tein accumulation and inhibition of tumor cell proliferation [89].

Although several experimental lines of evidence support theinvolvement of miR-223 in cell differentiation and tumor suppres-sion, miR-223 was also reported to be up-regulated in some tu-mors and to contribute to leukemogenesis in specific diseasesubtypes (T-ALL), indicating that the biological effect of miR-223strongly depends on the cellular context where this miRNA specif-ically performs its function [90,91].

1.4. miR-145 in tumor suppression

miR-145 is located at chromosome 5q33.1 and is usually tran-scribed in a bicistronic primary transcript with miR-143 [92]. Of

Please cite this article in press as: Blandino, G., et al. Tumor suppressor microRdx.doi.org/10.1016/j.febslet.2014.03.033

note, miR-145 represents one of the miRNAs that are highly ex-pressed in normal tissues but they are down-regulated in severalhuman cancers, including colorectal and breast cancer [41,93].Although the downregulation of miR-145 expression in humancancer may result from genetic aberrations, as occurs in hemato-logical malignancies associated with the 5q- syndrome phenotype[94], a major mechanism for the modulation of miR-145 expres-sion is represented by transcriptional and post-transcriptional con-trol. Interestingly, transcriptional contribution to the regulation ofmiR-145 expression was reported in breast and in colon cancer,cell lines where p53, by interacting with a consensus sequence inthe miR-145 promoter, transcriptionally induces miR-145 expres-sion, promoting the post-transcriptional downregulation of MYCand consequently the inhibition of tumor cell growth [95]. In linewith these results, in prostate cancer tissues and in cell lines,miR-145 was found to be silenced through the methylation of itspromoter, and miR-145 silencing was significantly correlated tothe status of the p53 gene [96]. A recent study also reported thatthe CCAAT/enhancer binding protein beta (C/EBPb) is able to coun-teract the p53-mediated induction of miR-145. In fact, C/EBPb in-duces the transcriptional downregulation of miR-145 expressionby interacting with a CCAAT binding site located within the miR-145 promoter; this downregulation seems to be independent fromp53 and it involves the AKT pathway [95,97].

Concerning miR-143/miR-145 processing, it was recently re-ported that p53 itself might be involved in the post-transcriptionalmaturation of several miRNAs with growth-suppressive functions,in response to DNA damage, including miR-143/miR-145. In partic-ular, the p53 protein, through an association with the DEAD-boxRNA helicase p68 (also known as DDX5), interacts with the Droshaprocessing complex, thus supporting the processing of primarymiRNAs to precursor miRNAs. On the contrary, p53 mutated pro-teins interfere with the functional assembly between the Droshacomplex and p68, inducing an attenuation of miRNA processing

NAs: A novel non-coding alliance against cancer. FEBS Lett. (2014), http://

Page 8: Tumor suppressor microRNAs: A novel non-coding alliance against cancer

510

511

512

513

514

515

516

517

518

519

520

521

522

523

524

525

526

527

528

529

530

531

532

533

534

535

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

580

581

582

583

584

585

586

587

588

589

590

591

592

593

594

595

596

597

598

599

8 G. Blandino et al. / FEBS Letters xxx (2014) xxx–xxx

FEBS 36528 No. of Pages 13, Model 5G

27 March 2014

activity [98]. Moreover, it was also recently reported that theDEAD-box RNA helicase 6 (DDX6), that is highly expressed in mostmalignant cells, post-transcriptionally down-regulates both miR-143 and miR-145 expression. In human gastric cancer cells,DDX6 protein, that is abundantly expressed and accumulated inprocessing bodies (P-bodies) containing many proteins involvedin mRNA turnover, negatively regulates the RNA stability of thebicistronic primary transcript resulting in the downregulation ofboth miR-143 and miR-145 [99]. Several reports, supporting miR-145’s action as a tumor suppressor in different tumor types, haverevealed that the downregulation of miR-145 expression is associ-ated with neoplastic cell growth and proliferation, as well as withcancer cell migration, invasion and metastasis (Table 5) [100–104],

In colon cancer, the tumor suppressor function of miR-145 wasinitially related to the downregulation of the insulin receptor sub-strate-1 (IRS-1). Of note, in human colon cancer cells miR-145, bytargeting the 30 UTR of IRS-1, dramatically down-regulates the IRS-1 protein, thereby inducing cell growth arrest [105]. In colon can-cer cell lines, YES and STAT1 factors were also evidenced as addi-tional miR-145 targets [106]. Interestingly, in colon cancertissues, low miR-145 expression level is inversely correlated top70S6K1 protein levels. On the contrary, the forced expression ofmiR-145, by targeting p70S6K1, resulted in the downregulationof two downstream molecules of p70S6K1 pathway, the VEGFAand HIF-1a proteins, thus inhibiting tumor growth and angiogene-sis [107].

The tumor suppressive function of miR-145 on cancer cell pro-liferation was also recently reported in lung cancer. By targetingthe EGFR and nucleoside diphosphate linked moiety X-type motif1 (NUDT1), miR-145 inhibits lung adenocarcinoma cell prolifera-tion and lung tumorigenesis [108]. Of note, the prognostic valueof miR-145 was originally evidenced in lung cancer, where thelow expression miR-145 is significantly correlated with a worseprognosis and survival of adenocarcinoma patients [109] (seeTable 6).

In breast cancer tissues, miR-145 expression is inversely corre-lated with the stage of malignancy. In particular, miR-145, throughthe post-transcriptional regulation of N-RAS and VEGFA expres-sion, exhibits inhibitory roles in tumor angiogenesis, invasionand tumor growth [110]. Accordingly, in the human breast cancercell line MCF-7, miR-145 induction was shown to promote theinhibition of cell growth and the induction of apoptosis by target-ing the Rho-effector rhotekin (RTKN) [111]. A death-promotingloop between miR-145 and TP53 was also identified in MCF7breast cancer cells expressing wild-type TP53. miR-145 indeed

Table 5miR-145 target genes.

Tissue Target Funct

Colon cancer IRS-1YES and STAT1c-Myc

Grow

p70S6K1 GrowJNK and MAPK pathways ColonCD44, KLF5, KRAS, BRAF Trans

Lung cancer EGFR, NUDT1 ProlifBreast cancer MUC1

JAM-A and fascinMigra

N-RAS and VEGFARTKN

Angio

p53 pathway and ER-alpha ProlifZeb2 and Klf4 Cance

Mesothelioma OCT4 GrowPancreatic cancer KRAS and RREB1 KRASProstate cancer BNIP3

FSCN1ERG

Prolif

Please cite this article in press as: Blandino, G., et al. Tumor suppressor microRdx.doi.org/10.1016/j.febslet.2014.03.033

activates the p53 pathway, resulting in the promotion of apoptosis,and sustaining in turn a further induction of miR-145 expression.In the same context, miR-145 may also down-regulate estrogenreceptor-alpha (ER-a) protein expression, whereby a miR-145 re-expression therapy was proposed, at least for the subgroup of pa-tients with ER-a-positive and/or TP53 wild-type tumors [112].Interestingly, the re-expression of miR-143/miR-145 was alsoshown to suppress cellular growth and to support the apoptosisof epithelial cancer cells by enhancing p53 activity via MDM2 turn-over [113].

The involvement of miR-145 in an integrated transcriptionalregulatory circuit together with TFs and chromatin-modifyingactivities that support the growth and function of breast cancerstem-like cells (CSCs) recently emerged [114]. Accordingly, previ-ous results evidenced that expression of miR-145 is low in self-renewing human embryonic stem cells (hESCs) but highly up-reg-ulated during differentiation. Increased miR-145 expression inhib-its hESC self-renewal, represses expression of pluripotency genes,such as OCT4, SOX2, and KLF4, and induces lineage-restricted dif-ferentiation [115].

In malignant pleural mesothelioma (MPM) cells, miR-145 hasthe potential to modulate many pro-tumorigenic features, includ-ing growth, clonogenicity and tumor engraftment in vivo. Interest-ingly, in MPM cells miR-145 targets directly OCT4 and, indirectly,the EMT-promoting target ZEB1. Of note, the levels of miR-145and OCT4 are inversely correlated in vivo. Promoter hyper-methylation may contribute to the low levels of miR-145 in bothMPM cells and in malignant MPM tissues. Importantly, miR-145downregulation has been proposed to differentiate benign frommalignant mesothelial tissues [116].

In pancreatic cancer cells, the activation of the KRAS signalingpathway consistently leads to the repression of the miR-143/miR-145 cluster and this is necessary to maintain the tumorigenicpotential of these cancer cells. The downregulation of miR-143/miR-145 expression requires the RAS-responsive element-bindingprotein (RREB1), which represses the miR-143/miR-145 promoter.Both KRAS and RREB1 transcripts are direct targets of these miR-NAs (see Table 5), demonstrating the existence of a feed-forwardpathway that potentiates KRAS-mediated tumorigenesis [117].

More recently, the RREB1 protein was found to be overexpres-sed in colorectal adenocarcinoma tumors and cell lines, wherethe expression of the miR-143/miR-145 primary transcript is in-versely related to that of RREB1. RREB1 negatively regulatesexpression of the miR-143/miR-145 cluster in a KRAS-dependentmanner, thus establishing a complex network of regulation

ion References

th and proliferation [105,106]

th and angiogenesis [95]y and tumor formation [107]formation properties [118]eration and tumorigenesis [108]tion and invasion [100,101]

genesis, invasion and growth [110]

eration and apoptosis [111]r stem-like cells function [112]th and clonogenicity [116]-mediated tumorigenesis [117]eration, migration, invasion, EMT and metastasis [120,104,122]

NAs: A novel non-coding alliance against cancer. FEBS Lett. (2014), http://

Page 9: Tumor suppressor microRNAs: A novel non-coding alliance against cancer

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

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

Table 6miR-204 target genes.

Tissue Target Function References

Eye Meis2Sox11

Lens and retina development [126,125]

Gastric cancer BCL-2SIRT-1

Apoptosis MET [127,130]

Head and neck cancer Multiple targets Inhibition of metastatization [131]Endometrial carcinoma TrKB Inhibition of clonogenic growth, migration and invasion [132]Glioma Sox4

EphB2Inhibition of stem-cell likephenotype and migration [133]

Renal cell carcinoma LC3B Inhibition of macroautophagy [134]Pancreatic cancer Mcl-1 Apoptosis Chen et al. (2013)Breast cancer IL-11 Inhibition of metastatization Pollari et al (2012)

G. Blandino et al. / FEBS Letters xxx (2014) xxx–xxx 9

FEBS 36528 No. of Pages 13, Model 5G

27 March 2014

through which the miR-143/miR-145 cluster is able to modulateKRAS signaling [118]. In line with this, additional direct and indi-rect miR-143/miR-145 target genes have been reported, and theybelong to the growth factor receptor-mitogen-activated protein ki-nase network, as well as to the p53 signaling pathway, further sup-porting a contribution of miR-143/miR-145 to the cell signalingpathways involved in colorectal tumorigenesis [119].

In prostate cancer miR-145 is consistently downregulated. Asignificant inverse correlation between the expression of miR-145 and that of the BNIP3 protein was observed in prostate cancerand in in benign prostate tissues. Accordingly, the overexpressionof miR-145 in prostate cancer PC-3 and DU145 cells significantlydown-regulated BNIP3, reduced cell growth, and increased celldeath. As aforementioned for breast and colon cancer, the overex-pression of wild-type p53 resulted in the up-regulation of miR-145expression also in PC-3 cells, with consequent pro-apoptotic effects[120]. Wild-type p53 induces up-regulation of miR-145 expressionand the inhibitory effects of wild-type p53 on migration, invasion,EMT and stemness of PC-3 cells were reversed by anti-miR-145.These results suggest that loss of wild-type p53 may promote bonemetastasis of prostate cancer, at least partially through miR-145downregulation, and resulting in increased EMT and stemness ofcancer cells [121]. Moreover, the ectopic expression of miR-145in LNCaP and DU145 cell lines led to a reduction in the expressionof the ERG protein, suggesting that downregulation of miR-145associated with prostate cancer may contribute to the increasedexpression of several ERG isoforms that are frequently observedin this tumor type [122] .

In conclusion, several lines of evidence indicate that miR-145may be largely considered a miRNA with tumor suppressor activ-ity, which is involved in the regulation of tumor growth, cell inva-sion and metastasis by targeting multiple cancer related genes,thus offering miR-145 as a novel therapeutic target for cancertherapy.

1.5. MiR-204: a key player in development and in tumor suppression

MiR-204 is encoded by the cancer associated genomic region(CAGR), the 9q21.1-q22.3 locus that exhibits high frequency of lossof heterozygosity in diverse human cancers [123]. MiR-204 is anintragenic miRNA, which is located within the transient receptorpotential melastatin-3 (TRPM3) gene belonging to the family oftransient receptor potential (TRP) channels [124].

MiR-204 activity is highly involved in vertebrate lens develop-ment and its loss is evidenced in different human cancers(Fig. 4). Banfi’s group has shown that miR-204 is highly expressedin retinal pigment epithelium, lens, ciliary body and neural retina[125]. Its activity is required for the proper development of lensand optic cups [125]. The TF Meis2 is a main target of the develop-mental activity of miR-204. Aberrant expression of Meis2 releasedby miR-204 downregulation leads to lens abnormalities, microph-

Please cite this article in press as: Blandino, G., et al. Tumor suppressor microRdx.doi.org/10.1016/j.febslet.2014.03.033

thalamia, and eye coloboma [125]. Interestingly, miR-204 and itshost gene TRMP3 are transcriptionally co-regulated by the devel-opmental TF PAX6 [126]. The analysis of genes aberrantly ex-pressed in PAX6 mutants during eye development revealed thatmiR-204 target genes are highly represented. This led to identifica-tion of novel developmental target genes of miR-204, such asSox11, a member of the SoxC family of neuronal TFs, which playsa critical role in normal eye development. Intriguingly, PAX6 exertsits transcriptional program either by modulating directly theexpression of its targets genes, or by indirectly modulating theexpression of miR-204 that controls that of mRNA targets. Thisgives rise to a complex regulatory network that tunes and inte-grates coding and non-coding gene expression to pursue properdevelopment.

Growing evidence indicates that miR-204 downregulation is acommon alteration in different types of human tumors. MiRNAexpression profiling of three different subsets of gastric cancer pa-tients revealed that miR-204 expression was down-regulated in tu-moral specimens when compared to matched peri-tumoral tissues[127]. TRPM3 gene loss was evidenced in a large fraction of theanalysed gastric patients. Hence, this might be one of the molecu-lar mechanisms underlying miR-204 downregulation in humancancers [127–129]. Notably, gastric cancer patients can be groupedaccording to the extent of miR-204 downregulation. Those charac-terized by a severe reduction (more than 0.5-fold) had the worstsurvival when compared to those with mild downregulation ofmiR-204 (less than 0.5-fold) [127]. The pivotal anti-apoptotic pro-tein, BCL-2, was shown to be a target of miR-204. Reduced expres-sion of miR-204 paired with increased staining of BCL-2 in gastriccancer patients [127]. BCL-2 ectopic expression counteracted thepro-apoptotic role of miR-204 in response to 5-Fluorouracil[127]. Downregulation of miR-204 in gastric cancers was also asso-ciated with increased expression of the class III histone deacetylaseSIRT1. The targeting of SIRT1 by miR-204 ectopic expression fa-vored mesenchymal to epithelial transition (MET) phenotypesand suppressed anoikis resistance of gastric cancer cells [130].Notably, network modelling performed in head and neck tumorsby the combination of gene expression data with inheritable can-cer traits and risk factor loci uncovered 18 targets of miR-204 thatare mainly involved in the development of metastasis [131]. Inter-estingly, ectopic expression of miR-204 not only reduced the levelsof its target genes but also resulted in the inhibition of metastaticphenotype of head and neck cancer cell lines [131]. Altogether thelocalization of miR-204 linked the 9q21.1-q22.3 CAGR locus, a verywell known risk factor for head and neck tumors, and the identifiedtarget genes, thus accounting for a bona-fide tumor suppressor mi-cro-RNA. Downregulation of miR-204-5p was also evidenced inendometrial carcinomas (EC), when compared to normal endome-trium [132]. This might result from the aberrant expression of theneurotrophic receptor tyrosine kinase B (TrkB), which is a target ofmiR-204 [132]. Ectopic expression of TrkB led to the accumulation

NAs: A novel non-coding alliance against cancer. FEBS Lett. (2014), http://

Page 10: Tumor suppressor microRNAs: A novel non-coding alliance against cancer

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

751

Fig. 4. miR-204 functions. miR- 204 plays important roles in both eye development (A) and in tumor suppression (B). During eye development, high levels of miR-204 inducea reduction in Meis2 and Sox11 protein expression, thereby modulating the expression of other target proteins that contribute to normal development of the lens and retina.In particular, miR-204 expression is positively regulated by Pax6, a regulator of multiple processes during eye development, that in turn is activated by Meis2. Hence, miR-204 and Pax6 co-regulate each other via a negative feedback loop. miR-204 is downregulated in different types of cancer. Its tumor suppressor activity is mediated by themodulation of the expression of different oncoproteins depending on the cellular context.

10 G. Blandino et al. / FEBS Letters xxx (2014) xxx–xxx

FEBS 36528 No. of Pages 13, Model 5G

27 March 2014

of phospho-STAT3 that was recruited to a specific binding site ofthe miR-204’s host gene, TRMP3, and might account for miR-204downregulation in EC [132]. Loss of miR-204 expression throughpromoter methylation was evidenced in both glioma and neuralstem cells [133]. This led to enhanced migration of glioma cellsand to the acquisition of a stem cell-like phenotype. Indeed, atten-uation of promoter methylation increased miR-204 expression inglioma cells and ectopic expression of miR-204 suppressed thetumorigenic potential of glioma cells. Mechanistic investigation re-vealed that miR-204 targets the expression of SOX4, a stemness TF,and EphB2, a receptor that promotes migration [133].

MiR-204 is a transcriptional target of the von Hippel-Lindau tu-mor suppressor gene (VHL) that is lost in the largest fraction ofclear cell renal cell carcinomas (ccRCC) [134]. VHL-induced expres-sion of miR-204 resulted in increased expression of short tran-scripts of TRMP3, thus indicating that miR-204 is not, at least inccRCC, transcriptionally co-regulated with the large transcriptencoding the full-length TRMP3 protein [134]. MiR-204 expressionis clearly reduced in ccRCC with known VHL status when comparedto matched normal kidney samples and its reconstitution led togrowth inhibition in vitro and in vivo [134]. The transcriptional tu-mor suppressor axis VHL/miR-204 inhibited macroautophagy. Thisoccurs through the ability of miR-204 to directly target LC3B andVHL-induced expression of the parolog, LC3C, which causedgrowth suppression.

It appears increasingly clear, that both in development and incancer the downregulation of miR-204 disables coordinated

Please cite this article in press as: Blandino, G., et al. Tumor suppressor microRdx.doi.org/10.1016/j.febslet.2014.03.033

transcriptional networks and instigates unscheduled signalingpathways. These might contribute to developmental diseases orto aberrant proliferation, metastasis and poor response to conven-tional anticancer treatments. The identification of additional miR-204 targets and the dissection of the epigenetic events regulatingits expression in normal and in malignant tissues might provideintriguing insights, which would make miR-204 an appealing andhopefully druggable target.

1.6. Perspectives and open questions

Collectively, the reviewed groups of tumor suppressor miRNAsare emerging as major players of the cellular response to differenttypes of oncogenic insults. Together with coding RNAs, epigeneticmodifications and other mechanisms, this response can lead to clo-nal expansion, migration and invasion, as well as chemoresistanceof a given tumor. Along with ever improving understanding of theconcerted alterations in miRNAs, many cardinal questions remainopen. For example, it is logical that tumor suppressor miRNAsmaintain complex crosstalks with protein coding tumor suppressorRNAs in order to fortify barriers to oncogenicity, but the details ofthis interplay are currently unknown. Also unknown are themolecular determinants underlying activation of intragenic andintergenic tumor suppressor miRNAs upon oncogenic insults. Themultiplicity of RNA targets of each tumor suppressor miRNA posesa pivotal challenge. Moreover, it is conceivable that the tumor cellcontext influences target specificity, and consequently determines

NAs: A novel non-coding alliance against cancer. FEBS Lett. (2014), http://

Page 11: Tumor suppressor microRNAs: A novel non-coding alliance against cancer

752

753

754

755

756

757

758

759

760

761

762

763

764

765

766767768769770771772773774775776777778779780781782783784785786787788789790791792793794795796797798799800801802803804805806807808809810811812813814815816817818819820821822823824825

826827828829Q6830831832833834835836837838839840841842843844845846847848849850851852853854855856857858859860861862863864865866867868869870871872873874875876877878879880881882883884885886887888889890891892893894Q7895896897898899900901902903904905906907908909910911

G. Blandino et al. / FEBS Letters xxx (2014) xxx–xxx 11

FEBS 36528 No. of Pages 13, Model 5G

27 March 2014

biological effects. These and additional issues will require deeper,perhaps more integrated understanding of tumor suppressormiRNAs.

Conflicts of interest

The authors declare that they have no conflict of interest.

Acknowledgements

Contribution of Associazione Italiana per la Ricerca sul Cancro-Rome Oncogenomic Center to G.B., of Epigen to G.B., and of FIRB(Investment Fund for Basic Research) to G.B. was greatly appreci-ated. Yarden’s research is supported by the National Cancer Insti-tute, the German-Israeli Project Cooperation (DIP), the IsraelCancer Research Fund and the Dr. Miriam and Sheldon G. AdelsonMedical Research Foundation.

References

[1] Chin, L. and Gray, J.W. (2008) Translating insights from the cancer genomeinto clinical practice. Nature 452, 553–563.

[2] Kloosterman, W.P. and Plasterk, R.H. (2006) The diverse functions ofmicroRNAs in animal development and disease. Dev. Cell 11, 441–450.

[3] Calin, G.A., Sevignani, C., Dumitru, C.D., Hyslop, T., Noch, E., Yendamuri, S.,Shimizu, M., Rattan, S., Bullrich, F., Negrini, M. and Croce, C.M. (2004) HumanmicroRNA genes are frequently located at fragile sites and genomic regionsinvolved in cancers. Proc. Natl. Acad. Sci. USA 101, 2999–3004.

[4] Lu, J., Getz, G., Miska, E.A., Alvarez-Saavedra, E., Lamb, J., Peck, D., Sweet-Cordero, A., Ebert, B.L., Mak, R.H., Ferrando, A.A., Downing, J.R., Jacks, T.,Horvitz, H.R. and Golub, T.R. (2005) MicroRNA expression profiles classifyhuman cancers. Nature 435, 834–838.

[5] Sevignani, C., Calin, G.A., Nnadi, S.C., Shimizu, M., Davuluri, R.V., Hyslop, T.,Demant, P., Croce, C.M. and Siracusa, L.D. (2007) MicroRNA genes arefrequently located near mouse cancer susceptibility loci. Proc. Natl. Acad. Sci.USA 104, 8017–8022.

[6] Lopez-Serra, P. and Esteller, M. (2012) DNA methylation-associated silencingof tumor-suppressor microRNAs in cancer. Oncogene 31, 1609–1622.

[7] Melo, S.A. and Kalluri, R. (2012) Molecular pathways: microRNAs as cancertherapeutics. Clin. Cancer Res. 18, 4234–4239.

[8] Melo, S.A., Moutinho, C., Ropero, S., Calin, G.A., Rossi, S., Spizzo, R., Fernandez,A.F., Davalos, V., Villanueva, A., Montoya, G., Yamamoto, H., Schwartz Jr., S.and Esteller, M. (2010) A genetic defect in exportin-5 traps precursormicroRNAs in the nucleus of cancer cells. Cancer Cell 18, 303–315.

[9] Melo, S.A., Ropero, S., Moutinho, C., Aaltonen, L.A., Yamamoto, H., Calin, G.A.,Rossi, S., Fernandez, A.F., Carneiro, F., Oliveira, C., Ferreira, B., Liu, C.G.,Villanueva, A., Capella, G., Schwartz Jr., S., Shiekhattar, R. and Esteller, M.(2009) A TARBP2 mutation in human cancer impairs microRNA processingand DICER1 function. Nat. Genet. 41, 365–370.

[10] Hill, J.M., Zhao, Y., Clement, C., Neumann, D.M. and Lukiw, W.J. (2009) HSV-1infection of human brain cells induces miRNA-146a and Alzheimer-typeinflammatory signaling. Neuroreport 20, 1500–1505.

[11] Rosenfeld, N., Aharonov, R., Meiri, E., Rosenwald, S., Spector, Y., Zepeniuk, M.,Benjamin, H., Shabes, N., Tabak, S., Levy, A., Lebanony, D., Goren, Y.,Silberschein, E., Targan, N., Ben-Ari, A., Gilad, S., Sion-Vardy, N., Tobar, A.,Feinmesser, M., Kharenko, O., Nativ, O., Nass, D., Perelman, M., Yosepovich, A.,Shalmon, B., Polak-Charcon, S., Fridman, E., Avniel, A., Bentwich, I., Bentwich,Z., Cohen, D., Chajut, A. and Barshack, I. (2008) MicroRNAs accurately identifycancer tissue origin. Nat. Biotechnol. 26, 462–469.

[12] Biagioni, F., Bossel Ben-Moshe, N., Fontemaggi, G., Canu, V., Mori, F.,Antoniani, B., Di Benedetto, A., Santoro, R., Germoni, S., De Angelis, F.,Cambria, A., Avraham, R., Grasso, G., Strano, S., Muti, P., Mottolese, M., Yarden,Y., Domany, E. and Blandino, G. (2012) miR-10b⁄, a master inhibitor of thecell cycle, is down-regulated in human breast tumours. EMBO Mol. Med. 4,1214–1229.

[13] Avraham, R., Sas-Chen, A., Manor, O., Steinfeld, I., Shalgi, R., Tarcic, G., Bossel,N., Zeisel, A., Amit, I., Zwang, Y., Enerly, E., Russnes, H.G., Biagioni, F.,Mottolese, M., Strano, S., Blandino, G., Borresen-Dale, A.L., Pilpel, Y., Yakhini,Z., Segal, E. and Yarden, Y. (2010) EGF decreases the abundance of microRNAsthat restrain oncogenic transcription factors. Sci. Signal. 3 (ra43).

[14] Agami, R. (2010) MicroRNAs, RNA binding proteins and cancer. Eur. J. Clin.Invest. 40, 370–374.

[15] Barker, A., Giles, K.M., Epis, M.R., Zhang, P.M., Kalinowski, F. and Leedman, P.J.(2010) Regulation of ErbB receptor signalling in cancer cells by microRNA.Curr. Opin. Pharmacol. 10, 655–661.

[16] Chou, Y.T., Lin, H.H., Lien, Y.C., Wang, Y.H., Hong, C.F., Kao, Y.R., Lin, S.C.,Chang, Y.C., Lin, S.Y., Chen, S.J., Chen, H.C., Yeh, S.D. and Wu, C.W. (2010) EGFRpromotes lung tumorigenesis by activating miR-7 through a Ras/ERK/Mycpathway that targets the Ets2 transcriptional repressor ERF. Cancer Res. 70,8822–8831.

Please cite this article in press as: Blandino, G., et al. Tumor suppressor microRdx.doi.org/10.1016/j.febslet.2014.03.033

[17] Inui, M., Martello, G. and Piccolo, S. (2010) MicroRNA control of signaltransduction. Nat. Rev. Mol. Cell. Biol. 11, 252–263.

[18] Shi, M., Liu, D., Duan, H., Shen, B. and Guo, N. (2010) Metastasis-relatedmiRNAs, active players in breast cancer invasion, and metastasis. CancerMetastasis Rev. 29, 785–799.

[19] Visone, R. and Croce, C.M. (2009) MiRNAs and cancer. Am. J. Pathol. 174,1131–1138.

[20] Rotllan, N. and Fernandez-Hernando, C. (2012) MicroRNA regulation ofcholesterol metabolism. Cholesterol 2012, 847849.

[21] Davalos, A., Goedeke, L., Smibert, P., Ramirez, C.M., Warrier, N.P., Andreo, U.,Cirera-Salinas, D., Rayner, K., Suresh, U., Pastor-Pareja, J.C., Esplugues, E.,Fisher, E.A., Penalva, L.O., Moore, K.J., Suarez, Y., Lai, E.C. and Fernandez-Hernando, C. (2011) MiR-33a/b contribute to the regulation of fatty acidmetabolism and insulin signaling. Proc. Natl. Acad. Sci. USA 108, 9232–9237.

[22] Vasilescu, C., Rossi, S., Shimizu, M., Tudor, S., Veronese, A., Ferracin, M.,Nicoloso, M.S., Barbarotto, E., Popa, M., Stanciulea, O., Fernandez, M.H.,Tulbure, D., Bueso-Ramos, C.E., Negrini, M. and Calin, G.A. (2009) MicroRNAfingerprints identify miR-150 as a plasma prognostic marker in patients withsepsis. PLoS One 4, e7405.

[23] Schmidt, W.M., Spiel, A.O., Jilma, B., Wolzt, M. and Muller, M. (2009) In vivoprofile of the human leukocyte microRNA response to endotoxemia.Biochem. Biophys. Res. Commun. 380, 437–441.

[24] Danger, R., Paul, C., Giral, M., Lavault, A., Foucher, Y., Degauque, N., Pallier, A.,Durand, M., Castagnet, S., Duong Van Huyen, J.P., Delahousse, M., Renaudin,K., Soulillou, J.P. and Brouard, S. (2013) Expression of miR-142-5p inperipheral blood mononuclear cells from renal transplant patients withchronic antibody-mediated rejection. PLoS One 8, e60702.

[25] Radom-Aizik, S., Zaldivar Jr., F., Oliver, S., Galassetti, P. and Cooper, D.M.(2010) Evidence for microRNA involvement in exercise-associated neutrophilgene expression changes. J. Appl. Physiol. 109, 252–261.

[26] Rager, J.E., Smeester, L., Jaspers, I., Sexton, K.G. and Fry, R.C. (2011) Epigeneticchanges induced by air toxics: formaldehyde exposure alters miRNAexpression profiles in human lung cells. Environ. Health Perspect. 119,494–500.

[27] Elamin, B.K., Callegari, E., Gramantieri, L., Sabbioni, S. and Negrini, M. (2011)MicroRNA response to environmental mutagens in liver. Mutat. Res. 717, 67–76.

[28] Izzotti, A., Calin, G.A., Arrigo, P., Steele, V.E., Croce, C.M. and De Flora, S. (2009)Downregulation of microRNA expression in the lungs of rats exposed tocigarette smoke. FASEB J. 23, 806–812.

[29] He, L., He, X., Lim, L.P., de Stanchina, E., Xuan, Z., Liang, Y., Xue, W., Zender, L.,Magnus, J., Ridzon, D., Jackson, A.L., Linsley, P.S., Chen, C., Lowe, S.W., Cleary,M.A. and Hannon, G.J. (2007) A microRNA component of the p53 tumoursuppressor network. Nature 447, 1130–1134.

[30] Mudduluru, G., Ceppi, P., Kumarswamy, R., Scagliotti, G.V., Papotti, M. andAllgayer, H. (2011) Regulation of Axl receptor tyrosine kinase expression bymiR-34a and miR-199a/b in solid cancer. Oncogene 30, 2888–2899.

[31] Valastyan, S. and Weinberg, R.A. (2010) MiR-31: a crucial overseer of tumormetastasis and other emerging roles. Cell Cycle 9, 2124–2129.

[32] Piovan, C., Palmieri, D., Di Leva, G., Braccioli, L., Casalini, P., Nuovo, G.,Tortoreto, M., Sasso, M., Plantamura, I., Triulzi, T., Taccioli, C., Tagliabue, E.,Iorio, M.V. and Croce, C.M. (2012) Oncosuppressive role of p53-induced miR-205 in triple negative breast cancer. Mol. Oncol. 6, 458–472.

[33] Zhou, J., Song, S., He, S., Zhu, X., Zhang, Y., Yi, B., Zhang, B., Qin, G. and Li, D.(2014) MicroRNA-375 targets PDK1 in pancreatic carcinoma and suppressescell growth through the Akt signaling pathway. Int. J. Mol. Med. 33, 950–956.

[34] Miao, L., Xiong, X., Lin, Y., Cheng, Y., Lu, J., Zhang, J. and Cheng, N. (2014) MiR-203 inhibits tumor cell migration and invasion via caveolin-1 in pancreaticcancer cells. Oncol. Lett. 7, 658–662.

[35] Wang, C., Zheng, X., Shen, C. and Shi, Y. (2012) MicroRNA-203 suppresses cellproliferation and migration by targeting BIRC5 and LASP1 in human triple-negative breast cancer cells. J. Exp. Clin. Cancer Res. 31, 58.

[36] Takeshita, N., Mori, M., Kano, M., Hoshino, I., Akutsu, Y., Hanari, N.,Yoneyama, Y., Ikeda, N., Isozaki, Y., Maruyama, T., Akanuma, N., Miyazawa,Y. and Matsubara, H. (2012) MiR-203 inhibits the migration and invasion ofesophageal squamous cell carcinoma by regulating LASP1. Int. J. Oncol. 41,1653–1661.

[37] Chen, F., Hou, S.K., Fan, H.J. and Liu, Y.F. (2014) MiR-15a-16 represses Criptoand inhibits NSCLC cell progression. Mol. Cell. Biochem..

[38] Bonci, D., Coppola, V., Musumeci, M., Addario, A., Giuffrida, R., Memeo, L.,D’Urso, L., Pagliuca, A., Biffoni, M., Labbaye, C., Bartucci, M., Muto, G., Peschle,C. and De Maria, R. (2008) The miR-15a-miR-16-1 cluster controls prostatecancer by targeting multiple oncogenic activities. Nat. Med. 14, 1271–1277.

[39] Cimmino, A., Calin, G.A., Fabbri, M., Iorio, M.V., Ferracin, M., Shimizu, M.,Wojcik, S.E., Aqeilan, R.I., Zupo, S., Dono, M., Rassenti, L., Alder, H., Volinia, S.,Liu, C.G., Kipps, T.J., Negrini, M. and Croce, C.M. (2005) MiR-15 and miR-16induce apoptosis by targeting BCL2. Proc. Natl. Acad. Sci. USA 102, 13944–13949.

[40] Biagioni, F., Bossel Ben-Moshe, N., Fontemaggi, G., Yarden, Y., Domany, E. andBlandino, G. (2013) The locus of microRNA-10b: a critical target for breastcancer insurgence and dissemination. Cell Cycle 12.

[41] Iorio, M.V., Ferracin, M., Liu, C.G., Veronese, A., Spizzo, R., Sabbioni, S., Magri,E., Pedriali, M., Fabbri, M., Campiglio, M., Menard, S., Palazzo, J.P., Rosenberg,A., Musiani, P., Volinia, S., Nenci, I., Calin, G.A., Querzoli, P., Negrini, M. andCroce, C.M. (2005) MicroRNA gene expression deregulation in human breastcancer. Cancer Res. 65, 7065–7070.

NAs: A novel non-coding alliance against cancer. FEBS Lett. (2014), http://

Page 12: Tumor suppressor microRNAs: A novel non-coding alliance against cancer

912913914915916917918919920921922923924925926927928929930931932933934935936937938939940941942943944945946947948949950951952953954955956957958959960961962963964965966967968969970971972973974975976977978979980981982983984985986987988989990991992993994995996

9979989991000100110021003100410051006100710081009101010111012101310141015101610171018101910201021102210231024102510261027102810291030103110321033103410351036103710381039104010411042104310441045104610471048104910501051105210531054105510561057105810591060106110621063106410651066106710681069107010711072107310741075107610771078107910801081

12 G. Blandino et al. / FEBS Letters xxx (2014) xxx–xxx

FEBS 36528 No. of Pages 13, Model 5G

27 March 2014

[42] Kim, K., Lee, H.C., Park, J.L., Kim, M., Kim, S.Y., Noh, S.M., Song, K.S., Kim, J.C.and Kim, Y.S. (2011) Epigenetic regulation of microRNA-10b and targeting ofoncogenic MAPRE1 in gastric cancer. Epigenetics 6, 740–751.

[43] Ma, L., Reinhardt, F., Pan, E., Soutschek, J., Bhat, B., Marcusson, E.G., Teruya-Feldstein, J., Bell, G.W. and Weinberg, R.A. (2010) Therapeutic silencing ofmiR-10b inhibits metastasis in a mouse mammary tumor model. Nat.Biotechnol. 28, 341–347.

[44] Ma, L., Teruya-Feldstein, J. and Weinberg, R.A. (2007) Tumour invasion andmetastasis initiated by microRNA-10b in breast cancer. Nature 449, 682–688.

[45] Plummer, P.N., Freeman, R., Taft, R.J., Vider, J., Sax, M., Umer, B.A., Gao, D.,Johns, C., Mattick, J.S., Wilton, S.D., Ferro, V., McMillan, N.A., Swarbrick, A.,Mittal, V. and Mellick, A.S. (2013) MicroRNAs regulate tumor angiogenesismodulated by endothelial progenitor cells. Cancer Res. 73, 341–352.

[46] Lin, J., Teo, S., Lam, D.H., Jeyaseelan, K. and Wang, S. (2012) MicroRNA-10bpleiotropically regulates invasion, angiogenicity and apoptosis of tumor cellsresembling mesenchymal subtype of glioblastoma multiforme. Cell DeathDis. 3, e398.

[47] Bossel Ben-Moshe, N., Avraham, R., Kedmi, M., Zeisel, A., Yitzhaky, A., Yarden,Y. and Domany, E. (2012) Context-specific microRNA analysis: identificationof functional microRNAs and their mRNA targets. Nucleic Acids Res. 40,10614–10627.

[48] Pasquinelli, A.E., Reinhart, B.J., Slack, F., Martindale, M.Q., Kuroda, M.I., Maller,B., Hayward, D.C., Ball, E.E., Degnan, B., Muller, P., Spring, J., Srinivasan, A.,Fishman, M., Finnerty, J., Corbo, J., Levine, M., Leahy, P., Davidson, E. andRuvkun, G. (2000) Conservation of the sequence and temporal expression oflet-7 heterochronic regulatory RNA. Nature 408, 86–89.

[49] Roush, S. and Slack, F.J. (2008) The let-7 family of microRNAs. Trends CellBiol. 18, 505–516.

[50] Calin, G.A., Cimmino, A., Fabbri, M., Ferracin, M., Wojcik, S.E., Shimizu, M.,Taccioli, C., Zanesi, N., Garzon, R., Aqeilan, R.I., Alder, H., Volinia, S., Rassenti,L., Liu, X., Liu, C.G., Kipps, T.J., Negrini, M. and Croce, C.M. (2008) MiR-15a andmiR-16-1 cluster functions in human leukemia. Proc. Natl. Acad. Sci. USA 105,5166–5171.

[51] Reinhart, B.J., Slack, F.J., Basson, M., Pasquinelli, A.E., Bettinger, J.C., Rougvie,A.E., Horvitz, H.R. and Ruvkun, G. (2000) The 21-nucleotide let-7 RNAregulates developmental timing in Caenorhabditis elegans. Nature 403, 901–906.

[52] Wulczyn, F.G., Smirnova, L., Rybak, A., Brandt, C., Kwidzinski, E., Ninnemann,O., Strehle, M., Seiler, A., Schumacher, S. and Nitsch, R. (2007) Post-transcriptional regulation of the let-7 microRNA during neural cellspecification. FASEB J. 21, 415–426.

[53] Yang, X., Rutnam, Z.J., Jiao, C., Wei, D., Xie, Y., Du, J., Zhong, L. and Yang, B.B.(2012) An anti-let-7 sponge decoys and decays endogenous let-7 functions.Cell Cycle 11, 3097–3108.

[54] Johnson, S.M., Grosshans, H., Shingara, J., Byrom, M., Jarvis, R., Cheng, A.,Labourier, E., Reinert, K.L., Brown, D. and Slack, F.J. (2005) RAS is regulated bythe let-7 microRNA family. Cell 120, 635–647.

[55] Cui, S.Y., Huang, J.Y., Chen, Y.T., Song, H.Z., Feng, B., Huang, G.C., Wang, R.,Chen, L.B. and De, W. (2013) Let-7c governs the acquisition of chemo- orradioresistance and epithelial-to-mesenchymal transition phenotypes indocetaxel-resistant lung adenocarcinoma. Mol. Cancer Res. 11, 699–713.

[56] Christensen, B.C., Moyer, B.J., Avissar, M., Ouellet, L.G., Plaza, S.L., McClean,M.D., Marsit, C.J. and Kelsey, K.T. (2009) A let-7 microRNA-binding sitepolymorphism in the KRAS 30 UTR is associated with reduced survival in oralcancers. Carcinogenesis 30, 1003–1007.

[57] Zhang, W., Winder, T., Ning, Y., Pohl, A., Yang, D., Kahn, M., Lurje, G., Labonte,M.J., Wilson, P.M., Gordon, M.A., Hu-Lieskovan, S., Mauro, D.J., Langer, C.,Rowinsky, E.K. and Lenz, H.J. (2011) A let-7 microRNA-binding sitepolymorphism in 30-untranslated region of KRAS gene predicts response inwild-type KRAS patients with metastatic colorectal cancer treated withcetuximab monotherapy. Ann. Oncol. 22, 104–109.

[58] Peng, Y., Laser, J., Shi, G., Mittal, K., Melamed, J., Lee, P. and Wei, J.J. (2008)Antiproliferative effects by Let-7 repression of high-mobility group A2 inuterine leiomyoma. Mol. Cancer Res. 6, 663–673.

[59] Tzur, G., Israel, A., Levy, A., Benjamin, H., Meiri, E., Shufaro, Y., Meir, K.,Khvalevsky, E., Spector, Y., Rojansky, N., Bentwich, Z., Reubinoff, B.E. andGalun, E. (2009) Comprehensive gene and microRNA expression profilingreveals a role for microRNAs in human liver development. PLoS One 4, e7511.

[60] Nadiminty, N., Tummala, R., Lou, W., Zhu, Y., Shi, X.B., Zou, J.X., Chen, H.,Zhang, J., Chen, X., Luo, J., deVere White, R.W., Kung, H.J., Evans, C.P. and Gao,A.C. (2012) MicroRNA let-7c is downregulated in prostate cancer andsuppresses prostate cancer growth. PLoS One 7, e32832.

[61] Brennan, E.P., Nolan, K.A., Borgeson, E., Gough, O.S., McEvoy, C.M., Docherty,N.G., Higgins, D.F., Murphy, M., Sadlier, D.M., Ali-Shah, S.T., Guiry, P.J., Savage,D.A., Maxwell, A.P., Martin, F., Godson, C. and Consortium, G. (2013) Lipoxinsattenuate renal fibrosis by inducing let-7c and suppressing TGFbetaR1. J. Am.Soc. Nephrol. 24, 627–637.

[62] Pelosi, A., Careccia, S., Lulli, V., Romania, P., Marziali, G., Testa, U., Lavorgna, S.,Lo-Coco, F., Petti, M.C., Calabretta, B., Levrero, M., Piaggio, G. and Rizzo, M.G.(2013) MiRNA let-7c promotes granulocytic differentiation in acute myeloidleukemia. Oncogene 32, 3648–3654.

[63] Shah, Y.M., Morimura, K., Yang, Q., Tanabe, T., Takagi, M. and Gonzalez, F.J.(2007) Peroxisome proliferator-activated receptor alpha regulates amicroRNA-mediated signaling cascade responsible for hepatocellularproliferation. Mol. Cell. Biol. 27, 4238–4247.

1082

Please cite this article in press as: Blandino, G., et al. Tumor suppressor microRdx.doi.org/10.1016/j.febslet.2014.03.033

[64] Zhao, B., Han, H., Chen, J., Zhang, Z., Li, S., Fang, F., Zheng, Q., Ma, Y., Zhang, J.,Wu, N. and Yang, Y. (2014) MicroRNA let-7c inhibits migration and invasionof human non-small cell lung cancer by targeting ITGB3 and MAP4K3. CancerLett. 342, 43–51.

[65] Han, H.B., Gu, J., Zuo, H.J., Chen, Z.G., Zhao, W., Li, M., Ji, D.B., Lu, Y.Y. andZhang, Z.Q. (2012) Let-7c functions as a metastasis suppressor by targetingMMP11 and PBX3 in colorectal cancer. J. Pathol. 226, 544–555.

[66] Yamada, H., Yanagisawa, K., Tokumaru, S., Taguchi, A., Nimura, Y., Osada, H.,Nagino, M. and Takahashi, T. (2008) Detailed characterization of ahomozygously deleted region corresponding to a candidate tumorsuppressor locus at 21q11-21 in human lung cancer. Genes ChromosomesCancer 47, 810–818.

[67] Piskounova, E., Polytarchou, C., Thornton, J.E., LaPierre, R.J., Pothoulakis, C.,Hagan, J.P., Iliopoulos, D. and Gregory, R.I. (2011) Lin28A and Lin28B inhibitlet-7 microRNA biogenesis by distinct mechanisms. Cell 147, 1066–1079.

[68] Heo, I., Joo, C., Cho, J., Ha, M., Han, J. and Kim, V.N. (2008) Lin28 mediates theterminal uridylation of let-7 precursor MicroRNA. Mol. Cell 32, 276–284.

[69] Heo, I., Joo, C., Kim, Y.K., Ha, M., Yoon, M.J., Cho, J., Yeom, K.H., Han, J. and Kim,V.N. (2009) TUT4 in concert with Lin28 suppresses microRNA biogenesisthrough pre-microRNA uridylation. Cell 138, 696–708.

[70] Thornton, J.E. and Gregory, R.I. (2012) How does Lin28 let-7 controldevelopment and disease? Trends Cell Biol. 22, 474–482.

[71] Moss, E.G. and Tang, L. (2003) Conservation of the heterochronic regulatorLin-28, its developmental expression and microRNA complementary sites.Dev. Biol. 258, 432–442.

[72] Guo, Y., Chen, Y., Ito, H., Watanabe, A., Ge, X., Kodama, T. and Aburatani, H.(2006) Identification and characterization of lin-28 homolog B (LIN28B) inhuman hepatocellular carcinoma. Gene 384, 51–61.

[73] Wang, P.Y., Sun, Y.X., Zhang, S., Pang, M., Zhang, H.H., Gao, S.Y., Zhang, C., Lv,C.J. and Xie, S.Y. (2013) Let-7c inhibits A549 cell proliferation throughoncogenic TRIB2 related factors. FEBS Lett. 587, 2675–2681.

[74] Thiery, J.P., Acloque, H., Huang, R.Y. and Nieto, M.A. (2009) Epithelial-mesenchymal transitions in development and disease. Cell 139, 871–890.

[75] Chen, C.Z., Li, L., Lodish, H.F. and Bartel, D.P. (2004) MicroRNAs modulatehematopoietic lineage differentiation. Science 303, 83–86.

[76] Fazi, F., Rosa, A., Fatica, A., Gelmetti, V., De Marchis, M.L., Nervi, C. andBozzoni, I. (2005) A minicircuitry comprised of microRNA-223 andtranscription factors NFI-A and C/EBPalpha regulates humangranulopoiesis. Cell 123, 819–831.

[77] Vian, L., Di Carlo, M., Pelosi, E., Fazi, F., Santoro, S., Cerio, A.M., Boe, A., Rotilio,V., Billi, M., Racanicchi, S., Testa, U., Grignani, F. and Nervi, C. (2014)Transcriptional fine-tuning of microRNA-223 levels directs lineage choice ofhuman hematopoietic progenitors. Cell Death Differ. 21, 290–301.

[78] Nervi, C., Fazi, F., Rosa, A., Fatica, A. and Bozzoni, I. (2007) Emerging role formicroRNAs in acute promyelocytic leukemia. Curr. Top. Microbiol. Immunol.313, 73–84.

[79] Zardo, G., Ciolfi, A., Vian, L., Starnes, L.M., Billi, M., Racanicchi, S., Maresca, C.,Fazi, F., Travaglini, L., Noguera, N., Mancini, M., Nanni, M., Cimino, G., Lo-Coco,F., Grignani, F. and Nervi, C. (2012) Polycombs and microRNA-223 regulatehuman granulopoiesis by transcriptional control of target gene expression.Blood 119, 4034–4046.

[80] Fazi, F., Racanicchi, S., Zardo, G., Starnes, L.M., Mancini, M., Travaglini, L.,Diverio, D., Ammatuna, E., Cimino, G., Lo-Coco, F., Grignani, F. and Nervi, C.(2007) Epigenetic silencing of the myelopoiesis regulator microRNA-223 bythe AML1/ETO oncoprotein. Cancer Cell 12, 457–466.

[81] Pulikkan, J.A., Dengler, V., Peramangalam, P.S., Peer Zada, A.A., Muller-Tidow,C., Bohlander, S.K., Tenen, D.G. and Behre, G. (2010) Cell-cycle regulator E2F1and microRNA-223 comprise an autoregulatory negative feedback loop inacute myeloid leukemia. Blood 115, 1768–1778.

[82] Valadi, H., Ekstrom, K., Bossios, A., Sjostrand, M., Lee, J.J. and Lotvall, J.O.(2007) Exosome-mediated transfer of mRNAs and microRNAs is a novelmechanism of genetic exchange between cells. Nat. Cell Biol. 9, 654–659.

[83] Ismail, N., Wang, Y., Dakhlallah, D., Moldovan, L., Agarwal, K., Batte, K., Shah,P., Wisler, J., Eubank, T.D., Tridandapani, S., Paulaitis, M.E., Piper, M.G. andMarsh, C.B. (2013) Macrophage microvesicles induce macrophagedifferentiation and miR-223 transfer. Blood 121, 984–995.

[84] Aucher, A., Rudnicka, D. and Davis, D.M. (2013) MicroRNAs transfer fromhuman macrophages to hepato-carcinoma cells and inhibit proliferation. J.Immunol. 191, 6250–6260.

[85] Jia, C.Y., Li, H.H., Zhu, X.C., Dong, Y.W., Fu, D., Zhao, Q.L., Wu, W. and Wu, X.Z.(2011) MiR-223 suppresses cell proliferation by targeting IGF-1R. PLoS One 6,e27008.

[86] Wong, Q.W., Lung, R.W., Law, P.T., Lai, P.B., Chan, K.Y., To, K.F. and Wong, N.(2008) MicroRNA-223 is commonly repressed in hepatocellular carcinomaand potentiates expression of Stathmin1. Gastroenterology 135, 257–269.

[87] Masciarelli, S., Fontemaggi, G., Di Agostino, S., Donzelli, S., Carcarino, E.,Strano, S. and Blandino, G. (2013) Gain-of-function mutant p53downregulates miR-223 contributing to chemoresistance of cultured tumorcells. Oncogene.

[88] Nian, W., Ao, X., Wu, Y., Huang, Y., Shao, J., Wang, Y., Chen, Z., Chen, F. andWang, D. (2013) MiR-223 functions as a potent tumor suppressor of theLewis lung carcinoma cell line by targeting insulin-like growth factor-1receptor and cyclin-dependent kinase 2. Oncol. Lett. 6, 359–366.

[89] Wu, L., Li, H., Jia, C.Y., Cheng, W., Yu, M., Peng, M., Zhu, Y., Zhao, Q., Dong, Y.W.,Shao, K., Wu, A. and Wu, X.Z. (2012) MicroRNA-223 regulates FOXO1expression and cell proliferation. FEBS Lett. 586, 1038–1043.

NAs: A novel non-coding alliance against cancer. FEBS Lett. (2014), http://

Page 13: Tumor suppressor microRNAs: A novel non-coding alliance against cancer

1083108410851086108710881089109010911092109310941095 Q81096109710981099110011011102110311041105110611071108110911101111111211131114111511161117111811191120112111221123112411251126112711281129113011311132113311341135113611371138113911401141114211431144114511461147114811491150115111521153115411551156115711581159116011611162116311641165

116611671168116911701171117211731174117511761177117811791180118111821183118411851186118711881189119011911192119311941195119611971198119912001201120212031204120512061207120812091210121112121213121412151216121712181219122012211222122312241225122612271228122912301231123212331234123512361237123812391240124112421243124412451246

1247

G. Blandino et al. / FEBS Letters xxx (2014) xxx–xxx 13

FEBS 36528 No. of Pages 13, Model 5G

27 March 2014

[90] Mansour, M.R., Sanda, T., Lawton, L.N., Li, X., Kreslavsky, T., Novina, C.D.,Brand, M., Gutierrez, A., Kelliher, M.A., Jamieson, C.H., von Boehmer, H.,Young, R.A. and Look, A.T. (2013) The TAL1 complex targets the FBXW7tumor suppressor by activating miR-223 in human T cell acute lymphoblasticleukemia. J. Exp. Med. 210, 1545–1557.

[91] Mavrakis, K.J., Van Der Meulen, J., Wolfe, A.L., Liu, X., Mets, E., Taghon, T.,Khan, A.A., Setty, M., Rondou, P., Vandenberghe, P., Delabesse, E., Benoit, Y.,Socci, N.B., Leslie, C.S., Van Vlierberghe, P., Speleman, F. and Wendel, H.G.(2011) A cooperative microRNA-tumor suppressor gene network in acute T-cell lymphoblastic leukemia (T-ALL). Nat. Genet. 43, 673–678.

[92] Iio, A., Nakagawa, Y., Hirata, I., Naoe, T. and Akao, Y. (2010) Identification ofnon-coding RNAs embracing microRNA-143/145 cluster. Mol. Cancer 9, 136.

[93] Michael, M.Z., O’ Connor, S.M., van Holst Pellekaan, N.G., Young, G.P. andJames, R.J. (2003) Reduced accumulation of specific microRNAs in colorectalneoplasia. Mol. Cancer Res. 1, 882–891.

[94] Starczynowski, D.T., Kuchenbauer, F., Argiropoulos, B., Sung, S., Morin, R.,Muranyi, A., Hirst, M., Hogge, D., Marra, M., Wells, R.A., Buckstein, R., Lam, W.,Humphries, R.K. and Karsan, A. (2010) Identification of miR-145 and miR-146a as mediators of the 5q- syndrome phenotype. Nat. Med. 16, 49–58.

[95] Sachdeva, M., Zhu, S., Wu, F., Wu, H., Walia, V., Kumar, S., Elble, R., Watabe, K.and Mo, Y.Y. (2009) P53 represses c-Myc through induction of the tumorsuppressor miR-145. Proc. Natl. Acad. Sci. USA 106, 3207–3212.

[96] Suh, S.O., Chen, Y., Zaman, M.S., Hirata, H., Yamamura, S., Shahryari, V., Liu, J.,Tabatabai, Z.L., Kakar, S., Deng, G., Tanaka, Y. and Dahiya, R. (2011)MicroRNA-145 is regulated by DNA methylation and p53 gene mutation inprostate cancer. Carcinogenesis 32, 772–778.

[97] Sachdeva, M., Liu, Q., Cao, J., Lu, Z. and Mo, Y.Y. (2012) Negative regulation ofmiR-145 by C/EBP-beta through the Akt pathway in cancer cells. NucleicAcids Res. 40, 6683–6692.

[98] Suzuki, H.I., Yamagata, K., Sugimoto, K., Iwamoto, T., Kato, S. and Miyazono, K.(2009) Modulation of microRNA processing by p53. Nature 460, 529–533.

[99] Iio, A., Takagi, T., Miki, K., Naoe, T., Nakayama, A. and Akao, Y. (2013) DDX6post-transcriptionally down-regulates miR-143/145 expression through hostgene NCR143/145 in cancer cells. Biochim. Biophys. Acta 1829, 1102–1110.

[100] Sachdeva, M. and Mo, Y.Y. (2010) MicroRNA-145 suppresses cell invasionand metastasis by directly targeting mucin 1. Cancer Res. 70, 378–387.

[101] Gotte, M., Mohr, C., Koo, C.Y., Stock, C., Vaske, A.K., Viola, M., Ibrahim, S.A.,Peddibhotla, S., Teng, Y.H., Low, J.Y., Ebnet, K., Kiesel, L. and Yip, G.W. (2010)MiR-145-dependent targeting of junctional adhesion molecule A andmodulation of fascin expression are associated with reduced breast cancercell motility and invasiveness. Oncogene 29, 6569–6580.

[102] Zhang, H., Pu, J., Qi, T., Qi, M., Yang, C., Li, S., Huang, K., Zheng, L. and Tong, Q.(2014) MicroRNA-145 inhibits the growth, invasion, metastasis andangiogenesis of neuroblastoma cells through targeting hypoxia-induciblefactor 2 alpha. Oncogene 33, 387–397.

[103] Gao, P., Xing, A.Y., Zhou, G.Y., Zhang, T.G., Zhang, J.P., Gao, C., Li, H. and Shi,D.B. (2013) The molecular mechanism of microRNA-145 to suppressinvasion-metastasis cascade in gastric cancer. Oncogene 32, 491–501.

[104] Fuse, M., Nohata, N., Kojima, S., Sakamoto, S., Chiyomaru, T., Kawakami, K.,Enokida, H., Nakagawa, M., Naya, Y., Ichikawa, T. and Seki, N. (2011)Restoration of miR-145 expression suppresses cell proliferation, migrationand invasion in prostate cancer by targeting FSCN1. Int. J. Oncol. 38, 1093–1101.

[105] Shi, B., Sepp-Lorenzino, L., Prisco, M., Linsley, P., deAngelis, T. and Baserga, R.(2007) Micro RNA 145 targets the insulin receptor substrate-1 and inhibitsthe growth of colon cancer cells. J. Biol. Chem. 282, 32582–32590.

[106] Gregersen, L.H., Jacobsen, A.B., Frankel, L.B., Wen, J., Krogh, A. and Lund, A.H.(2010) MicroRNA-145 targets YES and STAT1 in colon cancer cells. PLoS One5, e8836.

[107] Xu, Q., Liu, L.Z., Qian, X., Chen, Q., Jiang, Y., Li, D., Lai, L. and Jiang, B.H. (2012)MiR-145 directly targets p70S6K1 in cancer cells to inhibit tumor growth andangiogenesis. Nucleic Acids Res. 40, 761–774.

[108] Cho, W.C., Chow, A.S. and Au, J.S. (2011) MiR-145 inhibits cell proliferation ofhuman lung adenocarcinoma by targeting EGFR and NUDT1. RNA Biol. 8,125–131.

[109] Yanaihara, N., Caplen, N., Bowman, E., Seike, M., Kumamoto, K., Yi, M.,Stephens, R.M., Okamoto, A., Yokota, J., Tanaka, T., Calin, G.A., Liu, C.G., Croce,C.M. and Harris, C.C. (2006) Unique microRNA molecular profiles in lungcancer diagnosis and prognosis. Cancer Cell 9, 189–198.

[110] Zou, C., Xu, Q., Mao, F., Li, D., Bian, C., Liu, L.Z., Jiang, Y., Chen, X., Qi, Y., Zhang,X., Wang, X., Sun, Q., Kung, H.F., Lin, M.C., Dress, A., Wardle, F., Jiang, B.H. andLai, L. (2012) MiR-145 inhibits tumor angiogenesis and growth by N-RAS andVEGF. Cell Cycle 11, 2137–2145.

[111] Wang, S., Bian, C., Yang, Z., Bo, Y., Li, J., Zeng, L., Zhou, H. and Zhao, R.C. (2009)MiR-145 inhibits breast cancer cell growth through RTKN. Int. J. Oncol. 34,1461–1466.

[112] Spizzo, R., Nicoloso, M.S., Lupini, L., Lu, Y., Fogarty, J., Rossi, S., Zagatti, B.,Fabbri, M., Veronese, A., Liu, X., Davuluri, R., Croce, C.M., Mills, G., Negrini, M.and Calin, G.A. (2010) MiR-145 participates with TP53 in a death-promotingregulatory loop and targets estrogen receptor-alpha in human breast cancercells. Cell Death Differ. 17, 246–254.

[113] Zhang, J., Sun, Q., Zhang, Z., Ge, S., Han, Z.G. and Chen, W.T. (2013) Loss ofmicroRNA-143/145 disturbs cellular growth and apoptosis of human

Please cite this article in press as: Blandino, G., et al. Tumor suppressor microRdx.doi.org/10.1016/j.febslet.2014.03.033

epithelial cancers by impairing the MDM2-p53 feedback loop. Oncogene32, 61–69.

[114] Polytarchou, C., Iliopoulos, D. and Struhl, K. (2012) An integratedtranscriptional regulatory circuit that reinforces the breast cancer stem cellstate. Proc. Natl. Acad. Sci. USA 109, 14470–14475.

[115] Xu, N., Papagiannakopoulos, T., Pan, G., Thomson, J.A. and Kosik, K.S. (2009)MicroRNA-145 regulates OCT4, SOX2, and KLF4 and represses pluripotencyin human embryonic stem cells. Cell 137, 647–658.

[116] Cioce, M., Ganci, F., Canu, V., Sacconi, A., Mori, F., Canino, C., Korita, E., Casini,B., Alessandrini, G., Cambria, A., Carosi, M.A., Blandino, R., Panebianco, V.,Facciolo, F., Visca, P., Volinia, S., Muti, P., Strano, S., Croce, C.M., Pass, H.I. andBlandino, G. (2013) Protumorigenic effects of mir-145 loss in malignantpleural mesothelioma. Oncogene.

[117] Kent, O.A., Chivukula, R.R., Mullendore, M., Wentzel, E.A., Feldmann, G., Lee,K.H., Liu, S., Leach, S.D., Maitra, A. and Mendell, J.T. (2010) Repression of themiR-143/145 cluster by oncogenic Ras initiates a tumor-promoting feed-forward pathway. Genes Dev. 24, 2754–2759.

[118] Kent, O.A., Fox-Talbot, K. and Halushka, M.K. (2013) RREB1 repressed miR-143/145 modulates KRAS signaling through downregulation of multipletargets. Oncogene 32, 2576–2585.

[119] Pagliuca, A., Valvo, C., Fabrizi, E., di Martino, S., Biffoni, M., Runci, D., Forte, S.,De Maria, R. and Ricci-Vitiani, L. (2013) Analysis of the combined action ofmiR-143 and miR-145 on oncogenic pathways in colorectal cancer cellsreveals a coordinate program of gene repression. Oncogene 32, 4806–4813.

[120] Chen, X., Gong, J., Zeng, H., Chen, N., Huang, R., Huang, Y., Nie, L., Xu, M., Xia, J.,Zhao, F., Meng, W. and Zhou, Q. (2010) MicroRNA145 targets BNIP3 andsuppresses prostate cancer progression. Cancer Res. 70, 2728–2738.

[121] Ren, D., Wang, M., Guo, W., Zhao, X., Tu, X., Huang, S., Zou, X. and Peng, X.(2013) Wild-type p53 suppresses the epithelial-mesenchymal transition andstemness in PC-3 prostate cancer cells by modulating miR145. Int. J. Oncol.42, 1473–1481.

[122] Hart, M., Wach, S., Nolte, E., Szczyrba, J., Menon, R., Taubert, H., Hartmann, A.,Stoehr, R., Wieland, W., Grasser, F.A. and Wullich, B. (2013) The proto-oncogene ERG is a target of microRNA miR-145 in prostate cancer. FEBS J.280, 2105–2116.

[123] Wang, F.E., Zhang, C., Maminishkis, A., Dong, L., Zhi, C., Li, R., Zhao, J.,Majerciak, V., Gaur, A.B., Chen, S. and Miller, S.S. (2010) MicroRNA-204/211alters epithelial physiology. FASEB J. 24, 1552–1571.

[124] Oberwinkler, J. (2007) TRPM3, a biophysical enigma? Biochem. Soc. Trans.35, 89–90.

[125] Conte, I., Carrella, S., Avellino, R., Karali, M., Marco-Ferreres, R., Bovolenta, P.and Banfi, S. (2010) MiR-204 is required for lens and retinal development viaMeis2 targeting. Proc. Natl. Acad. Sci. USA 107, 15491–15496.

[126] Shaham, O., Gueta, K., Mor, E., Oren-Giladi, P., Grinberg, D., Xie, Q., Cvekl, A.,Shomron, N., Davis, N., Keydar-Prizant, M., Raviv, S., Pasmanik-Chor, M., Bell,R.E., Levy, C., Avellino, R., Banfi, S., Conte, I. and Ashery-Padan, R. (2013) Pax6regulates gene expression in the vertebrate lens through miR-204. PLoSGenet. 9, e1003357.

[127] Sacconi, A., Biagioni, F., Canu, V., Mori, F., Di Benedetto, A., Lorenzon, L.,Ercolani, C., Di Agostino, S., Cambria, A.M., Germoni, S., Grasso, G., Blandino,R., Panebianco, V., Ziparo, V., Federici, O., Muti, P., Strano, S., Carboni, F.,Mottolese, M., Diodoro, M., Pescarmona, E., Garofalo, A. and Blandino, G.(2012) MiR-204 targets Bcl-2 expression and enhances responsiveness ofgastric cancer. Cell Death Dis. 3, e423.

[128] Staub, E., Grone, J., Mennerich, D., Ropcke, S., Klamann, I., Hinzmann, B.,Castanos-Velez, E., Mann, B., Pilarsky, C., Brummendorf, T., Weber, B., Buhr,H.J. and Rosenthal, A. (2006) A genome-wide map of aberrantly expressedchromosomal islands in colorectal cancer. Mol. Cancer 5, 37.

[129] Balachandar, V., Lakshman Kumar, B., Sasikala, K., Manikantan, P., Sangeetha,R. and Mohana Devi, S. (2007) Identification of a high frequency ofchromosomal rearrangements in the centromeric regions of prostatecancer patients. J. Zhejiang Univ. Sci. B 8, 638–646.

[130] Zhang, L., Wang, X. and Chen, P. (2013) MiR-204 down regulates SIRT1 andreverts SIRT1-induced epithelial-mesenchymal transition, anoikis resistanceand invasion in gastric cancer cells. BMC Cancer 13, 290.

[131] Lee, Y., Yang, X., Huang, Y., Fan, H., Zhang, Q., Wu, Y., Li, J., Hasina, R., Cheng,C., Lingen, M.W., Gerstein, M.B., Weichselbaum, R.R., Xing, H.R. and Lussier,Y.A. (2010) Network modeling identifies molecular functions targeted bymiR-204 to suppress head and neck tumor metastasis. PLoS Comput. Biol. 6,e1000730.

[132] Bao, W., Wang, H.H., Tian, F.J., He, X.Y., Qiu, M.T., Wang, J.Y., Zhang, H.J.,Wang, L.H. and Wan, X.P. (2013) A TrkB-STAT3-miR-204-5p regulatorycircuitry controls proliferation and invasion of endometrial carcinoma cells.Mol. Cancer 12, 155.

[133] Ying, Z., Li, Y., Wu, J., Zhu, X., Yang, Y., Tian, H., Li, W., Hu, B., Cheng, S.Y. and Li,M. (2013) Loss of miR-204 expression enhances glioma migration and stemcell-like phenotype. Cancer Res. 73, 990–999.

[134] Mikhaylova, O., Stratton, Y., Hall, D., Kellner, E., Ehmer, B., Drew, A.F., Gallo,C.A., Plas, D.R., Biesiada, J., Meller, J. and Czyzyk-Krzeska, M.F. (2012) VHL-regulated MiR-204 suppresses tumor growth through inhibition of LC3B-mediated autophagy in renal clear cell carcinoma. Cancer Cell 21, 532–546.

NAs: A novel non-coding alliance against cancer. FEBS Lett. (2014), http://


Recommended