Skip to content
BY-NC-ND 3.0 license Open Access Published by De Gruyter Open Access July 19, 2016

Roles of long noncoding RNAs in Hepatocellular Carcinoma

  • Diyu Huang , Jie Fang and Gaojian Luo EMAIL logo
From the journal Open Life Sciences

Abstract

Long noncoding RNAs (lncRNAs) are nonprotein coding transcripts longer than 200 nucleotides. Aberrant expression of lncRNAs has been found to be associated with hepatocellular carcinoma, one of the most malignant tumors. In this paper, we give a systematic and comprehensive review of existing literature about the involvement of lncRNAs in hepatocellular carcinoma. To date, evidence suggests that a number of lncRNAs, including HEIH, H19, HOTAIR, MALAT1, and PVT1, may regulate the transcription of target genes by recruiting histone-modifying complexes. Under certain circumstances, lncRNAs form RNA-dsDNA triplexes. Certain lncRNAs, such as HULC, HOTAIR, H19, HOTTIP and PTENP1, exhibit their biological roles by associating with microRNAs (miRNAs). In addition, by complementary base pairing with mRNAs or forming complexes with RNA binding proteins (RBPs), lncRNA-ATB, MALAT1 and PCNA-AS1 may mediate mRNA stability and splicing. In conclusion, interactions with DNA, RNA and proteins appears to be involved in lncRNAs’ participation in tumorigenesis and developmental processes related to hepatocellular carcinoma.

1 Introduction

Hepatocellular carcinoma (HCC) is one of the ten most commonly occurring solid cancers worldwide and is the second most prevalent cause of death from malignancy [1,2]. Hepatitis B Virus (HBV) or Hepatitis C Virus (HCV) infection, smoking, alcohol use and liver cirrhosis are the major causes of HCC [3,4]. Despite advances in the understanding of the molecular mechanisms underlying HCC and improved treatments for HCC, the median survival time is short. More than 60% of initially detected HCC in Japan are early stage with a 43% 5-year survival rate and a median survival time of 50 months [5].

In recent years, advances in genome-wide analyses of the mammalian transcriptome have revealed a novel class of transcripts, long noncoding RNAs (lncRNAs), which are pervasively transcribed in the genome [6]. LncRNAs are arbitrarily defined as transcripts of more than 200 nucleotides (nt) in length that lack significant open reading frames (ORF) and can be localized to both the nucleus and cytoplasm [7,8]. According to their location and orientation, lncRNAs are classified as intergenic lncRNAs (lincRNAs), genic and intragenic lncRNAs [9]. In the nucleus, lncRNAs mainly modulate gene transcription and mRNA splicing, whereas they are involved in RNA stability and activation of microRNAs (miRNAs) in the cytoplasm [10].

It is noteworthy that an increasing number of studies have demonstrated that lncRNAs act as a new class of regulatory molecules that are involved in the development and progression of hepatocellular carcinoma. Indeed, lncRNAs affect cell proliferation, migration, apoptosis, invasion, tumorigenicity, cell cycle, and metastasis.[11-13] Furthermore, evidence is accumulating to suggest that the aberrant expressions of lncRNAs have a clinical impact on the diagnosis of hepatocellular carcinoma. In this regard lncRNAs are associated with clinicopathological features including metastasis, invasion, TNM stage, prognosis, tumor size, and differentiation of patients with hepatocellular carcinoma. Among these features the greatest proportion are involved in metastasis and invasion. Hence, it has been proposed that these hepatocellular carcinoma-associated lncRNAs may be used as biomarkers for indicating metastasis of hepatocellular carcinoma [11,14-16].

For the purposes of this review, we present (Table 1) a comprehensive listing of reported dysregulated lncRNAs in HCC and the molecular mechanisms of known lncRNAs on hepatocellular carcinoma. Importantly, lncRNAs play crucial roles in hepatocellular carcinoma tumorigenesis and development by interacting with DNA, RNA, and proteins.

Table 1

Reported Roles for lncRNAs in the development and progression of HCC.

FunctionlncRNAsNumberPercentage (%)
ProliferationANRIL, CCAT1, GAS5, HEIH, HOTTIP, HOTAIR, HULC, H19, LALR1, Linc00974, LINC-ROR, MALAT1, MEG3, MT1DP, PCNA-AS1, PTENP1, PVT1, SRHC, TUC338, TUC339, URHC, UFC1, ZNRD1-AS12362.16
MigrationHOTTIP, HOTAIR, HULC, H19, MALAT1, PTENP1616.22
ApoptosisANRIL, CCAT1, GAS5, HOTAIR, HULC, H19, Linc00974, MEG3, MALAT1, MVIH, PRAL, PTENP1, UFC1, uc002mbe.2, VLDLR1540.54
InvasionATB, H19, HOTAIR, HULC, MALAT1, PTENP1616.22
TumorigenicityANRIL, DANCR, GAS5, H19, HOTAIR, HOTTIP, PVT1, PANDAR, RERT, TCF-71027.03
Cell cycleCCAT1, GAS5, HEIH, KCNQ1OT1, Linc00974, MEG3, MALAT1, PVT1, PCNA-AS1, UFC11027.03
MetastasisATB, DREH, H19, HOTTIP, LET, Linc00152, Linc00974, ZFAS1821.62
Total37100

2 Interaction with DNA

Evidence suggest that lncRNAs combine with histonemodifying complexes and then target DNA. For example, HOX transcript antisense RNA (HOTAIR) occupies targeted double-strained DNA (dsDNA) [17]. Additionally, long noncoding RNA MEG3 regulates the TGF-ß pathway genes through formation of RNA-DNA triplex structures [18]. These triplex structures can serve as a specific recognition mechanism between lncRNA and genomic DNA. Indeed, it has been proposed that triplexes created between lncRNA and genomic DNA may decisively result in targeting specificity, as well as promoting favorable chromatin conformations that may contribute to the affinity.

3 Interaction with RNA

3.1 Interaction with mRNA

Accumulating evidence indicates that lncRNAs can influence mRNA processing and post-transcriptional regulation via forming lncRNA-mRNA duplexes. These effects depend on complementary base pairing, and involve the control of splicing, translation, and mRNA stability. Certain HCC-related lncRNAs have been demonstrated to bind directly and target mRNA to exert post-transcriptional regulation. For example, PCNA-AS1, antisense to PCNA, can increase PCNA mRNA stability via forming lncRNA–mRNA hybridization in HCC [19] (Figure 1a). Additionally, Yuan et al. found that lncRNA-ATB specifically increased the stability of IL-11 mRNA, which depends on the binding of IL-11 mRNA, in HCC [20] (Figure 1d).

Figure 1 Associations between mRNAs and miRNAs, lncRNAs regulate cell proliferation, migration, apoptosis, invasion, tumorigenicity, cell cycle, and metastasis. a. The effects of PCNA-AS1 rely on regulation of PCNA via forming RNA hybridization to increase PCNA mRNA stability. b. H19 modulates let-7 availability by acting as a molecular sponge, affecting let-7-mediated regulation of metastasis-promoting genes, including IGF2BP1. c. miR-17 family target PTENP1, which regulate autophagy genes as ULK1, ATG7 and p62. d. lncRNA-ATB specially increased the stability of IL-11 mRNA, which depends on the binding of IL-11 mRNA in HCC. e. HOTTIP is a negative target of miR-125b. f. miR-9 could directly bind MALAT1 RNA in vivo and regulate the MALAT1 in the nucleus in an AGO2-dependent manner. g. HULC directly binds to miR-372 and represses its expression and activity. The reduction of miR-372 induces increased level of Prkacb, a target mRNA of miR-372. Abbreviations: PCNA, Proliferating Cell Nuclear Antigen; IGF2BP1, Insulin-like growth factor 2 mRNA-binding protein 1; Prkacb, cAMP-dependent protein kinase catalytic subunit beta; IL-11, interleukin 11; ULK1, Unccoordinated (unc)-51-like kinase 1; ATG7, autophagy-related gene 7.
Figure 1

Associations between mRNAs and miRNAs, lncRNAs regulate cell proliferation, migration, apoptosis, invasion, tumorigenicity, cell cycle, and metastasis. a. The effects of PCNA-AS1 rely on regulation of PCNA via forming RNA hybridization to increase PCNA mRNA stability. b. H19 modulates let-7 availability by acting as a molecular sponge, affecting let-7-mediated regulation of metastasis-promoting genes, including IGF2BP1. c. miR-17 family target PTENP1, which regulate autophagy genes as ULK1, ATG7 and p62. d. lncRNA-ATB specially increased the stability of IL-11 mRNA, which depends on the binding of IL-11 mRNA in HCC. e. HOTTIP is a negative target of miR-125b. f. miR-9 could directly bind MALAT1 RNA in vivo and regulate the MALAT1 in the nucleus in an AGO2-dependent manner. g. HULC directly binds to miR-372 and represses its expression and activity. The reduction of miR-372 induces increased level of Prkacb, a target mRNA of miR-372. Abbreviations: PCNA, Proliferating Cell Nuclear Antigen; IGF2BP1, Insulin-like growth factor 2 mRNA-binding protein 1; Prkacb, cAMP-dependent protein kinase catalytic subunit beta; IL-11, interleukin 11; ULK1, Unccoordinated (unc)-51-like kinase 1; ATG7, autophagy-related gene 7.

Irrespective of the reported direct interactions, an indirect mechanism between lncRNAs and their targeted mRNAs is illustrated by metastasis-associated lung adenocarcinoma transcript 1 (MALAT1) [21,22]. By binding with associated proteins, MALAT1 alters splicing of mRNAs. The function of serine/arginine splicing factors (SF2/ASF) rely on MALAT1, and through the formation of a bipartite triple helix, MALAT1 promotes malignancy. A fragment near the 3’ end of MALAT1 may regulate the metastatic potential.

3.2 Interaction with miRNAs

It has been proposed that lncRNAs can function as competing endogenous RNAs (ceRNAs) and ‘sponges’ miRNAs, thus regulating the expression of target mRNAs. MicroRNAs regulate the expression of over 60% of protein coding genes by targeting their mRNAs to AGO2-containing complexes in the cytoplasm and promoting their translational inhibition and/or degradation [23]. For example, clinical studies have demonstrated a link of miR-182 expression to poor prognosis in liver cancer patients. Mechanistically, multiple downstream genes including missing-in-metastasis, microphthalm-associated transcription factor, FoxO1, cylindromatiosis, and others, can be targeted by miR-182 and mediate its roles in a variety of cancers [24].

A typical example is HULC, a highly up-regulated lncRNA in liver cancer, transcribed from human chromosome 6p24.3 and containing a conserved target site of miR-372. Phospho-CREB, activated through the protein kinase A pathway, binds to the CREB binding site and induces upregulation of HULC. Subsequently, HULC directly binds to miR-372 and represses its expression and activity. The reduction of miR-372 induces increased levels of Prkacb (cAMP-dependent protein kinase catalytic subunit beta), which is a target mRNA of miR-372. Interestingly, Prkacb can facilitate phosphorylation of CERB in the protein kinase A pathway. Finally, an autoregulatory loop (CREB-HULC-miR372-Prkacb) is formed [25,26] (Figure 1g). Metastasis-associated lung adenocarcinoma transcript 1 (MALAT1) is also known as nuclear-enriched abundant transcript 2 (NEAT2) and is transcribed from chromosome 11q13. It is upregulated in many solid carcinomas and is associated with cell proliferation and migration through the modulation of caspase-3, caspase-8, Bax and Bcl-2. Leucci et al. demonstrated that miR-9 could directly bind MALAT1 RNA in vivo and regulate the MALAT1 in the nucleus in an AGO2-dependent manner [27] (Figure 1f). In addition, Takahashi et al. revealed that linc-RoR functioned as a miRNA sponge to limit endogenous miR-145 that can modulate the expression of key effectors of the hypoxia response, such as HIF-1a expression, in HCC [28]. In addition, lncRNA-ATB up-regulated ZEB1 and ZEB2 by competitively binding the miR-200 family and then induced EMT and the invasiveness of HCC [20]. H19 harbors both canonical and non-canonical binding sites for the let-7 family of microRNAs, which plays key roles in development and cancer. It has been reported that H19 modulates let-7 availability by acting as a molecular sponge using H19 knockdown and overexpression, as well as in vivo crosslinking and genome-wide transcriptome analysis [29] (Figure 1b). These suggest that blocking the associations between lncRNAs and their partners (RNAs or proteins) may play key roles in development and cancer.

Further, the expression of lncRNAs may be repressed by miRNAs. The lncRNA PTENP1 is a pseudogene of the tumor suppressor gene PTEN, which suppresses the oncogenic PI3K/AKT pathway, inhibits cell proliferation and migration/invasion, as well as inducing autophagy and apoptosis. The PTENP1 decoyed oncomirs miR-17 family, which would otherwise target PTEN, PHLPP (a negative AKT regulator) and such autophagy genes as ULK1, ATG7 and p62 [30] (Figure 1c). HOTTIP is an antisense lncRNA mapped to the distal end of the HOXA gene cluster. Knockdown of HOTTIP significantly suppressed the expression of a number of HOXA genes Moreover, in human HCCs, HOTTIP expression negatively correlated with that of miR-125b, which is a post-transcriptional regulator of HOTTIP [31] (Figure 1e). miR-141 and miR-148a show an inhibitory effect on the expression of DNA methyltransferase 1 (DNMT-1) and thus induces the overexpression of MEG3 [32] (Figure 2c). The examples about miRNAs’ regulating lncRNAs via epigenetic modification, and vice versa, can also be found in other parts of this article. In conclusion, lncRNAs can target multiple miRNAs, also can be targeted by miRNAs re-establishing the expression of a single lncRNA, then forming a feedback loop. A better understanding of the mechanism of regulation of ncRNAs network would accelerate the development of novel therapeutic strategies for the arrest of primary hepatic tumors.

Figure 2 LncRNAs regulate cell proliferation, migration, apoptosis, invasion, tumorigenicity, cell cycle, and metastasis by histone modification. a. The P15/P16 is silenced by the occupancy of EZH2, which is recruited by PVT1. b. Through the formation of dsDNA/RNA triplex, HOTAIR recruits SUZ12 and EZH2 to the MiR-34A loci and then silences the transcription of miR-34a. HOTAIR also forms complexes with LSD1. HOTAIR attenuates Snail, N-cadherin and vimentin protein levels. All of these are targets of miR-34a. c. DNMT1-mediated MEG3 hypermethylation caused the loss of MEG3 expression. d. HOTAIR could promote migration and invasion of HCC cells by inhibiting RBM38. e. H19 guides EZH2 to the E-cadherin promoter loci. f. lncRNA-HEIH is found to interact with enhancer of zeste homolog 2 (EZH2), a key component of PRC2, then recruit PRC2 to p16 promoter and repress the expression of p16 gene, thereby contributing to cell cycle arrest. g. H19 associated with the protein complex hnRNP U/PCAF/RNAPol II, activating miR-200 family by increasing histone acetylation, thus contributing to the suppression of tumor metastasis. Abbreviations: RBM38, RNA binding motif protein 38; ZEB1/2, zinc finger E-box-bindingprotein1/2; LSD1, lysine-specific demethylase 1; PRC2, polycomb repressive complex 2; EZH2, enhancer of zeste homologue 2.
Figure 2

LncRNAs regulate cell proliferation, migration, apoptosis, invasion, tumorigenicity, cell cycle, and metastasis by histone modification. a. The P15/P16 is silenced by the occupancy of EZH2, which is recruited by PVT1. b. Through the formation of dsDNA/RNA triplex, HOTAIR recruits SUZ12 and EZH2 to the MiR-34A loci and then silences the transcription of miR-34a. HOTAIR also forms complexes with LSD1. HOTAIR attenuates Snail, N-cadherin and vimentin protein levels. All of these are targets of miR-34a. c. DNMT1-mediated MEG3 hypermethylation caused the loss of MEG3 expression. d. HOTAIR could promote migration and invasion of HCC cells by inhibiting RBM38. e. H19 guides EZH2 to the E-cadherin promoter loci. f. lncRNA-HEIH is found to interact with enhancer of zeste homolog 2 (EZH2), a key component of PRC2, then recruit PRC2 to p16 promoter and repress the expression of p16 gene, thereby contributing to cell cycle arrest. g. H19 associated with the protein complex hnRNP U/PCAF/RNAPol II, activating miR-200 family by increasing histone acetylation, thus contributing to the suppression of tumor metastasis. Abbreviations: RBM38, RNA binding motif protein 38; ZEB1/2, zinc finger E-box-bindingprotein1/2; LSD1, lysine-specific demethylase 1; PRC2, polycomb repressive complex 2; EZH2, enhancer of zeste homologue 2.

4 Interaction with proteins

4.1 Interaction with histone-modifying complexes

A large number of studies have revealed that many HCC-related lncRNAs exert their function through interaction with proteins or protein complexes, especially with epigenetic complexes, such as polycomb repressive complex 1 (PRC1) and polycomb repressive complex 2 (PRC2).

HOTAIR was the first demonstrated lncRNA shown to coordinate gene silencing via assembly of PRC2. The structural domains of HOTAIR formed in its 5’ region and 3’ region are bound to enhancer of zeste 2 (EZH2, PRC2 subunit) and lysine specific demethylase 1 (LSD1), respectively [33]. HOTAIR preferentially occupies a GA-rich DNA motif to enable the formation of a RNA-dsDNA triplex [17]. This occurs independently of EZH2. Simultaneously, HOTAIR is required for the occupancy of suppressor of zeste 12 homolog (SUZ12, PRC2 subunit) on MIR-34A loci [34], then epigenetically modify their targeted genes, including Snail and vimentin (Figure 2b). HOTAIR also could promote migration and invasion of HCC cells by inhibiting RBM38 [35] (Figure 2d). H19 can specifically associate with enhancer of zeste homolog 2 (EZH2), a key subunit of the PRC2 complex, and inhibit E-cad expression by directly suppressing E-cad transcription and by indirectly activating Wnt signaling [36] (Figure 2e); H19 associated with the protein complex hnRNP U/ PCAF/RNAPol II, activating miR-200 family by increasing histone acetylation. These results demonstrate that H19 can alter the miR-200 pathway [37], thus contributing to mesenchymal-to-epithelial transition and to the suppression of tumor metastasis] (Figure 2g). KCNQ1OT1 could interact with chromatin and with the H3K9- and H3K27-specific histone methyltransferases G9a and the PRC2 complex in a lineage-specific manner [38]. LncRNA-HEIH also can associate with EZH2, and this association is required for the repression of EZH2 target genes in HCC, involving p15, p16, p21 and p57 [39] (Figure 2f). Kaneko et al. demonstrated that MEG3 interacted with PRC2 mainly through the RBR of JARID2 and MEG3 acts in trans on PRC2 and JARID2 by facilitating their recruitment to a subset of target genes [40]. PVT1 silenced P15/P16 in in trans by recruiting EZH2 [41]. The upregulated EZH2 enforces hepatocellular carcinoma cell proliferation (Figure 2a). LncRNAs also recruit activating chromatin modifiers such as lysine (K)-specific methyltransferase 2A (MLL). LncRNA HOTTIP could bind the adaptor protein WDR5 directly and targets WDR5/MLL complexes across HOXA, driving H3K4 trimethylation and gene transcription [42]. LncRNA-LET could bind to NF90, a double-stranded RNA-binding protein that has been implicated in the stabilization, transport, and translational control of many target mRNAs, and decreases HIF1-a and CDC42 mRNA stability through its association with NF90 under hypoxic and normoxic conditions, respectively [43]. In these examples, lncRNAs are required for the assembly of different histonemodifying complexes or/and DNA methyltransferase, which in turn modify the neighboring chromatin. As many as 38% lincRNAs cooperate with at least one of multiple histone-modifying complexes. This suggests that one lncRNA may harbor several types of binding elements for chromatin modifiers. The specific locus is subject to the ensuing histone modification caused by the occupancy of histone-modifying complexes.

4.2 Interaction with other proteins

With the exception of the direct interaction with proteins, how do lncRNAs interact with other proteins is interesting. Evidence suggests that lncRNA-MVIH can activate angiogenesis by interacting with PGK1, a protein secreted by tumor cells which inhibits angiogenesis, and preventing its secretion [44]. LncRNA-DREH could specifically associate with the protein vimentin, a type III intermediate filament (IF) and the major cytoskeletal component of mesenchymal cells [13]. GAS5 positively influences YBX1 protein stability without increasing its transcription [45]. The putative stem-loop structure formed by exon 12 of GAS5 is responsible for its interplay with YBX1. YBX1 possesses the capacity of complexing with IGF2BP1, which combines with GHET1 and prevents mRNA degradation. The exon 12 is a GAS5’s predominant structural domain for mimicking binding domain of glucocorticoid receptor (GR). GAS5, GR, YBX1, and P53 may collaborate as a complex to influence cell cycle regulation.

5 Conclusion

LncRNAs are characterized by the complexity of their mechanisms. This review summarizes the potential impact of dysregulated lncRNAs in HCC and the known interactions between lncRNAs and DNA, RNA, and proteins in hepatocellular carcinoma. Sun et al. [46] proposed a global network-based computational framework to infer potential human lncRNA–disease associations by implementing the random walk with restart method on a lncRNA functional similarity network. In total, they observed 371 lncRNA–lncRNA functional associations between 117 lncRNAs in the lncRNA functional similarity network (LFSN).

Molecular-based tumor predictors are essential for individualized HCC diagnosis. Cancer-specific miRNAs have been widely regarded as potential biomarkers for the diagnosis and prognosis of cancers as they are easily detectable in the blood, urine, sputum and other biological fluids of patients. Similarly, lncRNAs that are found in body fluids have demonstrated a utility to be used as fluid-based non-invasive markers for clinical use. Therapeutic benefit can be obtained through RNA-based therapeutic strategies, such as siRNA and microRNA, or using small molecule compounds designed specifically to interact with target lncRNAs or ribonucleoprotein complexes [47]. It is tempting to speculate that a multitude of lncRNAs may interrupt specific steps in numerous tumor suppressive and oncogenic pathways. The revelation of the underlying mechanisms of lncRNAs may benefit our understanding of hepatocellular carcinoma’s pathogenesis.

Conflict of Interests

The authors declare that there is no conflict of interests regarding the publication of this paper.

References

[1] Ferlay J., Soerjomataram I., Dikshit R., Eser S., Mathers C., Rebelo M., et al., Cancer incidence and mortality worldwide: sources, methods and major patterns in GLOBOCAN 2012, Int J. Cancer, 2015, 136, E359-386.10.1002/ijc.29210Search in Google Scholar PubMed

[2] Wong L. L., Tsai N., Hepatocellular carcinoma in Filipinos, Asia-Pacific J. Clin. Oncol., 2007, 3, 84-88.10.1111/j.1743-7563.2007.00091.xSearch in Google Scholar

[3] El-Serag H. B., Rudolph K. L., Hepatocellular carcinoma: epidemiology and molecular carcinogenesis, Gastroenterology, 2007, 132, 2557-2576.10.1053/j.gastro.2007.04.061Search in Google Scholar PubMed

[4] J C. K., M S. P., H P. S., Prevalence and incidence of hepatitis C virus infection among Aboriginal young people who use drugs: results from the Cedar Project, Centr. Eur. J. Med., 2009, 3, e220-e227.Search in Google Scholar

[5] Gores G. J., Lieberman D., Good News-Bad News: Current Status of GI Cancers, Gastroenterology, (in press), 10.1053/j.gastro.2016.05.007.Search in Google Scholar PubMed

[6] Carninci P., Kasukawa T., Katayama S., Gough J., Frith M. C., Maeda N., et al., The transcriptional landscape of the mammalian genome, Science, 2005, 309, 1559-1563.10.1126/science.1112014Search in Google Scholar PubMed

[7] Kapranov P., Cheng J., Dike S., Nix D. A., Duttagupta R., Willingham A. T., et al., RNA maps reveal new RNA classes and a possible function for pervasive transcription, Science, 2007, 316, 1484-1488.10.1126/science.1138341Search in Google Scholar PubMed

[8] van Heesch S., van Iterson M., Jacobi J., Boymans S., Essers P. B., de Bruijn E., et al., Extensive localization of long noncoding RNAs to the cytosol and mono- and polyribosomal complexes, Genome Biol., 2014, 15, 1-12.10.1186/gb-2014-15-1-r6Search in Google Scholar PubMed PubMed Central

[9] Di Gesualdo F., Capaccioli S., Lulli M., A pathophysiological view of the long non-coding RNA world, Oncotarget, 2014, 5, 10976-10996.10.18632/oncotarget.2770Search in Google Scholar PubMed PubMed Central

[10] Qi P., Du X., The long non-coding RNAs, a new cancer diagnostic and therapeutic gold mine, Mod. Pathol. 2013, 26, 155-165.10.1038/modpathol.2012.160Search in Google Scholar PubMed

[11] Dickson I., Hepatocellular carcinoma: A role for lncRNA in liver cancer, Nat. Rev. Gastroenterol. Hepatol., 2016, 13, 122-123.10.1038/nrgastro.2016.21Search in Google Scholar PubMed

[12] Tao R., Hu S., Wang S., Zhou X., Zhang Q., Wang C., et al., Association between indel polymorphism in the promoter region of lncRNA GAS5 and the risk of hepatocellular carcinoma, Carcinogenesis, 2015, 36, 1136-1143.10.1093/carcin/bgv099Search in Google Scholar PubMed

[13] Huang J. F., Guo Y. J., Zhao C. X., Yuan S. X., Wang Y., Tang G. N., et al., Hepatitis B virus X protein (HBx)-related long noncoding RNA (lncRNA) down-regulated expression by HBx (Dreh) inhibits hepatocellular carcinoma metastasis by targeting the intermediate filament protein vimentin, Hepatology, 2013, 57, 1882-1892.10.1002/hep.26195Search in Google Scholar PubMed

[14] Zhu Z., Gao X., He Y., Zhao H., Yu Q., Jiang D., et al., An insertion/deletion polymorphism within RERT-lncRNA modulates hepatocellular carcinoma risk, Cancer Res., 2012, 72, 6163-6172.10.1158/0008-5472.CAN-12-0010Search in Google Scholar PubMed

[15] Zhang D., Cao C., Liu L., Wu D., Up-regulation of LncRNA SNHG20 Predicts Poor Prognosis in Hepatocellular Carcinoma, J. Cancer, 2016, 7, 608-617.10.7150/jca.13822Search in Google Scholar PubMed PubMed Central

[16] Li S. P., Xu H. X., Yu Y., He J. D., Wang Z., Xu Y. J., et al., LncRNA HULC enhances epithelial-mesenchymal transition to promote tumorigenesis and metastasis of hepatocellular carcinoma via the miR-200a-3p/ZEB1 signaling pathway, Oncotarget, (in press), 10.18632/oncotarget.9883.Search in Google Scholar PubMed PubMed Central

[17] Chu C., Qu K., Zhong F. L., Artandi S. E., Chang H. Y., Genomic maps of long noncoding RNA occupancy reveal principles of RNA-chromatin interactions, Mol. Cell., 2011, 44, 667-678.10.1016/j.molcel.2011.08.027Search in Google Scholar PubMed PubMed Central

[18] Mondal T., Subhash S., Vaid R., Enroth S., Uday S., Reinius B., et al., MEG3 long noncoding RNA regulates the TGF-beta pathway genes through formation of RNA-DNA triplex structures, Nat. Commun., 2015, 6, 7743-7743.10.1038/ncomms8743Search in Google Scholar PubMed PubMed Central

[19] Yuan S. X., Tao Q. F., Wang J., Yang F., Liu L., Wang L. L., et al., Antisense long non-coding RNA PCNA-AS1 promotes tumor growth by regulating proliferating cell nuclear antigen in hepatocellular carcinoma, Cancer Lett., 2014, 349, 87-94.10.1016/j.canlet.2014.03.029Search in Google Scholar PubMed

[20] Yuan J. H., Yang F., Wang F., Ma J. Z., Guo Y. J., Tao Q. F., et al., A long noncoding RNA activated by TGF-beta promotes the invasion-metastasis cascade in hepatocellular carcinoma, Cancer Cell., 2014, 25, 666-681.10.1016/j.ccr.2014.03.010Search in Google Scholar PubMed

[21] Lai M. C., Yang Z., Zhou L., Zhu Q. Q., Xie H. Y., Zhang F., et al., Long non-coding RNA MALAT-1 overexpression predicts tumor recurrence of hepatocellular carcinoma after liver transplantation, Med. Oncol., 2012, 29, 1810-1816.10.1007/s12032-011-0004-zSearch in Google Scholar PubMed

[22] Tripathi V., Ellis J. D., Shen Z., Song D. Y., Pan Q., Watt A. T., et al., The nuclear-retained noncoding RNA MALAT1 regulates alternative splicing by modulating SR splicing factor phosphorylation, Mol. Cell., 2010, 39, 925-938.10.1016/j.molcel.2010.08.011Search in Google Scholar PubMed PubMed Central

[23] Kumar R. and Xi Y., MicroRNA, epigenetic machinery and lung cancer, Thoracic Cancer, 2011, 2, 35-44.10.1111/j.1759-7714.2011.00043.xSearch in Google Scholar PubMed

[24] Wei Q., Lei R. and Hu G., Roles of miR-182 in sensory organ development and cancer, Thoracic Cancer, 2015, 6, 2-9.10.1111/1759-7714.12164Search in Google Scholar PubMed PubMed Central

[25] Wang J., Liu X., Wu H., Ni P., Gu Z., Qiao Y., et al., CREB up-regulates long non-coding RNA, HULC expression through interaction with microRNA-372 in liver cancer, Nucleic Acids Res., 2010, 38, 5366-5383.10.1093/nar/gkq285Search in Google Scholar PubMed PubMed Central

[26] Panzitt K., Tschernatsch M. M., Guelly C., Moustafa T., Stradner M., Strohmaier H. M., et al., Characterization of HULC, a novel gene with striking up-regulation in hepatocellular carcinoma, as noncoding RNA, Gastroenterology, 2007, 132, 330-342.10.1053/j.gastro.2006.08.026Search in Google Scholar PubMed

[27] Leucci E., Patella F., Waage J., Holmstrom K., Lindow M., Porse B., et al., microRNA-9 targets the long non-coding RNA MALAT1 for degradation in the nucleus, Sci. Rep., 2013, 3, 2535-2535.10.1038/srep02535Search in Google Scholar PubMed PubMed Central

[28] Takahashi Y., Nishikawa M., Takakura Y., Suppression of tumor growth by intratumoral injection of short hairpin RNA-expressing plasmid DNA targeting beta-catenin or hypoxia-inducible factor 1alpha, J. Control Release, 2006, 116, 90-95.10.1016/j.jconrel.2006.09.002Search in Google Scholar PubMed

[29] Kallen A. N., Zhou X. B., Xu J., Qiao C., Ma J., Yan L., et al., The imprinted H19 lncRNA antagonizes let-7 microRNAs, Mol. Cell., 2013, 52, 101-112.10.1016/j.molcel.2013.08.027Search in Google Scholar PubMed PubMed Central

[30] Chen C. L., Tseng Y. W., Wu J. C., Chen G. Y., Lin K. C., Hwang S. M., et al., Suppression of hepatocellular carcinoma by baculovirus-mediated expression of long non-coding RNA PTENP1 and MicroRNA regulation, Biomaterials, 2015, 44, 71-81.10.1016/j.biomaterials.2014.12.023Search in Google Scholar PubMed

[31] Tsang F. H., Au S. L., Wei L., Fan D. N., Lee J. M., Wong C. C., et al., Long non-coding RNA HOTTIP is frequently up-regulated in hepatocellular carcinoma and is targeted by tumour suppressive miR-125b, Liver Int., 2015, 35, 1597-1606.10.1111/liv.12746Search in Google Scholar PubMed

[32] Yuan K., Lian Z., Sun B., Clayton M. M., Ng I. O., Feitelson M. A., Role of miR-148a in hepatitis B associated hepatocellular carcinoma, PLoS One, 2012, 7, e35331-e35331.10.1371/journal.pone.0035331Search in Google Scholar PubMed PubMed Central

[33] Yoon J. H., Abdelmohsen K., Kim J., Yang X., Martindale J. L., Tominaga-Yamanaka K., et al., Scaffold function of long non-coding RNA HOTAIR in protein ubiquitination, Nat. Commun., 2013, 4, 2939-2939.10.1038/ncomms3939Search in Google Scholar PubMed PubMed Central

[34] Yang Z., Zhou L., Wu L. M., Lai M. C., Xie H. Y., Zhang F., et al., Overexpression of long non-coding RNA HOTAIR predicts tumor recurrence in hepatocellular carcinoma patients following liver transplantation, Ann. Surg. Oncol., 2011, 18, 1243-1250.10.1245/s10434-011-1581-ySearch in Google Scholar PubMed

[35] Ding C., Cheng S., Yang Z., Lv Z., Xiao H., Du C., et al., Long non-coding RNA HOTAIR promotes cell migration and invasion via down-regulation of RNA binding motif protein 38 in hepatocellular carcinoma cells, Int. J. Mol. Sci., 2014, 15, 4060-4076.10.3390/ijms15034060Search in Google Scholar PubMed PubMed Central

[36] Luo M., Li Z., Wang W., Zeng Y., Liu Z., Qiu J., Long non-coding RNA H19 increases bladder cancer metastasis by associating with EZH2 and inhibiting E-cadherin expression, Cancer Lett., 2013, 333, 213-221.10.1016/j.canlet.2013.01.033Search in Google Scholar PubMed

[37] Zhang L., Yang F., Yuan J. H., Yuan S. X., Zhou W. P., Huo X. S., et al., Epigenetic activation of the MiR-200 family contributes to H19-mediated metastasis suppression in hepatocellular carcinoma, Carcinogenesis, 2013, 34, 577-586.10.1093/carcin/bgs381Search in Google Scholar PubMed

[38] Wan J., Huang M., Zhao H., Wang C., Zhao X., Jiang X., et al., A novel tetranucleotide repeat polymorphism within KCNQ1OT1 confers risk for hepatocellular carcinoma, DNA Cell Biol., 2013, 32, 628-634.10.1089/dna.2013.2118Search in Google Scholar PubMed

[39] Yang F., Zhang L., Huo X. S., Yuan J. H., Xu D., Yuan S. X., et al., Long noncoding RNA high expression in hepatocellular carcinoma facilitates tumor growth through enhancer of zeste homolog 2 in humans, Hepatology, 2011, 54, 1679-1689.10.1002/hep.24563Search in Google Scholar PubMed

[40] Kaneko S., Bonasio R., Saldana-Meyer R., Yoshida T., Son J., Nishino K., et al., Interactions between JARID2 and noncoding RNAs regulate PRC2 recruitment to chromatin, Mol. Cell., 2014, 53, 290-300.10.1016/j.molcel.2013.11.012Search in Google Scholar PubMed PubMed Central

[41] Ding C., Yang Z., Lv Z., Du C., Xiao H., Peng C., et al., Long non-coding RNA PVT1 is associated with tumor progression and predicts recurrence in hepatocellular carcinoma patients, Oncol. Lett., 2015, 9, 955-963.10.3892/ol.2014.2730Search in Google Scholar PubMed PubMed Central

[42] Quagliata L., Matter M. S., Piscuoglio S., Arabi L., Ruiz C., Procino A., et al., Long noncoding RNA HOTTIP/HOXA13 expression is associated with disease progression and predicts outcome in hepatocellular carcinoma patients, Hepatology, 2014, 59, 911-923.10.1002/hep.26740Search in Google Scholar PubMed PubMed Central

[43] Yang F., Huo X. S., Yuan S. X., Zhang L., Zhou W. P., Wang F.., et al.,. Repression of the long noncoding RNA-LET by histone deacetylase 3 contributes to hypoxia-mediated metastasis, Mol. Cell., 2013, 49, 1083-109610.1016/j.molcel.2013.01.010Search in Google Scholar PubMed

[44] Yuan S. X., Yang F., Yang Y., Tao Q. F., Zhang J., Huang G.., et al., Long noncoding RNA associated with microvascular invasion in hepatocellular carcinoma promotes angiogenesis and serves as a predictor for hepatocellular carcinoma patients’ poor recurrence-free survival after hepatectomy, Hepatology, 2012, 56, 2231-2241.10.1002/hep.25895Search in Google Scholar PubMed

[45] Chang L., Li C., Lan T., Wu L., Yuan Y., Liu Q., et al., Decreased expression of long non-coding RNA GAS5 indicates a poor prognosis and promotes cell proliferation and invasion in hepatocellular carcinoma by regulating vimentin, Mol. Med. Rep., 2016, 13, 1541-1550.10.3892/mmr.2015.4716Search in Google Scholar PubMed PubMed Central

[46] Sun J., Shi H., Wang Z., Zhang C., Liu L., Wang L., et al., Inferring novel lncRNA-disease associations based on a random walk model of a lncRNA functional similarity network, Mol. Biosyst., 2014, 10, 2074-2081.10.1039/C3MB70608GSearch in Google Scholar

[47] Lesko E., Miekus K., Grabowska A., Gładys A., Majka M., Optimization of a synthetic siRNA delivery for the treatment of rhabdomyosarcoma, Centr. Eur. J. Biol., 2008, 3, 371-379.10.2478/s11535-008-0042-5Search in Google Scholar

Received: 2016-4-28
Accepted: 2016-6-3
Published Online: 2016-7-19
Published in Print: 2016-1-1

© 2016 Diyu Huang et al., published by De Gruyter Open

This work is licensed under the Creative Commons Attribution-NonCommercial-NoDerivatives 3.0 License.

Downloaded on 24.4.2024 from https://www.degruyter.com/document/doi/10.1515/biol-2016-0012/html
Scroll to top button