Next Article in Journal
Role of Genetic Variations in the Hepatic Handling of Drugs
Next Article in Special Issue
Pathogenic Roles of Autoantibodies and Aberrant Epigenetic Regulation of Immune and Connective Tissue Cells in the Tissue Fibrosis of Patients with Systemic Sclerosis
Previous Article in Journal
C5-Substituted 2-Selenouridines Ensure Efficient Base Pairing with Guanosine; Consequences for Reading the NNG-3′ Synonymous mRNA Codons
Previous Article in Special Issue
Potential Roles of Long Noncoding RNAs as Therapeutic Targets in Renal Fibrosis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

LncRNAs Act as a Link between Chronic Liver Disease and Hepatocellular Carcinoma

Department of Pathology, Catholic University of Daegu School of Medicine, Daegu 42472, Korea
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2020, 21(8), 2883; https://doi.org/10.3390/ijms21082883
Submission received: 1 April 2020 / Revised: 16 April 2020 / Accepted: 18 April 2020 / Published: 20 April 2020
(This article belongs to the Special Issue Fibrosis-Related lncRNA)

Abstract

:
Long non-coding RNAs (lncRNAs) are emerging as important contributors to the biological processes underlying the pathophysiology of various human diseases, including hepatocellular carcinoma (HCC). However, the involvement of these molecules in chronic liver diseases, such as nonalcoholic fatty liver disease (NAFLD) and viral hepatitis, has only recently been considered in scientific research. While extensive studies on the pathogenesis of the development of HCC from hepatic fibrosis have been conducted, their regulatory molecular mechanisms are still only partially understood. The underlying mechanisms related to lncRNAs leading to HCC from chronic liver diseases and cirrhosis have not yet been entirely elucidated. Therefore, elucidating the functional roles of lncRNAs in chronic liver disease and HCC can contribute to a better understanding of the molecular mechanisms, and may help in developing novel diagnostic biomarkers and therapeutic targets for HCC, as well as in preventing the progression of chronic liver disease to HCC. Here, we comprehensively review and briefly summarize some lncRNAs that participate in both hepatic fibrosis and HCC.

1. Introduction

Long non-coding RNAs (lncRNAs) are a subgroup of non-coding RNA (ncRNAs) transcripts, which are greater than 200 nucleotides in length, with little or no protein-coding potential [1,2]. While lncRNAs are emerging as important contributors to the biological processes underlying the pathophysiology of various human diseases, including inflammation and neoplasia [3], the involvement of these molecules in chronic liver diseases, such as viral hepatitis and nonalcoholic fatty liver disease (NAFLD), has only been considered in scientific research [3].
Hepatocellular carcinoma (HCC) is one of the most common malignant tumors, with a poor prognosis and low survival rates [4,5]. Despite advances in the understanding of the molecular mechanisms of HCC, the overall low survival time, high rates of metastasis, postsurgical recurrence, and chemoresistance are still unsolved problems [5,6,7,8]. Hepatic fibrosis is a major risk factor for HCC, and it is a continuous wound-healing process that leads to the dysregulation of extracellular matrix (ECM) proteins and the distortion of normal liver architecture [9]. Many chronic liver diseases, such as viral hepatitis, alcohol toxicity, drug abuse, metabolic syndrome, and autoimmune hepatitis result in hepatic fibrosis or cirrhosis [10], which is the primary stage of HCC [11]. While extensive studies on the pathogenesis of the development of HCC from hepatic fibrosis have been performed, their regulatory molecular mechanisms are still only partially understood [11].
The recent application of next-generation sequencing (NGS), particularly RNA-sequencing (RNA-Seq), has improved our knowledge of lncRNAs related to various types of diseases [2]. With the application of NGS and high-resolution microarray techniques, many studies suggest that HCC-related lncRNAs could influence the initiation, progression and suppression of HCC [5]. However, the underlying mechanisms related to lncRNAs leading to HCC from chronic liver diseases and cirrhosis have not been entirely elucidated yet [12]. Therefore, elucidating the functional roles of lncRNAs in chronic liver disease and HCC can contribute to a better understanding of the molecular mechanisms and may help in developing novel diagnostic biomarkers and therapeutic targets for HCC as well as in preventing the progression of HCC from chronic liver disease [2]. In this review, some well-documented lncRNAs that participate in both hepatic fibrosis and HCC are comprehensively summarized.

2. The Classifications and Functions of lncRNAs

Numerous ncRNA molecules have been identified by RNA microarrays and the NGS of transcriptomes [13]. NcRNAs are important regulators involved in most cellular functions and modulation [14]. There have been three kinds of ncRNA classifications so far in the literature. First, they are classified into three types based on their relative sizes [11,15]. There are long ncRNAs (lncRNAs with more than 200 nucleotides), small ncRNAs (with less than 200 nucleotides) and microRNAs (miRNAs with 18–22 nucleotides) [15]. Second, lncRNAs are classified into five classes based on their genomic location relative to neighboring protein-coding genes: (1) sense lncRNAs overlap with the sense strand of a protein-coding gene; (2) antisense lncRNAs overlap one or more exons of a protein-coding gene on the opposite strand and initiate 3′ of a protein-coding gene; (3) bidirectional lncRNAs, in which the expression of a lncRNA and a protein-coding gene on the opposite strand are initiated <1000 base pairs away in close genomic proximity; (4) intronic lncRNAs are initiated completely within an intron of a protein-coding gene, without overlapping exons; and (5) intergenic lncRNAs (also termed large intervening ncRNAs or lincRNAs), are located near no other protein-coding loci [16,17,18]. Third, lncRNAs can also be classified according to their targeting mechanisms: signal, decoy, guide, and scaffold [3]. Increasing evidence has revealed that lncRNAs can interact with miRNAs. Indeed, lncRNAs can play a role of miRNA sponges (the so-called competing endogenous RNAs, ceRNAs), reducing their regulatory effect. In turn, miRNAs may directly interact with lncRNAs and silence their expression [19,20].
Many lncRNAs could not be easily classified into any particular category, and it is likely that the same lncRNAs may be listed in different groups in all classifications [21,22]. To date, the biological functions and molecular mechanisms of most lncRNAs remain largely unclear, with only a few being partially characterized [2]. Nevertheless, existing evidence demonstrates that these molecules play critical roles in regulating some cellular processes, specifically in the expression of protein-coding genes at the epigenetic, transcriptional and post-transcriptional levels [23,24,25]. Epigenetic alterations include changes in DNA methylation and histone modifications, as well as ncRNA-mediated gene silencing [26].
LncRNAs are considered to play regulatory roles in the pathogenesis and progression of various human diseases, including cell proliferation, differentiation, apoptosis, and tumorigenesis [27,28,29]. Many lncRNAs are frequently aberrantly expressed in human cancers in which they may act as oncogenes or tumor suppressors, indicating that they may serve as novel drivers of tumorigenesis [2,30]. Compared with protein-coding genes, the alterations of lncRNAs are highly tumor- and cell line-specific [31], and this characteristic of specificity makes lncRNAs promising diagnostic biomarkers [2].
Some studies suggested that HCC-related lncRNAs play critical regulatory roles in the development and progression of HCC, while their dysregulation is associated with diverse biological processes including proliferation, differentiation, apoptosis, invasion, and metastasis [22,32]. LncRNAs may also play a role in chemo-sensitivity or radio-resistance by blocking the cell cycle, suppressing apoptosis and reinforcing DNA injury repair [33]. For this reason, they can be used as potential targets for discovering new approaches to chemotherapy and radiotherapy in HCC-affected subjects, as demonstrated by Huang et al. [34].

3. LncRNAs Associated with both Chronic Liver Disease and HCC

Numerous lncRNAs that were increased in conjunction with inflammation and fibrosis were observed, and analyses of these transcripts found many pathways, including those involved in TGF-β1 and TNF signaling, ECM deposition, and insulin resistance [35]. LncRNAs found in animal models were also highly expressed in fibrosis-related NAFLD, including NEAT1, MALAT1, and PVT1. Another study suggested that some lncRNAs identified in a CCl4 mouse model, namely, APTR, MALAT1, NEAT1, and HOTAIR, may also be related to NAFLD fibrosis in humans [36].
In a study of hepatitis B virus (HBV) persistent carriers and HBV-positive HCC patients, Liu et al. [37] found two single nucleotide polymorphisms, rs7763881 in HULC and rs619586 in MALAT1. The data showed that mutations of both rs7763881 in HULC and rs619586 in MALAT1 were related to a decreased HCC risk. LncRNA levels were also increased in the patients with hepatitis C virus (HCV) infection [1]. HCV infection stimulates GAS5 and HOTAIR in HCV-infected liver cells [38,39]. Yuan et al. [40] reported that the plasma levels of LINC00152, RP11-160H22.5, and XLOC_014172 could effectively distinguish patients with HCC from chronic hepatitis patients and healthy controls. Another study also showed that serum levels of MALAT1 could differentiate HCC patients from HCV-induced liver cirrhosis patients and healthy controls [41].
An association of lncRNAs with metabolic disease has also been reported [1]. High expressions of lncRNAs in non-alcoholic steatohepatitis (NASH) patients showed that increased expressions of liver-specific lncRNA lnc18q22.2 were positively correlated with NASH grade and NAFLD score [42]. The deletion of LncLSTR, the liver-specific triglyceride regulator, reduced plasma triglyceride content by modulating metabolic pathways in mice [43]. In addition, mice with a genetic knock-out of lncRNA SRA, the steroid receptor RNA activator, were resistant to high fat diet-induced obesity, and SRA suppressed adipose triglyceride lipase, reducing free fatty acid β-oxidation and promoting hepatic steatosis [44].
An analysis of human lncRNAs from peripheral blood RNA identified that lncRNAs, such as AK128652 and AK054921, expressed in normal human plasma and liver, were significantly increased in patients with alcoholic cirrhosis [1]. These levels were inversely associated with the survival rate of patients with alcoholic cirrhosis [45].
As mentioned above, HCC-related lncRNAs play critical regulatory roles in the progression of HCC, while their dysregulation is related to diverse biological processes [32]. As a type of regulator of cellular processes including proliferation, apoptosis, and carcinogenesis, lncRNAs play important roles in the tumorigenesis and development of HCC [5]. An increased expression of lncRNAs in HCC is thought to have an oncogenic function, whereas lncRNAs exhibiting a decreased expression in HCC may act as tumor suppressors [2]. The dysregulation of HULC, HOTAIR, MALAT1 and other genes has been identified in HCC [5,46,47,48,49,50,51,52,53,54]. In the meta-analysis, high expression levels of 27 types of lncRNAs, such as AFAP-AS1, HOTTIP, and ZEB-1-AS1 were found to be related to a poor prognosis [55], and a low expression of 18 types of lncRNAs, such as GAS5, MEG3, and XIST, were found to be associated with a worse prognosis. LINC00052, ZEB1-AS1 and LINC01225 displayed oncogenic properties by facilitating the invasiveness and metastasis induction of HCC cells [39]. Some other lncRNAs, such as XIST, HOST2, HOXA-AS2, CCHE1, and AFAP1-AS1, can induce and accelerate cell proliferation, while inhibiting the apoptosis of HCC cells [33]. This feature might be useful as a therapeutic target of HCC, especially as an alternative to the patients who show chemoresistance for the chemotherapeutic drugs [6,7,8,56].
While the underlying mechanism of HCC-related lncRNAs remains largely elusive, understanding the differential expression and potential functional roles of lncRNAs in HCC, as well as the relationship in expression between chronic liver disease and HCC, is mandatory. Therefore, we searched manuscripts by the PubMed, using key words such as ‘lncRNA’, ‘liver’, ‘fibrosis’, and/or ‘HCC’. We found 32 lncRNAs associated with chronic liver diseases and/or HCCs. Among them, nine lncRNAs were selected which were associated with both fibrosis and carcinogenesis in the liver. Here, we briefly summarize nine well-documented lncRNAs that participate in both hepatic fibrosis and HCC. A few lncRNAs associated with hepatic fibrosis with unknown effects in HCC are also discussed. The detailed information on the lncRNAs described below is also summarized in Table 1.

3.1. HULC

Panzitt et al. [82] reported a ‘highly up-regulated in liver cancer’ (HULC) located on chromosome 6p24.3, as a novel ncRNA. HULC is implicated in hepatocarcinogenesis as an oncogene by regulating multiple biological processes [2]. HULC promotes the proliferation of HCC cells via the regulation of regulating cell cycle-related genes in HCC cells [59]. HULC is negatively associated with PTEN or miR-15a expression in HCC patients, which promotes malignant progression [83]. HULC contributes to malignant development by acting as a sponge of miRNAs such as miR-9, miR-107, and miR-372, which induce PPARα, E2F1, and cAMP response element-binding protein (CREB) respectively [47,48,49]. In addition, HULC is responsible for perturbations of the circadian rhythm by up-regulating the circadian oscillator CLOCK, clock circadian regulator, in HCC cells, leading to the promotion of hepatocarcinogenesis [84].
Zhao et al. [57] confirmed the effects of HULC on Tregs differentiation in HBV-related liver cirrhosis. They found that circulating Tregs and HULC were significantly up-regulated in HBV-related cirrhosis patients, and HULC regulates the function of Tregs by directly down-regulating the level of p18 [57]. An increased expression of HULC was also found in the liver tissue of high--fat-diet NAFLD rats. Shen et al. [58] investigated the role of HULC in hepatic fibrosis and hepatocyte apoptosis by inhibiting the MAPK signaling pathway in rats with NAFLD.
The levels of HULC were positively correlated with HBV X protein (HBx) in HCC patients [1]. The activation of HULC promotes HBx-mediated cell proliferation by inducing p18 [46]. Du et al. [46] reported that the up-regulation of HULC, mediated by HBx, promoted the proliferation of HCC through the down-regulation of the tumor suppressor gene, CDKN2C (p18).
HULC also induced epithelial mesenchymal transition (EMT), promoting tumor progression and metastasis by its competition with miR-200a [38,39]. In these studies, HULC expression was related to TNM stage, intrahepatic metastases, HCC recurrence, and postoperative survival [1,38,39]. Furthermore, HULC acts as a ceRNA to activate the EMT process through the HULC/miR-200a-3p/ZEB1 signaling pathway and stimulates HCC progression and metastasis [38,60]. Li et al. [85] demonstrates that HULC specifically binds to Y-box protein-1 (YB-1) to promote its phosphorylation through ERK kinase and then regulates the interaction of YB-1 with certain oncogenic mRNAs, consequently accelerating the translation of these oncogenic mRNAs in hepatic carcinogenesis.

3.2. MALAT1 (NEAT2)

Metastasis-associated lung adenocarcinoma transcript 1 (MALAT1), which is also known as nuclear-enriched abundant transcript 2 (NEAT2), is expressed in both human and mouse tissues and is located at chromosome 11q13 [2,5]. It is up-regulated in many malignant tumors and is involved in cell proliferation and migration via modulating caspase-3, caspase-8, Bax, Bcl-2, and Bcl-xL [51]. Aberrant MALAT1 expression promotes tumor metastasis by regulating gene expression and alternative pre-mRNA splicing [86,87].
In CCl4-treated mice, hepatic expression levels of MALAT1 were elevated in hepatic stellate cells (HSCs) and hepatocytes, respectively [62]. MALAT1 can promote the activation of HSCs by blocking the silent information regulator 1 (SIRT1)-induced inhibition of the TGF-β1 signaling pathway in the progression of liver fibrosis [11,61]. A recent study reported that hepatic expression of MALAT1 was higher in NASH patients than in those NAFLD patients with simple steatosis and in healthy controls [88]. It is also reported that MALAT1 acts as a ceRNA for miR-101b to regulate RAS-related C3 botulinum substrate 1 (Rac1), promoting proliferation, cell cycle progression, and HSC activation and contributing to hepatic fibrosis [62].
MALAT1 is up-regulated in HCC, and the overexpression of MALAT1 promotes cell proliferation, migration, and invasion in HCC. It is also correlated with the expression of HBx [2,5,89]. Furthermore, MALAT1 was identified to act as a highly sensitive marker of human HCCs, suggesting that MALAT1 can be used as a potential tool for HCC diagnosis [2,51]. Lai et al. [90] examined the role of MALAT1 in HCC prognosis. They observed that siRNA knockdown of MALAT1 reduced cell proliferation and repressed migration and invasion as well as apoptosis, indicating that blocking MALAT1 activity in HCC might be a vital anticancer therapy [90]. Huang et al. [91] showed the role of specificity protein (Sp) 1/3 in the transcriptional regulation of MALAT1 in HCC cells. They found that Sp1 and Sp3 are participated in the up-regulation of MALAT1 expression [91]. MALAT1 was identified to be up-regulated in HCC and to play a role of a proto-oncogene to enhance HCC cell growth via the activation of Wnt and mTOR pathway and the up-regulation of serine/arginine-rich splicing factor 1 (SRSF1) [2,63]. Therefore, the inhibition of SRSF1 expression or mTOR activity is able to block the oncogenic effects of MALAT1 and HCC development.
Given the presence of elevated MALAT1 levels in HCC, the biological functions of MALAT1 remain largely unclear and require further studies regarding their roles in the progression of chronic liver disease, from cirrhosis to HCC.

3.3. HOTAIR

Homeobox (HOX) transcript antisense intergenic RNA (HOTAIR) is a lncRNA that resides on a boundary of the HOXC locus on chromosome 12q13.13, which is co-expressed with HOXC genes [2,92,93]. An increasing number of studies have investigated whether HOTAIR is up-regulated in multiple cancers, including breast cancer, lung adenocarcinoma, renal cell carcinoma, pancreatic cancer, and HCC [94,95,96,97,98].
Yang et al. [98] reported that the expression of HOTAIR in HCC tissues is significantly increased, when compared to that in adjacent non-cancerous tissues. In addition, the levels of HOTAIR expression in HCC cell lines were elevated compared to those in normal liver cell lines [98]. Notably, HCC patients with a high expression of HOTAIR had significantly worse prognoses than those without expression of HOTAIR [50]. HCC patients with HOTAIR overexpression have an increased risk of recurrence after hepatectomy, and HOTAIR overexpression is also correlated with increased risk of lymph node metastasis [99]. Furthermore, patients with a high expression of HOTAIR have a significantly shorter recurrence--free survival than patients with a low expression of HOTAIR [100].
To understand the oncogenic functions of HOTAIR in HCC, various mechanisms have been suggested [2]. The up-regulation of HOTAIR induces the proliferation, migration, and invasion of human HCC cells through the activation of autophagy [51,53,54]. Ding et al. [53] suggested that HOTAIR plays a critical role in the progression of HCC via inhibition of RNA binding motif protein 38 (RBM38). A regulatory network between miR-218 and HOTAIR was identified, whereby HOTAIR inactivates P16 (Ink4a) and P14 (ARF) signaling by down-regulating the expression of miR-218, resulting in hepatocarcinogenesis [52].
HOTAIR expression was up-regulated in the livers of CCl4-treated mice [64]. The expression levels of HOTAIR were also increased in cirrhotic liver tissues from HBV patients and colocalized with ACTA2, indicating that HSCs may be the primary source for HOTAIR in fibrotic liver [36]. A functional characterization of the lncRNA demonstrated that the overexpression of HOTAIR induced cell proliferation and elevated levels of ACTA2 and COL1A1, as well as fibrosis-related genes, such as matrix metalloproteinase 2 (MMP2) and MMP9 [64]. In addition, HOTAIR functions as a ceRNA to sponge miR-29b and then represses DNA methyltransferase 3b (DNMT3b), resulting in up-regulation of PTEN methylation, which contributes to hepatic fibrosis [65]. These results suggest that HOTAIR may promote fibrosis in liver by regulating DNMT1, MEG3, and the p53 pathway in HSCs, although further studies on this lncRNA is necessary in order to identify the role of HOTAIR in the process of hepatic fibrogenesis and carcinogenesis.

3.4. MEG3

Maternally expressed gene 3 (MEG3), is a lncRNA located at human chromosome 14q32 [101]. MEG3 could activate p53 to induce caspase-3-dependent apoptosis and suppress the expression of col1α1 and α-smooth muscle actin (α-SMA) in activated HSCs [66]. It is suggested that MEG3 plays a critical role in the activation of HSCs and fibrogenesis in liver [11].
One of the first studies of lncRNAs in NAFLD fibrosis identified that the expression of MEG3 was decreased in the livers of CCl4-treated mice, when compared to those of oil-fed control animals, and that the expression of MEG3 reduced concordantly with the progression of fibrosis [66]. In contrast to these results, hepatic MEG3 levels were significantly increased in liver fibrosis and NASH cirrhosis in human patients [102].
Numerous investigations have evaluated the functional role of MEG3 as a tumor suppressor in various types of human cancers, such as gastric cancer, lung cancer, glioma, cervical cancer, bladder cancer, and HCC [103,104,105,106,107,108,109]. The expression of MEG3 was found to be markedly decreased in HCC tissues [67,110]. Furthermore, ectopic MEG3 expression in HCC cells significantly inhibits proliferation and mediates apoptosis [67,105,110]. Several studies have also determined that MEG3 overexpression results in an increase of p53 protein and stimulates its transcriptional activity in HCC cells [67,111].
MEG3 is suggested to be an independent prognostic factor for HCC because the low expression of MEG3 was associated with a worse prognosis, compared to the high expression of MEG3 in HCC patients [111]. The biological roles of MEG3 also remain largely unknown and require further investigation regarding its function in the progression of chronic liver disease to HCC.

3.5. lncRNAp21

The long intergenic non-coding RNA-p21 (lncRNAp21), which is located 15 Kb upstream of the gene encoding the critical cell cycle regulator Cdkn1a (also known as p21), contains two exons comprising 3.1 Kb, together [112]. LncRNA-p21 acts as a transcriptional suppressor in the p53 pathway by activating p53 to promote apoptosis [113]. It has been investigated that lncRNAp21 is deregulated in various human diseases, such as skin tumors, prostatic cancers, and HCCs [114,115,116]. It also functions as a tumor suppressor in malignancies, but the mechanism of the process remains elusive [115].
Serum levels of lncRNA-p21 were negatively correlated with liver fibrosis in HBV patients [117]. Yu et al. [68] determined that lncRNA-p21 enhanced the expression of PTEN by sequestering miR-181b as a ceRNA and inhibited HSC activation through the PTEN/Akt pathway in liver fibrosis [68]. It was also found that the lncRNA-p21 competitively binds miR-17-5p for the inhibition of WIF1 via the Wnt/β-catenin pathway leading to the suppression of HSC activation [69].

3.6. GAS5

Growth arrest-specific transcript (GAS) 5 was initially discovered in a screen for potential tumor suppressor genes that are highly expressed during growth arrest. GAS5 was originally isolated from mouse embryo NIH 3T3 cells [118].
GAS5 has been reported as a tumor suppressor in some cancers, and it is related to the proliferation, apoptosis, and migration of tumor cells in breast cancer, stomach cancer and prostate cancer [119,120,121]. GAS5 directly binds to miR-21 to down-regulate its expression and negatively regulate the expression of miR-21 in HCC [71]. In liver fibrosis, GAS5 interacts with miR-222 and promotes the expression of p27 protein, thereby inhibiting the activation and proliferation of HSCs [70].

3.7. PVT1

Plasmacytoma variant translocation 1 (PVT1) is transcribed from a locus adjacent to the MYC locus on human chromosome 8q24 [72]. The PVT1 is known to be up-regulated in some human tumors, such as HCC, ovarian cancer, malignant pleural mesothelioma, non-small cell lung cancer and renal cancer [122,123,124,125,126].
The silencing of PVT1 in primary HSCs not only reduced the cell proliferation, but also decreased the protein levels of Actα2 and Col1α1 [36]. The PVT1 knockdown in primary HSCs was also associated with changes in the markers of the EMT process, indicating a possible mechanism by which this lncRNA promotes hepatic fibrosis [127].
PVT1 also activates the Hedgehog pathway by enhancing the methylation of Patched1 (PTCH1) and down-regulating PTCH1 expression through competitively binding miR-152, which is a driver of EMT and HSC activation in hepatic fibrosis [72].

3.8. NEAT1

Nuclear-enriched abundant transcript 1 (NEAT1) is located in 11q13 [128]. The expression of NEAT1 was found to be increased in primary HSCs derived from CCl4-treated mice compared to control mice [74]. NEAT1 overexpression promoted the activation of HSCs and increased levels of Actα2 and Col1α1, indicating that this lncRNA plays a role in HSC activation [36]. NEAT1 overexpression corresponded with decreased levels of miR-122, which was identified to regulate NEAT1 effects on the activation of HSCs by a mechanism associated with Kruppel-like factor 6 (Klf6) [36,74]. The expression of NEAT1 was up-regulated in HCC, while its knockdown was correlated with decreases not only in HCC cell proliferation, but with also invasion, and migration via regulating heterogeneous nuclear ribonucleoprotein A2 (hnRNP A2) [75]. However, the role of NEAT1 expression in cancer is controversial, because NEAT1 is strongly up-regulated by p53 in several cancers, and appears to play a tumor-suppressive role in the presence of wild-type p53 activation [129,130,131]. To understand the molecular mechanism of NEAT1 in chronic liver disease and HCC properly, the effect of p53 activation and mutation status always needs to be considered [131].

3.9. LncRNA-ATB

Qiu et al. [132] propose that lncRNA-ATB, activated by TGF-β1, could be used as a novel diagnostic biomarker to identify the severity of inflammation and fibrosis. Some studies investigated lncRNA-ATB, a lncRNA associated with liver fibrosis and HCC, in HCV patients and identified that plasma levels of this lncRNA were significantly correlated with the stage of liver fibrosis [133,134]. LncRNA-ATB was also significantly increased in HCC tissues and positively correlated with intrahepatic or extrahepatic metastases [135].

3.10. Other lncRNAs Related to Liver Fibrosis

The liver fibrosis-associated lncRNA1 (lnc-LFAR1) is a 734 nt transcript, which was originally demonstrated as a liver-enriched lncRNA in the fibrotic livers of mice [11]. Zhang et al. [77] used microarray analysis to profile lncRNAs in CCl4-treated mice and identified 266 up-regulated and 447 down-regulated lncRNAs [77]. Of these, a single lncRNA was identified, and it was most abundantly expressed in hepatocytes, followed by HSCs and Kupffer cells [77]. The lnc-LFAR1 of mice is located in chromosome 4q25, and it is adjacent to the CYP2U1 and HADH genes, which also exist in humans [77]. A recent analysis showed that three potential Smad2/3 binding sites (SBE) are present in the promoter of lnc-LFAR1, which means that Smad2/3 can bind to the promoter of lnc-LFAR1 to increase its expression [11,77]. Lnc-LFAR1 binds directly to Smad2/3 to control the transcription of a number of genes, including TGFB1, PAI, ACTA2, COL1A1, Smad2, Smad3, Notch2, and Notch3, resulting in the activation of the TGF-β1 and Notch pathways [36,77]. Together, lnc-LFAR1 was found to exert effects on HSC activation, hepatocyte apoptosis, and hepatic fibrogenesis in a mouse model [36,77].
Negishi et al. [136] identified a novel lncRNA, the Alu-mediated p21 transcriptional regulator (APTR), which was demonstrated to modulate cell cycle progression and cell proliferation. In an independent study, the expression of APTR was observed to be more than twofold higher in the fibrotic livers of animal models and cirrhotic livers in humans [79]. The silencing of APTR in primary HSCs also reduced ACTA2 and COL1A1 mRNA and protein expression and suppressed the TGF-β1-mediated up-regulation of ACTA2 [36]. In individuals with a cirrhotic liver, serum levels of APTR were approximately four-fold higher than individuals with a normal histology, and two-fold higher in patients with decompensated cirrhosis than those with compensated cirrhosis [79], indicating that serum levels of APTR may also serve as biomarkers of the severity for liver fibrosis [36].
LncRNA HIF1A-AS1 interacts with the partner TET3, one member of the ten-eleven translocation (TET) protein family, which is associated with DNA methylation, to inhibit the HSC activation [11,78]. The levels of both Cox2 and lncRNA-Cox2 were elevated in CCl4-treated mice, when compared to control animals, and those two transcripts were positively correlated with the amount of tissue affected by fibrosis [80,81].

4. Conclusions

It is believed that numerous lncRNAs are correlated with the progression of both liver cirrhosis and HCC from a review of the literature. Therefore, we can suppose that such lncRNAs play a crosslinking role between chronic liver disease and HCC. With the rapid development of high technology, such as NGS and bioinformatics, an increasing number of lncRNAs are emerging as novel biomarkers for early diagnosis, better prognostic evaluation and efficient therapeutic targets for HCC in future clinical applications [5]. It has been demonstrated that LncRNAs found in body fluids can be used as fluid-based non-invasive biological markers for clinical trials. In addition, lncRNAs can influence the sensitivity of HCC to chemo- or radiation therapy [5]. Thus, in the future, a better understanding of the molecular mechanism of lncRNAs associated with the initiation and progression of HCC will provide a rationale for novel effective lncRNA-based targeted therapies. Genomic/epigenetic directed stratifications in clinical trial design and enrollments in the era of lncRNAs could be considered. One of the underlying messages is that more precision and more individualized approaches need to be tested or considered in well-designed clinical trials [137,138]. Further studies are required, in particular, to examine the toxicity and the pharmacokinetics of lncRNAs and, additionally, to evaluate their biological properties for chronic liver disease. Nevertheless, these lncRNAs have the potential to be novel promising tools for HCC diagnosis and prognosis, as well as the protection of HCC development from chronic liver disease.

Author Contributions

Writing—original draft preparation, Y.-A.K.; writing—review and editing, S.-J.L.; visualization, K.-K.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the grant of Research Institute of Medical Science, Catholic University of Daegu (2019).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Heo, M.J.; Yun, J.; Kim, S.G. Role of non-coding RNAs in liver disease progression to hepatocellular carcinoma. Arch. Pharm. Res. 2019, 42, 48–62. [Google Scholar] [CrossRef]
  2. Niu, Z.S.; Niu, X.J.; Wang, W.H. Long non-coding RNAs in hepatocellular carcinoma: Potential roles and clinical implications. World J. Gastroenterol. 2017, 23, 5860–5874. [Google Scholar] [CrossRef]
  3. Wang, K.C.; Chang, H.Y. Molecular mechanisms of long noncoding RNAs. Mol. Cell 2011, 43, 904–914. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Wakayama, K.; Kamiyama, T.; Yokoo, H.; Orimo, T.; Shimada, S.; Einama, T.; Kamachi, H.; Taketomi, A. Huge hepatocellular carcinoma greater than 10 cm in diameter worsens prognosis by causing distant recurrence after curative resection. J. Surg. Oncol. 2017, 115, 324–329. [Google Scholar] [CrossRef]
  5. Li, C.; Chen, J.; Zhang, K.; Feng, B.; Wang, R.; Chen, L. Progress and Prospects of Long Noncoding RNAs (lncRNAs) in Hepatocellular Carcinoma. Cell Physiol. Biochem. 2015, 36, 423–434. [Google Scholar] [CrossRef]
  6. Brunetti, O.; Gnoni, A.; Licchetta, A.; Longo, V.; Calabrese, A.; Argentiero, A.; Delcuratolo, S.; Solimando, A.G.; Casadei-Gardini, A.; Silvestris, N. Predictive and Prognostic Factors in HCC Patients Treated with Sorafenib. Medicina (Kaunas) 2019, 55, 707. [Google Scholar] [CrossRef] [Green Version]
  7. Hu, X.; Jiang, J.; Xu, Q.; Ni, C.; Yang, L.; Huang, D. A Systematic Review of Long Noncoding RNAs in Hepatocellular Carcinoma: Molecular Mechanism and Clinical Implications. BioMed Res. Int. 2018, 2018, 8126208. [Google Scholar] [CrossRef] [Green Version]
  8. Ghidini, M.; Braconi, C. Non-Coding RNAs in Primary Liver Cancer. Front. Med. (Lausanne) 2015, 2, 36. [Google Scholar] [CrossRef] [Green Version]
  9. Wang, J.; Chu, E.S.; Chen, H.-Y.; Man, K.; Go, M.Y.; Huang, X.R.; Lan, H.Y.; Sung, J.J.; Yu, J. microRNA-29b prevents liver fibrosis by attenuating hepatic stellate cell activation and inducing apoptosis through targeting PI3K/AKT pathway. Oncotarget 2015, 6, 7325. [Google Scholar] [CrossRef]
  10. Wang, X.; Hu, F.; Hu, X.; Chen, W.; Huang, Y.; Yu, X. Proteomic identification of potential Clonorchis sinensis excretory/secretory products capable of binding and activating human hepatic stellate cells. Parasitol. Res. 2014, 113, 3063–3071. [Google Scholar] [CrossRef]
  11. Peng, H.; Wan, L.Y.; Liang, J.J.; Zhang, Y.Q.; Ai, W.B.; Wu, J.F. The roles of lncRNA in hepatic fibrosis. Cell Biosci. 2018, 8, 63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Liu, Y.; Yang, Y.; Wang, T.; Wang, L.; Wang, X.; Li, T.; Shi, Y.; Wang, Y. Long non-coding RNA CCAL promotes hepatocellular carcinoma progression by regulating AP-2alpha and Wnt/beta-catenin pathway. Int. J. Biol. Macromol. 2018, 109, 424–434. [Google Scholar] [CrossRef] [PubMed]
  13. Koh, W.; Pan, W.; Gawad, C.; Fan, H.C.; Kerchner, G.A.; Wyss-Coray, T.; Blumenfeld, Y.J.; El-Sayed, Y.Y.; Quake, S.R. Noninvasive in vivo monitoring of tissue-specific global gene expression in humans. Proc. Natl. Acad. Sci. USA 2014, 111, 7361–7366. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. He, X.; Sun, F.; Guo, F.; Wang, K.; Gao, Y.; Feng, Y.; Song, B.; Li, W.; Li, Y. Knockdown of Long Noncoding RNA FTX Inhibits Proliferation, Migration, and Invasion in Renal Cell Carcinoma Cells. Oncol. Res. 2017, 25, 157–166. [Google Scholar] [CrossRef]
  15. Zhang, L.-G.; Zhou, X.-K.; Zhou, R.-J.; Lv, H.-Z.; Li, W.-P. Long non-coding RNA LINC00673 promotes hepatocellular carcinoma progression and metastasis through negatively regulating miR-205. Am. J. Cancer Res. 2017, 7, 2536. [Google Scholar] [PubMed]
  16. Ponting, C.P.; Oliver, P.L.; Reik, W. Evolution and functions of long noncoding RNAs. Cell 2009, 136, 629–641. [Google Scholar] [CrossRef] [Green Version]
  17. He, Y.; Meng, X.M.; Huang, C.; Wu, B.M.; Zhang, L.; Lv, X.W.; Li, J. Long noncoding RNAs: Novel insights into hepatocelluar carcinoma. Cancer Lett. 2014, 344, 20–27. [Google Scholar] [CrossRef]
  18. Derrien, T.; Johnson, R.; Bussotti, G.; Tanzer, A.; Djebali, S.; Tilgner, H.; Guernec, G.; Martin, D.; Merkel, A.; Knowles, D.G.; et al. The GENCODE v7 catalog of human long noncoding RNAs: analysis of their gene structure, evolution, and expression. Genome Res. 2012, 22, 1775–1789. [Google Scholar] [CrossRef] [Green Version]
  19. Yoon, J.H.; Abdelmohsen, K.; Gorospe, M. Functional interactions among microRNAs and long noncoding RNAs. Semin. Cell Dev. Biol. 2014, 34, 9–14. [Google Scholar] [CrossRef] [Green Version]
  20. Liz, J.; Esteller, M. lncRNAs and microRNAs with a role in cancer development. Biochim. Biophys. Acta 2016, 1859, 169–176. [Google Scholar] [CrossRef] [Green Version]
  21. Chen, Y.; Li, C.; Pan, Y.; Han, S.; Feng, B.; Gao, Y.; Chen, J.; Zhang, K.; Wang, R.; Chen, L. The Emerging Role and Promise of Long Noncoding RNAs in Lung Cancer Treatment. Cell Physiol. Biochem. 2016, 38, 2194–2206. [Google Scholar] [CrossRef]
  22. Mercer, T.R.; Dinger, M.E.; Mattick, J.S. Long non-coding RNAs: insights into functions. Nat. Rev. Genet. 2009, 10, 155–159. [Google Scholar] [CrossRef] [PubMed]
  23. Cao, J. The functional role of long non-coding RNAs and epigenetics. Biol. Proced. Online 2014, 16, 11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Ulitsky, I.; Bartel, D.P. lincRNAs: Genomics, evolution, and mechanisms. Cell 2013, 154, 26–46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Iyer, M.K.; Niknafs, Y.S.; Malik, R.; Singhal, U.; Sahu, A.; Hosono, Y.; Barrette, T.R.; Prensner, J.R.; Evans, J.R.; Zhao, S. The landscape of long noncoding RNAs in the human transcriptome. Nat. Genet. 2015, 47, 199. [Google Scholar] [CrossRef] [PubMed]
  26. Toiyama, Y.; Okugawa, Y.; Goel, A. DNA methylation and microRNA biomarkers for noninvasive detection of gastric and colorectal cancer. Biochem. Biophys. Res. Commun. 2014, 455, 43–57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Isin, M.; Dalay, N. LncRNAs and neoplasia. Clin. Chim. Acta 2015, 444, 280–288. [Google Scholar] [CrossRef] [PubMed]
  28. Serviss, J.T.; Johnsson, P.; Grander, D. An emerging role for long non-coding RNAs in cancer metastasis. Front. Genet. 2014, 5, 234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Harries, L.W. Long Non-Coding RNAs and Human Disease; Portland Press Ltd.: London, UK, 2012. [Google Scholar]
  30. Amicone, L.; Citarella, F.; Cicchini, C. Epigenetic regulation in hepatocellular carcinoma requires long noncoding RNAs. Biomed Res. Int. 2015, 2015, 473942. [Google Scholar] [CrossRef] [Green Version]
  31. Yan, X.; Hu, Z.; Feng, Y.; Hu, X.; Yuan, J.; Zhao, S.D.; Zhang, Y.; Yang, L.; Shan, W.; He, Q.; et al. Comprehensive Genomic Characterization of Long Non-coding RNAs across Human Cancers. Cancer Cell 2015, 28, 529–540. [Google Scholar] [CrossRef] [Green Version]
  32. Gutschner, T.; Diederichs, S. The hallmarks of cancer: a long non-coding RNA point of view. RNA Biol. 2012, 9, 703–719. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Peng, L.; Yuan, X.Q.; Zhang, C.Y.; Peng, J.Y.; Zhang, Y.Q.; Pan, X.; Li, G.C. The emergence of long non-coding RNAs in hepatocellular carcinoma: an update. J. Cancer 2018, 9, 2549–2558. [Google Scholar] [CrossRef] [Green Version]
  34. Huang, H.; Chen, J.; Ding, C.M.; Jin, X.; Jia, Z.M.; Peng, J. LncRNA NR2F1-AS1 regulates hepatocellular carcinoma oxaliplatin resistance by targeting ABCC1 via miR-363. J. Cell Mol. Med. 2018, 22, 3238–3245. [Google Scholar] [CrossRef]
  35. Leti, F.; Legendre, C.; Still, C.D.; Chu, X.; Petrick, A.; Gerhard, G.S.; DiStefano, J.K. Altered expression of MALAT1 lncRNA in nonalcoholic steatohepatitis fibrosis regulates CXCL5 in hepatic stellate cells. Transl. Res. 2017, 190, 25–39.e21. [Google Scholar] [CrossRef] [PubMed]
  36. Hanson, A.; Wilhelmsen, D.; DiStefano, J.K. The Role of Long Non-Coding RNAs (lncRNAs) in the Development and Progression of Fibrosis Associated with Nonalcoholic Fatty Liver Disease (NAFLD). Noncoding RNA 2018, 4, 18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Liu, Y.; Pan, S.; Liu, L.; Zhai, X.; Liu, J.; Wen, J.; Zhang, Y.; Chen, J.; Shen, H.; Hu, Z. A genetic variant in long non-coding RNA HULC contributes to risk of HBV-related hepatocellular carcinoma in a Chinese population. PLoS ONE 2012, 7, e35145. [Google Scholar] [CrossRef]
  38. Li, S.-P.; Xu, H.-X.; Yu, Y.; He, J.-D.; Wang, Z.; Xu, Y.-J.; Wang, C.-Y.; Zhang, H.-M.; Zhang, R.-X.; Zhang, J.-J. LncRNA HULC enhances epithelial-mesenchymal transition to promote tumorigenesis and metastasis of hepatocellular carcinoma via the miR-200a-3p/ZEB1 signaling pathway. Oncotarget 2016, 7, 42431. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Li, T.; Xie, J.; Shen, C.; Cheng, D.; Shi, Y.; Wu, Z.; Deng, X.; Chen, H.; Shen, B.; Peng, C. Upregulation of long noncoding RNA ZEB1-AS1 promotes tumor metastasis and predicts poor prognosis in hepatocellular carcinoma. Oncogene 2016, 35, 1575–1584. [Google Scholar] [CrossRef] [PubMed]
  40. Yuan, W.; Sun, Y.; Liu, L.; Zhou, B.; Wang, S.; Gu, D. Circulating LncRNAs Serve as Diagnostic Markers for Hepatocellular Carcinoma. Cell Physiol. Biochem. 2017, 44, 125–132. [Google Scholar] [CrossRef] [PubMed]
  41. Toraih, E.A.; Ellawindy, A.; Fala, S.Y.; Al Ageeli, E.; Gouda, N.S.; Fawzy, M.S.; Hosny, S. Oncogenic long noncoding RNA MALAT1 and HCV-related hepatocellular carcinoma. Biomed Pharmacother. 2018, 102, 653–669. [Google Scholar] [CrossRef] [PubMed]
  42. Atanasovska, B.; Rensen, S.S.; van der Sijde, M.R.; Marsman, G.; Kumar, V.; Jonkers, I.; Withoff, S.; Shiri-Sverdlov, R.; Greve, J.W.M.; Faber, K.N.; et al. A liver-specific long noncoding RNA with a role in cell viability is elevated in human nonalcoholic steatohepatitis. Hepatology 2017, 66, 794–808. [Google Scholar] [CrossRef] [PubMed]
  43. Li, P.; Ruan, X.; Yang, L.; Kiesewetter, K.; Zhao, Y.; Luo, H.; Chen, Y.; Gucek, M.; Zhu, J.; Cao, H. A liver-enriched long non-coding RNA, lncLSTR, regulates systemic lipid metabolism in mice. Cell Metab. 2015, 21, 455–467. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Chen, G.; Yu, D.; Nian, X.; Liu, J.; Koenig, R.J.; Xu, B.; Sheng, L. LncRNA SRA promotes hepatic steatosis through repressing the expression of adipose triglyceride lipase (ATGL). Sci. Rep. 2016, 6, 35531. [Google Scholar] [CrossRef] [PubMed]
  45. Yang, Z.; Ross, R.A.; Zhao, S.; Tu, W.; Liangpunsakul, S.; Wang, L. LncRNA AK054921 and AK128652 are potential serum biomarkers and predictors of patient survival with alcoholic cirrhosis. Hepatol. Commun. 2017, 1, 513–523. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Du, Y.; Kong, G.; You, X.; Zhang, S.; Zhang, T.; Gao, Y.; Ye, L.; Zhang, X. Elevation of highly up-regulated in liver cancer (HULC) by hepatitis B virus X protein promotes hepatoma cell proliferation via down-regulating p18. J. Biol. Chem. 2012, 287, 26302–26311. [Google Scholar] [CrossRef] [Green Version]
  47. Wang, J.; Liu, X.; Wu, H.; Ni, P.; Gu, Z.; Qiao, Y.; Chen, N.; Sun, F.; Fan, Q. CREB up-regulates long non-coding RNA, HULC expression through interaction with microRNA-372 in liver cancer. Nucleic Acids Res. 2010, 38, 5366–5383. [Google Scholar] [CrossRef] [Green Version]
  48. Cui, M.; Xiao, Z.; Wang, Y.; Zheng, M.; Song, T.; Cai, X.; Sun, B.; Ye, L.; Zhang, X. Long noncoding RNA HULC modulates abnormal lipid metabolism in hepatoma cells through an miR-9-mediated RXRA signaling pathway. Cancer Res. 2015, 75, 846–857. [Google Scholar] [CrossRef] [Green Version]
  49. Lu, Z.; Xiao, Z.; Liu, F.; Cui, M.; Li, W.; Yang, Z.; Li, J.; Ye, L.; Zhang, X. Long non-coding RNA HULC promotes tumor angiogenesis in liver cancer by up-regulating sphingosine kinase 1 (SPHK1). Oncotarget 2016, 7, 241–254. [Google Scholar] [CrossRef] [Green Version]
  50. Ishibashi, M.; Kogo, R.; Shibata, K.; Sawada, G.; Takahashi, Y.; Kurashige, J.; Akiyoshi, S.; Sasaki, S.; Iwaya, T.; Sudo, T.; et al. Clinical significance of the expression of long non-coding RNA HOTAIR in primary hepatocellular carcinoma. Oncol. Rep. 2013, 29, 946–950. [Google Scholar] [CrossRef] [Green Version]
  51. Lin, R.; Maeda, S.; Liu, C.; Karin, M.; Edgington, T.S. A large noncoding RNA is a marker for murine hepatocellular carcinomas and a spectrum of human carcinomas. Oncogene 2007, 26, 851–858. [Google Scholar] [CrossRef] [Green Version]
  52. Fu, W.M.; Zhu, X.; Wang, W.M.; Lu, Y.F.; Hu, B.G.; Wang, H.; Liang, W.C.; Wang, S.S.; Ko, C.H.; Waye, M.M.; et al. Hotair mediates hepatocarcinogenesis through suppressing miRNA-218 expression and activating P14 and P16 signaling. J. Hepatol. 2015, 63, 886–895. [Google Scholar] [CrossRef] [PubMed]
  53. Ding, C.; Cheng, S.; Yang, Z.; Lv, Z.; Xiao, H.; Du, C.; Peng, C.; Xie, H.; Zhou, L.; Wu, J.; 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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Su, D.N.; Wu, S.P.; Chen, H.T.; He, J.H. HOTAIR, a long non-coding RNA driver of malignancy whose expression is activated by FOXC1, negatively regulates miRNA-1 in hepatocellular carcinoma. Oncol. Lett. 2016, 12, 4061–4067. [Google Scholar] [CrossRef] [PubMed]
  55. Zheng, C.; Liu, X.; Chen, L.; Xu, Z.; Shao, J. lncRNAs as prognostic molecular biomarkers in hepatocellular carcinoma: a systematic review and meta-analysis. Oncotarget 2017, 8, 59638–59647. [Google Scholar] [CrossRef] [Green Version]
  56. Pennisi, G.; Celsa, C.; Giammanco, A.; Spatola, F.; Petta, S. The Burden of Hepatocellular Carcinoma in Non-Alcoholic Fatty Liver Disease: Screening Issue and Future Perspectives. Int. J. Mol. Sci. 2019, 20, 5613. [Google Scholar] [CrossRef] [Green Version]
  57. Zhao, J.; Fan, Y.; Wang, K.; Ni, X.; Gu, J.; Lu, H.; Lu, Y.; Lu, L.; Dai, X.; Wang, X. LncRNA HULC affects the differentiation of Treg in HBV-related liver cirrhosis. Int. Immunopharmacol. 2015, 28, 901–905. [Google Scholar] [CrossRef]
  58. Shen, X.; Guo, H.; Xu, J.; Wang, J. Inhibition of lncRNA HULC improves hepatic fibrosis and hepatocyte apoptosis by inhibiting the MAPK signaling pathway in rats with nonalcoholic fatty liver disease. J. Cell Physiol. 2019, 234, 18169–18179. [Google Scholar] [CrossRef]
  59. Zhang, Y.; Li, Z.; Zhang, Y.; Zhong, Q.; Chen, Q.; Zhang, L. Molecular mechanism of HEIH and HULC in the proliferation and invasion of hepatoma cells. Int. J. Clin. Exp. Med. 2015, 8, 12956–12962. [Google Scholar]
  60. Tao, H.; Cao, W.; Yang, J.J.; Shi, K.H.; Zhou, X.; Liu, L.P.; Li, J. Long noncoding RNA H19 controls DUSP5/ERK1/2 axis in cardiac fibroblast proliferation and fibrosis. Cardiovasc. Pathol. 2016, 25, 381–389. [Google Scholar] [CrossRef]
  61. Wu, Y.; Liu, X.; Zhou, Q.; Huang, C.; Meng, X.; Xu, F.; Li, J. Silent information regulator 1 (SIRT1) ameliorates liver fibrosis via promoting activated stellate cell apoptosis and reversion. Toxicol. Appl. Pharmacol. 2015, 289, 163–176. [Google Scholar] [CrossRef]
  62. Yu, F.; Lu, Z.; Cai, J.; Huang, K.; Chen, B.; Li, G.; Dong, P.; Zheng, J. MALAT1 functions as a competing endogenous RNA to mediate Rac1 expression by sequestering miR-101b in liver fibrosis. Cell Cycle 2015, 14, 3885–3896. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Malakar, P.; Shilo, A.; Mogilevsky, A.; Stein, I.; Pikarsky, E.; Nevo, Y.; Benyamini, H.; Elgavish, S.; Zong, X.; Prasanth, K.V.; et al. Long Noncoding RNA MALAT1 Promotes Hepatocellular Carcinoma Development by SRSF1 Upregulation and mTOR Activation. Cancer Res. 2017, 77, 1155–1167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Bian, E.B.; Wang, Y.Y.; Yang, Y.; Wu, B.M.; Xu, T.; Meng, X.M.; Huang, C.; Zhang, L.; Lv, X.W.; Xiong, Z.G.; et al. Hotair facilitates hepatic stellate cells activation and fibrogenesis in the liver. Biochim. Biophys. Acta Mol. Basis Dis. 2017, 1863, 674–686. [Google Scholar] [CrossRef] [PubMed]
  65. Yu, F.; Chen, B.; Dong, P.; Zheng, J. HOTAIR Epigenetically Modulates PTEN Expression via MicroRNA-29b: A Novel Mechanism in Regulation of Liver Fibrosis. Mol. Ther. 2017, 25, 205–217. [Google Scholar] [CrossRef] [Green Version]
  66. He, Y.; Wu, Y.T.; Huang, C.; Meng, X.M.; Ma, T.T.; Wu, B.M.; Xu, F.Y.; Zhang, L.; Lv, X.W.; Li, J. Inhibitory effects of long noncoding RNA MEG3 on hepatic stellate cells activation and liver fibrogenesis. Biochim. Biophys. Acta 2014, 1842, 2204–2215. [Google Scholar] [CrossRef] [Green Version]
  67. Zhu, J.; Liu, S.; Ye, F.; Shen, Y.; Tie, Y.; Zhu, J.; Wei, L.; Jin, Y.; Fu, H.; Wu, Y.; et al. Long Noncoding RNA MEG3 Interacts with p53 Protein and Regulates Partial p53 Target Genes in Hepatoma Cells. PLoS ONE 2015, 10, e0139790. [Google Scholar] [CrossRef]
  68. Yu, F.; Lu, Z.; Chen, B.; Dong, P.; Zheng, J. Identification of a Novel lincRNA-p21-miR-181b-PTEN Signaling Cascade in Liver Fibrosis. Mediators Inflamm. 2016, 2016, 9856538. [Google Scholar] [CrossRef] [Green Version]
  69. Yu, F.; Guo, Y.; Chen, B.; Shi, L.; Dong, P.; Zhou, M.; Zheng, J. LincRNA-p21 Inhibits the Wnt/β-Catenin Pathway in Activated Hepatic Stellate Cells via Sponging MicroRNA-17-5p. Cell Physiol. Biochem. 2017, 41, 1970–1980. [Google Scholar] [CrossRef]
  70. Yu, F.; Zheng, J.; Mao, Y.; Dong, P.; Lu, Z.; Li, G.; Guo, C.; Liu, Z.; Fan, X. Long Non-coding RNA Growth Arrest-specific Transcript 5 (GAS5) Inhibits Liver Fibrogenesis through a Mechanism of Competing Endogenous RNA. J. Biol. Chem. 2015, 290, 28286–28298. [Google Scholar] [CrossRef] [Green Version]
  71. Zhang, Z.; Zhu, Z.; Watabe, K.; Zhang, X.; Bai, C.; Xu, M.; Wu, F.; Mo, Y.Y. Negative regulation of lncRNA GAS5 by miR-21. Cell Death Differ. 2013, 20, 1558–1568. [Google Scholar] [CrossRef] [Green Version]
  72. Barsotti, A.M.; Beckerman, R.; Laptenko, O.; Huppi, K.; Caplen, N.J.; Prives, C. p53-Dependent induction of PVT1 and miR-1204. J. Biol. Chem. 2012, 287, 2509–2519. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Guo, J.; Hao, C.; Wang, C.; Li, L. Long noncoding RNA PVT1 modulates hepatocellular carcinoma cell proliferation and apoptosis by recruiting EZH2. Cancer Cell Int. 2018, 18, 98. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Yu, F.; Jiang, Z.; Chen, B.; Dong, P.; Zheng, J. NEAT1 accelerates the progression of liver fibrosis via regulation of microRNA-122 and Kruppel-like factor 6. J. Mol. Med. (Berl.) 2017, 95, 1191–1202. [Google Scholar] [CrossRef]
  75. Mang, Y.; Li, L.; Ran, J.; Zhang, S.; Liu, J.; Li, L.; Chen, Y.; Liu, J.; Gao, Y.; Ren, G. Long noncoding RNA NEAT1 promotes cell proliferation and invasion by regulating hnRNP A2 expression in hepatocellular carcinoma cells. OncoTargets Ther. 2017, 10, 1003–1016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Wang, C.Z.; Yan, G.X.; Dong, D.S.; Xin, H.; Liu, Z.Y. LncRNA-ATB promotes autophagy by activating Yes-associated protein and inducing autophagy-related protein 5 expression in hepatocellular carcinoma. World J. Gastroenterol. 2019, 25, 5310–5322. [Google Scholar] [CrossRef]
  77. Zhang, K.; Han, X.; Zhang, Z.; Zheng, L.; Hu, Z.; Yao, Q.; Cui, H.; Shu, G.; Si, M.; Li, C.; et al. The liver-enriched lnc-LFAR1 promotes liver fibrosis by activating TGFβ and Notch pathways. Nat. Commun. 2017, 8, 144. [Google Scholar] [CrossRef] [Green Version]
  78. Zhang, Q.Q.; Xu, M.Y.; Qu, Y.; Hu, J.J.; Li, Z.H.; Zhang, Q.D.; Lu, L.G. TET3 mediates the activation of human hepatic stellate cells via modulating the expression of long non-coding RNA HIF1A-AS1. Int. J. Clin. Exp. Pathol. 2014, 7, 7744–7751. [Google Scholar] [PubMed]
  79. Yu, F.; Zheng, J.; Mao, Y.; Dong, P.; Li, G.; Lu, Z.; Guo, C.; Liu, Z.; Fan, X. Long non-coding RNA APTR promotes the activation of hepatic stellate cells and the progression of liver fibrosis. Biochem. Biophys. Res. Commun. 2015, 463, 679–685. [Google Scholar] [CrossRef]
  80. Jeong, S.W.; Jang, J.Y.; Lee, S.H.; Kim, S.G.; Cheon, Y.K.; Kim, Y.S.; Cho, Y.D.; Kim, H.S.; Lee, J.S.; Jin, S.Y.; et al. Increased expression of cyclooxygenase-2 is associated with the progression to cirrhosis. Korean J. Intern. Med. 2010, 25, 364–371. [Google Scholar] [CrossRef]
  81. Tong, Q.; Gong, A.Y.; Zhang, X.T.; Lin, C.; Ma, S.; Chen, J.; Hu, G.; Chen, X.M. LincRNA-Cox2 modulates TNF-α-induced transcription of Il12b gene in intestinal epithelial cells through regulation of Mi-2/NuRD-mediated epigenetic histone modifications. FASEB J. 2016, 30, 1187–1197. [Google Scholar] [CrossRef] [Green Version]
  82. Panzitt, K.; Tschernatsch, M.M.; Guelly, C.; Moustafa, T.; Stradner, M.; Strohmaier, H.M.; Buck, C.R.; Denk, H.; Schroeder, R.; Trauner, M.; et al. Characterization of HULC, a novel gene with striking up-regulation in hepatocellular carcinoma, as noncoding RNA. Gastroenterology 2007, 132, 330–342. [Google Scholar] [CrossRef] [PubMed]
  83. Xin, X.; Wu, M.; Meng, Q.; Wang, C.; Lu, Y.; Yang, Y.; Li, X.; Zheng, Q.; Pu, H.; Gui, X.; et al. Long noncoding RNA HULC accelerates liver cancer by inhibiting PTEN via autophagy cooperation to miR15a. Mol. Cancer 2018, 17, 94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Cui, M.; Zheng, M.; Sun, B.; Wang, Y.; Ye, L.; Zhang, X. A long noncoding RNA perturbs the circadian rhythm of hepatoma cells to facilitate hepatocarcinogenesis. Neoplasia 2015, 17, 79–88. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Li, D.; Liu, X.; Zhou, J.; Hu, J.; Zhang, D.; Liu, J.; Qiao, Y.; Zhan, Q. Long noncoding RNA HULC modulates the phosphorylation of YB-1 through serving as a scaffold of extracellular signal-regulated kinase and YB-1 to enhance hepatocarcinogenesis. Hepatology 2017, 65, 1612–1627. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Yang, L.; Lin, C.; Liu, W.; Zhang, J.; Ohgi, K.A.; Grinstein, J.D.; Dorrestein, P.C.; Rosenfeld, M.G. ncRNA- and Pc2 methylation-dependent gene relocation between nuclear structures mediates gene activation programs. Cell 2011, 147, 773–788. [Google Scholar] [CrossRef] [Green Version]
  87. Yoshimoto, R.; Mayeda, A.; Yoshida, M.; Nakagawa, S. MALAT1 long non-coding RNA in cancer. Biochim. Biophys. Acta 2016, 1859, 192–199. [Google Scholar] [CrossRef]
  88. Sookoian, S.; Flichman, D.; Garaycoechea, M.E.; San Martino, J.; Castaño, G.O.; Pirola, C.J. Metastasis-associated lung adenocarcinoma transcript 1 as a common molecular driver in the pathogenesis of nonalcoholic steatohepatitis and chronic immune-mediated liver damage. Hepatol. Commun. 2018, 2, 654–665. [Google Scholar] [CrossRef]
  89. Hou, Z.; Xu, X.; Fu, X.; Tao, S.; Zhou, J.; Liu, S.; Tan, D. HBx-related long non-coding RNA MALAT1 promotes cell metastasis via up-regulating LTBP3 in hepatocellular carcinoma. Am. J. Cancer Res. 2017, 7, 845–856. [Google Scholar]
  90. Lai, M.C.; Yang, Z.; Zhou, L.; Zhu, Q.Q.; Xie, H.Y.; Zhang, F.; Wu, L.M.; Chen, L.M.; Zheng, S.S. Long non-coding RNA MALAT-1 overexpression predicts tumor recurrence of hepatocellular carcinoma after liver transplantation. Med. Oncol. 2012, 29, 1810–1816. [Google Scholar] [CrossRef]
  91. Huang, Z.; Huang, L.; Shen, S.; Li, J.; Lu, H.; Mo, W.; Dang, Y.; Luo, D.; Chen, G.; Feng, Z. Sp1 cooperates with Sp3 to upregulate MALAT1 expression in human hepatocellular carcinoma. Oncol. Rep. 2015, 34, 2403–2412. [Google Scholar] [CrossRef] [Green Version]
  92. Rinn, J.L.; Kertesz, M.; Wang, J.K.; Squazzo, S.L.; Xu, X.; Brugmann, S.A.; Goodnough, L.H.; Helms, J.A.; Farnham, P.J.; Segal, E.; et al. Functional demarcation of active and silent chromatin domains in human HOX loci by noncoding RNAs. Cell 2007, 129, 1311–1323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Gupta, R.A.; Shah, N.; Wang, K.C.; Kim, J.; Horlings, H.M.; Wong, D.J.; Tsai, M.C.; Hung, T.; Argani, P.; Rinn, J.L.; et al. Long non-coding RNA HOTAIR reprograms chromatin state to promote cancer metastasis. Nature 2010, 464, 1071–1076. [Google Scholar] [CrossRef] [PubMed]
  94. Lu, L.; Zhu, G.; Zhang, C.; Deng, Q.; Katsaros, D.; Mayne, S.T.; Risch, H.A.; Mu, L.; Canuto, E.M.; Gregori, G.; et al. Association of large noncoding RNA HOTAIR expression and its downstream intergenic CpG island methylation with survival in breast cancer. Breast Cancer Res. Treat. 2012, 136, 875–883. [Google Scholar] [CrossRef] [PubMed]
  95. Liu, Y.; Jiang, H.; Zhou, H.; Ying, X.; Wang, Z.; Yang, Y.; Xu, W.; He, X.; Li, Y. Lentivirus-mediated silencing of HOTAIR lncRNA restores gefitinib sensitivity by activating Bax/Caspase-3 and suppressing TGF-α/EGFR signaling in lung adenocarcinoma. Oncol. Lett. 2018, 15, 2829–2838. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Dasgupta, P.; Kulkarni, P.; Majid, S.; Shahryari, V.; Hashimoto, Y.; Bhat, N.S.; Shiina, M.; Deng, G.; Saini, S.; Tabatabai, Z.L.; et al. MicroRNA-203 Inhibits Long Noncoding RNA HOTAIR and Regulates Tumorigenesis through Epithelial-to-mesenchymal Transition Pathway in Renal Cell Carcinoma. Mol. Cancer Ther. 2018, 17, 1061–1069. [Google Scholar] [CrossRef] [Green Version]
  97. Kim, K.; Jutooru, I.; Chadalapaka, G.; Johnson, G.; Frank, J.; Burghardt, R.; Kim, S.; Safe, S. HOTAIR is a negative prognostic factor and exhibits pro-oncogenic activity in pancreatic cancer. Oncogene 2013, 32, 1616–1625. [Google Scholar] [CrossRef] [Green Version]
  98. Yang, Z.; Zhou, L.; Wu, L.M.; Lai, M.C.; Xie, H.Y.; Zhang, F.; Zheng, S.S. Overexpression of long non-coding RNA HOTAIR predicts tumor recurrence in hepatocellular carcinoma patients following liver transplantation. Ann. Surg. Oncol. 2011, 18, 1243–1250. [Google Scholar] [CrossRef]
  99. Gao, J.Z.; Li, J.; Du, J.L.; Li, X.L. Long non-coding RNA HOTAIR is a marker for hepatocellular carcinoma progression and tumor recurrence. Oncol. Lett. 2016, 11, 1791–1798. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Xu, Z.Y.; Yu, Q.M.; Du, Y.A.; Yang, L.T.; Dong, R.Z.; Huang, L.; Yu, P.F.; Cheng, X.D. Knockdown of long non-coding RNA HOTAIR suppresses tumor invasion and reverses epithelial-mesenchymal transition in gastric cancer. Int. J. Biol. Sci. 2013, 9, 587–597. [Google Scholar] [CrossRef] [Green Version]
  101. Miyoshi, N.; Wagatsuma, H.; Wakana, S.; Shiroishi, T.; Nomura, M.; Aisaka, K.; Kohda, T.; Surani, M.A.; Kaneko-Ishino, T.; Ishino, F. Identification of an imprinted gene, Meg3/Gtl2 and its human homologue MEG3, first mapped on mouse distal chromosome 12 and human chromosome 14q. Genes Cells 2000, 5, 211–220. [Google Scholar] [CrossRef]
  102. Zhang, L.; Yang, Z.; Trottier, J.; Barbier, O.; Wang, L. Long noncoding RNA MEG3 induces cholestatic liver injury by interaction with PTBP1 to facilitate shp mRNA decay. Hepatology 2017, 65, 604–615. [Google Scholar] [CrossRef] [PubMed]
  103. Zhang, J.; Lin, Z.; Gao, Y.; Yao, T. Downregulation of long noncoding RNA MEG3 is associated with poor prognosis and promoter hypermethylation in cervical cancer. J. Exp. Clin. Cancer Res. 2017, 36, 5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Kruer, T.L.; Dougherty, S.M.; Reynolds, L.; Long, E.; de Silva, T.; Lockwood, W.W.; Clem, B.F. Expression of the lncRNA Maternally Expressed Gene 3 (MEG3) Contributes to the Control of Lung Cancer Cell Proliferation by the Rb Pathway. PLoS ONE 2016, 11, e0166363. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Anwar, S.L.; Krech, T.; Hasemeier, B.; Schipper, E.; Schweitzer, N.; Vogel, A.; Kreipe, H.; Lehmann, U. Loss of imprinting and allelic switching at the DLK1-MEG3 locus in human hepatocellular carcinoma. PLoS ONE 2012, 7, e49462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Yan, J.; Guo, X.; Xia, J.; Shan, T.; Gu, C.; Liang, Z.; Zhao, W.; Jin, S. MiR-148a regulates MEG3 in gastric cancer by targeting DNA methyltransferase 1. Med. Oncol. 2014, 31, 879. [Google Scholar] [CrossRef]
  107. Lu, K.H.; Li, W.; Liu, X.H.; Sun, M.; Zhang, M.L.; Wu, W.Q.; Xie, W.P.; Hou, Y.Y. Long non-coding RNA MEG3 inhibits NSCLC cells proliferation and induces apoptosis by affecting p53 expression. BMC Cancer 2013, 13, 461. [Google Scholar] [CrossRef] [Green Version]
  108. Wang, P.; Ren, Z.; Sun, P. Overexpression of the long non-coding RNA MEG3 impairs in vitro glioma cell proliferation. J. Cell Biochem. 2012, 113, 1868–1874. [Google Scholar] [CrossRef]
  109. Ying, L.; Huang, Y.; Chen, H.; Wang, Y.; Xia, L.; Chen, Y.; Liu, Y.; Qiu, F. Downregulated MEG3 activates autophagy and increases cell proliferation in bladder cancer. Mol. Biosyst 2013, 9, 407–411. [Google Scholar] [CrossRef]
  110. Braconi, C.; Kogure, T.; Valeri, N.; Huang, N.; Nuovo, G.; Costinean, S.; Negrini, M.; Miotto, E.; Croce, C.M.; Patel, T. microRNA-29 can regulate expression of the long non-coding RNA gene MEG3 in hepatocellular cancer. Oncogene 2011, 30, 4750–4756. [Google Scholar] [CrossRef] [Green Version]
  111. Zhuo, H.; Tang, J.; Lin, Z.; Jiang, R.; Zhang, X.; Ji, J.; Wang, P.; Sun, B. The aberrant expression of MEG3 regulated by UHRF1 predicts the prognosis of hepatocellular carcinoma. Mol. Carcinog 2016, 55, 209–219. [Google Scholar] [CrossRef]
  112. Riley, T.; Sontag, E.; Chen, P.; Levine, A. Transcriptional control of human p53-regulated genes. Nat. Rev. Mol. Cell Biol. 2008, 9, 402–412. [Google Scholar] [CrossRef] [PubMed]
  113. Huarte, M.; Guttman, M.; Feldser, D.; Garber, M.; Koziol, M.J.; Kenzelmann-Broz, D.; Khalil, A.M.; Zuk, O.; Amit, I.; Rabani, M.; et al. A large intergenic noncoding RNA induced by p53 mediates global gene repression in the p53 response. Cell 2010, 142, 409–419. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Jiang, Y.J.; Bikle, D.D. LncRNA profiling reveals new mechanism for VDR protection against skin cancer formation. J. Steroid. Biochem. Mol. Biol. 2014, 144 Pt A, 87–90. [Google Scholar] [CrossRef]
  115. Işın, M.; Uysaler, E.; Özgür, E.; Köseoğlu, H.; Şanlı, Ö.; Yücel Ö, B.; Gezer, U.; Dalay, N. Exosomal lncRNA-p21 levels may help to distinguish prostate cancer from benign disease. Front. Genet. 2015, 6, 168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Yang, N.; Fu, Y.; Zhang, H.; Sima, H.; Zhu, N.; Yang, G. LincRNA-p21 activates endoplasmic reticulum stress and inhibits hepatocellular carcinoma. Oncotarget 2015, 6, 28151–28163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Yu, F.; Zhou, G.; Huang, K.; Fan, X.; Li, G.; Chen, B.; Dong, P.; Zheng, J. Serum lincRNA-p21 as a potential biomarker of liver fibrosis in chronic hepatitis B patients. J. Viral Hepat. 2017, 24, 580–588. [Google Scholar] [CrossRef]
  118. Schneider, C.; King, R.M.; Philipson, L. Genes specifically expressed at growth arrest of mammalian cells. Cell 1988, 54, 787–793. [Google Scholar] [CrossRef]
  119. Mourtada-Maarabouni, M.; Pickard, M.R.; Hedge, V.L.; Farzaneh, F.; Williams, G.T. GAS5, a non-protein-coding RNA, controls apoptosis and is downregulated in breast cancer. Oncogene 2009, 28, 195–208. [Google Scholar] [CrossRef] [Green Version]
  120. Sun, M.; Jin, F.Y.; Xia, R.; Kong, R.; Li, J.H.; Xu, T.P.; Liu, Y.W.; Zhang, E.B.; Liu, X.H.; De, W. Decreased expression of long noncoding RNA GAS5 indicates a poor prognosis and promotes cell proliferation in gastric cancer. BMC Cancer 2014, 14, 319. [Google Scholar] [CrossRef] [Green Version]
  121. Yacqub-Usman, K.; Pickard, M.R.; Williams, G.T. Reciprocal regulation of GAS5 lncRNA levels and mTOR inhibitor action in prostate cancer cells. Prostate 2015, 75, 693–705. [Google Scholar] [CrossRef] [Green Version]
  122. Wang, F.; Yuan, J.H.; Wang, S.B.; Yang, F.; Yuan, S.X.; Ye, C.; Yang, N.; Zhou, W.P.; Li, W.L.; Li, W.; et al. Oncofetal long noncoding RNA PVT1 promotes proliferation and stem cell-like property of hepatocellular carcinoma cells by stabilizing NOP2. Hepatology 2014, 60, 1278–1290. [Google Scholar] [CrossRef] [PubMed]
  123. Liu, E.; Liu, Z.; Zhou, Y. Carboplatin-docetaxel-induced activity against ovarian cancer is dependent on up-regulated lncRNA PVT1. Int. J. Clin. Exp. Pathol. 2015, 8, 3803–3810. [Google Scholar] [PubMed]
  124. Riquelme, E.; Suraokar, M.B.; Rodriguez, J.; Mino, B.; Lin, H.Y.; Rice, D.C.; Tsao, A.; Wistuba, I.I. Frequent coamplification and cooperation between C-MYC and PVT1 oncogenes promote malignant pleural mesothelioma. J. Thorac. Oncol. 2014, 9, 998–1007. [Google Scholar] [CrossRef] [Green Version]
  125. Yang, Y.R.; Zang, S.Z.; Zhong, C.L.; Li, Y.X.; Zhao, S.S.; Feng, X.J. Increased expression of the lncRNA PVT1 promotes tumorigenesis in non-small cell lung cancer. Int. J. Clin. Exp. Pathol. 2014, 7, 6929–6935. [Google Scholar] [PubMed]
  126. Wu, Q.; Yang, F.; Yang, Z.; Fang, Z.; Fu, W.; Chen, W.; Liu, X.; Zhao, J.; Wang, Q.; Hu, X.; et al. Long noncoding RNA PVT1 inhibits renal cancer cell apoptosis by up-regulating Mcl-1. Oncotarget 2017, 8, 101865–101875. [Google Scholar] [CrossRef] [PubMed]
  127. Zheng, J.; Yu, F.; Dong, P.; Wu, L.; Zhang, Y.; Hu, Y.; Zheng, L. Long non-coding RNA PVT1 activates hepatic stellate cells through competitively binding microRNA-152. Oncotarget 2016, 7, 62886–62897. [Google Scholar] [CrossRef]
  128. Yu, X.; Li, Z.; Zheng, H.; Chan, M.T.; Wu, W.K. NEAT1: A novel cancer-related long non-coding RNA. Cell Prolif 2017, 50. [Google Scholar] [CrossRef] [Green Version]
  129. Adriaens, C.; Standaert, L.; Barra, J.; Latil, M.; Verfaillie, A.; Kalev, P.; Boeckx, B.; Wijnhoven, P.W.; Radaelli, E.; Vermi, W.; et al. p53 induces formation of NEAT1 lncRNA-containing paraspeckles that modulate replication stress response and chemosensitivity. Nat. Med. 2016, 22, 861–868. [Google Scholar] [CrossRef]
  130. Idogawa, M.; Ohashi, T.; Sasaki, Y.; Nakase, H.; Tokino, T. Long non-coding RNA NEAT1 is a transcriptional target of p53 and modulates p53-induced transactivation and tumor-suppressor function. Int. J. Cancer 2017, 140, 2785–2791. [Google Scholar] [CrossRef] [Green Version]
  131. Idogawa, M.; Nakase, H.; Sasaki, Y.; Tokino, T. Prognostic Effect of Long Noncoding RNA NEAT1 Expression Depends on p53 Mutation Status in Cancer. J. Oncol. 2019, 2019, 4368068. [Google Scholar] [CrossRef] [PubMed]
  132. Qiu, J.; Chen, Y.; Huang, G.; Zhang, Z.; Chen, L.; Na, N. Transforming growth factor-β activated long non-coding RNA ATB plays an important role in acute rejection of renal allografts and may impacts the postoperative pharmaceutical immunosuppression therapy. Nephrology (Carlton) 2017, 22, 796–803. [Google Scholar] [CrossRef] [PubMed]
  133. Yuan, J.H.; Yang, F.; Wang, F.; Ma, J.Z.; Guo, Y.J.; Tao, Q.F.; Liu, F.; Pan, W.; Wang, T.T.; Zhou, C.C.; et al. A long noncoding RNA activated by TGF-β promotes the invasion-metastasis cascade in hepatocellular carcinoma. Cancer Cell 2014, 25, 666–681. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Fu, N.; Niu, X.; Wang, Y.; Du, H.; Wang, B.; Du, J.; Li, Y.; Wang, R.; Zhang, Y.; Zhao, S.; et al. Role of LncRNA-activated by transforming growth factor beta in the progression of hepatitis C virus-related liver fibrosis. Discov. Med. 2016, 22, 29–42. [Google Scholar] [PubMed]
  135. Jang, S.Y.; Kim, G.; Park, S.Y.; Lee, Y.R.; Kwon, S.H.; Kim, H.S.; Yoon, J.S.; Lee, J.S.; Kweon, Y.O.; Ha, H.T.; et al. Clinical significance of lncRNA-ATB expression in human hepatocellular carcinoma. Oncotarget 2017, 8, 78588–78597. [Google Scholar] [CrossRef]
  136. Negishi, M.; Wongpalee, S.P.; Sarkar, S.; Park, J.; Lee, K.Y.; Shibata, Y.; Reon, B.J.; Abounader, R.; Suzuki, Y.; Sugano, S.; et al. A new lncRNA, APTR, associates with and represses the CDKN1A/p21 promoter by recruiting polycomb proteins. PLoS ONE 2014, 9, e95216. [Google Scholar] [CrossRef] [Green Version]
  137. Luo, Y.; Yang, J.; Yu, J.; Liu, X.; Yu, C.; Hu, J.; Shi, H.; Ma, X. Long Non-coding RNAs: Emerging Roles in the Immunosuppressive Tumor Microenvironment. Front. Oncol. 2020, 10, 48. [Google Scholar] [CrossRef] [Green Version]
  138. Longo, V.; Brunetti, O.; Gnoni, A.; Licchetta, A.; Delcuratolo, S.; Memeo, R.; Solimando, A.G.; Argentiero, A. Emerging role of Immune Checkpoint Inhibitors in Hepatocellular Carcinoma. Medicina (Kaunas) 2019, 55, 698. [Google Scholar] [CrossRef] [Green Version]
Table 1. Long non-coding RNAs (lncRNAs) associated with both chronic liver disease and hepatocellular carcinoma (HCC).
Table 1. Long non-coding RNAs (lncRNAs) associated with both chronic liver disease and hepatocellular carcinoma (HCC).
LncRNAsLocationDys-RegulationAssociated HepatitisIn Liver FibrosisIn Hepatocellular CarcinomaReferences
Function Mechanism Function Mechanism
HULC6p24.3Uphepatitis B-Treg differentiation
-hepatocyte apoptosis
-p18 downregulation
-inhibit MAPK pathway
oncogene;
proliferation, EMT
- regulate cell cycle-related genes
- miR-200a-3p/ZEB1 pathway
[38,46,47,48,57,58,59,60]
MALAT111q13UpCCl4-Tx mice
NASH
hepatitis B
HSC activationactivate TGF-β pathway and Rac1oncogene- modulate apoptosis
- mTOR and Wnt pathway
[51,61,62,63]
HOTAIR12q13.13UpCCl4-Tx mice
hepatitis B
inhibit apoptosis
express fibrosis-related genes
- DNMT1-MEG3-p53 pathway
- PTEN methylation
oncogene - repress RBM38
- inactivate p16 (Ink4a) and p14 (ARF)
[51,52,53,54,64,65]
MEG314q32Down 1CCl4-Tx miceapoptosisactivate p53tumor suppressor; activate apoptosisp53 target gene expression[66,67]
LncRNAp216p21Downhepatitis BSuppress
HSC activation
-enhance PTEN/Akt pathway
-inhibit Wnt/β
-catenin pathway
tumor suppressorUnclear[68,69]
GAS51q25Downprimary HSCinhibit HSC activation and proliferationpromote p27 expressiontumor suppressorbind to miR-21[70,71]
PVT18q24Upprimary HSCEMT and
HSC activation
hedgehog pathwayoncogene- recruit EZH2,
- inhibit P53 expression
[72,73]
NEAT111q13Up 1CCl4-Tx miceHSC activationNeat1-miR-122-Klf6 axis proliferation and migration regulate hnRNP A2 [74,75]
LncRNA-ATB14Uphepatitis CHSC activationregulate TGF-β pathwayautophagyYAP and ATG5 expression[76]
Lnc-LFAR1 4q25 2UpCCl4-Tx miceHSC activation, hepatocyte apoptosisTGF-β and Notch pathwayunknownnot established[77]
HIF1A-AS114q23.2Downprimary HSCHSC inactivationinteract with TET3unknownnot established[78]
APTR7q11.23UpCCl4 and
BDL Mice 3
HSC activationTGF-β pathwayunknownnot established[79]
LncRNA-Cox2 1q25UpCCl4-Tx miceunknownnot establishedunknownnot established[80,81]
1 Controversial, 2 mouse chromosome, 3 including undisclosed etiology of human liver fibrosis; ATG5; autophagy-related gene 5, BDL; bile duct ligation, CCl4-Tx; CCl4-treated, DNMT1; DNA methyltransferase 1, EMT; epithelial mesenchymal transition, EZH2; enhancer of zeste homolog 2, hnRNP A2; heterogeneous nuclear ribonucleoprotein A2, HSC; hepatic stellate cell, Klf6; Kruppel-like factor, RBM38; RNA binding motif protein 38, TET3; ten-eleven translocation 3, YAP; yes-associated protein.

Share and Cite

MDPI and ACS Style

Kim, Y.-A.; Park, K.-K.; Lee, S.-J. LncRNAs Act as a Link between Chronic Liver Disease and Hepatocellular Carcinoma. Int. J. Mol. Sci. 2020, 21, 2883. https://doi.org/10.3390/ijms21082883

AMA Style

Kim Y-A, Park K-K, Lee S-J. LncRNAs Act as a Link between Chronic Liver Disease and Hepatocellular Carcinoma. International Journal of Molecular Sciences. 2020; 21(8):2883. https://doi.org/10.3390/ijms21082883

Chicago/Turabian Style

Kim, Young-Ah, Kwan-Kyu Park, and Sun-Jae Lee. 2020. "LncRNAs Act as a Link between Chronic Liver Disease and Hepatocellular Carcinoma" International Journal of Molecular Sciences 21, no. 8: 2883. https://doi.org/10.3390/ijms21082883

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop