Next Article in Journal
Mesenchymal Stem Cell Use in Acute Tendon Injury: In Vitro Tenogenic Potential vs. In Vivo Dose Response
Next Article in Special Issue
Mechanical Characterisation and Numerical Modelling of TPMS-Based Gyroid and Diamond Ti6Al4V Scaffolds for Bone Implants: An Integrated Approach for Translational Consideration
Previous Article in Journal
Initial Healing Effects of Platelet-Rich Plasma (PRP) Gel and Platelet-Rich Fibrin (PRF) in the Deep Corneal Wound in Rabbits
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Function of the Long Noncoding RNAs in Hepatocellular Carcinoma: Classification, Molecular Mechanisms, and Significant Therapeutic Potentials

College of Life Sciences, Zhejiang University, Hangzhou 310058, China
*
Author to whom correspondence should be addressed.
Bioengineering 2022, 9(8), 406; https://doi.org/10.3390/bioengineering9080406
Submission received: 2 July 2022 / Revised: 15 August 2022 / Accepted: 16 August 2022 / Published: 21 August 2022
(This article belongs to the Special Issue Scaffolds for Tissue Engineering and Regenerative Medicines)

Abstract

:
Hepatocellular carcinoma (HCC) is the most common and serious type of primary liver cancer. HCC patients have a high death rate and poor prognosis due to the lack of clear signs and inadequate treatment interventions. However, the molecular pathways that underpin HCC pathogenesis remain unclear. Long non-coding RNAs (lncRNAs), a new type of RNAs, have been found to play important roles in HCC. LncRNAs have the ability to influence gene expression and protein activity. Dysregulation of lncRNAs has been linked to a growing number of liver disorders, including HCC. As a result, improved understanding of lncRNAs could lead to new insights into HCC etiology, as well as new approaches for the early detection and treatment of HCC. The latest results with respect to the role of lncRNAs in controlling multiple pathways of HCC were summarized in this study. The processes by which lncRNAs influence HCC advancement by interacting with chromatin, RNAs, and proteins at the epigenetic, transcriptional, and post-transcriptional levels were examined. This critical review also highlights recent breakthroughs in lncRNA signaling pathways in HCC progression, shedding light on the potential applications of lncRNAs for HCC diagnosis and therapy.

Graphical Abstract

1. Introduction

Cancer, a major public health issue and one of the world’s lethal illnesses [1,2,3,4,5,6], is a multifaceted disorder characterized by uncontrolled cellular growth through genetic variations, epigenetic changes, chromosomal rearrangements, and amplification [7,8,9]. Various cancers have been linked to increased causes of death and mortality, despite the best efforts of experts to conduct comprehensive investigations to develop more effective therapeutic techniques [10,11]. As a consequence, finding effective screening tools, diagnostic biomarkers, and more effective treatment approaches to increase tumor patients’ long-term survival and treatment rates is essential [12,13,14].
Hepatocellular carcinoma (HCC), a severe specific type of primary liver cancer [15], accounts for 75–85% of cases of death [16,17,18,19,20]. The low survival rate of HCC is due to asymptomatic initiation in premature stages and loss of optimum treatment period after diagnosis in middle or late stages [17]. Hepatitis B or C virus infection, aflatoxin B1, drugs, alcohol, and metabolic diseases are the most common risk factors for HCC [21,22,23,24,25]. Moreover, there are a few uncommon diseases that increase HCC risks, including alpha1-antitrypsin deficiency (A1AT), tyrosinemia, and Wilson disease. Tyrosinemia is caused by deficiency of fumarylacetoacetate hydrolase, whereas Wilson’s disease is caused by a mutation in the ATP7B gene. All of these are complex multifactorial diseases that cause hepatotoxicity via various mechanisms, which can eventually lead to cirrhosis and HCC [24].
The initiation and persistence of HCC are complicated and are influenced by multiple variables [26,27]. HCC is associated with high levels of tumor growth, postoperative relapse, and chemo resistance [28,29,30,31,32]. Hepatic fibrosis, a wound-healing condition that involves the dysregulation of extracellular matrix proteins and alteration of normal hepatic architecture, is a significant risk factor for HCC [33,34,35]. The regulatory mechanisms involved in HCC are still hot topics [36,37]. HCC progression is a complex mediated through accumulation of genetic and epigenetic modifications [38,39], accumulating the necessary amount of genetic and epigenetic variations, leading to the formation of dysplastic foci and nodules, eventually progressing into HCC (Figure 1). To elucidate the HCC progression, landscape, and biology, a complete transcriptome study of the specimens demonstrating diverse disease stages may offer a higher-resolution view of the essential mechanisms of progression [40]. Nonalcoholic fatty liver disease (NAFLD) has been increasing in prevalence and is defined as excessive fat accumulation in the liver and steatosis presence in >5% of hepatocytes [41]. NAFLD can develop to nonalcoholic steatohepatitis (NASH), with inflammation and ballooning with or without fibrosis. NASH further develops to liver cirrhosis in a significant proportion of the patients and eventually progresses into HCC [41]. Generally, NAFLD HCC patients have poor prognosis and are associated with a more progressive stage of the disease [42]. Improved control of virus C illnesses and latent liver cancer has resulted in an increasing number of patients with restored liver function in response to cirrhosis [43]. Systemic chemotherapy, molecule-targeted therapy, trans catheter artery chemoembolization, and immunotherapy are the most common and effective treatments [27].
To formulate new diagnostic and therapeutic approaches against HCC and to enhance the prognostic value of diagnosed patients, it is important to reveal the relationship among signs, symptoms, and molecular alterations [44]. Advances in biomedical technology to date, such as live transplantation, surgical excision, and radiofrequency ablation, has increased the 5-year survival rates of HCC patients [45,46,47]. Additional molecular mechanisms and the development of reliable biological indicators for HCC detection are critical at the initial stage of HCC development [26,48,49,50,51,52].
Carcinogenesis is frequently caused by abnormal expressions of genes [53,54]. Recent evidence suggests that RNA processing has been consistently changed in cancer [55,56,57], revealing the critical role of RNA in tumor genesis and cancer development [54,58]. The long noncoding RNAs (lncRNAs) are mainly categorized according to their positional relationship with adjacent coding genes [59]. Several reports have been published in the scientific literature, highlighting a potential role for lncRNAs in tissue pathophysiology and development [20,49,60,61]. Evidence has shown that lncRNAs are mostly dysregulated as tumor suppressors in various cancers [20,62], and many lncRNAs are intricately linked to the progression of cancer, including HCC [25,63,64,65,66,67], signifying that lncRNAs are potential therapeutic targets in HCC [68].
In this review article, we present an overview of the existing knowledge on lncRNAs in HCC progression and analyze their mechanisms in the cancer phenotype. We also discuss the prospective application of lncRNAs as prognostic and therapeutic targets for HCC patients with future prospective to recognize diverse mechanisms of lncRNAs in HCC.

2. Characteristics and Classification of RNAs

RNA sequencing technology has identified more than one hundred thousand (100,000) distinct RNA molecules of mammalian species [69,70,71]. Coding RNAs and noncoding RNAs (ncRNAs) are the two types of RNAs [72,73]. Based on the length of transcripts, ncRNAs can be divided into two classes (small ncRNAs and long ncRNAs). The miRNAs (microRNAs), snRNAs (small nucleolar RNAs), PIWI-interacting RNAs, and other endogenous RNAs are examples of small ncRNAs [53,54], which have a nucleotide number of less than 200 [74]. LncRNAs (long noncoding RNAs), lincRNAs (long intergenic noncoding RNAs), NATs (natural antisense transcripts), T-UCRs (transcribed ultra-conserved regions), long enhancer ncRNAs, and noncoding repeat sequences, as pseudo genes, are examples of long ncRNAs with more than 200 nucleotides (Figure 2) [74].
Small ncRNAs were first identified by exogenous RNA interference (RNAi) in plants and nematodes and were found to exist endogenously, functioning mostly as gene regulators through pairing to the target genes, hence directing their post-transcriptional activities in animals and plants [75]. It is well known that ncRNAs account for the majority of the human transcriptome, including miRNAs, lncRNAs, and circRNAs. MicroRNAs are single-stranded RNAs and participate in a series of physiological and pathological processes by facilitating post-transcriptional regulation of the target genes [76]. Numerous abnormally expressing miRNAs are associated with HCC initiation and progression [76,77]. Various studies have exposed the biological roles of lncRNAs as regulators of transcription, modulators of mRNA processing, and organizers of nuclear domains [76,77,78,79]. Compared with linear RNAs, circRNAs are more stable to exonuclease and ribonuclease, with conserved structure and stable sequence and tissue specificity [78,79]. It has been shown that circRNAs play significant pathophysiological roles in the existence and development of alcoholic liver injury; hepatic fibrosis, HCC, and other liver diseases CircRNAs have also been confirmed to exert effects with respect to regulation of cellular metabolisms of HCC [78].
For example, small ncRNAs, siRNAs, and/or miRNAs, have been well characterized [74,80,81]. LncRNAs, in comparison with small ncRNAs, are less understood in terms of their mechanisms and functions [74]. The prevalence of various forms of RNAs is altered in most eukaryotic cells. Ribosomal RNAs are responsible for approximately 80–85% of cellular RNA mass, accompanied mostly by tRNAs and mRNAs [82]. Although ncRNAs are not translated into proteins, they play important roles in the physiological functions of organisms [69,83]. In particular, lncRNAs are essential controllers of chromatin dynamics, growth, differentiation, and gene development [20,84]. At present, with the advancement of high-throughput sequencing and DNA tiling array technology, a number of investigations are concentrating on ncRNAs [85,86,87]. The functions of ncRNA-encoding peptides and proteins have prospective applications in cancers, with some potential challenges [88].

3. Characteristics and Functions of lncRNAs

Relying on the genetic position concerning neighboring protein-coding genomes, lncRNAs have been classified into five categories (Figure 2) [89,90]. The first category is the sense lncRNAs, which interact with protein coding gene. The specific genes on the sense strand are transcribed from the sense strand of the genome concerning protein-coding genes, such as COLDAIR [9,91]. The second group of lncRNAs is the antisense lncRNAs, which are transcribed from the antisense strand of the genome, such as lncRNA ANRIL. These lncRNAs interact with one and sometimes most exons of the protein-coding genome upon its reverse strand [9,91]. The third category is the bidirectional lncRNAs; for example, the lncRNA-enhancing eNOS (endothelial nitric oxide synthase) expression (LEENE) and lncRNA HCCL5, in this category the lncRNA and a protein-coding gene are located on the opposite sides of the genome and are derived from different directions of protein-coding genes [9,17,91]. The fourth intronic lncRNAs are generated entirely within the introns of the protein-coding genes, with no exons overlapping [9,91]. The fifth group of intergenic lncRNAs is found nearby almost no protein-coding genes [9,17,91,92,93,94].
LncRNAs can also be classified according to their targeting mechanisms, including signal, decoy, and scaffolds [9,95]. The lncRNA signal can control cell-specific expression in response to numerous stimuli [9,96]. Some lncRNAs function as decoys to negatively regulate target expression, acting as a molecular basin to dilute the cellular level of protein or other miRNAs [95,97]. Some lncRNAs act as scaffolds to a prearranged telomerase complex by accumulating modular binding sites for telomeric regulatory proteins [17,95]. Many investigations have found that lncRNAs mainly interact with miRNAs to execute their biological functions as competing endogenous RNAs (ceRNAs) [28,98,99]. In turn, miRNAs may directly interact with lncRNAs to silence their expressions. Various lncRNAs are difficult to classify in specific classifications [20,28,49,98,99].
The amount of illustrated lncRNAs has changed dramatically in recent years due to sequencing technologies. More than 50,000 lncRNAs have been described, with almost 58,000 lncRNA transcripts assembled in the Encyclopaedia of DNA Elements (ENCODE), and Project Consortium (GENCODE release 36), with 27,919 lncRNAs of humans and the elevated 50 ending in the Functional Annotation of Mammalian Genome (FANTOM5) [25]. There may be more than 15,000 lncRNAs in the human genome. Their expression is highly regulated by transcription factors and methylated lysines, such as mRNAs [100]. A more specific definition of lncRNA is an RNA molecule that cannot code for proteins and has a length of 200 bp to 100 kbp [9,54]. LncRNAs can have an open reading frame of more than 100 amino acids [101]. Polypeptides with fewer than 100 amino acids can be useful in species and are not considered byproducts of authoritative proteins [101]. RNA polymerase II transcribes the largest portion of lncRNAs, which is most often capped and polyadenylated [102], unlike mRNAs, which are highly conserved between humans and rodents [103].
LncRNAs have species- and tissue-specific expression patterns, which may relate to their key roles [73,103]. The three fundamental levels of lncRNA structure and sequence composition are primary, secondary, and tertiary [104]. The structural properties of lncRNAs assist researchers to improve their understanding of the chemical mechanisms that enable lncRNAs to perform their roles. Secondary structures of lncRNAs typically include duplexes, internal loops, junctions, and bulges, which can serve as protein-binding sites and are important components of operational lncRNAs, such as Watson-Crick complementary base pairing and stability of unpaired locations [105,106]. Terminal differentiation-induced noncoding RNA is conserved at its 5′ ends across vertebrates other than mice [107], but the 3′ end indicates the difference in sequence in vertebrates [73,108]. The triple helix at the 3′ ends of lncRNAs can stabilize the poly (A) tail-lacking lncRNAs. It also contributes to the structure of lncRNAs by providing interactive interfaces and preserving lncRNA stabilization [29,109].
LncRNAs have been found to play important roles not only in the normal biological functions of cells but also in the pathophysiological behaviors of various illnesses. Particularly tumors, through chromosome alteration, splicing, transcription factor activation, mRNA fragmentation, and other mechanisms (Figure 3) [25,29,110,111,112,113,114,115,116]. LncRNAs are considered to have significant regulative functions in pathogenesis with respect to the development of various human diseases. Proliferation, apoptosis, differentiation, and tumor growth are only a few examples that describe the functions of lncRNAs [117,118,119]. LncRNAs are sometimes expressed abnormally in tumors [120]. They can function as oncogenes or tumor suppressor drivers [88,110,112]. Compared to protein-coding genes, lncRNA modifications are particular to tumors. This particularity provides lncRNAs with important diagnostic biomarkers [99,121,122,123,124,125]. In HCC, some lncRNAs play important controlling roles in the growth and metastasis of HCC [126,127,128] by halting the cell cycle, preventing cell death, and enhancing DNA injury repair. LncRNAs can perform significant functions with respect to chemo- and radio resistance of tumors [129], which could be used to identify possible targets and explore novel strategies for chemo- and radiotherapy in HCC [20,28,130].

4. Cancer-Associated lncRNAs

In comparison to healthy controls, most lncRNAs are found in patients with malignant tumors [25,49]. Due to high expression levels of lncRNAs in tumors, lncRNAs can be found in body fluids such as blood, saliva, and plasma, suggesting that circulating lncRNAs may be employed as non-invasive tools for diagnosis of various cancers, including HCC [25,49,131,132].
A number of mechanisms involving genetic, as well as environmental, changes are involved in transforming normal cells into cancer cells, with which they share some common characteristics [133,134,135]. To alter the cell physiology and regulate cancerous development, healthy cells must introduce new capabilities. Biochemical capabilities that are gained during the multiphase production of human tumors are considered the hallmarks of cancer [135]. Maintaining proliferative signaling, escaping progression suppressors, avoiding apoptosis, initiating angiogenesis, and inducing invasion and metastasis are all examples of such alterations [136]. LncRNAs are related to nearly all cancer hallmarks [90,137]. The manipulation of diverse mechanisms is responsible for the effect of such lncRNAs in cancer hallmarks [71,135,138].
A strong relationship between tumors and lncRNAs has been identified [139]. Differential expression of lncRNAs in nearby normal and tumor tissues, as well as in normal and malignant cell lines, makes lncRNAs potential cancer biomarkers [67,125,139]. Because lncRNAs can change cell growth by modifying expressions of genes, dysregulated expressions of lncRNAs may contribute to cancer pathophysiology [74]. In such situations, the changes within lncRNAs are linked to cancer. LncRNAs can be used as potential therapeutic biomarkers [140,141,142,143,144,145,146]. Cancer-causing and anticancer lncRNAs are two types of lncRNAs in tumors (Table 1). Due to their ability to interact with molecules of DNA, protein, and RNA, as well as the ability to alter many cancer hallmarks, lncRNAs play important roles in tumor progression [73,138,141,142]. The splicing of precursor mRNAs is affected by lncRNA-mediated gene expression control, which occurs in the post-transcriptional stage. The stability of mRNAs and proteins, as well as nuclear trafficking, is factors to be considered [147,148,149,150]. More than 8000 lncRNAs have been discovered in cancer cells [151,152]. Owing to their considerable quantity and specificity of expression, such lncRNAs are effective biomarkers and strong therapeutic targets (Table 1).

5. LncRNAs in HCC

LncRNAs perform important functions with respect to the induction and development of HCC [69], with increased expression levels of 27 kinds of lncRNAs. Actin filamentin-1 antisense RNA (AFAP-AS1), zinc finger E-box binding homeobox 1-antisense 1 (ZEB-1-AS1), and HOX transcript antisense intergenic RNA (HOTAIR) are correlated with poor prognosis of HCC, whereas reduced expressions of 18 lncRNAs, including growth arrest-specific transcript 5 (GAS5), XIST, and maternally expressed gene 3 (MEG3), are correlated with an even worse prognosis of HCC [28]. By inducing the invasion and initiation of metastatic spread of HCC cells, lncRNA SNHG8 (small nucleolar RNA host gene 8), LINC00052, lncRNA W42 [67], LINC01225, PITPNA antisense RNA 1 (PITPNA-AS1), and ZEB1-AS1 exhibit oncogenic characteristics [120,174,175]. LncRNA-hPVT1, AFAP1-AS1, XIST, HOXA cluster antisense RNA 2 (HOXA-AS2), HOST2, and cervical carcinoma high-expressed lncRNA 1 (CCHE1) are examples of lncRNAs that can cause and promote cell proliferation, suppressing apoptosis of HCC cells [28,176]. LncRNA SNHG17 is considerably upregulated in tissues and cell lines of HCC and associated with large tumor size, poor differentiation, and the presence of vascular invasion [177]. The lncRNA TUG1-miR328-3p-SRSF9 mRNA axis works as a unique ceRNA regulator axis related to HCC malignancies [178]. LINC01194 is upregulated in the HCC cell line and controls the proliferation and migration of HCC cells by interacting with the miR-655-3p/SMAD5 axis, which provides new biomarkers for HCC diagnosis and treatment [179]. The increasing appearance of lncRNAs in HCC is assumed to be oncogenic, and lncRNAs with low expression in HCC are considered to be tumor-suppressor lncRNAs [28,99]. Those characteristics may be effective as potential therapeutic targets for HCC, especially for patients who may have already developed resistance to chemotherapeutic drugs [28,30,31,32,180]. The functions of lncRNAs in HCC are to promote cancer cell growth and invasion, repress cancer cell growth and invasion, estimate prognosis and efficacy, and act as potential biomarkers (Table 2).
The biosynthesis of lncRNAs is similar to that of protein-coding transcripts. Epigenetic modification, transcription complex recruitment, and RNA processing are all important activities that influence lncRNA production. Aberrant lncRNA biosynthesis is related to the pathogenesis of various diseases, including HCC. In comparison to non-cancerous liver tissues, high-throughput techniques such as RNA sequencing and microarray have characterized distinct lncRNA expression patterns within HCC tissues, demonstrating that lncRNA production is dysregulated throughout HCC progression [189,239]. Aberrant biogenesis activities include epigenetic activation of tumor suppression. LncRNA transcriptional repression through certain tumor-suppressive transcription factors, special processing patterns that associate lncRNAs with oncogenic activities and the binding of lncRNAs with miRNAs affect lncRNA stability [195].

5.1. Regulation and Modification of Chromatin by lncRNAs

Increasing evidence has shown that lncRNAs can perform a variety of functions, including epigenetic modifications in HCC [240,241] (Figure 4a). Methylation of histone and DNA is an essential epigenetic modulation that regulates gene expressions [24,242,243,244]. Inappropriate chromatin alterations of lncRNA genes, such as methylation of DNA histone modification, have generally been described throughout HCC development [245], which can cause a reduction in repressive lncRNAs of HCC and an increase in cancer-promoting lncRNAs related to HCC [195]. Linc-GALH (Gankyrin-associated LincRNA in HCC), with respect to judgment of HCC metastasis, can promote DNMT1 (DNA methyltransferase1) degradation by enthusing ubiquitination and appearance of Gankyrin (PSMD10) and decreasing HCC methylation [246]. EMT (epithelial-mesenchymal transition) is thought to be essential for tumor metastasis and relapse [247]. The up regulation of linc00441 increases H3K27 acetylation [248]. In contrast, abundantly expressed linc00441 induces DNA methyltransferases 3 alpha (DNMT3A) to methylation, deactivating the neighborhood RB1 gene to induce HCC cell proliferation [195,248]. Significantly increasing lncRNAs has been demonstrated to show their interaction with epigenetic regulator enhancer of zest homolog 2 (EZH2) to stimulate gene expression, influencing HCC metastasis [195].

5.2. Transcriptional Regulation and Activation

LncRNAs can regulate transcription by binding to promoters of nearby or distinct genes to recruit transcription factors to further regulate transcriptional activation [95] (Figure 4b). LncMAPK6, a mitogen-activated protein kinase 6 (MAPK6) lncRNA, has been abundantly expressed in association with liver tumor development [249]. Its interaction with RNA polymerase II recruits MAPK6 promoter, thus activating MAPK6 transcription [249]. YAP, c-Myc, and catenin are the oncogenic transcription factors that are highly expressed in HCC [250]. These transcription factors enable lncRNAs to be involved in necroptosis and cell cycle arrest in HCC, demonstrating that lncRNA transcription is critical throughout HCC [195,251].

5.3. Interaction with mRNAs

Certain lncRNAs affect the stabilization and translating procedures of mRNAs (Figure 4c). Through intermodulation with lncRNA-mRNA, lncRNAs activated by transforming growth factor-β (lncRNA-ATB) regulate and maximize mRNA of interleukin-11, encouraging the proliferation of circulated HCC cells in distant parts of the body [252]. Only a few primary lncRNA transcripts are affected by the exon-insertion scenario. During the formation of HCC, such process modules can play an oncogenic role. The splicing factor muscle blind-like 3 (MBNL3) was found to be overexpressed throughout the fetal liver in HCC tissues that were lacking in adults, causing LncRNA PXN-AS1 (PXN antisense RNA 1) exon 4 inclusions. Due to the splicing alteration, the lncRNA PXN-AS1 is able to interact with PXN mRNA in HCC [253]. HCC progression and metastatic spread are inhibited through lncRNA LINC01093 specific to the liver, which also acts as protein scaffolding on the way to induce insulin-like growth factor mRNA binding protein1 (IGF2BP1) and to further promote the degradation of Glioma-associated oncogene homologue 1 (GLI1) mRNA [254].

5.4. Sponge of MicroRNAs

LncRNAs can function as miRNA sponges, reducing deficit miRNA activity (Figure 4d). LncRNAs have been used as an extra layer of post-transcriptional regulation of gene expression [234,255]. Numerous lncRNAs have been implicated in controlling the expression of genes by interfering with miRNAs and prohibiting particular miRNAs from binding with the target mRNAs [256,257,258,259]. The HCC-associated lncRNA (HCAL) stimulates HCC metastasis by binding miR-196a, miR-196b, and miR-15a [260]. LncRNAMALAT1 (metastasis-associated lung adenocarcinoma transcription 1) can promotes migration and invasion of HCC via sponging of miR-204 [261]. In patients with HCC, a sufficient proportion of lncRNA HOXD-AS1 (HOXD cluster antisense RNA 1) expression is related to the high tumor nod [200,262]. HOXD-AS1 conservatively binds with miR-130a-3p, which can inhibit SOX4 (sex-determining region Y-related high-mobility group box transcription factor 4) to miRNA intermediated destruction, stimulating the expression of EZH2 and MMP2 to promote HCC metastasis [200]. HOXD-AS1 can also regulate the expression of Rho GTPase-activating protein 11A through highly competitive interaction with miR-19a, resulting in HCC tumor growth [262]. LncRNA HULC (highly upregulated in liver cancer) enhances HCC progression and metastasis by increasing epithelial–mesenchymal transition (EMT) progression in the miR-200a-3p/ZEB1 signaling pathway [263]. LncRNA MALAT1 enhances HCC development by sponging miR-143-3p to control ZEB1 expression [190]. MiR-34a has been reported to bind specifically with lncRNA-UFC1, causing its half-life to be reduced, thus preventing HCC growth mediated by lncRNA-UFC1 [189]. Because miRNAs are downregulated generally during HCC production, oncogenic lncRNAs are likely to be reactivated, leading to abnormal lncRNA expression profiles [195].

5.5. Protein Binding and/or Modification

Except for binding with miRNAs, lncRNAs are subject to biochemical processes that include protein modification (Figure 4e). Many reports have suggested that lncRNAs perform roles in protein phosphorylation modulation [264,265]. LncRNA TSLNC8 (tumor-suppressive lncRNA on chromosome 8p12) inhibits phosphorylation of STAT3 (signal transducer and activator of transcription 3) in HCC by inactivating the IL-6/STAT3 signaling pathway [264,265]. RNA-binding proteins (RBPs) have been discovered to manipulate lncRNA stabilization via physical interaction [266]. IGF2BP1/3 (insulin-like growth factor 2 mRNA-binding protein 1/3) is an RBP that binds and remains stable with long intergenic non-protein coding RNA 1138 (LINC01138) on its 220-1560-nt fragment, which is essential for HCC invasion progression [266]. Furthermore, lncRNA UFC1 can interact with another RBP, called human antigen R (HuR), via its fragment (1102-1613-nt), which is required for HCC [189]. These findings indicate that RBP-controlled lncRNA decay occurs to compensate for unusual lncRNA biogenesis in HCC.
Some reports have shown that lncRNA HNF1A-AS1 (HNF1A antisense RNA 1) prevents HCC invasion and spread by directly attaching to the C terminal of SHP-1 (SH2-containing protein tyrosine phosphatase 1), thereby stimulating phosphatase [267]. The effect of lncRNAs on the expression of genes by modification of protein is not restricted to target protein phosphorylation. In HCC, LINC01138 can exert oncogenesis behavior by interfering with arginine methyltransferases 5 (PRMT5), strengthening the stability of protein by stopping ubiquitin degradation [266]. LncRNA miR503HG comes into contact mostly with heterogeneous nuclear ribonucleoprotein A2/B1 (hnRNPA2B1) and represses metastatic tumor repression by controlling the ubiquitination status of hnRNPA2B1 [231]. By hindering CUL4A (cullin4A) intermediated ubiquitination and degradation of LATS1 (long-acting thyroid stimulator 1) within the cytoplasm, lncRNA uc.134 can disrupt HCC invasion and metastasis [224]. In particular, lncRNAs affect protein acetylation, which is an essential post-translational modification of protein control degradation [268]. Histone deacetylase 3 (HDAC3) governs lncRNA-LET (low expression in the tumor), which may be implicated in hypoxia-induced cell death [268]. LncRNA-LET inhibits Nuclear Factor 90 (NF90) protein degradation but is essential for hypoxia-induced cellular penetration [268]. LncRNAs can have a variety of effects on the formation of HCC.

5.6. Other Mechanisms and Pathways of lncRNAs in HCC

LncRNAs have significant effects on transcriptional, as well as post-transcriptional, regulation, relying on their subcellular localization [25]. Trans-acting nuclear lncRNAs control gene transcription epigenetically by interacting with tissue-specific chromatin modifications, such as histone-modifying complexes and DNA methyltransferases [269] (Figure 4f). Certain lncRNAs manage to sustain nuclear architecture through the scaffolding structure of the DNA-RNA-protein framework at unique sites [270,271]. Due to the genomic similarity toward their targets, the cis-acting lncRNAs may become capable of controlling gene expression inside the locus with an allele-specific method [270,271].
In cancer, lncRNAs are involved in tumor proliferation and metastasis signaling pathways [272]. The crucial mediator throughout the development of cancer is significant in HCC development and progression [273,274]. According to increasing prevalence, triggering of the catenin cascade can play a vital role in HCC [275]. Several lncRNAs play key roles in the stimulation and repression of the catenin pathway in HCC [271]. Overexpression of long intergenic non-protein-coding RNA 00210 (LINC00210) in liver tumor tissues interferes with catenin beta-interacting protein 1 (CTNNBIP1) to block the inhibitory function of CTNNBIP1 in catenin stimulation and enhance the association of catenin and TCF/LEF (T-cell factor/lymphoid enhancer factor family) complex, thereby triggering catenin signaling and liver tumor growth [276]. Some other pathways seem to be lncRNA-activated through TGF (lncRNA-ATB), further inducing EMT and aggression via highly competitive binding the miR-200 family but also modulating ZEB1 and ZEB2 [277]. In HCC, lncRNA-HEIH (HCC upregulated EZH2-associated lncRNA), in combination with enhancer of zeste homolog 2 (EZH2), performs very significant roles in G0/G1 arrest, usually requiring suppression of the EZH2 target gene [44]. Higher URHC, upregulated in HCC, can induce cell proliferation and prevent cell death by suppressing the sterile alpha motif and leucine zipper-containing kinase AZK (also known as ZAK (zipper-containing kinase) [278]. Two single nucleotide polymorphisms, rs7763881 within HULC and rs619586 within MALAT1, exist in 1344 HBV-persistent drivers and 1300 HBV-positive HCC patients [279]. Interactions of lncRNAs with some other significant signaling pathways participating in HCC metastasis and growth have been identified [280,281]. Through upregulation of PTTG1 (pituitary tumor-transforming gene 1) to trigger the PI3K/AKT signaling pathway, lncRNA PTTG3P (pituitary tumor-transforming 3 pseudo gene) enhances HCC development, as well as tumor growth [282]. HCC metastasis-promoting linc-GALH is known to be implicated in the regulation of the AKT signaling pathway [246,283]. Linc00974 also encourages the growth and migration in HCC by interfering in KRT19 (Keratin 19) [284]. LncRNA uc.134 stimulates hippo kinase signaling by preventing CUL4A from moving to the cytoplasm from the nucleus [224]. These findings illustrate that lncRNAs can function as mediating variables of the oncogenesis signaling pathways such, as Hippo kinase, Wnt, JAK/STAT, and PI3K/AKT. Although it is still unknown how lncRNAs affect HCC development, the relationship between lncRNAs and signaling pathways has paved the way for both the identification of innovative diagnostics and therapy in HCC [244,285].

6. Importance of Gene Expression Regulation in HCC Progression

HCC onset and development can be assessed using global genomic research due to genetic alterations that alter the expression of thousands of cancer-related genes. Hepatocarcinogenesis and the molecular pathways that underpin complicated clinical features have been studied using HCC gene regulation analysis [286,287]. The development of phenotypic expression gene profiling could revolutionize how HCC is identified and treated [286,287]. Complementary DNA microarrays for analysis of global gene expression, single-nucleotide polymorphism genotyping for identification of mutations that significantly alter gene expression and abnormal protein activities, chromosome instability mapping, and DNA–protein interactions are all widely accepted genomic data analysis technologies. In addition, several functional groups are used to develop new HCC serum diagnostic markers and therapy targets [287]. Although cancer cells disrupt EMT, it is a straightforward physiological activity that involves development and wound repair. In HCC, EMT effectors, such as fibronectin, cadherins, integrins, and vimentin, have been found to be altered, allowing for a much more mesenchymal phenotype [39,288,289,290]. In HCC, transcription factors that promote EMT, such as slug, twist, Snail, and Zeb, are upregulated [39,288,289,290]. Furthermore, the majority of studies on miRNAs, exosomes, lncRNAs, and regulatory cellular processes have been associated with EMT and found to be important in the advancement of HCC [39,288,289,290]. During primary HCC, the hypoxic microenvironment is significantly related to cancer development and angiogenesis [291]. Cancer cells interact with the aberrant microenvironment, ECM, cytokines, and chemokines and elevate the growth factors, resulting in enhanced angiogenesis [292,293]. Hepatic cells play a significant role in hepatocarcinogenesis, and the transformation of all such cells can result in cancer stem cells (CSCs) with various intrinsic factors (genetics and autoimmune diseases) and various extrinsic factors (HBV, HCV, alcohol, and AFB1), accounting for about 70–90% of the conversion of tiny hepatocyte-like progenitor cells into cancer cells [294]. Several potential surface markers of liver CSCs, such as epithelial cell adhesion molecule (EpCAM) [295], CD90 [296], CD133 [297], CD44 [298], and CD13 [299], have been identified. However, an improved understanding as to how molecular categorization and mutational confirmations influence HCC progression is required before it can be used as a targeted therapy in a medical context [296].

7. LncRNAs as Diagnostic and Therapeutic Markers in HCC

In HCC patients who are diagnosed later in the disease process, curative medications are no longer valuable [265]. Currently, ultrasound imaging and alpha-fetoprotein (AFP) analysis are used to diagnose HCC. Ultrasound scanning and testing are recommended in high-risk populations, and patients who undergo increasingly regular imaging have been associated with improved prognosis [300]. Nevertheless, with 47% sensitivity, surveillance imaging is insufficient to detect early-stage HCC [301]. The commonly used HCC biomarker AFP (alpha-fetoprotein) seems to have a sensitivity of 52.9%, as well as a specificity of 93.3%, which can be strengthened when combined with ultrasound imaging [302]. In the absence of HCC, some variables, including HCV infection, have also been reported to increase AFP levels [303]. However, neither ultrasound imaging nor AFP analysis reduces HCC patient mortality [304]. In early HCC, surgical procedures, including resection and liver transplantation, remain the only therapeutic choices, whereas late-stage HCC is essentially untreatable. To develop the diagnosis and treatment of HCC, new biomarkers and targeted therapies are critically required [305]. Metastasis seems to be a significant factor affecting long-term survival in patients with severe HCC [306].

7.1. LncRNAs as a Potential Biomarker of HCC

Patients with HCC who are diagnosed early have an increased chance of survival. Because of their tissue specificity, lncRNAs are intriguing as biomarkers [265]. It would be more appropriate to use circulating lncRNAs throughout the body fluid instead of some in malignant tissues as non-invasive markers for cancer diagnosis and surveillance [265]. However, most lncRNAs have been shown to exhibit uneven expression levels in some cancers and non-cancerous illnesses, such as cirrhosis and liver damage, resulting in diminished consistency [307]. As a result, combining lncRNAs with other chemicals, such as the well-known HCC biomarker AFP, makes a successful HCC diagnosis considerably more likely. Multiple lncRNAs, for example, UCA1 and WRAP53, in combination with AFP, ensure up to 100 percent responsiveness [307]. Similarly, combining two lncRNAs, PVT1 and uc002mbe.2, along with AFP, has been shown to serve well in the diagnosis of HCC relative to AFP alone [195,308].
As reported, lncRNA ZFAS1 (zinc finger antisense 1) is a new serum diagnostic marker for the detection of HCC [309]. The extracellular vesicle long RNAs (exLRs), which were found only in blood samples of 104 patients with HCC, can effectively distinguish HCC from non-tumor controls [310]. Consequently, a combination of serum exosomal ENSG00000258332.1 and LINC00635 with AFP is a reliable tool for HCC diagnosis [311]. LncRNA associated with micro vascular invasion in HCC (LncRNA MVIH) up regulation has been found to significantly predict persistent relapse in initial HCC patients, indicating that MVIH might be a useful marker for the early detection and individual care assessment of HCC patients [244,312,313]. The combination of XLOC014172, LINC00152, and RP11-160H22.5 could differentiate HCC patients from hepatitis patients [314]. Furthermore, the lncRNA gene polymorphism is important for HCC diagnosis [315].

7.2. LncRNAs as Promising Therapeutic Potentials for HCC

In addition to their potential use as diagnostic biomarkers, lncRNAs have important therapeutic techniques for new treatments of HCC [195]. The base-pairing paradigm RNA-targeting methods are simpler to implement than protein-targeting approaches. Antisense oligonucleotides (ASOs) and RNAi are the most widespread oncogenic lncRNA-targeting techniques for the treatment of HCC [316,317,318,319]. The infusion of ASOs, such as MALAT1, inhibits tumor growth in HCC-bearing nude mice [265,276,320]. ASO-mediated linc00210 absence inhibits HCC cell self-renewal and aggression, but knockout of lncRNA CASC9 (cancer susceptibility candidate 9) by RNAi decreases cancer progression in HCC [194,276]. Using precisely constructed siRNAs against lncRNAs is a technique for influencing lncRNA efficiency. The use of artificial lncRNA has been proposed to specifically target many miRNAs and may be a useful approach for resolving Sorafenib resistance in the HCC medication [321].
Discovery of more operative treatments is imperative. Recent findings have shown that a combination of atezolizumab and bevacizumab results in antitumor activity in patients with unresectable HCC [322]. Taraxacum officinale (L.) Weber ex F.H. Wigg, a perennial member of the Compositae family, has antitumor properties in HCC cells and has long been conventionally used as Chinese herbal medicine for liver, breast and gallbladder, hepatitis, as well as digestive, diseases [323]. According to the US Food and Drug Administration, the medication for HCC first-line therapies are bevacizumab in combination with atezolizumab, Sorafenib, and Lenvatinib; the second-line therapies include cabozantinib, pembrolizumab, ramucirumab, and regorafenib, in addition to other agents, such as bevacizumab, nivolumab, and nivolumab in combination with ipilimumab [324].
Among first-line treatments, atezolizumab in combination with bevacizumab has the highest overall survival (OS) value, although lenvatinib has the highest objective response rate (ORR) value. Among second-line treatments, cabozantinib has the highest progression-free survival (PFS) value, as well as ORR value, compared to placebo [325]. Sorafenib, the RTK-targeting drug, is perhaps the most commonly used effective medication for the treatment of HCC. LncRNA-targeting methods have certain benefits over protein-targeting approaches for the treatment of HCC [326].
Recent advancements in molecular cell biology have significantly contributed to our awareness of the molecular mechanisms of tumor genesis and its development, which, in turn, offers prospects for finding of new molecularly targeted agents to prevent molecular irregularities as promising cancer treatments [135]. Molecularly targeted treatment generally includes TKIs (tyrosine kinase inhibitors), as well as monoclonal antibodies. Five targeted therapies have been approved for treatment of progressed HCC. Among these five therapies, four are small-molecule kinase inhibitors, and the one is a monoclonal antibody against VEGFR2 (vascular endothelial growth factor receptor) [327]. In addition to the mentioned appropriate targeted therapies, various targeted therapies are in clinical trials.
Knocking out oncogenic lncRNAs and injection tumor-suppressor lncRNA may be acceptable strategies for HCC treatment. As reported, lncRNA PRAL, a tumor suppressor that acts by stabilizing p53, dramatically prevents HCC development in tumor-bearing mice [239]. ASOs and RNAi function depending on a variety of factors, such as the subcellular positioning of the target lncRNAs. ASOs perform better than RNAi in nuclei, but RNAi performs better than ASOs when it targets cytoplasmic lncRNAs [328], which may be why RNaseH is primarily found in the nucleus, although RISC is primarily found in the cytoplasm [329,330].

8. Future Prospects and Conclusions

Cancer-related lncRNAs are slowly but steadily becoming the most widely discussed themes, even in RNA biology, as well as oncology. According to the existing data, abnormal transcription and processing activities may result in up regulation of the tumor-promoting lncRNAs that mostly interact with DNA, RNA, and proteins. As a consequence, lncRNAs can control expression, function, and some similar characteristics of their partner binding sites, causing various cancerous phenotypes, including recurrent proliferation, irregular metabolism, and tumor growth. All of these contribute to HCC carcinogenesis and development. Given their critical functions, a subclass of lncRNAs found in body fluid may be used as HCC biomarkers, either alone or in association with other metabolites to increase specificity.
As a result, altering lncRNA expression could be a new diagnostic and treatment technique for HCC [195]. According to the US food and Drug Administration, tyrosine kinase inhibitors Sorafenib and Lenvatinib have been proven as first-line treatments, and now, bevacizumab, in combination with atezolizumab, Sorafenib, and Lenvatinib, is considered the first-line treatment for accelerated HCC [324,331].
Some lncRNAs linked to inflammatory signaling pathways, such as the IL-6/STAT3 and NF-B pathways, have been discovered. However, the exact regulatory systems that govern development from inflammation to neoplasia remain unknown. The liver is responsible for lipid metabolism and is the primary location for endogenous cholesterol metabolism. Abnormalities in these metabolic pathways promote HCC etiology, as indicated by the increased risk of HCC in patients with diabetes, extreme obesity, and hepatic steatosis [265]. Whereas significant progress has been made, the activities of lncRNAs remain unknown. LncRNAs are often questioned based on the lack of functional analyses that may be attributed to the lower sequence conservation in comparison to protein-coding genes [332]. LncRNAs prefer to sustain highly preserved secondary structures [333]. The important problem at present is thoroughly attempting to understand the main aspects of lncRNAs, such as their structures, functions, expressions, and related mechanisms. Improved statistical techniques for lncRNA biological activities can assist in identifying their significance with respect to various cancers. This knowledge may open the way for lncRNAs as potential prognostic markers and possibly even targeted therapies. In particular, strategies to target lncRNAs, including the use of siRNAs to initiate lncRNA deterioration and CRISPR/Cas9-mediated editing of the gene, must be regarded and improved. It is difficult to determine how to get the perfect molecules into appropriate cells [334]. RNA-seq is used to determine the differential expression of lncRNAs amongst tumor and non-tumor cells in an effort to explain active lncRNAs in HCC. Activities of lncRNAs are not always reflected in their variable expression patterns. Various genetic strategies are needed to illustrate lncRNA activities, which appears to be a challenging task, given thousands of lncRNAs that can only be identified simultaneously. The CRISPR sequencing technique is used to investigate the roles of protein-coding genes and lncRNAs related to screening phenotypes, proliferation, and drug resistance [335,336,337]. CRISPR analysis not only allows for identification of new functional lncRNAs that affect phenotypes of concern but makes it easier to create lncRNA-based potential therapies for a variety of human diseases [25].

9. Conclusions

Translating lncRNA studies into potential treatments is complex. LncRNA-based identification strategies are slowly emerging. The involvement of lncRNAs in regulation is related to the development of HCC. There are many unanswered questions at present. Future studies should concentrate on the functions and molecular pathways of lncRNAs in stimulating HCC development rather than just the concise recognition of differentially regulated and expressed lncRNAs. The main objective of gaining an improved appreciation of lncRNAs in HCC is to find new targeted therapies and biomarkers for HCC.

Author Contributions

A.K. collected and analyzed the data and wrote the paper. A.K. and X.Z. analyzed the data and wrote the paper. X.Z. conceived and designed this study and analyzed the data. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data generated or analyzed during this study are included in the review manuscript.

Acknowledgments

The authors wish to thank their parent institutions for providing the necessary facilities to complete the current research.

Conflicts of Interest

The authors declare that they have no competing interest.

Abbreviations

HCChepatocellular carcinoma
LncRNAslong noncoding RNAs
snRNAssmall nuclear RNAs
ncRNAsnoncoding RNAs
siRNAssmall interfering RNA
mRNAsmessenger RNAs
tRNAstransport RNAs
rRNAsribosomal RNAs
ceRNAscompeting endogenous RNAs
DNMT1DNA methyltransferase1
MAPK6lncMAPK6, mitogen-activated protein kinase 6
HULChighly upregulated in liver cancer
LATS1long-acting thyroid stimulator 1
HDAC3histone deacetylase 3
EMTepithelial-mesenchymal transformation
URHCupregulated in HCC
PTTG3Ppituitary tumor-transforming 3 pseudo gene
AFPalpha-fetoprotein
ASOsantisense oligonucleotides
MALAT1metastasis-associated in lung adenocarcinoma transcript
TINCRterminal differentiation-induced noncoding RNA

References

  1. Bray, F.; Ferlay, J.; Soerjomataram, I.; Siegel, R.L.; Torre, L.A.; Jemal, A. Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2018, 68, 394–424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Fitzmaurice, C.; Akinyemiju, T.F.; Al Lami, F.H.; Alam, T.; Alizadeh-Navaei, R.; Allen, C.; Yonemoto, N. Global, Regional, and National Cancer Incidence, Mortality, Years of Life Lost, Years Lived with Disability, and Disability-Adjusted Life-Years for 29 Cancer Groups, a Systematic Analysis for the Global Burden of Disease Study. JAMA Oncol. 2018, 4, 1553–1568. [Google Scholar] [PubMed]
  3. Ferlay, J.; Foucher, E.; Tieulent, J.; Rosso, S.; Coebergh, J.W.; Comber, H.; Forman, D.; Bray, F. Cancer incidence and mortality patterns in Europe: Estimates for 40 countries in 2012. Eur. J. Cancer 2013, 49, 1374–1403. [Google Scholar] [CrossRef] [Green Version]
  4. Ferlay, J.; Foucher, E.; Tieulent, J.; Rosso, S.; Coebergh, J.W.; Comber, H.; Forman, D.; Bray, F. Cancer incidence and mortality patterns in Europe: Estimates for 40 countries and 25 major cancers. Eur. J. Cancer 2018, 103, 356–387. [Google Scholar] [CrossRef] [PubMed]
  5. Döbrössy, L. Cancer mortality in central-eastern Europe: Facts behind the figures. Lancet Oncol. 2002, 3, 374–381. [Google Scholar] [CrossRef]
  6. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Jemal, A. Cancer statistics. CA Cancer J. Clin. 2021, 71, 7–33. [Google Scholar] [CrossRef]
  7. Bhan, A.; Soleimani, M.; Mandal, S.S. Long noncoding RNA and cancer: A new paradigm. Cancer Res. 2017, 77, 3965–3981. [Google Scholar] [CrossRef] [Green Version]
  8. Glassman, M.L.; de Groot, N.; Hochberg, A. Relaxation of imprinting in carcinogenesis. Cancer Genet. Cytogenet. 1996, 89, 69–73. [Google Scholar] [CrossRef]
  9. Nandwani, A.; Rathore, S.; Datta, M. LncRNAs in cancer: Regulatory and therapeutic implications. Cancer Lett. 2021, 501, 162–171. [Google Scholar] [CrossRef]
  10. Chen, W.; Sauer, A.M.G.; Chen, M.S., Jr.; Kagawa-Singer, M.; Jemal, A.; Siegel, R.L. Cancer statistics in China. CA Cancer J. Clin. 2016, 66, 115–132. [Google Scholar] [CrossRef] [Green Version]
  11. Torre, L.A.; Siegel, R.L.; Ward, E.M.; Jemal, A. Global cancer incidence and mortality rates and trends—An update. Cancer Epidemiol. Biomark. Prev. 2016, 25, 16–27. [Google Scholar] [CrossRef] [Green Version]
  12. Gotwals, P.; Cameron, S.; Cipolletta, D.; Cremasco, V.; Crystal, A.; Hewes, B.; Mueller, B. Prospects for combining targeted and conventional cancer therapy with immunotherapy. Nat. Rev. Cancer 2017, 17, 286–301. [Google Scholar] [CrossRef] [PubMed]
  13. Baudino, A.T. Targeted cancer therapy: The next generation of cancer treatment. Curr. Drug Discov. Technol. 2015, 12, 3–20. [Google Scholar] [CrossRef] [PubMed]
  14. Da, C.M.; Gong, C.Y.; Nan, W.; Zhou, K.S.; Zuo-Long, W.U.; Zhang, H.H. The role of long non-coding RNA MIAT in cancers. Biomed. Pharmacother. 2020, 129, 110359. [Google Scholar] [CrossRef]
  15. Siegel, R.L.; Miller, K.D.; Fedewa, S.A.; Ahnen, D.J.; Mester, R.G.; Barzi, A.; Jemal, A. Colorectal cancer statistics. CA Cancer J. Clin. 2017, 67, 177–193. [Google Scholar] [CrossRef] [PubMed]
  16. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef]
  17. Han, X.; Yao, Y.; Li, J.; Li, Z.; Han, X. Role of LncRNAs in the Epithelial-Mesenchymal Transition in Hepatocellular Carcinoma. Front. Oncol. 2021, 11, 2014. [Google Scholar]
  18. Han, L.-L.; Lv, Y.; Guo, H.; Ruan, Z.-P.; Nan, K.-J. Implications of biomarkers in human hepatocellular carcinoma pathogenesis and therapy. World J. Gastroenterol. WJG 2014, 20, 10249. [Google Scholar] [CrossRef]
  19. El–Serag, H.B.; Rudolph, K.L. Hepatocellular carcinoma: Epidemiology and molecular carcinogenesis. Gastroenterology 2007, 132, 2557–2576. [Google Scholar] [CrossRef]
  20. Zhang, X.; Zhu, Y. Research Progress on regulating LncRNAs of hepatocellular carcinoma stem cells. OncoTargets Ther. 2021, 14, 917. [Google Scholar] [CrossRef] [PubMed]
  21. Singal, A.G.; El-Serag, H.B. Hepatocellular carcinoma from epidemiology to prevention: Translating knowledge into practice. J. Clin. Gastroenterol. 2015, 13, 2140–2141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. McGlynn, K.A.; Petrick, J.L.; London, W.T. Global epidemiology of hepatocellular carcinoma: An emphasis on demographic and regional variability. Clin. Liver Dis. 2015, 19, 223–238. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Ledda, C.; Loreto, C.; Zammit, C.; Marconi, A.; Fago, L.; Matera, S.; Constanzo, V.; Fuccio, G.; Palmucci, S.; Ferrante, M.; et al. Non-infective occupational risk factors for hepatocellular carcinoma: A review. Mol. Med. Rep. 2017, 15, 511–533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Nagaraju, G.P.; Dariya, B.; Kasa, P.; Peela, S.; El-Rayes, B.F. Epigenetics in hepatocellular carcinoma. In Seminars in Cancer Biology; Elsevier: Amsterdam, The Netherlands, 2021. [Google Scholar]
  25. Wong, L.-S.; Wong, C.-M. Decoding the Roles of Long Noncoding RNAs in Hepatocellular Carcinoma. Int. J. Mol. Sci. 2021, 22, 3137. [Google Scholar] [CrossRef] [PubMed]
  26. Lin, P.; Wen, D.Y.; Li, Q.; He, Y.; Yang, H.; Chen, G. Genome-wide analysis of prognostic lncRNAs, miRNAs, and mRNAs forming a competing endogenous RNA network in hepatocellular carcinoma. Cell. Physiol. Biochem. 2018, 48, 1953–1967. [Google Scholar] [CrossRef] [PubMed]
  27. Longo, D.; Villanueva, A. Hepatocellular Carcinoma. N. Engl. J. Med. 2019, 380, 1450–1462. [Google Scholar]
  28. 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. [Google Scholar] [CrossRef] [Green Version]
  29. 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] [PubMed]
  30. Brunetti, O.; Gnoni, A.; Licchetta, A.; Longo, V.; Calabrese, A.; Argentiero, A.; Delcuratoro, S.; Solimando, A.; Gardini, A.; Silvestris, A. Predictive and prognostic factors in HCC patients treated with sorafenib. Medicina 2019, 55, 707. [Google Scholar] [CrossRef] [Green Version]
  31. 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]
  32. Ghidini, M.; Braconi, C. Non-coding RNAs in primary liver cancer. Front. Med. 2015, 2, 36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Wang, J.; Chu, E.S.H.; Chen, H.Y.; Man, K.; Go, M.Y.Y.; Huang, X.R.; Lan, H.Y.; Sung, J.J.Y.; 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] [PubMed] [Green Version]
  34. 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] [PubMed]
  35. Intini, C.; Hodgkinson, T.; Casey, S.M.; Gleeson, J.P.; O’Brien, F.J. Highly Porous Type II Collagen-Containing Scaffolds for Enhanced Cartilage Repair with Reduced Hypertrophic Cartilage Formation. Bioengineering 2022, 9, 232. [Google Scholar] [CrossRef] [PubMed]
  36. 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]
  37. Piscaglia, F.; Svegliati-Baroni, G.; Barchetti, A.; Pecorelli, A.; Marinelli, S.; Tiribelli, C.; Bellentani, S.; HCC-NAFLD Italian Study Group. Clinical patterns of hepatocellular carcinoma in nonalcoholic fatty liver disease: A multicenter prospective study. Hepatology 2016, 63, 827–838. [Google Scholar] [CrossRef]
  38. Juhling, F.; Hamdane, N.; Crouchet, E.; Li, S.; El Saghire, H.; Mukherji, A.; Fujiwara, N.; Oudot, M.A.; Thumann, C.; Saviano, A.; et al. Targeting clinical epigenetic reprogramming for chemoprevention of metabolic and viral hepatocellular carcinoma. Gut 2021, 70, 157–169. [Google Scholar] [CrossRef]
  39. Ogunwobi, O.O.; Harricharran, T.; Huaman, J.; Galuza, A.; Odumuwagun, O.; Tan, Y.; Ma, G.X.; Nguyen, M.T. Mechanisms of hepatocellular carcinoma progression. World J. Gastroenterol. 2019, 25, 2279–2293. [Google Scholar] [CrossRef]
  40. Okrah, K.; Tarighat, S.; Liu, B.; Koeppen, H.; Wagle, M.C.; Cheng, G.; Sun, C.; Dey, A.; Chang, M.T.; Sumiyoshi, T.; et al. Transcriptomic analysis of hepatocellular carcinoma reveals molecular features of disease progression and tumor immune biology. NPJ Precis. Oncol. 2018, 2, 25. [Google Scholar] [CrossRef] [PubMed]
  41. Safcak, D.; Drazilova, S.; Gazda, J.; Andrasina, I.; Adamcova-Selcanova, S.; Barila, R.; Mego, M.; Rac, M.; Skladany, L.; Zigrai, M.; et al. Nonalcoholic Fatty Liver Disease-Related Hepatocellular Carcinoma: Clinical Patterns, Outcomes, and Prognostic Factors for Overall Survival—A Retrospective Analysis of a Slovak Cohort. J. Clin. Med. 2021, 10, 3186. [Google Scholar] [CrossRef] [PubMed]
  42. Geh, D.; Derek, M.; Manas, H.; Reeves, L. Hepatocellular carcinoma in non-alcoholic fatty liver disease—A review of an emerging challenge facing clinicians. Hepatobiliary Surg. Nutr. 2021, 10, 59. [Google Scholar] [CrossRef] [PubMed]
  43. Vesque, A.D.; Decraecker, M.; Blanc, J.-F. Profile of Cabozantinib for the Treatment of Hepatocellular Carcinoma: Patient Selection and Special Considerations. J. Hepatocell. Carcinoma 2020, 7, 91. [Google Scholar] [CrossRef]
  44. Yang, F.U.; Zhang, L.; Huo, X.S.; Yuan, J.H.; Xu, D.; Yuan, S.X.; Zhu, N.; Zhou, W.P.; Yang, G.S.; Wang, Y.Z.; 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. [Google Scholar] [CrossRef]
  45. Kudo, M. Systemic therapy for hepatocellular carcinoma: 2017 update. Oncology 2017, 93 (Suppl. S1), 135–146. [Google Scholar] [CrossRef]
  46. 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] [Green Version]
  47. Ghanem, I.; Riveiro, M.E.; Paradis, V.; Faivre, S.; de Parga, P.M.V.; Raymond, E. Insights on the CXCL12-CXCR4 axis in hepatocellular carcinoma carcinogenesis. Am. J. Transl. Res. 2014, 6, 340. [Google Scholar]
  48. Villanueva, A.; Minguez, B.; Forner, A.; Reig, M.; Llovet, J.M. Hepatocellular carcinoma: Novel molecular approaches for diagnosis, prognosis, and therapy. Annu. Rev. Med. 2010, 61, 317–328. [Google Scholar] [CrossRef] [Green Version]
  49. Feng, Y.; Dramani Maman, S.T.; Zhu, X.; Liu, X.; Bongolo, C.C.; Liang, C.; Tu, J. Clinical value and potential mechanisms of LINC00221 in hepatocellular carcinoma based on integrated analysis. Epigenomics 2021, 13, 299–317. [Google Scholar] [CrossRef]
  50. Hu, X.; Chen, R.; Wei, Q.; Xu, X. The Landscape of Alpha Fetoprotein in Hepatocellular Carcinoma: Where Are We? Int. J. Biol. Sci. 2022, 18, 536–551. [Google Scholar] [CrossRef] [PubMed]
  51. Bao, H.; Su, H. Long noncoding RNAs act as novel biomarkers for hepatocellular carcinoma: Progress and prospects. Biomed. Res. Int. 2017, 2017, 6049480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Wu, J.-L.; Su, T.-H.; Chen, P.-J.; Chen, Y.-R. Acute-phase serum amyloid A for early detection of hepatocellular carcinoma in cirrhotic patients with low AFP level. Sci. Rep. 2022, 12, 5799. [Google Scholar] [CrossRef]
  53. Shi, Y. Mechanistic insights into precursor messenger RNA splicing by the spliceosome. Nat. Rev. Mol. Cell Biol. 2017, 18, 655. [Google Scholar] [CrossRef] [PubMed]
  54. Goodall, G.J.; Wickramasinghe, V.O. RNA in cancer. Nat. Rev. Cancer 2021, 21, 22–36. [Google Scholar] [CrossRef]
  55. Vo, J.N.; Cieslik, M.; Zhang, Y.; Shukla, S.; Xiao, L.; Zhang, Y.; Wu, Y.-M.; Dhanasekaran, S.M.; Engelke, C.G.; Cao, X.; et al. The landscape of circular RNA in cancer. Cell 2019, 176, 869–881.e13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Gruber, A.J.; Zavolan, M. Alternative cleavage and polyadenylation in health and disease. Nat. Rev. Genet. 2019, 20, 599–614. [Google Scholar] [CrossRef] [PubMed]
  57. Dvinge, H.; Guenthoer, J.; Porter, P.L.; Bradley, R.K. RNA components of the spliceosome regulate tissue-and cancer-specific alternative splicing. Genome Res. 2019, 29, 1591–1604. [Google Scholar] [CrossRef] [Green Version]
  58. Ramos-Rodriguez, D.H.; Pashneh-Tala, S.; Bains, A.K.; Moorehead, R.D.; Kassos, N.; Kelly, A.L.; Paterson, T.E.; Orozco-Diaz, C.A.; Gill, A.A.; Asencio, I.O. Demonstrating the Potential of Using Bio-Based Sustainable Polyester Blends for Bone Tissue Engineering Applications. Bioengineering 2022, 9, 163. [Google Scholar] [CrossRef]
  59. Ning, J.; Sun, K.; Fan, X.; Jia, K.; Wang, X.; Ma, C.; Wei, L. Necroptosis-related lncRNAs and Hepatocellular Carcinoma Undoubtedly Secret. Res. Sq. 2022, 25. [Google Scholar] [CrossRef]
  60. Amaral, P.P.; Leonardi, T.; Han, N.; Vire, E.; Gascoigne, D.K.; Arias-Carrasco, R.; Büscher, M.; Pandolfini, L.; Zhang, A.; Pluchino, S.; et al. Genomic positional conservation identifies topological anchor point RNAs linked to developmental loci. Genome Biol. 2018, 19. [Google Scholar] [CrossRef] [PubMed]
  61. Anastasiadou, E.; Jacob, L.S.; Slack, F.J. Non-coding RNA networks in cancer. Nat. Rev. Cancer 2018, 18, 5–18. [Google Scholar] [CrossRef]
  62. Fernandez-Ruiz, I. A new role for lncRNAs in atherosclerosis. Nat. Rev. Cardiol. 2018, 15, 195. [Google Scholar] [CrossRef] [PubMed]
  63. Wong, C.-M.; Tsang, F.H.-C.; Ng, I.O.-L. Non-coding RNAs in hepatocellular carcinoma: Molecular functions and pathological implications. Nat. Rev. Gastroenterol. Hepatol. 2018, 15, 137–151. [Google Scholar] [CrossRef] [PubMed]
  64. Chen, X.; Wang, S. LncRNA-ZFAS1 contributes to colon cancer progression through the miR-150-5p/VEGFA axis. Ann. Oncol. 2018, 29, v52–v53. [Google Scholar] [CrossRef]
  65. Kirchhoff, T.; Simpson, D.; Hekal, T.; Ferguson, R.; Kazlow, E.; Moran, U.; Lee, Y.; Izsak, A.; Wilson, M.; Shapiro, R.; et al. Discovery of novel germline genetic biomarkers of melanoma recurrence impacting exonic and long non-coding RNA (lncRNA) transcripts. Ann. Oncol. 2018, 29, viii463. [Google Scholar] [CrossRef]
  66. Zhang, J.; Han, X.; Jiang, L.; Han, Z.; Wang, Z. LncRNA CERNA2 is an independent predictor for clinical prognosis and is related to tumor development in gastric cancer. Int. J. Clin. Exp. Pathol. 2018, 11, 5783. [Google Scholar] [PubMed]
  67. Lei, G.L.; Niu, Y.; Cheng, S.J.; Li, Y.Y.; Bai, Z.F.; Yu, L.X.; Hong, Z.X.; Liu, H.; Liu, H.H.; Yan, J.; et al. Upregulation of long noncoding RNA W42 promotes tumor development by binding with DBN1 in hepatocellular carcinoma. World J. Gastroenterol. 2021, 27, 2586. [Google Scholar] [CrossRef]
  68. Li, Y.; Guo, D.; Lu, G.; Chowdhury, A.T.M.M.; Zhang, D.; Ren, M.; Chen, Y.; Wang, R.; He, S. LncRNA SNAI3-AS1 promotes PEG10-mediated proliferation and metastasis via decoying of miR-27a-3p and miR-34a-5p in hepatocellular carcinoma. Cell Death Dis. 2020, 11, 685. [Google Scholar] [CrossRef]
  69. Guttman, M.; Amit, I.; Garber, M.; French, C.; Lin, M.F.; Feldser, D.; Huarte, M.; Zuk, O.; Carey, B.W.; Cassady, J.P.; et al. Chromatin signature reveals over a thousand highly conserved large non-coding RNAs in mammals. Nature 2009, 458, 223–227. [Google Scholar] [CrossRef]
  70. Hu, W.; Alvarez-Dominguez, J.R.; Lodish, H.F. Regulation of mammalian cell differentiation by long non-coding RNAs. EMBO Rep. 2012, 13, 971–983. [Google Scholar] [CrossRef] [Green Version]
  71. Garcia, L.; Zambalde, E.; Mathias, C.; Barazetti, J.; Gradia, D.; Oliveira, J. lncRNAs in Hallmarks of Cancer and Clinical Applications. In Non-Coding RNAs; IntechOpen: London, UK, 2019. [Google Scholar]
  72. Guttman, M.; Rinn, J.L. Modular regulatory principles of large non-coding RNAs. Nature 2012, 482, 339–346. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Taniue, K.; Akimitsu, N. The functions and unique features of lncrnas in cancer development and tumorigenesis. Int. J. Mol. Sci. 2021, 22, 632. [Google Scholar] [CrossRef] [PubMed]
  74. Parasramka, M.A.; Maji, S.; Matsuda, A.; Yan, I.K.; Patel, T. Long non-coding RNAs as novel targets for therapy in hepatocellular carcinoma. Pharmacol. Ther. 2016, 161, 67–78. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Xiao, Z.; Shen, J.; Zhang, L.; Li, M.; Hu, W.; Cho, C. Therapeutic targeting of noncoding RNAs in hepatocellular carcinoma: Recent progress and future prospects. Oncol. Lett. 2018, 15, 3395–3402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Chen, S.; Zhang, Y.; Ding, X.; Li, W. Identification of lncRNA/circRNA-miRNA-mRNA ceRNA Network as Biomarkers for Hepatocellular Carcinoma. Front. Genet. 2022, 538. [Google Scholar] [CrossRef] [PubMed]
  77. Han, T.-S.; Hur, K.; Cho, H.-S.; Ban, H.S. Epigenetic associations between lncRNA/circRNA and miRNA in hepatocellular Carcinoma. Cancers 2020, 12, 2622. [Google Scholar] [CrossRef]
  78. Meng, H.; Niu, R.; Huang, C.; Li, J. Circular RNA as a Novel Biomarker and Therapeutic Target for HCC. Cells 2022, 11, 1948. [Google Scholar] [CrossRef] [PubMed]
  79. Sato, K.; Glaser, S.; Francis, H.; Alpini, G. Concise review: Functional roles and therapeutic potentials of long non-coding RNAs in cholangiopathies. Front. Med. 2020, 7, 48. [Google Scholar] [CrossRef] [PubMed]
  80. Jopling, C.L.; Yi, M.; Lancaster, A.M.; Lemon, S.M.; Sarnow, P. Modulation of hepatitis C virus RNA abundance by a liver-specific MicroRNA. Science 2005, 309, 1577–1581. [Google Scholar] [CrossRef] [Green Version]
  81. Kitabayashi, J.; Shirasaki, T.; Shimakami, T.; Nishiyama, T.; Welsch, C.; Funaki, M.; Murai, K.; Sumiyadorj, A.; Takatori, H.; Kitamura, K.; et al. Upregulation of the long non-coding RNA HULC by hepatitis C virus and its regulation of viral replication. J. Infect. Dis. 2020. [Google Scholar] [CrossRef]
  82. Wang, D.; Chen, F.; Zeng, T.; Tang, Q.; Chen, B.; Chen, L.; Dong, Y.; Li, X. Comprehensive biological function analysis of lncRNAs in hepatocellular carcinoma. Genes Dis. 2020, 8, 157–167. [Google Scholar] [CrossRef]
  83. Khalil, A.M.; Guttman, M.; Huarte, M.; Garber, M.; Raj, A.; Morales, D.R.; Thomas, K.; Presser, A.; Bernstein, B.E.; van Oudenaarden, A.; et al. Many human large intergenic noncoding RNAs associate with chromatin-modifying complexes and affect gene expression. Proc. Natl. Acad. Sci. USA 2009, 106, 11667–11672. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Bhan, A.; Mandal, S.S. LncRNA HOTAIR: A master regulator of chromatin dynamics and cancer. Biochim. Biophys. Acta Rev. Cancer 2015, 1856, 151–164. [Google Scholar] [CrossRef] [Green Version]
  85. Zheng, J.; Lin, Z.; Dong, P.; Lu, Z.; Gao, S.; Chen, X.; Wu, C.; Yu, F. Activation of hepatic stellate cells is suppressed by microRNA-150. Int. J. Mol. Med. 2013, 32, 17–24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Zheng, J.; Dong, P.; Gao, S.; Wang, N.; Yu, F. High expression of serum miR-17-5p associated with poor prognosis in patients with hepatocellular Carcinoma. Hepatogastroenterology 2013, 60, 549–552. [Google Scholar] [PubMed]
  87. Zheng, J.; Wu, C.; Lin, Z.; Guo, Y.; Shi, L.; Dong, P.; Lu, Z.; Gao, S.; Liao, Y.; Chen, B.; et al. Curcumin up-regulates phosphatase and tensin homologue deleted on chromosome 10 through micro RNA-mediated control of DNA methylation—A novel mechanism suppressing liver fibrosis. FEBS J. 2014, 281, 88–103. [Google Scholar] [CrossRef]
  88. Zhou, B.; Yang, H.; Bao, Y.-L.; Yang, S.-M.; Liu, J.; Xiao, Y.-F. Translation of noncoding RNAs and cancer. Cancer Lett. 2021, 497, 89–99. [Google Scholar] [CrossRef]
  89. 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]
  90. de Oliveira, J.C.; Oliveira, L.C.; Mathias, C.; Pedroso, G.A.; Lemos, D.S.; Salviano-Silva, A.; Jucoski, T.S.; Lobo-Alves, S.C.; Zambalde, E.P.; Cipolla, G.A.; et al. Long non-coding RNAs in cancer: Another layer of complexity. J. Gene Med. 2019, 21, e3065. [Google Scholar] [PubMed] [Green Version]
  91. ZZhang, T.; Hu, H.; Yan, G.; Wu, T.; Liu, S.; Chen, W.; Ning, Y.; Lu, Z. Long non-coding RNA and breast cancer. Technol. Cancer Res. Treat. 2019, 18, 1533033819843889. [Google Scholar]
  92. 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]
  93. 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] [PubMed]
  94. 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] [PubMed] [Green Version]
  95. Wang, K.C.; Chang, H.Y. Molecular mechanisms of long noncoding RNAs. Mol. Cell 2011, 43, 904–914. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Pandey, R.R.; Mondal, T.; Mohammad, F.; Enroth, S.; Redrup, L.; Komorowski, J.; Nagano, T.; Mancini-DiNardo, D.; Kanduri, C. Kcnq1ot1 antisense noncoding RNA mediates lineage-specific transcriptional silencing through chromatin-level rulation. Mol. Cell 2008, 32, 232–246. [Google Scholar] [CrossRef] [PubMed]
  97. Hung, T.; Wang, Y.; Lin, M.F.; Koegel, A.K.; Kotake, Y.; Grant, G.D.; Horlings, H.M.; Shah, N.; Umbricht, C.; Wang, P.; et al. Extensive and coordinated transcription of noncoding RNAs within cell-cycle promoters. Nat. Genet. 2011, 43, 621–629. [Google Scholar] [CrossRef] [Green Version]
  98. Liz, J.; Esteller, M. lncRNAs and microRNAs with a role in cancer development. Biochim. Biophys. Acta Gene Regul. Mech. 2016, 1859, 169–176. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. 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. [Google Scholar] [CrossRef] [PubMed]
  100. O’Brien, A.; Zhou, T.; Tan, C.; Alpini, G.; Glaser, S. Role of Non-Coding RNAs in the Progression of Liver Cancer: Evidence from Experimental Models. Cancers 2019, 11, 1652. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Yu, F.-J.; Zheng, J.-J.; Dong, P.-H.; Fan, X.-M. Long non-coding RNAs and hepatocellular carcinoma. Mol. Clin. Oncol. 2015, 3, 13–17. [Google Scholar] [CrossRef] [Green Version]
  102. Engreitz, J.M.; Ollikainen, N.; Guttman, M. Long non-coding RNAs: Spatial amplifiers that control nuclear structure and gene expression. Nat. Rev. Mol. Cell Biol. 2016, 17, 756–770. [Google Scholar] [CrossRef] [Green Version]
  103. Zhao, Y.; Wu, J.; Liangpunsakul, S.; Wang, L. Long non-coding RNA in liver metabolism and disease: Current status. Liver Res. 2017, 1, 163–167. [Google Scholar] [CrossRef]
  104. Li, C.H.; Chen, Y. Targeting long non-coding RNAs in cancers: Progress and prospects. Int. J. Biochem. Cell Biol. 2013, 45, 1895–1910. [Google Scholar] [CrossRef]
  105. Wan, Y.; Kertesz, M.; Spitale, R.C.; Segal, E.; Chang, H.Y. Understanding the transcriptome through RNA structure. Nat. Rev. Genet. 2011, 12, 641–655. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Tripathi, V.; Ellis, J.D.; Shen, Z.; Song, D.Y.; Pan, Q.; Watt, A.T.; Freier, S.M.; Bennett, C.F.; Sharma, A.; Bubulya, P.A.; et al. The nuclear-retained noncoding RNA MALAT1 regulates alternative splicing by modulating SR splicing factor phosphorylation. Mol. Cell 2010, 39, 925–938. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Kretz, M.; Siprashvili, Z.; Chu, C.; Webster, D.; Zehnder, A.; Qu, K.; Lee, C.S.; Flockhart, R.J.; Groff, A.F.; Chow, J.; et al. Control of somatic tissue differentiation by the long non-coding RNA TINCR. Nature 2013, 493, 231–235. [Google Scholar] [CrossRef]
  108. Ransohoff, J.D.; Wei, Y.; Khavari, P.A. The functions and unique features of long intergenic non-coding RNA. Nat. Rev. Mol. Cell Biol. 2018, 19, 143–157. [Google Scholar] [CrossRef]
  109. Jayaraj, G.G.; Pandey, S.; Scaria, V.; Maiti, S. Potential G-quadruplexes in the human long non-coding transcriptome. RNA Biol. 2012, 9, 81–89. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Li, P.; Zhang, G.; Li, J.; Yang, R.; Chen, S.; Wu, S.; Zhang, F.; Bai, Y.; Zhao, H.; Wang, Y.; et al. Long noncoding RNA RGMB-AS1 indicates a poor prognosis and modulates cell proliferation, migration and invasion in lung adenocarcinoma. PLoS ONE 2016, 11, e0150790. [Google Scholar] [CrossRef]
  111. Wilk, R.; Hu, J.; Blotsky, D.; Krause, H.M. Diverse and pervasive subcellular distributions for both coding and long noncoding RNAs. Genes Dev. 2016, 30, 594–609. [Google Scholar] [CrossRef] [Green Version]
  112. Chen, K.; Zhao, B.S.; He, C. Nucleic acid modifications in regulation of gene expression. Cell Chem. Biol. 2016, 23, 74–85. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Wang, J.; Sun, J.; Wang, J.; Song, Y.; Gao, P.; Shi, J.; Chen, P.; Wang, Z. Long noncoding RNAs in gastric cancer: Functions and clinical applications. OncoTargets Ther. 2016, 9, 681. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Wang, J.; Xu, A.; Zhang, J.; He, X.; Pan, Y.; Cheng, G.; Qin, C.; Hua, L.; Wang, Z. Prognostic significance of long non-coding RNA MALAT-1 in various human carcinomas: A meta-analysis. Genet. Mol. Res. 2016, 15, 15017433. [Google Scholar]
  115. Sun, L.; Guo, Y.; He, P.; Xu, X.; Zhang, X.; Wang, H.; Tang, T.; Zhou, W.; Xu, P.; Xie, P. Genome-wide profiling of long noncoding RNA expression patterns and CeRNA analysis in mouse cortical neurons infected with different strains of borna disease virus. Genes Dis. 2019, 6, 147–158. [Google Scholar] [PubMed]
  116. Rinn, J.L.; Chang, H.Y. Genome regulation by long noncoding RNAs. Annu. Rev. Biochem. 2012, 81, 145–166. [Google Scholar] [PubMed] [Green Version]
  117. Isin, M.; Dala, Y.N. LncRNAs and neoplasia. Clin. Chim. Acta 2015, 444, 280–288. [Google Scholar]
  118. Serviss, J.T.; Johnsson, P.; Grandér, D. An emerging role for long non-coding RNAs in cancer metastasis. Front. Genet. 2014, 5, 234. [Google Scholar] [PubMed] [Green Version]
  119. Harries, L.W. Long non-coding RNAs and human disease. Biochem. Soc. Trans. 2012, 40, 902–906. [Google Scholar] [PubMed]
  120. Wang, Q.-f.; Wang, Q.-l.; Cao, M.-b. LncRNA PITPNA-AS1 as a potential diagnostic marker and therapeutic target promotes hepatocellular carcinoma progression via modulating miR-448/ROCK1 axis. Front. Med. 2021, 8, 668787. [Google Scholar] [CrossRef]
  121. Amicone, L.; Citarella, F.; Cicchini, C. Epigenetic regulation in hepatocellular carcinoma requires long noncoding RNAs. Biomed. Res. Int. 2015. [Google Scholar]
  122. Yari, H.; Jin, L.; Teng, L.; Wang, Y.; Wu, Y.; Liu, G.; Gao, W.; Liang, J.; Xi, Y.; Feng, Y.C.; et al. LncRNA REG1CP promotes tumorigenesis through an enhancer complex to recruit FANCJ helicase for REG3A transcription. Nat. Commun. 2019, 10, 5334. [Google Scholar] [PubMed] [Green Version]
  123. 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] [PubMed] [Green Version]
  124. Jin, T. LncRNA DRAIR is a novel prognostic and diagnostic biomarker for gastric cancer. Mamm. Genome 2021, 32, 503–507. [Google Scholar] [CrossRef] [PubMed]
  125. Zheng, Q.-X.; Wang, J.; Gu, X.-Y.; Huang, C.-H.; Chen, C.; Hong, M.; Chen, Z. TTN-AS1 as a potential diagnostic and prognostic biomarker for multiple cancers. Biomed. Pharmacother. 2021, 135, 111169. [Google Scholar] [CrossRef] [PubMed]
  126. 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]
  127. 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] [Green Version]
  128. DiStefano, J.K.; Gerhard, G.S. Long Noncoding RNAs and Human Liver Disease. Annu. Rev. Pathol. 2021, 17, 1–21. [Google Scholar] [CrossRef] [PubMed]
  129. 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. [Google Scholar] [CrossRef] [Green Version]
  130. Huang, H.; Chen, J.; Ding, C.; Jin, X.; Jia, Z.; Peng, J. Lnc RNA NR 2F1-AS 1 regulates hepatocellular carcinoma oxaliplatin resistance by targeting ABCC 1 via miR-363. J. Cell. Mol. Med. 2018, 22, 3238–3245. [Google Scholar] [CrossRef]
  131. El-Ashmawy, N.E.; Hussien, F.Z.; El-Feky, O.A.; Hamouda, S.M.; Al-Ashmawy, G.M. Serum LncRNA-ATB and FAM83H-AS1 as diagnostic/prognostic non-invasive biomarkers for breast cancer. Life Sci. 2020, 259, 118193. [Google Scholar] [CrossRef]
  132. Maier, I.M.; Maier, A.C. miRNAs and lncRNAs: Potential non-invasive biomarkers for endometriosis. Biomedicines 2021, 9, 1662. [Google Scholar] [CrossRef]
  133. Herceg, Z.; Hainaut, P. Genetic and epigenetic alterations as biomarkers for cancer detection, diagnosis and prognosis. Mol. Oncol. 2007, 1, 26–41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Nagy, Á.; Munkácsy, G.; Győrffy, B. Pancancer survival analysis of cancer hallmark genes. Sci. Rep. 2021, 11, 6047. [Google Scholar] [CrossRef] [PubMed]
  136. Hanahan, D.; Weinberg, R.A. The hallmarks of cancer. Cell 2000, 100, 57–70. [Google Scholar] [CrossRef] [Green Version]
  137. Menyhárt, O.; Harami-Papp, H.; Sukumar, S.; Schäfer, R.; Magnani, L.; de Barrios, O.; Győrffy, B. Guidelines for the selection of functional assays to evaluate the hallmarks of cancer. Biochim. Biophys. Acta Rev. Cancer 2016, 1866, 300–319. [Google Scholar] [CrossRef] [Green Version]
  138. Chen, Y.; Zitello, E.; Guo, R.; Deng, Y. The function of LncRNAs and their role in the prediction, diagnosis, and prognosis of lung cancer. Clin. Transl. Med. 2021, 11, e367. [Google Scholar] [CrossRef]
  139. Zhang, R.; Xia, L.Q.; Lu, W.W.; Zhang, J.; Zhu, J.S. LncRNAs and cancer. Oncol. Lett. 2016, 12, 1233–1239. [Google Scholar] [CrossRef] [Green Version]
  140. Ahmad, S.; Abbas, M.; Ullah, M.F.; Aziz, M.H.; Beylerli, O.; Alam, M.A.; Syed, M.A.; Uddin, S.; Ahmad, A. Long non-coding RNAs regulated NF-κB signaling in cancer metastasis: Micromanaging by not so small non-coding RNAs. In Seminars in Cancer Biology; Elsevier: Amsterdam, The Netherlands, 2021. [Google Scholar]
  141. Tanne, A.; Muniz, L.R.; Puzio-Kuter, A.; Leonova, K.I.; Gudkov, A.V.; Ting, D.T.; Monasson, R.; Cocco, S.; Levine, A.J.; Bhardwaj, N.; et al. Distinguishing the immunostimulatory properties of noncoding RNAs expressed in cancer cells. Proc. Natl. Acad. Sci. USA 2015, 112, 15154–15159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Rooney, M.S.; Shukla, S.A.; Wu, C.J.; Getz, G.; Hacohen, N. Molecular and genetic properties of tumors associated with local immune cytolytic activity. Cell 2015, 160, 48–61. [Google Scholar] [CrossRef] [Green Version]
  143. Pan, L.; Liang, W.; Gu, J.; Zang, X.; Huang, Z.; Shi, H.; Chen, J.; Fu, M.; Zhang, P.; Xiao, X.; et al. Long noncoding RNA DANCR is activated by SALL4 and promotes the proliferation and invasion of gastric cancer cells. Oncotarget 2018, 9, 1915. [Google Scholar] [CrossRef] [Green Version]
  144. Li, S.; Li, J.; Chen, C.; Zhang, R.; Wang, K. Pan-cancer analysis of long non-coding RNA NEAT1 in various cancers. Genes Dis. 2018, 5, 27–35. [Google Scholar] [CrossRef] [PubMed]
  145. Zhou, D.; Ren, K.; Wang, M.; Wang, J.; Li, E.; Hou, C.; Su, Y.; Jin, Y.; Zou, Q.; Zhou, P.; et al. Long non-coding RNA RACGAP1P promotes breast cancer invasion and metastasis via miR-345-5p/RACGAP1-mediated mitochondrial fission. Mol. Oncol. 2021, 15, 543. [Google Scholar] [CrossRef]
  146. Wang, X.; Wang, Y.; Lin, F.; Xu, M.; Zhao, X. Long non-coding RNA LINC00665 promotes melanoma cell growth and migration via regulating the miR-224-5p/VMA21 axis. Exp. Dermatol. 2020, 31, 64–73. [Google Scholar] [CrossRef] [PubMed]
  147. He, R.-Z.; Luo, D.-X.; Mo, Y.-Y. Emerging roles of lncRNAs in the post-transcriptional regulation in cancer. Genes Dis. 2019, 6, 6–15. [Google Scholar] [CrossRef] [PubMed]
  148. Liang, W.Q.; Zeng, D.; Chen, C.F.; Sun, S.M.; Lu, X.F.; Peng, C.Y.; Lin, H.Y. Long noncoding RNA H19 is a critical oncogenic driver and contributes to epithelial-mesenchymal transition in papillary thyroid Carcinoma. Cancer Manag. Res. 2019, 11, 2059. [Google Scholar] [CrossRef] [Green Version]
  149. Wang, P.; Liu, G.; Xu, W.; Liu, H.; Bu, Q.; Sun, D. Long noncoding RNA H19 inhibits cell viability, migration, and invasion via downregulation of IRS-1 in thyroid cancer cells. Technol. Cancer Res. 2017, 16, 1102–1112. [Google Scholar] [CrossRef] [Green Version]
  150. Wu, Z.R.; Yan, L.; Liu, Y.T.; Cao, L.; Guo, Y.H.; Zhang, Y.; Yao, H.; Cai, L.; Shang, H.B.; Rui, W.W.; et al. Inhibition of mTORC1 by lncRNA H19 via disrupting 4E-BP1/Raptor interaction in pituitary tumours. Nat. Commun. 2018, 9, 4624. [Google Scholar] [CrossRef] [Green Version]
  151. Gao, H.; Hao, G.; Sun, Y.; Li, L.; Wang, Y. Long noncoding RNA H19 mediated the chemosensitivity of breast cancer cells via Wnt pathway and EMT process. OncoTargets Ther. 2018, 11, 8001. [Google Scholar] [CrossRef] [Green Version]
  152. Gao, L.M.; Xu, S.F.; Zheng, Y.; Wang, P.; Zhang, L.; Shi, S.S.; Wu, T.; Li, Y.; Zhao, J.; Tian, Q.; et al. Long non-coding RNA H19 is responsible for the progression of lung adenocarcinoma by mediating methylation-dependent repression of CDH1 promoter. J. Cell. Mol. Med. 2019, 23, 6411–6428. [Google Scholar] [CrossRef] [Green Version]
  153. Yang, Q.; Dong, Y.-J. LncRNA SNHG20 promotes migration and invasion of ovarian cancer via modulating the microRNA-148a/ROCK1 axis. J. Ovarian Res. 2021, 14, 168. [Google Scholar] [CrossRef]
  154. Yan, Q.; Tian, Y.; Hao, F. Downregulation of lncRNA UCA1 inhibits proliferation and invasion of cervical cancer cells through miR-206 expression. Oncol. Res. Featur. Preclin. Clin. Cancer Ther. 2021. [Google Scholar] [CrossRef] [PubMed]
  155. Feng, L.; Li, J.; Li, F.; Li, H.; Bei, S.; Zhang, X.; Yang, Z. Long noncoding RNA VCAN-AS1 contributes to the progression of gastric cancer via regulating p53 expression. J. Cell. Physiol. 2020, 235, 4388–4398. [Google Scholar] [CrossRef]
  156. Lou, C.; Zhao, J.; Gu, Y.; Li, Q.; Tang, S.; Wu, Y.; Tang, J.; Zhang, C.; Li, Z.; Zhang, Y. LINC01559 accelerates pancreatic cancer cell proliferation and migration through YAP-mediated pathway. J. Cell. Physiol. 2020, 235, 3928–3938. [Google Scholar] [CrossRef] [PubMed]
  157. Wang, Z.Y.; Duan, Y.; Wang, P. SP1-mediated upregulation of lncRNA SNHG4 functions as a ceRNA for miR-377 to facilitate prostate cancer progression through regulation of ZIC5. J. Cell. Physiol. 2020, 235, 3916–3927. [Google Scholar] [CrossRef]
  158. Wang, Y.; Jiang, F.; Xiong, Y.; Cheng, X.; Qiu, Z.; Song, R. LncRNA TTN-AS1 sponges miR-376a-3p to promote colorectal cancer progression via upregulating KLF15. Life Sci. 2020, 244, 116936. [Google Scholar] [CrossRef] [PubMed]
  159. Qiao, K.; Ning, S.; Wan, L.; Wu, H.; Wang, Q.; Zhang, X.; Xu, S.; Pang, D. LINC00673 is activated by YY1 and promotes the proliferation of breast cancer cells via the miR-515-5p/MARK4/Hippo signaling pathway. J. Exp. Clin. Cancer Res. 2019, 38, 418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Rossi, T.; Pistoni, M.; Sancisi, V.; Gobbi, G.; Torricelli, F.; Donati, B.; Ribisi, S.; Gugnoni, M.; Ciarrocchi, A. RAIN is a novel Enhancer-associated lncRNA that controls RUNX2 expression and promotes breast and thyroid cancer. Mol. Cancer Res. 2020, 18, 140–152. [Google Scholar] [CrossRef] [Green Version]
  161. Chang, Q.-Q.; Chen, C.-Y.; Chen, Z.; Chang, S. LncRNA PVT1 promotes proliferation and invasion through enhancing Smad3 expression by sponging miR-140-5p in cervical cancer. Radiol. Oncol. 2019, 53, 443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Li, H.; Han, Q.; Chen, Y.; Chen, X.; Ma, R.; Chang, Q.; Yin, D. Upregulation of the long non-coding RNA FOXD2-AS1 is correlated with tumor progression and metastasis in papillary thyroid cancer. Am. J. Transl. Res. 2019, 11, 5457. [Google Scholar]
  163. Ouyang, T.; Zhang, Y.; Tang, S.; Wang, Y. Long non-coding RNA LINC00052 regulates miR-608/EGFR axis to promote progression of head and neck squamous cell Carcinoma. Exp. Mol. Pathol. 2019, 111, 104321. [Google Scholar] [CrossRef]
  164. Tang, C.; Wang, Y.; Zhang, L.; Wang, J.; Wang, W.; Han, X.; Mu, C.; Gao, D. Identification of novel LncRNA targeting Smad2/PKCα signal pathway to negatively regulate malignant progression of glioblastoma. J. Cell. Physiol. 2020, 235, 3835–3848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Liu, D.; Wu, K.; Yang, Y.; Zhu, D.; Zhang, C.; Zhao, S. Long noncoding RNA ADAMTS9-AS2 suppresses the progression of esophageal cancer by mediating CDH3 promoter methylation. Mol. Carcinog. 2020, 59, 32–44. [Google Scholar] [CrossRef]
  166. Wang, W.; Xia, S.; Zhan, W. The long non-coding RNA ENST00000489676 influences papillary thyroid cancer cell proliferation and invasion through regulating MiR-922. J. Cancer 2019, 10, 5434. [Google Scholar] [CrossRef] [PubMed]
  167. Fan, J.; Zhang, J.; Huang, S.; Li, P. lncRNA OSER1-AS1 acts as a ceRNA to promote tumorigenesis in hepatocellular carcinoma by regulating miR-372-3p/Rab23 axis. Biochem. Biophys. Res. Commun. 2020, 521, 196–203. [Google Scholar] [CrossRef] [PubMed]
  168. Lin, S.; Zhang, R.; An, X.; Li, Z.; Fang, C.; Pan, B.; Chen, W.; Xu, G.; Han, W. LncRNA HOXA-AS3 confers cisplatin resistance by interacting with HOXA3 in non-small-cell lung Carcinoma cells. Oncogenesis 2019, 8, 60. [Google Scholar] [CrossRef] [PubMed]
  169. Yan, Y.; Xu, Z.; Chen, X.; Wang, X.; Zeng, S.; Zhao, Z.; Qian, L.; Li, Z.; Wei, J.; Huo, L.; et al. Novel function of lncRNA ADAMTS9-AS2 in promoting temozolomide resistance in glioblastoma via upregulating the FUS/MDM2 ubiquitination axis. Front. Cell Dev. Biol. 2019, 7, 217. [Google Scholar]
  170. Yokoyama, Y.; Sakatani, T.; Wada, R.; Ishino, K.; Kudo, M.; Koizumi, M.; Yamada, T.; Yoshida, H.; Naito, Z. In vitro and in vivo studies on the association of long non-coding RNAs H19 and urothelial cancer associated 1 with the susceptibility to 5-fluorouracil in rectal cancer. Int. J. Oncol. 2019, 55, 1361–1371. [Google Scholar] [CrossRef]
  171. Tamang, S.; Acharya, V.; Roy, D.; Sharma, R.; Aryaa, A.; Sharma, U.; Khandelwal, A.; Prakash, H.; Vasquez, K.M.; Jain, A. SNHG12: An LncRNA as a potential therapeutic target and biomarker for human cancer. Front. Oncol. 2019, 9, 901. [Google Scholar] [CrossRef]
  172. Avazpour, N.; Hajjari, M.; Birgani, M.T. HOTAIR: A promising long non-coding RNA with potential role in breast invasive carcinoma. Front. Genet. 2017, 8, 170. [Google Scholar] [CrossRef] [Green Version]
  173. Xu, W.; Zhou, G.; Wang, H.; Liu, Y.; Chen, B.; Chen, W.; Lin, C.; Wu, S.; Gong, A.; Xu, M. Circulating lncRNA SNHG11 as a novel biomarker for early diagnosis and prognosis of colorectal cancer. Int. J. Cancer 2020, 146, 2901–2912. [Google Scholar] [CrossRef]
  174. Li, T.; Xie, J.; Shen, C.; Cheng, D.; Shi, Y.; Wu, Z.; Deng, X.; Chen, H.; Shen, B.; Peng, C.; et al. 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]
  175. Tong, H.; He, S.; Li, K.; Zhang, K.; Jin, H.; Shi, J.; Cheng, Y.; Wang, L.; Liu, P. LncRNA SNHG8 Promotes Liver Cancer Proliferation and Metastasis by Sponging miR-542-3p and miR-4701-5p. Res. Sq. 2021, 25. [Google Scholar] [CrossRef]
  176. Wang, F.; Yuan, J.-H.; Wang, S.-B.; Yang, F.; Yuan, S.-X.; Ye, C.; Yang, N.; Zhou, W.-P.; Li, W.-L.; Sun, S.-H. 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]
  177. Luo, Y.; Lin, J.; Zhang, J.; Song, Z.; Zheng, D.; Chen, F.; Zhuang, X.; Li, A.; Liu, X. LncRNA SNHG17 contributes to proliferation, migration, and poor prognosis of hepatocellular carcinoma. Can. J. Gastroenterol. Hepatol. 2021; 11, 9990338. [Google Scholar]
  178. Liu, Y.; Mao, X.; Ma, Z.; Chen, W.; Guo, X.; Yu, L.; Deng, X.; Jiang, F.; Li, T.; Lin, N.; et al. Aberrant regulation of LncRNA TUG1-microRNA-328-3p-SRSF9 mRNA Axis in hepatocellular carcinoma: A promising target for prognosis and therapy. Mol. Cancer 2022, 21, 36. [Google Scholar] [CrossRef]
  179. Liu, Y.; Liu, J.; Cui, J.; Zhong, R.; Sun, G. Role of lncRNA LINC01194 in hepatocellular carcinoma via the miR-655-3p/SMAD family member 5 axis. Bioengineered 2022, 13, 1115–1125. [Google Scholar] [CrossRef] [PubMed]
  180. 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] [PubMed] [Green Version]
  181. Tian, Q.; Yan, X.; Yang, L.; Liu, Z.; Yuan, Z.; Zhang, Y. lncRNA CYTOR promotes cell proliferation and tumor growth via miR-125b/SEMA4C axis in hepatocellular carcinoma. Oncol. Lett. 2021, 22, 796. [Google Scholar] [CrossRef]
  182. Fu, C.; Li, J.; Li, P.; Cheng, D. LncRNA DNAJC3-AS1 Promotes Hepatocellular Carcinoma (HCC) Progression via Sponging Premature miR-27b. Cancer Manag. Res. 2021, 13, 8575. [Google Scholar] [CrossRef]
  183. Wang, Y.; Yang, L.; Chen, T.; Liu, X.; Guo, Y.; Zhu, Q.; Tong, X.; Yang, W.; Xu, Q.; Huang, D.; et al. A novel lncRNA MCM3AP-AS1 promotes the growth of hepatocellular carcinoma by targeting miR-194-5p/FOXA1 axis. Mol. Cancer 2019, 18, 28. [Google Scholar] [CrossRef]
  184. Hu, X.; Li, Q.; Zhang, J. The long noncoding RNA LINC00908 facilitates hepatocellular carcinoma progression via interaction with Sox-4. Cancer Manag. Res. 2019, 11, 8789. [Google Scholar] [CrossRef] [Green Version]
  185. Dai, W.; Dai, J.L.; Tang, M.H.; Ye, M.S.; Fang, S. lncRNA-SNHG15 accelerates the development of hepatocellular carcinoma by targeting miR-490-3p/histone deacetylase 2 axis. World J. Gastroenterol. 2019, 25, 5789. [Google Scholar] [CrossRef] [PubMed]
  186. Sui, C.-J.; Zhou, Y.-M.; Shen, W.-F.; Dai, B.-H.; Lu, J.-J.; Zhang, M.-F.; Yang, J.-M. Long noncoding RNA GIHCG promotes hepatocellular carcinoma progression through epigenetically regulating miR-200b/a/429. J. Mol. Med. 2016, 94, 1281–1296. [Google Scholar] [CrossRef] [PubMed]
  187. Huang, M.D.; Chen, W.M.; Qi, F.Z.; Xia, R.; Sun, M.; Xu, T.P.; Yin, L.; Zhang, E.B.; De, W.; Shu, Y.Q. Long non-coding RNA ANRIL is upregulated in hepatocellular carcinoma and regulates cell apoptosis by epigenetic silencing of KLF2. J. Hematol. Oncol. 2015, 8, 50. [Google Scholar] [CrossRef] [Green Version]
  188. Huang, M.D.; Chen, W.M.; Qi, F.Z.; Sun, M.; Xu, T.P.; Ma, P.; Shu, Y.Q. Long non-coding RNA TUG1 is up-regulated in hepatocellular Carcinoma and promotes cell growth and apoptosis by epigenetically silencing of KLF2. Mol. Cancer. 2015, 14, 165. [Google Scholar] [CrossRef] [Green Version]
  189. Cao, C.; Sun, J.; Zhang, D.; Guo, X.; Xie, L.; Li, X.; Wu, D.; Liu, L. The long intergenic noncoding RNA UFC1, a target of MicroRNA 34a, interacts with the mRNA stabilizing protein HuR to increase levels of beta-catenin in HCC cells. Gastroenterology 2015, 148, 415–426.e18. [Google Scholar] [CrossRef]
  190. Chen, L.; Yao, H.; Wang, K.; Liu, X. Long non-coding RNA MALAT1 regulates ZEB1 expression by sponging miR-143-3p and promotes hepatocellular carcinoma progression. J. Cell. Biochem. 2017, 118, 4836–4843. [Google Scholar] [CrossRef] [PubMed]
  191. Guo, W.; Liu, S.; Cheng, Y.; Lu, L.; Shi, J.; Xu, G.; Li, N.; Cheng, K.; Wu, M.; Cheng, S.; et al. ICAM-1-related noncoding RNA in cancer stem cells maintains ICAM-1 expression in hepatocellular carcinoma. Clin Cancer Res. 2016, 22, 2041–2050. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Li, T.; Xie, J.; Shen, C.; Cheng, D.; Shi, Y.; Wu, Z.; Deng, X.; Chen, H.; Shen, B.; Peng, C.; et al. Amplification of long noncoding RNA ZFAS1 promotes metastasis in hepatocellular carcinoma. Cancer Res. 2015, 75, 3181–3191. [Google Scholar] [CrossRef] [Green Version]
  193. Yuan, S.-X.; Yang, F.; Yang, Y.; Tao, Q.-F.; Zhang, J.; Huang, G.; Wang, R.-Y.; Yang, S.; Huo, X.-S.; Zhang, L.; 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. [Google Scholar] [CrossRef] [PubMed]
  194. Klingenberg, M.; Gross, M.; Goyal, A.; Polycarpou-Schwarz, M.; Miersch, T.; Ernst, A.S.; Leupold, J.; Patil, N.; Warnken, U.; Allgayer, H.; et al. The lncRNA CASC9 and RNA binding protein HNRNPL form a complex and co-regulate genes linked to AKT signaling. Hepatology 2018, 68, 1817–1832. [Google Scholar] [CrossRef]
  195. Huang, Z.; Zhou, J.; Peng, Y.; He, W.; Huang, C. The role of long noncoding RNAs in hepatocellular carcinoma. Mol. Cancer 2020, 19, 77. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Ding, L.-J.; Li, Y.; Wang, S.-D.; Wang, X.-S.; Fang, F.; Wang, W.-Y.; Lv, P.; Zhao, D.-H.; Wei, F.; Qi, L. Long noncoding RNA lncCAMTA1 promotes proliferation and cancer stem cell-like properties of liver cancer by inhibiting CAMTA1. Int. J. Mol. Sci. 2016, 17, 1617. [Google Scholar] [CrossRef] [PubMed]
  197. Li, X.; Zhao, Q.; Qi, J.; Wang, W.; Zhang, D.; Li, Z.; Qin, C. lncRNA Ftx promotes aerobic glycolysis and tumor progression through the PPARγ pathway in hepatocellular carcinoma. Int. J. Oncol. 2018, 53, 551–566. [Google Scholar] [CrossRef] [PubMed]
  198. 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. [Google Scholar] [CrossRef]
  199. Pan, W.; Li, W.; Zhao, J.; Huang, Z.; Zhao, J.; Chen, S.; Wang, C.; Xue, Y.; Huang, F.; Fang, Q.; et al. lnc RNA-PDPK 2P promotes hepatocellular carcinoma progression through the PDK 1/AKT/Caspase 3 pathway. Mol. Oncol. 2019, 13, 2246–2258. [Google Scholar] [CrossRef] [Green Version]
  200. Wang, H.; Huo, X.; Yang, X.R.; He, J.; Cheng, L.; Wang, N.; Deng, X.; Jin, H.; Wang, N.; Wang, C.; et al. STAT3-mediated upregulation of lncRNA HOXD-AS1 as a ceRNA facilitates liver cancer metastasis by regulating SOX4. Mol. Cancer 2017, 16, 136. [Google Scholar] [CrossRef] [Green Version]
  201. Chen, Z.; Yu, W.; Zhou, Q.; Zhang, J.; Jiang, H.; Hao, D.; Wang, J.; Zhou, Z.; He, C.; Xiao, Z. A novel lncRNA IHS promotes tumor proliferation and metastasis in HCC by regulating the ERK-and AKT/GSK-3β-signaling pathways. Mol. Ther. Nucleic Acids 2019, 16, 707–720. [Google Scholar] [CrossRef] [Green Version]
  202. Wu, Y.; Xiong, Q.; Li, S.; Yang, X.; Ge, F. Integrated proteomic and transcriptomic analysis reveals long noncoding RNA HOX transcript antisense intergenic RNA (HOTAIR) promotes hepatocellular carcinoma cell proliferation by regulating opioid growth factor receptor (OGFr). Mol. Cell. Proteom. 2018, 17, 146–159. [Google Scholar] [CrossRef] [Green Version]
  203. Cheng, D.; Deng, J.; Zhang, B.; He, X.; Meng, Z.; Li, G.; Ye, H.; Zheng, S.; Wei, L.; Deng, X.; et al. LncRNA HOTAIR epigenetically suppresses miR-122 expression in hepatocellular carcinoma via DNA methylation. EBioMedicine 2018, 36, 159–170. [Google Scholar] [CrossRef] [Green Version]
  204. Yao, Y.; Li, J.; Wang, L. Large intervening non-coding RNA HOTAIR is an indicator of poor prognosis and a therapeutic target in human cancers. Int. J. Mol. Sci. 2014, 15, 18985–18999. [Google Scholar] [CrossRef] [Green Version]
  205. Wu, L.; Zhang, L.; Zheng, S. Role of the long non-coding RNA HOTAIR in hepatocellular Carcinoma. Oncol. Lett. 2017, 14, 1233–1239. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Li, H.; An, J.; Wu, M.; Zheng, Q.; Gui, X.; Li, T.; Pu, H.; Lu, D. LncRNA HOTAIR promotes human liver cancer stem cell malignant growth through downregulation of SETD2. Oncotarget 2015, 6, 27847. [Google Scholar] [CrossRef] [PubMed]
  207. You, L.-N.; Tai, Q.-W.; Xu, L.; Hao, Y.; Guo, W.-J.; Zhang, Q.; Tong, Q.; Zhang, H.; Huang, W.-K. Exosomal LINC00161 promotes angiogenesis and metastasis via regulating miR-590-3p/ROCK axis in hepatocellular carcinoma. Cancer Gene Ther. 2021, 28, 719–736. [Google Scholar] [CrossRef]
  208. Lin, Y.; Jian, Z.; Jin, H.; Wei, X.; Zou, X.; Guan, R.; Huang, J. Long non-coding RNA DLGAP1-AS1 facilitates tumorigenesis and epithelial–mesenchymal transition in hepatocellular carcinoma via the feedback loop of miR-26a/b-5p/IL-6/JAK2/STAT3 and Wnt/β-catenin pathway. Cell Death Dis. 2020, 11, 1–17. [Google Scholar] [CrossRef] [PubMed]
  209. Yi, T.; Wang, T.; Shi, Y.; Peng, X.; Tang, S.; Zhong, L.; Chen, Y.; Li, Y.; He, K.; Wang, M.; et al. Long noncoding RNA 91H overexpression contributes to the growth and metastasis of HCC by epigenetically positively regulating IGF2 expression. Liver Int. 2020, 40, 456–467. [Google Scholar] [CrossRef]
  210. Teng, F.; Zhang, J.-X.; Chang, Q.-M.; Wu, X.-B.; Tang, W.-G.; Wang, J.-F.; Feng, J.-F.; Zhang, Z.-P.; Hu, Z.-Q. LncRNA MYLK-AS1 facilitates tumor progression and angiogenesis by targeting miR-424-5p/E2F7 axis and activating VEGFR-2 signaling pathway in hepatocellular Carcinoma. J. Exp. Clin. Cancer Res. 2020, 39, 1–18. [Google Scholar]
  211. Tian, C.; Abudoureyimu, M.; Lin, X.; Chu, X.; Wang, R. Linc-ROR facilitates progression and angiogenesis of hepatocellular carcinoma by modulating DEPDC1 expression. Cell Death Dis. 2021, 12, 1047. [Google Scholar] [CrossRef]
  212. Li, P.; Li, Y.; Ma, L. Long noncoding RNA highly upregulated in liver cancer promotes the progression of hepatocellular carcinoma and attenuates the chemosensitivity of oxaliplatin by regulating miR-383-5p/vesicle-associated membrane protein-2 axis. Pharmacol. Res. Perspect. 2021, 9, e00815. [Google Scholar] [CrossRef]
  213. Qian, H.; Wu, Q.; Wu, J.H.; Tian, X.Y.; Xu, W.; Hao, C.Y. Long noncoding LINC00238 restrains Hepatocellular Carcinoma Malignant Phenotype via Sponging miR-522. Res. Sq. 2021, 21. [Google Scholar] [CrossRef]
  214. Liu, Y.; Liu, R.; Zhao, J.; Zeng, Z.; Shi, Z.; Lu, Q.; Guo, J.; Li, L.; Yao, Y.; Liu, X.; et al. LncRNA TMEM220-AS1 suppresses hepatocellular carcinoma cell proliferation and invasion by regulating the TMEM220/β-catenin axis. J. Cancer 2021, 12, 6805. [Google Scholar] [CrossRef]
  215. Sheng, J.-Q.; Wang, M.-R.; Fang, D.; Liu, L.; Huang, W.-J.; Tian, D.-A.; He, X.-X.; Li, P.-Y. LncRNA NBR2 inhibits tumorigenesis by regulating autophagy in hepatocellular carcinoma. Biomed. Pharmacother. 2021, 133, 111023. [Google Scholar] [CrossRef]
  216. Lei, G.-L.; Fan, H.-X.; Wang, C.; Niu, Y.; Li, T.-L.; Yu, L.-X.; Hong, Z.-X.; Yan, J.; Wang, X.-L.; Zhang, S.-G.; et al. Long non-coding ribonucleic acid W5 inhibits progression and predicts favorable prognosis in hepatocellular carcinoma. World J. Gastroenterol. 2021, 27, 55. [Google Scholar] [CrossRef]
  217. Pan, W.; Zhang, N.; Liu, W.; Liu, J.; Zhou, L.; Liu, Y.; Yang, M. The long noncoding RNA GAS8-AS1 suppresses hepatocarcinogenesis by epigenetically activating the tumor suppressor GAS8. J. Biol. Chem. 2018, 293, 17154–17165. [Google Scholar] [CrossRef] [Green Version]
  218. Wu, Y.; Zhou, Y.; Huan, L.; Xu, L.; Shen, M.; Huang, S.; Liang, L. LncRNA MIR22HG inhibits growth, migration and invasion through regulating the miR-10a-5p/NCOR2 axis in hepatocellular carcinoma cells. Cancer Sci. 2019, 110, 973–984. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Yan, S.; Tang, Z.; Chen, K.; Liu, Y.; Yu, G.; Chen, Q.; Dang, H.; Chen, F.; Ling, J.; Zhu, L.; et al. Long noncoding RNA MIR31HG inhibits hepatocellular carcinoma proliferation and metastasis by sponging microRNA-575 to modulate ST7L expression. J. Exp. Clin. Cancer Res. 2018, 37, 214. [Google Scholar] [CrossRef] [PubMed]
  220. Chen, F.; Li, Y.; Li, M.; Wang, L. Long noncoding RNA GAS5 inhibits metastasis by targeting miR-182/ANGPTL1 in hepatocellular Carcinoma. Am. J. Cancer Res. 2019, 9, 108. [Google Scholar] [PubMed]
  221. Yu, Z.; Zhao, H.; Feng, X.; Li, H.; Qiu, C.; Yi, X.; Tang, H.; Zhang, J. Long non-coding RNA FENDRR acts as a miR-423-5p sponge to suppress the Treg-mediated immune escape of hepatocellular Carcinoma cells. Mol. Ther. Nucleic Acids 2019, 17, 516–529. [Google Scholar] [CrossRef] [Green Version]
  222. Wang, Y.-G.; Wang, T.; Shi, M.; Zhai, B. Long noncoding RNA EPB41L4A-AS2 inhibits hepatocellular carcinoma development by sponging miR-301a-5p and targeting FOXL1. Exp. Clin. Cancer Res. 2019, 38, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Yu, S.; Li, N.; Huang, Z.; Chen, R.; Yi, P.; Kang, R.; Tang, D.; Hu, X.; Fan, X. A novel lncRNA, TCONS_00006195, represses hepatocellular carcinoma progression by inhibiting enzymatic activity of ENO1. Cell Death Dis. 2018, 9, 153. [Google Scholar] [CrossRef]
  224. Ni, W.; Zhang, Y.; Zhan, Z.; Ye, F.; Liang, Y.; Huang, J.; Chen, K.; Chen, L.; Ding, Y. A novel lncRNA uc. 134 represses hepatocellular carcinoma progression by inhibiting CUL4A-mediated ubiquitination of LATS1. J. Hematol. Oncol. 2017, 10, 91. [Google Scholar] [CrossRef] [Green Version]
  225. Hu, J.; Song, C.; Duan, B.; Zhang, X.; Li, D.; Zhu, L.; Gao, H. LncRNA-SVUGP2 suppresses progression of hepatocellular carcinoma. Oncotarget 2017, 8, 97835. [Google Scholar] [CrossRef] [Green Version]
  226. Jiang, X.; Wang, G.; Liu, Y.; Mei, C.; Yao, Y.; Wu, X.; Chen, X.; Ma, W.; Li, K.; Zhang, Z.; et al. A novel long non-coding RNA RP11-286H15. 1 represses hepatocellular carcinoma progression by promoting ubiquitination of PABPC4. Cancer Lett. 2021, 499, 109–121. [Google Scholar] [CrossRef]
  227. Huang, J.-F.; Guo, Y.-J.; Zhao, C.-X.; Yuan, S.-X.; Wang, Y.; Tang, G.-N.; Zhou, W.-P.; Sun, S.-H. 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. [Google Scholar] [CrossRef]
  228. Zhuang, L.K.; Yang, Y.T.; Ma, X.; Han, B.; Wang, Z.S.; Zhao, Q.Y.; Wu, L.Q.; Qu, Z.Q. MicroRNA-92b promotes hepatocellular Carcinoma progression by targeting Smad7 and is mediated by long non-coding RNA XIST. Cell Death Dis. 2016, 7, e2203. [Google Scholar] [CrossRef] [Green Version]
  229. Yang, L.; Deng, W.-L.; Zhao, B.-G.; Xu, Y.; Wang, X.-W.; Fang, Y.; Xiao, H.-J. FOXO3-induced lncRNA LOC554202 contributes to hepatocellular carcinoma progression via the miR-485-5p/BSG axis. Cancer Gene Ther. 2021, 29, 326–340. [Google Scholar] [CrossRef]
  230. Wan, T.; Zheng, J.; Yao, R.; Yang, S.; Zheng, W.; Zhou, P. LncRNA DDX11-AS1 accelerates hepatocellular carcinoma progression via the miR-195-5p/MACC1 pathway. Ann. Hepatol. 2021, 20, 100258. [Google Scholar] [CrossRef]
  231. Wang, H.; Liang, L.; Dong, Q.; Huan, L.; He, J.; Li, B.; Yang, C.; Jin, H.; Wei, L.; Yu, C.; et al. Long noncoding RNA miR503HG, a prognostic indicator, inhibits tumor metastasis by regulating the HNRNPA2B1/NF-κB pathway in hepatocellular carcinoma. Theranostics 2018, 8, 2814. [Google Scholar] [CrossRef] [PubMed]
  232. Li, G.; Zhang, H.; Wan, X.; Yang, X.; Zhu, C.; Wang, A.; He, L.; Miao, R.; Chen, S.; Zhao, H. Long noncoding RNA plays a key role in metastasis and prognosis of hepatocellular Carcinoma. Biomed. Res. Int. 2014, 2014, 780521. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Hu, W.Y.; Wei, H.Y.; Li, K.M.; Wang, R.B.; Xu, X.Q.; Feng, R. LINC00511 as a ceRNA promotes cell malignant behaviors and correlates with prognosis of hepatocellular carcinoma patients by modulating miR-195/EYA1 axis. Biomed. Pharmacother. 2020, 121, 109642. [Google Scholar] [CrossRef] [PubMed]
  234. Yu, J.; Han, J.; Zhang, J.; Li, G.; Liu, H.; Cui, X.; Xu, Y.; Li, T.; Liu, J.; Wang, C. The long noncoding RNAs PVT1 and uc002mbe. 2 in sera provide a new supplementary method for hepatocellular carcinoma diagnosis. Medicine 2016, 95, e4436. [Google Scholar] [CrossRef] [PubMed]
  235. Zheng, Z.-K.; Pang, C.; Yang, Y.; Duan, Q.; Zhang, J.; Liu, W.-C. Serum long noncoding RNA urothelial carcinoma-associated 1: A novel biomarker for diagnosis and prognosis of hepatocellular carcinoma. Int. J. Med. Res. 2018, 46, 348–356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Li, G.; Shi, H.; Wang, X.; Wang, B.; Qu, Q.; Geng, H.; Sun, H. Identification of diagnostic long non-coding RNA biomarkers in patients with hepatocellular carcinoma. Mol. Med. Rep. 2019, 20, 1121–1130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Wang, Z.; Hu, R.; Pang, J.; Zhang, G.; Yan, W.; Li, Z. Serum long noncoding RNA LRB1 as a potential biomarker for predicting the diagnosis and prognosis of human hepatocellular carcinoma. Oncol. Lett. 2018, 16, 1593–1601. [Google Scholar] [CrossRef] [Green Version]
  238. Luo, T.; Chen, M.; Zhao, Y.; Wang, D.; Liu, J.; Chen, J.; Luo, H.; Li, L. Macrophage-associated lncRNA ELMO1-AS1: A novel therapeutic target and prognostic biomarker for hepatocellular carcinoma. OncoTargets Ther. 2019, 12, 6203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  239. Zhou, C.-C.; Yang, F.; Yuan, S.-X.; Ma, J.-Z.; Liu, F.; Yuan, J.-H.; Bi, F.-R.; Lin, K.-Y.; Yin, J.-H.; Cao, G.-W.; et al. Systemic genome screening identifies the outcome associated focal loss of long noncoding RNA PRAL in hepatocellular carcinoma. Hepatology 2016, 63, 850–863. [Google Scholar] [CrossRef]
  240. Wang, Z.; Yang, B.; Zhang, M.; Guo, W.; Wu, Z.; Wang, Y.; Jia, L.; Li, S.; Caesar-Johnson, S.J.; Demchok, J.A.; et al. lncRNA epigenetic landscape analysis identifies EPIC1 as an oncogenic lncRNA that interacts with MYC and promotes cell-cycle progression in cancer. Cancer Cell 2018, 33, 706–720.e9. [Google Scholar] [CrossRef] [Green Version]
  241. Xing, Z.; Lin, A.; Li, C.; Liang, K.; Wang, S.; Liu, Y.; Park, P.K.; Qin, L.; Wei, Y.; Hawke, D.H.; et al. lncRNA directs cooperative epigenetic regulation downstream of chemokine signals. Cell 2014, 159, 1110–1125. [Google Scholar] [CrossRef] [Green Version]
  242. Chiappinelli, K.B.; Strissel, P.L.; Desrichard, A.; Li, H.; Henke, C.; Akman, B.; Hein, A.; Rote, N.S.; Cope, L.M.; Snyder, A.; et al. Inhibiting DNA methylation causes an interferon response in cancer via dsRNA including endogenous retroviruses. Cell 2015, 162, 974–986. [Google Scholar] [CrossRef] [Green Version]
  243. Guccione, E.; Bassi, C.; Casadio, F.; Martinato, F.; Cesaroni, M.; Schuchlautz, H.; Lüscher, B.; Amati, B. Methylation of histone H3R2 by PRMT6 and H3K4 by an MLL complex are mutually exclusive. Nature 2007, 449, 933–937. [Google Scholar] [CrossRef]
  244. Chen, X.; Tang, F.-R.; Arfuso, F.; Cai, W.-Q.; Ma, Z.; Yang, J.; Sethi, G. The emerging role of long non-coding RNAs in the metastasis of hepatocellular carcinoma. Biomolecules 2020, 10, 66. [Google Scholar] [CrossRef] [Green Version]
  245. Yuan, S.-X.; Zhang, J.; Xu, Q.-G.; Yang, Y.; Zhou, W.-P. Long noncoding RNA, the methylation of genomic elements and their emerging crosstalk in hepatocellular carcinoma. Cancer Lett. 2016, 379, 239–244. [Google Scholar] [CrossRef]
  246. Xu, X.; Lou, Y.; Tang, J.; Teng, Y.; Zhang, Z.; Yin, Y.; Zhuo, H.; Tan, Z. The long non-coding RNA Linc-GALH promotes hepatocellular carcinoma metastasis via epigenetically regulating Gankyrin. Cell Death Dis. 2019, 10, 1–13. [Google Scholar] [CrossRef]
  247. Nebel, S.; Lux, M.; Kuth, S.; Bider, F.; Dietrich, W.; Egger, D.; Boccaccini, A.R.; Kasper, C. Alginate Core–Shell Capsules for 3D Cultivation of Adipose-Derived Mesenchymal Stem Cells. Bioengineering 2022, 9, 66. [Google Scholar] [CrossRef] [PubMed]
  248. Tang, J.; Xie, Y.; Xu, X.; Yin, Y.; Jiang, R.; Deng, L.; Tan, Z.; Gangarapu, V.; Tang, J.; Sun, B. Bidirectional transcription of Linc00441 and RB1 via H3K27 modification-dependent way promotes hepatocellular carcinoma. Cell Death Dis. 2017, 8, e2675. [Google Scholar] [CrossRef] [Green Version]
  249. Chen, Z.; Gao, Y.; Yao, L.; Liu, Y.; Huang, L.; Yan, Z.; Zhao, W.; Zhu, P.; Weng, H. LncFZD6 initiates Wnt/β-catenin and liver TIC self-renewal through BRG1-mediated FZD6 transcriptional activation. Oncogene 2018, 37, 3098–3112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  250. Zheng, G.X.; Do, B.T.; Webster, D.E.; Khavari, P.A.; Chang, H.Y. Dicer-microRNA-Myc circuit promotes transcription of hundreds of long noncoding RNAs. Nat. Struct. Mol. Biol. 2014, 21, 585–590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  251. Tran, D.D.H.; Kessler, C.; E Niehus, S.; Mahnkopf, M.; Koch, A.; Tamura, T. Myc target gene, long intergenic noncoding RNA, Linc00176 in hepatocellular carcinoma regulates cell cycle and cell survival by titrating tumor suppressor microRNAs. Oncogene 2018, 37, 75–85. [Google Scholar] [CrossRef]
  252. Li, W.; Kang, Y. A new Lnc in metastasis: Long noncoding RNA mediates the prometastatic functions of TGF-β. Cancer Cell 2014, 25, 557–559. [Google Scholar] [CrossRef] [Green Version]
  253. Yuan, J.H.; Liu, X.N.; Wang, T.T.; Pan, W.; Tao, Q.F.; Zhou, W.P.; Wang, F.; Sun, S.H. The MBNL3 splicing factor promotes hepatocellular carcinoma by increasing PXN expression through the alternative splicing of lncRNA-PXN-AS1. Nat. Cell Biol. 2017, 19, 820–832. [Google Scholar] [CrossRef] [PubMed]
  254. He, J.; Zuo, Q.; Hu, B.; Jin, H.; Wang, C.; Cheng, Z.; Deng, X.; Yang, C.; Ruan, H.; Yu, C.; et al. A novel, liver-specific long noncoding RNA LINC01093 suppresses HCC progression by interaction with IGF2BP1 to facilitate decay of GLI1 mRNA. Cancer Lett. 2019, 450, 98–109. [Google Scholar] [CrossRef]
  255. Moore, J.B., IV; Uchida, S. Functional characterization of long noncoding RNAs. Curr. Opin. Cardiol. 2020, 35, 199–206. [Google Scholar] [CrossRef] [PubMed]
  256. Salmena, L.; Poliseno, L.; Tay, Y.; Kats, L.; Pandolfi, P.P. A ceRNA hypothesis: The Rosetta Stone of a hidden RNA language? Cell 2011, 146, 353–358. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Yoon, J.-H.; Abdelmohsen, K.; Gorospe, M. Functional interactions among microRNAs and long noncoding RNAs. In Seminars in Cell & Developmental Biology; Elsevier: Amsterdam, The Netherlands, 2014. [Google Scholar]
  258. Cesana, M.; Cacchiarelli, D.; Legnini, I.; Santini, T.; Sthandier, O.; Chinappi, M.; Tramontano, A.; Bozzoni, I. A long noncoding RNA controls muscle differentiation by functioning as a competing endogenous RNA. Cell 2011, 147, 358–369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  259. Uchida, S.; Kauppinen, S. Long non-coding RNAs in liver cancer and nonalcoholic steatohepatitis. Non-Coding RNA 2020, 6, 34. [Google Scholar]
  260. Xie, C.-R.; Wang, F.; Zhang, S.; Wang, F.-Q.; Zheng, S.; Li, Z.; Lv, J.; Qi, H.-Q.; Fang, Q.-L.; Wang, X.-M.; et al. Long noncoding RNA HCAL facilitates the growth and metastasis of hepatocellular carcinoma by acting as a ceRNA of LAPTM4B. Mol. Ther. Nucleic Acids 2017, 9, 440–451. [Google Scholar] [CrossRef] [Green Version]
  261. Hou, Z.; Xu, X.; Zhou, L.; Fu, X.; Tao, S.; Zhou, J.; Tan, D.; Liu, S. The long non-coding RNA MALAT1 promotes the migration and invasion of hepatocellular carcinoma by sponging miR-204 and releasing SIRT1. Tumor Biol. 2017, 39, 1010428317718135. [Google Scholar] [CrossRef] [Green Version]
  262. Lu, S.; Zhou, J.; Sun, Y.; Li, N.; Miao, M.; Jiao, B.; Chen, H. The noncoding RNA HOXD-AS1 is a critical regulator of the metastasis and apoptosis phenotype in human hepatocellular carcinoma. Mol. Cancer 2017, 16, 125. [Google Scholar] [CrossRef] [PubMed]
  263. 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.; 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 2016, 7, 42431. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Zhang, J.; Li, Z.; Liu, L.; Wang, Q.; Li, S.; Chen, D.; Hu, Z.; Yu, T.; Ding, J.; Li, J.; et al. Long noncoding RNA TSLNC8 is a tumor suppressor that inactivates the interleukin-6/STAT3 signaling pathway. Hepatology 2018, 67, 171–187. [Google Scholar] [CrossRef] [Green Version]
  265. Xie, C.; Li, S.-Y.; Fang, J.-H.; Zhu, Y.; Yang, J.-E. Functional long non-coding RNAs in hepatocellular carcinoma. Cancer Lett. 2020, 500, 281–291. [Google Scholar] [CrossRef]
  266. Li, Z.; Zhang, J.; Liu, X.; Li, S.; Wang, Q.; Chen, D.; Hu, Z.; Yu, T.; Ding, J.; Li, J.; et al. The LINC01138 drives malignancies via activating arginine methyltransferase 5 in hepatocellular carcinoma. Nat. Commun. 2018, 9, 1572. [Google Scholar] [CrossRef] [PubMed]
  267. Ding, C.H.; Yin, C.; Chen, S.J.; Wen, L.Z.; Ding, K.; Lei, S.J.; Liu, J.P.; Wang, J.; Chen, K.X.; Jiang, H.L.; et al. The HNF1α-regulated lncRNA HNF1A-AS1 reverses the malignancy of hepatocellular carcinoma by enhancing the phosphatase activity of SHP-1. Mol. Cancer 2018, 17, 63. [Google Scholar] [CrossRef] [PubMed]
  268. Yang, F.; Huo, X.-S.; Yuan, S.-X.; Zhang, L.; Zhou, W.-P.; Wang, F.; Sun, S.-H. Repression of the long noncoding RNA-LET by histone deacetylase 3 contributes to hypoxia-mediated metastasis. Mol. Cell 2013, 49, 1083–1096. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  269. Tao, G.-Z.; Lehwald, N.; Jang, K.Y.; Baek, J.; Xu, B.; Omary, M.B.; Sylvester, K.G. Wnt/β-catenin signaling protects mouse liver against oxidative stress-induced apoptosis through the inhibition of forkhead transcription factor FoxO3. J. Biol. Chem. 2013, 288, 17214–17224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  270. Chen, X.; Xu, C.; Liu, Y. Involvement of ERK1/2 signaling in proliferation of eight liver cell types during hepatic regeneration in rats. Genet. Mol. Res. 2013, 12, 665–677. [Google Scholar] [CrossRef]
  271. Xu, D.; Yang, F.; Yuan, J.-H.; Zhang, L.; Bi, H.-S.; Zhou, C.-C.; Liu, F.; Wang, F.; Sun, S.-H. Long noncoding RNAs associated with liver regeneration 1 accelerates hepatocyte proliferation during liver regeneration by activating Wnt/β-Catenin signaling. Hepatology 2013, 58, 739–751. [Google Scholar] [CrossRef]
  272. Schmitt, A.M.; Chang, H.Y. Long noncoding RNAs in cancer pathways. Cancer Cell 2016, 29, 452–463. [Google Scholar] [CrossRef] [Green Version]
  273. Clevers, H. Wnt/β-catenin signaling in development and disease. Cell 2006, 127, 469–480. [Google Scholar] [CrossRef] [Green Version]
  274. Clevers, H.; Nusse, R. Wnt/β-catenin signaling and disease. Cell 2012, 149, 1192–1205. [Google Scholar] [CrossRef] [Green Version]
  275. Monga, S.P. β-catenin signaling and roles in liver homeostasis, injury, and tumorigenesis. Gastroenterology 2015, 148, 1294–1310. [Google Scholar] [CrossRef] [Green Version]
  276. Fu, X.; Zhu, X.; Qin, F.; Zhang, Y.; Lin, J.; Ding, Y.; Yang, Z.; Shang, Y.; Wang, L.; Zhang, Q.; et al. Linc00210 drives Wnt/β-catenin signaling activation and liver tumor progression through CTNNBIP1-dependent manner. Mol. Cancer 2018, 17, 73. [Google Scholar] [CrossRef] [Green Version]
  277. 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]
  278. Xu, W.-H.; Zhang, J.-B.; Dang, Z.; Li, X.; Zhou, T.; Liu, J.; Wang, D.-S.; Song, W.-J.; Dou, K.-F. Long non-coding RNA URHC regulates cell proliferation and apoptosis via ZAK through the ERK/MAPK signaling pathway in hepatocellular carcinoma. Int. J. Biol. Sci. 2014, 10, 664. [Google Scholar] [CrossRef] [Green Version]
  279. Zhang, L.; Yang, F.U.; Yuan, J.H.; Yuan, S.X.; Zhou, W.P.; Huo, X.S.; Xu, D.; Bi, H.S.; Wang, F.; Sun, S.H. Epigenetic activation of the MiR-200 family contributes to H19-mediated metastasis suppression in hepatocellular carcinoma. Carcinogenesis 2013, 34, 577–586. [Google Scholar] [CrossRef] [Green Version]
  280. Dai, X.; Ahn, K.S.; Kim, C.; Siveen, K.S.; Ong, T.H.; Shanmugam, M.K.; Li, F.; Shi, J.; Kumar, A.P.; Wang, L.Z.; et al. Ascochlorin, an isoprenoid antibiotic inhibits growth and invasion of hepatocellular carcinoma by targeting STAT3 signaling cascade through the induction of PIAS3. Mol. Oncol. 2015, 9, 818–833. [Google Scholar] [CrossRef] [PubMed]
  281. Rajendran, P.; Li, F.; Shanmugam, M.K.; Vali, S.; Abbasi, T.; Kapoor, S.; Ahn, K.S.; Kumar, A.P.; Sethi, G. Honokiol inhibits signal transducer and activator of transcription-3 signaling, proliferation, and survival of hepatocellular carcinoma cells via the protein tyrosine phosphatase SHP-1. J. Cell. Physiol. 2012, 227, 2184–2195. [Google Scholar] [CrossRef] [PubMed]
  282. Huang, J.-L.; Cao, S.-W.; Ou, Q.-S.; Yang, B.; Zheng, S.-H.; Tang, J.; Chen, J.; Hu, Y.-W.; Zheng, L.; Wang, Q. The long non-coding RNA PTTG3P promotes cell growth and metastasis via up-regulating PTTG1 and activating PI3K/AKT signaling in hepatocellular carcinoma. Mol. Cancer 2018, 17, 93. [Google Scholar] [CrossRef]
  283. Zhao, X.; Liu, Y.; Yu, S. Long noncoding RNA AWPPH promotes hepatocellular carcinoma progression through YBX1 and serves as a prognostic biomarker. Biochim. Biophys. Acta Mol. Basis Dis. 2017, 1863, 1805–1816. [Google Scholar] [CrossRef]
  284. Tang, J.; Zhuo, H.; Zhang, X.; Jiang, R.; Ji, J.; Deng, L.; Qian, X.; Zhang, F.; Sun, B. A novel biomarker Linc00974 interacting with KRT19 promotes proliferation and metastasis in hepatocellular carcinoma. Cell Death Dis. 2014, 5, e1549. [Google Scholar] [CrossRef] [Green Version]
  285. Zhi, Y.; Huang, S.; Lina, Z. Suppressor of Cytokine Signaling 6 in cancer development and therapy: Deciphering its emerging and suppressive roles. Cytokine Growth Factor Rev. 2022, 64, 21–32. [Google Scholar]
  286. Zhang, L.-H.; Ji, J.-F. Molecular profiling of hepatocellular carcinomas by cDNA microarray. World J. Gastroenterol. 2005, 11, 463. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Aravalli, R.N.; Steer, C.J.; Cressman, E.N. Molecular mechanisms of hepatocellular carcinoma. Hepatology 2008, 48, 2047–2063. [Google Scholar] [CrossRef]
  288. Ogunwobi, O.O.; Puszyk, W.; Dong, H.-J.; Liu, C. Epigenetic upregulation of HGF and c-Met drives metastasis in hepatocellular carcinoma. PLoS ONE 2013, 8, e63765. [Google Scholar] [CrossRef] [PubMed]
  289. Long, L.; Xiang, H.; Liu, J.; Zhang, Z.; Sun, L. ZEB1 mediates doxorubicin (Dox) resistance and mesenchymal characteristics of hepatocarcinoma cells. Exp. Mol. Pathol. 2019, 106, 116–122. [Google Scholar] [CrossRef] [PubMed]
  290. Qu, W.; Wen, X.; Su, K.; Gou, W. MiR-552 promotes the proliferation, migration and EMT of hepatocellular carcinoma cells by inhibiting AJAP 1 expression. J. Cell Mol. Med. 2019, 23, 1541–1552. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  291. Lu, Y.I.; Lin, N.A.N.; Chen, Z.; Xu, R. Hypoxia-induced secretion of platelet-derived growth factor-BB by hepatocellular carcinoma cells increases activated hepatic stellate cell proliferation, migration and expression of vascular endothelial growth factor-A. Mol. Med. Rep. 2015, 11, 691–697. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  292. Galiè, M.; Sorrentino, C.; Montani, M.; Micossi, L.; Di Carlo, E.; D’Antuono, T.; Calderan, L.; Marzola, P.; Benati, D.; Merigo, F.; et al. Mammary carcinoma provides highly tumourigenic and invasive reactive stromal cells. Carcinogenesis 2005, 26, 1868–1878. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  293. Hernandez–Gea, V.; Toffanin, S.; Friedman, S.L.; Llovet, J.M. Role of the microenvironment in the pathogenesis and treatment of hepatocellular Carcinoma. Gastroenterology 2013, 144, 512–527. [Google Scholar] [CrossRef] [Green Version]
  294. Wu, S.; E Powers, S.; Zhu, W.; Hannun, Y.A. Substantial contribution of extrinsic risk factors to cancer development. Nature 2016, 529, 43–47. [Google Scholar] [CrossRef]
  295. Yamashita, T.; Ji, J.; Budhu, A.; Forgues, M.; Yang, W.; Wang, H.Y.; Jia, H.; Ye, Q.; Qin, L.X.; Wauthier, E.; et al. EpCAM-positive hepatocellular carcinoma cells are tumor-initiating cells with stem/progenitor cell features. Gastroenterology 2009, 136, 1012–1024.e4. [Google Scholar] [CrossRef] [Green Version]
  296. Yang, Z.F.; Ho, D.W.; Ng, M.N.; Lau, C.K.; Yu, W.C.; Ngai, P.; Chu, P.W.; Lam, C.T.; Poon, R.T.; Fan, S.T. Significance of CD90+ cancer stem cells in human liver cancer. Cancer Cell 2008, 13, 153–166. [Google Scholar] [CrossRef] [Green Version]
  297. Ma, S.; Chan, K.-W.; Hu, L.; Lee, T.K.-W.; Wo, J.Y.-H.; Ng, I.O.-L.; Zheng, B.-J.; Guan, X.-Y. Identification and characterization of tumorigenic liver cancer stem/progenitor cells. Gastroenterology 2007, 132, 2542–2556. [Google Scholar] [CrossRef] [PubMed]
  298. Zhu, Z.; Hao, X.; Yan, M.; Yao, M.; Ge, C.; Gu, J.; Li, J. Cancer stem/progenitor cells are highly enriched in CD133+ CD44+ population in hepatocellular carcinoma. Int. J. Cancer 2010, 126, 2067–2078. [Google Scholar] [CrossRef] [PubMed]
  299. Haraguchi, N.; Ishii, H.; Mimori, K.; Tanaka, F.; Ohkuma, M.; Kim, H.M.; Akita, H.; Takiuchi, D.; Hatano, H.; Nagano, H.; et al. CD13 is a therapeutic target in human liver cancer stem cells. J. Clin. Investig. 2010, 120, 3326–3339. [Google Scholar] [CrossRef] [Green Version]
  300. Wu, C.-Y.; Hsu, Y.-C.; Ho, H.J.; Chen, Y.-J.; Lee, T.-Y.; Lin, J.-T. Association between ultrasonography screening and mortality in patients with hepatocellular carcinoma: A nationwide cohort study. Gut 2016, 65, 693–701. [Google Scholar] [CrossRef]
  301. Tzartzeva, K.; Obi, J.; Rich, N.E.; Parikh, N.D.; Marrero, J.A.; Yopp, A.; Waljee, A.K.; Singal, A.G. Surveillance imaging and alpha fetoprotein for early detection of hepatocellular carcinoma in patients with cirrhosis: A meta-analysis. Gastroenterology 2018, 154, 1706–1718.e1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  302. Chang, T.-S.; Wu, Y.-C.; Tung, S.-Y.; Wei, K.-L.; Hsieh, Y.-Y.; Huang, H.-C.; Chen, W.-M.; Shen, C.-H.; Lu, C.-H.; Wu, C.-S.; et al. Alpha-fetoprotein measurement benefits hepatocellular carcinoma surveillance in patients with cirrhosis. Am. J. Gastroenterol. 2015, 110, 836–844. [Google Scholar] [CrossRef] [PubMed]
  303. Bisceglie, A.M.; Sterling, R.K.; Chung, R.T.; Everhart, J.E.; Dienstag, J.L.; Bonkovsky, H.L.; Wright, E.C.; Everson, G.T.; Lindsay, K.L.; Lok, A.S.; et al. Serum alpha-fetoprotein levels in patients with advanced hepatitis C: Results from the HALT-C. Trial. J. Hepatol. 2005, 43, 434–441. [Google Scholar] [CrossRef] [PubMed]
  304. Moon, A.M.; Weiss, N.S.; Beste, L.A.; Su, F.; Ho, S.B.; Jin, G.-Y.; Lowy, E.; Berry, K.; Ioannou, G.N. No association between screening for hepatocellular carcinoma and reduced cancer-related mortality in patients with cirrhosis. Gastroenterology 2018, 155, 1128–1139.e6. [Google Scholar] [CrossRef]
  305. Xu, C.; Xu, Z.; Zhang, Y.; Evert, M.; Calvisi, D.F.; Chen, X. β-Catenin signaling in hepatocellular carcinoma. J. Clin. Investig. 2022, 132, e154515. [Google Scholar] [CrossRef]
  306. Li, J.; Yang, X.; Huang, L.; Zhu, X.; Qiu, M.; Yan, J.; Yan, Y.; Wei, S. Treatment Strategy for Post-hepatectomy Recurrent Hepatocellular Carcinoma Within the Milan Criteria: Repeat Resection, Local Ablative Therapy or Transarterial Chemoembolization? Indian J. Surg. 2022, 26, 1–6. [Google Scholar] [CrossRef]
  307. Kamel, M.M.; Matboli, M.; Sallam, M.; Montasser, I.F.; Saad, A.S.; El-Tawdi, A.H. Investigation of long noncoding RNAs expression profile as potential serum biomarkers in patients with hepatocellular carcinoma. Transl. Res. 2016, 168, 134–145. [Google Scholar] [CrossRef]
  308. Kodidela, S.; Behera, A.; Reddy, A.B.M. Risk factors and clinical aspects associated with hepatocellular carcinoma: Role of long noncoding RNAs. In Theranostics and Precision Medicine for the Management of Hepatocellular Carcinoma; Elsevier: Amsterdam, The Netherlands, 2022; pp. 341–356. [Google Scholar]
  309. Luo, P.; Liang, C.; Zhang, X.; Liu, X.; Wang, Y.; Wu, M.; Feng, X.; Tu, J. Identification of long non-coding RNA ZFAS1 as a novel biomarker for diagnosis of HCC. Biosci. Rep. 2018, 31, 38. [Google Scholar] [CrossRef] [Green Version]
  310. Li, Y.; Zhao, J.; Yu, S.; Wang, Z.; He, X.; Su, Y.; Guo, T.; Sheng, H.; Chen, J.; Zheng, Q.; et al. Extracellular vesicles long RNA sequencing reveals abundant mRNA, circRNA, and lncRNA in human blood as potential biomarkers for cancer diagnosis. Clin. Chem. 2019, 65, 798–808. [Google Scholar] [CrossRef] [PubMed]
  311. Xu, H.; Chen, Y.; Dong, X.; Wang, X. Serum exosomal long noncoding RNAs ENSG00000258332. 1 and LINC00635 for the diagnosis and prognosis of hepatocellular carcinoma. Cancer Epidemiol. Biomark. Prev. 2018, 27, 710–716. [Google Scholar] [CrossRef] [Green Version]
  312. Sasaki, R.; Kanda, T.; Yokosuka, O.; Kato, N.; Matsuoka, S.; Moriyama, M. Exosomes and hepatocellular carcinoma: From bench to bedside. Int. J. Mol. Sci. 2019, 20, 1406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  313. Zhang, C.; Ji, Q.; Yang, Y.; Li, Q.; Wang, Z. Exosome: Function and role in cancer metastasis and drug resistance. Technol. Cancer Res. Treat. 2018, 17, 1533033818763450. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  314. 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]
  315. Wu, E.-R.; Hsieh, M.-J.; Chiang, W.-L.; Hsueh, K.-C.; Yang, S.-F.; Su, S.-C. Association of lncRNA CCAT2 and CASC8 gene polymorphisms with hepatocellular carcinoma. Int. J. Environ. Res. 2019, 16, 2833. [Google Scholar] [CrossRef] [Green Version]
  316. Tonus, C.; Cloquette, K.; Ectors, F.; Piret, J.; Gillet, L.; Antoine, N.; Desmecht, D.; Vanderplasschen, A.; Waroux, O.; Grobet, L. Long term-cultured and cryopreserved primordial germ cells from various chicken breeds retain high proliferative potential and gonadal colonisation competency. Reprod. Fertil. Dev. Rep. 2016, 28, 628–639. [Google Scholar] [CrossRef]
  317. Hong, D.; Kurzrock, R.; Kim, Y.; Woessner, R.; Younes, A.; Nemunaitis, J.; Fowler, N.; Zhou, T.; Schmidt, J.; Jo, M.; et al. AZD9150, a next-generation antisense oligonucleotide inhibitor of STAT3 with early evidence of clinical activity in lymphoma and lung cancer. Sci. Transl. Med. 2015, 7, 314ra185. [Google Scholar] [CrossRef] [Green Version]
  318. Arun, G.; Diermeier, S.; Akerman, M.; Chang, K.-C.; Wilkinson, J.E.; Hearn, S.; Kim, Y.; MacLeod, A.R.; Krainer, A.R.; Norton, L.; et al. Differentiation of mammary tumors and reduction in metastasis upon Malat1 lncRNA loss. Genes Dev. 2016, 30, 34–51. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  319. Esposito, R.; Bosch, N.; Lanzós, A.; Polidori, T.; Pulido-Quetglas, C.; Johnson, R. Hacking the cancer genome: Profiling therapeutically actionable long non-coding RNAs using CRISPR-Cas9 screening. Cancer Cell 2019, 35, 545–557. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  320. Huang, M.; Wang, H.; Hu, X.; Cao, X. lncRNA MALAT1 binds chromatin remodeling subunit BRG1 to epigenetically promote inflammation-related hepatocellular carcinoma progression. Oncoimmunology 2019, 8, e1518628. [Google Scholar] [CrossRef] [PubMed]
  321. Tang, S.; Tan, G.; Jiang, X.; Han, P.; Zhai, B.; Dong, X.; Qiao, H.; Jiang, H.; Sun, X. An artificial lncRNA targeting multiple miRNAs overcomes sorafenib resistance in hepatocellular carcinoma cells. Oncotarget 2016, 7, 73257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  322. Finn, R.S.; Qin, S.; Ikeda, M.; Galle, P.R.; Ducreux, M.; Kim, T.-Y.; Kudo, M.; Breder, V.; Merle, P.; Kaseb, A.O.; et al. Atezolizumab plus bevacizumab in unresectable hepatocellular carcinoma. N. Engl. J. Med. 2020, 382, 1894–1905. [Google Scholar] [CrossRef]
  323. Ren, F.; Zhang, Y.; Qin, Y.; Shang, J.; Wang, Y.; Wei, P.; Guo, J.; Jia, H.; Zhao, T. Taraxasterol prompted the anti-tumor effect in mice burden hepatocellular carcinoma by regulating T lymphocytes. Cell Death Discov. 2022, 8, 264. [Google Scholar] [CrossRef]
  324. Su, G.L. AGA Clinical Practice Guideline on Systemic Therapy for Hepatocellular Carcinoma. Gastroenterology 2022, 162, 920–934. [Google Scholar] [CrossRef]
  325. Park, R.; da Silva, L.L.; Nissaisorakarn, V.; Riano, I.; Williamson, S.; Sun, W.; Saeed, A. Comparison of Efficacy of Systemic Therapies in Advanced Hepatocellular Carcinoma: Updated Systematic Review and Frequentist Network Meta-Analysis of Randomized Controlled Trials. J. Hepatocell Carcinoma 2021, 8, 145–154. [Google Scholar] [CrossRef]
  326. Marron, T.U.; Fiel, M.I.; Hamon, P.; Fiaschi, N.; Kim, E.; Ward, S.C.; Zhao, Z.; Kim, J.; Kennedy, P.; Gunasekaran, G.; et al. Neoadjuvant cemiplimab for resectable hepatocellular carcinoma: A single-arm, open-label, phase 2 trial. Lancet Gastroenterol. Hepatol. 2022, 7, 219–229. [Google Scholar] [CrossRef]
  327. Zhang, H.; Zhang, W.; Jiang, L.; Chen, Y. Recent advances in systemic therapy for hepatocellular carcinoma. Biomark. Res. 2022, 10, 3. [Google Scholar] [CrossRef] [PubMed]
  328. Lennox, K.A.; Behlke, M.A. Cellular localization of long non-coding RNAs affects silencing by RNAi more than by antisense oligonucleotides. Nucleic Acids Res. 2016, 44, 863–877. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  329. Peng, Y.; Croce, C.M. The role of MicroRNAs in human cancer. Signal Transduct. Target. Ther. 2016, 1, 1–9. [Google Scholar] [CrossRef] [Green Version]
  330. Wheeler, T.M.; Leger, A.J.; Pandey, S.K.; MacLeod, A.R.; Nakamori, M.; Cheng, S.H.; Wentworth, B.M.; Bennett, C.F.; Thornton, C.A. Targeting nuclear RNA for in vivo correction of myotonic dystrophy. Nature 2012, 488, 111–115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  331. Li, D.; Sedano, S.; Allen, R.; Gong, J.; Cho, M.; Sharma, S. Current treatment landscape for advanced hepatocellular carcinoma: Patient outcomes and the impact on quality of life. Cancers 2019, 11, 841. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  332. Necsulea, A.; Soumillon, M.; Warnefors, M.; Liechti, A.; Daish, T.; Zeller, U.; Baker, J.C.; Grützner, F.; Kaessmann, H. The evolution of lncRNA repertoires and expression patterns in tetrapods. Nature 2014, 505, 635–640. [Google Scholar] [CrossRef]
  333. Juan, V.; Crain, C.; Wilson, C. Evidence for evolutionarily conserved secondary structure in the H19 tumor suppressor RNA. Nucleic Acids Res. 2000, 28, 1221–1227. [Google Scholar] [CrossRef] [Green Version]
  334. Zhou, L.; Zhu, Y.; Sun, D.; Zhang, Q. Emerging roles of long non-coding RNAs in the tumor microenvironment. Int. J. Biol. Sci. 2020, 16, 2094. [Google Scholar] [CrossRef]
  335. Wei, L.; Lee, D.; Law, C.-T.; Zhang, M.S.; Shen, J.; Chin, D.W.-C.; Zhang, A.; Tsang, F.H.-C.; Wong, C.L.-S.; Ng, I.O.-L.; et al. Genome-wide CRISPR/Cas9 library screening identified PHGDH as a critical driver for Sorafenib resistance in HCC. Nat. Commun. 2019, 10, 4681. [Google Scholar] [CrossRef] [Green Version]
  336. Bester, A.C.; Lee, J.D.; Chavez, A.; Lee, Y.-R.; Nachmani, D.; Vora, S.; Victor, J.; Sauvageau, M.; Monteleone, E.; Rinn, J.L.; et al. An integrated genome-wide CRISPRa approach to functionalize lncRNAs in drug resistance. Cell 2018, 173, 649–664.e20. [Google Scholar] [CrossRef]
  337. Liu, S.J.; Horlbeck, M.A.; Weissman, J.S.; Lim, D.A. Genome-Scale Perturbation of Long Noncoding RNA Expression Using CRISPR Interference. In Functional Analysis of Long Non-Coding RNAs; Springer: New York, NY, USA, 2021; pp. 323–338. [Google Scholar]
Figure 1. Genetic alterations in HCC. After gaining necessary genetic and epigenetic variations, cirrhosis develops into dysplastic foci and nodules to form HCC.
Figure 1. Genetic alterations in HCC. After gaining necessary genetic and epigenetic variations, cirrhosis develops into dysplastic foci and nodules to form HCC.
Bioengineering 09 00406 g001
Figure 2. Noncoding RNAs classified into small noncoding RNAs and long noncoding RNAs. Small noncoding RNAs include miRNAs (microRNAs), snRNAs (small nucleolar RNAs), PIWI-interacting RNAs, and endogenous small interfering RNAs. Long noncoding RNAs (lncRNAs) are composed of sense, antisense, bidirectional, enhancer, intergenic, and intronic lncRNAs based on their localizations as compared to the nearby protein-coding genes. LncRNAs could function as competing endogenous RNAs (ceRNAs).
Figure 2. Noncoding RNAs classified into small noncoding RNAs and long noncoding RNAs. Small noncoding RNAs include miRNAs (microRNAs), snRNAs (small nucleolar RNAs), PIWI-interacting RNAs, and endogenous small interfering RNAs. Long noncoding RNAs (lncRNAs) are composed of sense, antisense, bidirectional, enhancer, intergenic, and intronic lncRNAs based on their localizations as compared to the nearby protein-coding genes. LncRNAs could function as competing endogenous RNAs (ceRNAs).
Bioengineering 09 00406 g002
Figure 3. The functions of lncRNAs. LncRNAs perform a key function in gene regulation via a variety of processes, including transcriptional regulation, post-transcriptional regulation, and other mechanisms.
Figure 3. The functions of lncRNAs. LncRNAs perform a key function in gene regulation via a variety of processes, including transcriptional regulation, post-transcriptional regulation, and other mechanisms.
Bioengineering 09 00406 g003
Figure 4. The roles of lncRNAs in HCC. (a) Regulation and modification of chromatin. (b) Transcriptional activation. (c) Interaction with mRNAs. (d) Sponging of microRNAs. (e) Protein binding and modification. (f) Other mechanisms and pathways of lncRNAs.
Figure 4. The roles of lncRNAs in HCC. (a) Regulation and modification of chromatin. (b) Transcriptional activation. (c) Interaction with mRNAs. (d) Sponging of microRNAs. (e) Protein binding and modification. (f) Other mechanisms and pathways of lncRNAs.
Bioengineering 09 00406 g004
Table 1. Biological functions of lncRNAs in cancers.
Table 1. Biological functions of lncRNAs in cancers.
LncRNAsTarget Pathways/
Mechanisms
Biological Functions in CancersType of CancerReference
LncRNA 00665miR-224-5p/VMA21Promoting proliferation, invasion, and migration of cancer cellsMelanoma[146]
lncRNA RACGAP1PmiR-345-5p/RACGAP1Breast cancer[145]
SNHG20miR-148a/ROCK1Ovarian cancer[153]
UCA1miR-206Cervical cancer[154]
VCAN-AS1p53Gastric cancer[155]
LINC01559YAPPancreatic cancer[156]
SNHG4ZIC5Prostate cancer[157]
TTN-AS1KLF15Colorectal cancer[158]
LINC00673miR-515-5p/MARK4/HippoBreast cancer[159]
RAINRUNX2Breast/thyroid[160]
PVT1Smad3/miR-140-5pCervical cancer[161]
FOXD2-AS1miR-185-5pThyroid cancer[162]
LINC00052miR-608/EGFRHead/neck cancer[163]
TCONS-00020456Smad2/PKCaSuppression of proliferation and invasion of cancer cellsGlioblastoma cancer[164]
ADAMTS9-AS2CDH3Esophageal cancer[165]
ENST00000489676MiR-922Thyroid cancer[166]
OSER1-AS1miR-372-3p/Rab23Hepatocellular carcinoma[167]
HOXA-AS3HOXA3Prognosis and efficacyNSCL cancer[168]
ADAMTS9-AS2FUS/MDM2Glioblastoma cancer[169]
UCA1, H195-fluorouracilRectal cancer[170]
SNHG12------Potential biomarkersPan-cancer[171]
HOTAIR------Breast cancer[172]
SNHG11------Colorectal cancer[173]
Table 2. Biological functions of lncRNAs in hepatocellular carcinoma (HCC).
Table 2. Biological functions of lncRNAs in hepatocellular carcinoma (HCC).
LncRNAsTarget Pathways/MechanismsBiological Functions in HCCReference
LncRNA CYTORmiR-125b/SEMA4CPromoting proliferation, invasion, and migration of cancer cells
Angiogenesis and metastasis
Tumorigenesis and EMT
Growth and metastasis
Progression and angiogenesis
[181]
DNAJC3-AS1miR-27b[182]
LncRNA SNHG8miR-542-3p and miR-4701-5p[175]
MCM3AP-AS1miR-194-5p/FOXA1 axis[183]
RNA LINC00908Sox-4[184]
SNHG15miR-490-3p/histone deacetylase 2 axis[185]
GIHCGmiR-200b/a/429 PPAR gamma[186]
ANRILEZH2 protein Target gene DNA[187]
TUG1EZH2 protein Target gene DNA[188]
UFC1β-catenin mRNA HuR protein[189]
MALAT1miR-143-3p[190]
ICRICAM-1 mRNA[191]
ZFAS1miR-150[192]
MVIHPGK1 protein[193]
CASC9HNRNPL protein[194,195]
LncCAMTA1CAMTA1[196]
FtxPPAR gamma[197]
ATBAutophagy-related protein[198]
PDPK2PPDK1/AKT/Caspase 3[199]
HOXD-AS1SOX4[200]
HISERK&AKT/GSK-3b[201]
HOTAIROGFr, miR-122, SETD2[202,203,204,205,206]
LINC00161Activate ROCK2, miR-590-3p[207]
DLGAP1-AS1miR-26a/b-5p/IL-6/JAK2/STAT3[208]
91HIGF2[209]
MYLK-AS1miR-424-5p/E2F7 & activating VEGFR-2[210]
Linc-RORDEPDC1[211]
HULCHULC/miR-383-5p/VAMP2[212]
LINC00238miR-522/SFRP2/DKK1Suppression of proliferation, invasion, and migration
Suppression of HCC progression
[213]
TMEM220-AS1TMEM220/β-catenin[214]
NBR2JNK/ERK[215]
lncRNA W5---------[216]
GAS8-AS1GAS8[217]
MIR22HGmiR-10a-5p/NCOR2[218]
MIR31HGmicroRNA-575[219]
GAS5miR182/ANGPTL1[220]
FENDRRmiR-423-5p[221]
EPB41L4A-AS2miR301a-5p/FOXL1[222]
TCONS_00006195ENO1[223]
Uc.134LATS1[224]
SVUGP2MMP2 and 9[225]
RP11-286H15.1PABPC4 Ubiquitination[226]
LncRNA-DrehVimentin protein[227]
XISTmiR-92b[228]
LINC00221lncRNA–miRNA–mRNA
miR-485-5p/BSG
miR-195-5p/MACC1
----
HNRNPA2B1/NF-KB
Caspase-8/LSD1/H3K9me3
------
------
miR-195/EYA1 axis
Prognosis and efficacy[49]
[229]
[230]
[231]
[232]
[204]
[82]
[233]
LOC554202
LncRNA DDX11-AS1
RP11-464I1.1
miR503HG
MALAT1, HOTAIR, MDG
HOTAIR
MIR22HG, CTC-297N7.9,
CTD-2139B15.2, RP11-589N15.2,
RP11-343N15.5, and
RP11-479G22.8
LINC00511
lncRNA W42DBN1
miR-448/ROCK1
-----
-----
-----
-----
-----
Potential biomarkers[67]
[120]
[234]
[235]
[236]
[237]
[238]
PITPNA-AS1
PVT1, uc002mbe.2 e
UCA1
RP11-486O12.2, RP11-273G15.2, RP11 863K10.7 and LINC01093
LRB1
ELMO1-AS1
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Khan, A.; Zhang, X. Function of the Long Noncoding RNAs in Hepatocellular Carcinoma: Classification, Molecular Mechanisms, and Significant Therapeutic Potentials. Bioengineering 2022, 9, 406. https://doi.org/10.3390/bioengineering9080406

AMA Style

Khan A, Zhang X. Function of the Long Noncoding RNAs in Hepatocellular Carcinoma: Classification, Molecular Mechanisms, and Significant Therapeutic Potentials. Bioengineering. 2022; 9(8):406. https://doi.org/10.3390/bioengineering9080406

Chicago/Turabian Style

Khan, Ahmad, and Xiaobo Zhang. 2022. "Function of the Long Noncoding RNAs in Hepatocellular Carcinoma: Classification, Molecular Mechanisms, and Significant Therapeutic Potentials" Bioengineering 9, no. 8: 406. https://doi.org/10.3390/bioengineering9080406

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