Next Article in Journal
Effect of Low Testosterone Levels on the Expression of Proliferator-Activated Receptor Alpha in Female Patients with Primary Biliary Cholangitis
Previous Article in Journal
Vitamin D Influences the Activity of Mast Cells in Allergic Manifestations and Potentiates Their Effector Functions against Pathogens
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A New Understanding of Long Non-Coding RNA in Hepatocellular Carcinoma—From m6A Modification to Blood Biomarkers

1
Department of Gastroenterology, Ajou University School of Medicine, 164 World cup-ro, Yeongtong-gu, Suwon 16499, Republic of Korea
2
Department of Biochemistry, College of Medicine, Kosin University, Seo-gu, Busan 49267, Republic of Korea
3
Institute for Medical Science, College of Medicine, Kosin University, Seo-gu, Busan 49267, Republic of Korea
*
Author to whom correspondence should be addressed.
Cells 2023, 12(18), 2272; https://doi.org/10.3390/cells12182272
Submission received: 21 August 2023 / Revised: 12 September 2023 / Accepted: 13 September 2023 / Published: 14 September 2023
(This article belongs to the Section Cellular Immunology)

Abstract

:
With recent advancements in biological research, long non-coding RNAs (lncRNAs) with lengths exceeding 200 nucleotides have emerged as pivotal regulators of gene expression and cellular phenotypic modulation. Despite initial skepticism due to their low sequence conservation and expression levels, their significance in various biological processes has become increasingly apparent. We provided an overview of lncRNAs and discussed their defining features and modes of operation. We then explored their crucial function in the hepatocarcinogenesis process, elucidating their complex involvement in hepatocellular carcinoma (HCC). The influential role of lncRNAs within the HCC tumor microenvironment is emphasized, illustrating their potential as key modulators of disease dynamics. We also investigated the significant influence of N6-methyladenosine (m6A) modification on lncRNA function in HCC, enhancing our understanding of both their roles and their upstream regulators. Additionally, the potential of lncRNAs as promising biomarkers was discussed in liver cancer diagnosis, suggesting a novel avenue for future research and clinical application. Finally, our work underscored the dual potential of lncRNAs as both contributors to HCC pathogenesis and innovative tools for its diagnosis. Existing challenges and prospective trajectories in lncRNA research are also discussed, emphasizing their potential in advancing liver cancer research.

1. Introduction

In recent decades, advancements in biological technology have profoundly evolved our comprehension of genomic information. This shift was largely attributed to the discovery of non-coding RNAs (ncRNAs), which, while not encoding proteins, regulate various biological processes [1]. The categories of ncRNAs—microRNAs (miRNAs), circular RNAs (circRNAs), PIWI-interacting RNAs (piRNAs), and small nucleolar RNAs (snoRNAs)—each have unique roles, such as gene expression regulation, miRNA sponging, genome stability maintenance, and guiding chemical RNA modifications, respectively [2,3].
Among these ncRNAs, long non-coding RNAs (lncRNAs)—transcripts longer than 200 nucleotides—have emerged as crucial regulators in determining cellular fate. They serve as critical molecular players in diverse biological processes, including X-chromosome inactivation and genomic imprinting [4]. These lncRNAs were initially considered as non-functional transcripts due to their generally low sequence conservation and expression levels. However, the increasing number of publications highlighting the dynamic expression and biological functions of lncRNAs, together with the advent of novel technologies facilitating their identification and characterization, have redefined our understanding of this perception [5]. Moreover, their aberrant expression and function have been implicated in various cancers, particularly hepatocellular carcinoma (HCC) [6].
This review explored the recent advances in research on the complex roles and mechanisms of lncRNA dysregulation in HCC. We provided a comprehensive exploration of their implications in disease pathogenesis and potential as diagnostic markers. Additionally, we included information specifically on the N6-methyladenosine (m6A) modification, offering a more detailed understanding of its influence. Luo et al. concluded a critical facet of lncRNA biology is the m6A modification; it represents the most predominant internal modification observed in eukaryotic RNAs, including lncRNAs, exerting a pivotal role in RNA metabolism and functionality [7]. The modulation of lncRNA structures and functions by m6A introduces additional layers of regulatory intricacy [7,8]. Importantly, Sivasudhan et al. highlighted in their review the significance of lncRNA alterations via m6A modifications, emphasizing their potential influence on disease progression and therapeutic outcomes in HCC [9]. Finally, we identified current challenges and discussed potential future research directions in this compelling field. Despite the uncertainty and controversy that have involved lncRNA research, these molecules play pivotal roles in regulating cellular functions, with significant implications for HCC, highlighting their significance in biomedical research.

2. Understanding lncRNA: Its Definition and Mechanism of Action

LncRNAs play a pivotal role in the intricate regulation of gene expression. As previously reviewed, lncRNAs’ influence spans multiple levels of gene regulation, from reshaping chromatin structures to guiding post-transcriptional modifications [10,11]. A fundamental mechanism underlying their function is their ability to interact with various cellular components, such as DNA, RNA, and proteins [12]. Through these interactions, lncRNAs become essential modulators of both cellular architecture and activity.

2.1. Signal

One of the primary modes of lncRNA function is acting as signals (Figure 1A). These molecular indicators are transcribed in response to diverse cellular stimuli, serving as indicators of specific cellular states or events. Their transcription often reflects changes in the cellular environment, such as stress, differentiation, or developmental cues. For instance, the lncRNA HOTAIR is transcribed in response to certain oncogenic signals and plays a pivotal role in regulating gene expression patterns associated with cancer progression [13]. Similarly, XIST is transcribed during early female embryonic development, signaling the initiation of X-chromosome inactivation [4]. These lncRNAs, by acting as signals, provide the cell with a dynamic mechanism to rapidly respond to internal or external changes, ensuring appropriate cellular reactions and adaptations. Their signaling role underscores the complexity and versatility of the non-coding genome in cellular regulation and function.

2.2. Decoy

Another well-known function of lncRNAs is acting as decoys (Figure 1B). In this capacity, lncRNAs can bind to transcription factors or other proteins, effectively sequestering them away from their target genomic loci [14,15]. For instance, the lncRNA Growth Arrest-Specifc 5 (GAS5) serves as a decoy by binding to the glucocorticoid receptor (GR). Under conditions of growth arrest, GAS5 accumulates and binds to the DNA-binding domain of the GR. This prevents the GR from binding to its glucocorticoid response elements (GREs) on DNA, thereby inhibiting the transcription of glucocorticoid-responsive genes. As a result, GAS5 plays a crucial role in modulating cell growth and the cellular response to stress through its decoy function [16]. These lncRNA–protein interactions are responsive to cellular conditions, with factors such as environmental stress potentially affecting lncRNA expression levels and, in turn, their decoy function.

2.3. Guide

LncRNAs can also act as guides, directing the localization of chromatin-modifying enzymes to specific genomic regions (Figure 1C). This targeted recruitment enables precise epigenetic modifications, which can subsequently lead to changes in gene expression profiles. As an example, a recent study demonstrated that TARID (TCF21 antisense RNA inducing demethylation) functioned by partnering with growth arrest and DNA-damage-inducible alpha protein (GADD45A). This partnership steered the DNA demethylation machinery to specific gene loci in cancer cells, thereby regulating gene expression [17].

2.4. Scaffold

In their function as scaffolds, lncRNAs can facilitate the formation of multi-protein complexes, providing a structural platform for the assembly of these complexes. This function aids in the organization and coordination of various cellular processes, including signal transduction and the regulation of gene expression. For instance, the RNA-binding NONO-PSF heterodimer and NEAT1 are both involved in enhancing the processing of primary miRNAs (pri-miRNAs) in HeLa cells. They interact with each other and other RNA-binding proteins, facilitating the microprocessor’s access to pri-miRNAs. This suggests that lncRNA may play a central role in regulating small noncoding RNAs in the nucleus [18].

2.5. Enhancer RNA

LncRNAs can function as enhancers, augmenting the transcription of nearby genes by looping the DNA, bringing distant regions into close proximity for transcriptional activation (Figure 1E). They can modulate transcription either in cis, regulating neighboring genes on the same chromosome, or in trans, influencing genes on different chromosomes [19]. Miao et al. identified the role of LEENE (a lncRNA that enhances eNOS expression) in enhancing the expression of endothelial nitric oxide synthase (eNOS), a key factor in vascular function. LEENE facilitated RNA Pol II’s binding to the eNOS promoter, positively modulating the synthesis of eNOS mRNA and promoting endothelial function [20].
Beyond their known roles, lncRNAs possess a plethora of unexplored functions, validating the perspective that they might perform a limitless array of tasks within a biological context. Not only are lncRNAs pivotal for cellular differentiation and development, but they also influence an extensive range of physiological processes. These encompass DNA damage response, immune system regulation, metabolic processes, and synapse function [21,22,23,24]. The initial skepticism about their importance, due to their lack of protein-coding potential, has been substantially overcome by a growing body of research. This shift in perspective underlines not only the importance of understanding the full range of lncRNA functions, but also their potential implications in health and disease.

2.6. MiRNA Sponge

One of the prominent functions of lncRNAs is their ability to act as miRNA sponges. By binding to miRNAs, these lncRNAs effectively inhibit the miRNAs from associating with their target mRNAs, thereby modulating post-transcriptional regulation (Figure 1F).
A classic example is the lncRNA HOTAIR, which has been reported to sponge miR-34a, a tumor suppressor miRNA. By sequestering miR-34a, HOTAIR can promote oncogenic pathways in certain cancer types, pointing to the significance of lncRNA–miRNA interactions in disease progression [25,26,27]. Furthermore, lncRNA–miRNA interactions introduce an added layer of post-transcriptional regulation complexity. These interactions not only modulate individual gene expressions, but also impact broader cellular pathways and processes. Such processes include cell proliferation, differentiation, and apoptosis, where the balance between lncRNAs and miRNAs plays a crucial role [28,29,30].

3. Pivotal Role of lncRNA in Hepatocarcinogenesis

Due to the increasing demand for better diagnostic and therapeutic approaches, the molecular underpinnings of HCC have become a focal point of research. As Ghafouri-Fard et al. and Abbastabar et al. concluded in their reviews, multiple lncRNAs have been shown to either promote or inhibit HCC progression, affecting processes such as cell proliferation, migration, and invasion [31,32]. The following lncRNAs exemplify these characteristics in the context of HCC.

3.1. HOTAIR (HOX Transcript Antisense RNA)

HOTAIR is an oncogenic lncRNA found to be upregulated in HCC. It originates from the antisense strand of the Homeobox (HOX) gene cluster and plays a significant role in cancer progression [33]. HOTAIR was reported to play a crucial role in HCC by regulating cell growth, migration, invasion, and apoptosis. Its overexpression was often associated with greater tumor size, metastasis, and poor prognosis [34,35,36].
Specifically, HOTAIR plays a key role through its interactions with Polycomb Repressive Complex 2 (PRC2) and Lysine Specific Demethylase 1 (LSD1), which are crucial players in gene silencing [37]. More specifically, HOTAIR acts as a scaffold, binding to EZH2, a subunit of PRC2, thereby facilitating PRC2’s role in repressing gene transcription through the trimethylation of histone H3 at lysine 27 (H3K27me3) [38]. At the same time, HOTAIR can bind to LSD1, which in cooperation with CoREST/REST forms a multi-protein complex involved in gene silencing. This interaction allows HOTAIR to contribute to the silencing of miRNA through its association with PRC2 and LSD1 [39]. Furthermore, HOTAIR is implicated in the regulation of SUZ12, a key binding subunit of PRC2. Overexpression of HOTAIR accelerates the proteasome degradation of SUZ12 and enhances the ubiquitination of SUZ12, facilitated by PLK1. Another interaction of HOTAIR is with the DEAD-box helicase protein 5 (DDX5). This interaction stabilizes SUZ12, reinforcing SUZ12- and PRC2-mediated gene silencing. DDX5 replaces the original Mex-3 RNA binding family member B (Mex3b) linked to HOTAIR, thereby stabilizing the HOTAIR–PRC2 interaction [40]. In HCC patients, overexpression of HOTAIR and PLK1, more than twice the normal levels, was associated with a significant increase in the expression of PRC2 target genes and EPCAM, underscoring the impact of HOTAIR on the epigenetic regulation in HCC [39].
Moreover, HOTAIR also functions as a molecular sponge, sequestering miRNAs like miR-218, which suppresses tumorigenesis. By sponging miRNAs, HOTAIR prevented their anti-cancer effects, leading to enhanced cell proliferation and metastasis [34].

3.2. NEAT1 (Nuclear Enriched Abundant Transcript 1)

NEAT1 is a central component of paraspeckles, specialized sub-nuclear bodies, and plays a pivotal role in their formation and integrity [41]. Over the years, research has shown that NEAT1 is aberrantly upregulated in a variety of cancers. This heightened expression often correlates with a poorer prognosis for patients, making it a potential biomarker for disease progression [42,43,44]. In the context of liver diseases, NEAT1’s role is multifaceted. It has been implicated in accelerating the progression of non-alcoholic fatty liver disease (NAFLD), liver fibrosis, and HCC. However, in conditions characterized by an acute deterioration of liver function in patients with pre-existing chronic liver disease, NEAT1 assumes a protective role by mitigating the inflammatory response [45].
Additional studies in HCC confirmed that NEAT1 is typically overexpressed, promoting cell proliferation, migration, and invasion [46]. Mechanistically, NEAT1 forms a complex with U2AF65, which in turn boosts the expression of hnRNP A2, a known driver in HCC [47]. Another layer of regulation involves HIF-2α, which enhances NEAT1 expression, subsequently influencing the epithelial-mesenchymal transition, a critical process in cancer metastasis [48]. A significant aspect of NEAT1’s function is its ability to act as a molecular sponge for a range of miRNAs. By sequestering these miRNAs, NEAT1 restores the expression of specific genes that these miRNAs would otherwise inhibit [49,50,51].
Apart from its role as a miRNA sponge, NEAT1 also plays a pivotal role in ferroptosis, a unique form of cell death driven by iron-dependent lipid peroxidation, crucial for tumor development and drug resistance [52]. Recent studies have shown that two ferroptosis inducers, erastin and RSL3, elevate NEAT1 expression by enhancing p53’s binding to the NEAT1 promoter. Once upregulated, NEAT1 boosts MIOX expression by competitively binding to miR-362-3p. This leads to an increase in ROS production and a decrease in intracellular NADPH and GSH levels, amplifying the effects of erastin and RSL3. Notably, overexpressing NEAT1 enhances the anti-tumor effects by intensifying ferroptosis in vitro and in vivo [51].
Additionally, NEAT1’s role extends to influencing drug resistance in HCC. It had been shown to synergistically enhance cisplatin resistance in certain liver cancer cells [53]. Furthermore, its involvement in the resistance to sorafenib, a primary therapeutic agent for HCC, has been shown. To be specific, inhibition of NEAT1 amplifies the efficacy of sorafenib, resulting in increased drug-induced cell death and notably smaller tumors in nude mice than with just sorafenib treatment [54].

3.3. HULC (Highly Upregulated in Liver Cancer)

In liver cancer, an increase in the lncRNA HULC, driven by the protein CREB, significantly influences cellular mechanisms by altering YB-1 phosphorylation patterns, which is key in hepatocarcinogenesis [55,56]. HULC also blocks the programmed cell death, or apoptosis, in these cancer cells, specifically when triggered by miR-9 [57]. Another protein, Hepatitis B virus X (HBX), also induces increased levels of HULC and reduces p18, which in turn helps HCC to grow [58]. Given these roles, the overexpression of HULC suggests its potential as a noninvasive biomarker for diagnosis and prognosis [59]. This previous research highlighted the role of HULC in the growth and progression of HCC cells.

3.4. MALAT1 (Metastasis Associated Lung Adenocarcinoma Transcript 1)

MALAT1 is a recognized oncogenic lncRNA that plays a crucial role in the progression of HCC. This lncRNA functions through various pathways, particularly serving as a molecular sponge, and has been observed to be overexpressed in HCC [60]. To be specific, the function of MALAT1 is its ability to bind and sequester various miRNAs, thereby influencing their target function. For instance, MALAT1 reduces the expression of miR-204, leading to an increase in SIRT1 levels and the facilitation of the epithelial–mesenchymal transition (EMT) [61]. Additionally, MALAT1 sequesters miR-143-3p, leading to the upregulation of FGF1- and EMT-promoting proteins [62]. Furthermore, MALAT1’s interaction with miR-200a leads to increased levels of proteins involved in EMT and cell proliferation [63]. Moreover, MALAT1’s interaction with miRNAs such as miR-124-3p and miR-195 results in an upregulation of cell proliferation and invasion facilitating proteins, further highlighting its role in HCC progression [64,65]. Lastly, through downregulation of miR-22, MALAT1 promotes EMT and recruits EZH2 to suppress E-cadherin and miR-22 expression [66].
Beyond the aforementioned lncRNAs, a vast number of additional lncRNAs have been recently discovered and their functions are deeply intertwined with the onset and progression of HCC (Table 1). This has solidified their position as a distinct subject of study within the field of oncology.

4. LncRNAs as Key Modulators of the HCC Tumor Microenvironment

The evolving landscape of cancer biology has recognized the pivotal role of lncRNAs in modulating the tumor microenvironment (TME). Park et al. reviewed that these lncRNAs, by orchestrating intricate interactions within the TME, contribute to various aspects of cancer progression including uncontrolled growth, metastasis, and immune evasion [100].
The immune landscape within HCC is a complex interplay of multiple cell types, influenced and orchestrated by lncRNAs. Jiang et al. demonstrated that the polarization of macrophages, a crucial cellular component in the environment of HCC, is indicative of this interplay [101].
To be specific, the overexpression of MALAT1 in HCC cells promotes angiogenesis and fosters an immunosuppressive environment. This occurs through MALAT1’s interaction with miR-140, inhibiting miR-140’s activity, and consequently increasing VEGF-A production which aids HCC progression by promoting angiogenesis and favoring the polarization of macrophages towards the M2 immunosuppressive subset [102].
Beyond macrophages, lncRNAs also play significant roles in modulating T-cell functions within HCC, impacting the disease’s progression and immune escape mechanisms. For example, lncRNA epidermal growth factor receptor (lnc-EGFR) is highly expressed in regulatory T cells (Tregs) in HCC. Lnc-EGFR interacts with EGFR, inhibits its ubiquitination by c-CBL, and amplifies downstream signaling via AP-1/NFAT1, promoting Treg-cell differentiation and immune evasion [103].
On the contrary, the lncRNA fetal-lethal non-coding developmental regulatory RNA (FENDRR) acts as a sponge for miR-423-5p, impeding Tregs’ immune-suppressive activities. Overexpressed FENDRR competitively binds miR-423-5p and upregulates growth arrest and DNA-damage-inducible beta protein (GADD45B), which inversely correlate with Treg-cell number, thereby reducing immunosuppressive cytokines TGF-β and IL-10 and inducing tumor-cell apoptosis [96]. Moreover, lncRNAs Tims and lncNNT-AS1 are associated with reduced infiltration of tumor CD4 and CD8 T cells, influencing clinical outcomes and responses to immunotherapies [104].
Another lncRNA, Myocardial Infarction Associated Transcript (MIAT), shows elevated expression in various cells implicated in the disease, including tumor cells themselves: FoxP3+ Tregs, PD-1+ CD8+ T cells, and GZMK+ CD8+ T cells. Furthermore, the upregulation of MIAT is associated with how well patients respond to sorafenib. This lncRNA’s expression also has a significant correlation with the presence of PD-L1, a protein involved in immune evasion by cancer cells [105].

5. Implications of m6A Modification on lncRNA in HCC

N6-methyladenosine (m6A) is the most prevalent internal modification in eukaryotic RNA. It plays critical roles in various biological processes, including mRNA splicing, export, stability, and translation efficiency [106,107,108]. The addition of the m6A modification is catalyzed by an enzyme complex known as the m6A methyltransferase complex, including METTL3, METTL14, KIAA1429, RBM15, and WTAP [109,110,111,112]. On the other hand, demethylases, such as FTO and ALKBH5, remove m6A modifications [99,113,114]. Proteins that recognize m6A modifications, often referred to as reader proteins, such as YTH domain-containing proteins, recognize these modifications and influence the fate of m6A-modified RNAs [115,116,117]. Figure 2 depicts the mode of action for the m6A modification mechanism of this protein complex.
Lately, many studies have demonstrated the intricate roles of m6A modification on lncRNAs in HCC progression. One such example is LINC00958, a lncRNA found to be overexpressed in HCC. LINC00958 acted as a molecular sponge for miR-3619-5p, leading to the upregulation of hepatoma-derived growth factor (HDGF), thereby promoting HCC progression. Its overexpression was facilitated by the m6A methyltransferase METTL3, suggesting a critical interplay between m6A modification and lncRNA functionality [118].
An m6A-related lncRNA prognostic signature involving LINC02362, SNHG20, and SNHG6 was identified as a powerful predictor of patient survival, reflecting the close connection between the m6A modification landscape, lncRNA dynamics, and patient outcomes [119]. Additionally, MEG3 demonstrated tumor-suppressive roles through the miR-544b/BTG2 signaling pathway upon upregulation by m6A modification [94].
In a related study, ALKBH5-mediated m6A demethylation downregulated LINC02551, a crucial lncRNA for HCC growth and metastasis, indicating the importance of the balance between methylation and demethylation processes in the m6A–lncRNA axis [86]. Further investigations revealed the promoting role of METTL16 in HCC progression through the downregulation of the tumor suppressor RAB11B-AS1 via an m6A–YTHDF2-dependent mechanism [99].
The role of m6A modification was also underscored in immune evasion, with lipopolysaccharide (LPS) found to increase PD-L1 expression through the m6A modification of MIR155HG, a process essential for HCC immune evasion [120]. The lncRNA ARHGAP5-AS1, which exhibited elevated m6A levels on its transcript, is overexpressed in HCC. It was modulated by METTL14, which functions as its m6A writer, and by IGF2BP2, which acts as the m6A reader. Interestingly, the oncogenic ARHGAP5-AS1 diminished the interactions between CSDE1 and TRIM28, effectively preventing the proteasomal degradation of CSDE1. As a consequence of this interaction, CSDE1 was enabled to coordinate oncogenic RNA regulons, which in turn activate the ERK pathway, a critical player in the prognosis of HCC [121].
Lastly, the lncRNA miR4458HG was discovered to influence HCC-cell proliferation, activate the glycolysis pathway, and promote tumor-associated macrophages’ polarization, highlighting its oncogenic role in HCC patients with high glucose metabolisms [122]. The lncRNAs regulated by m6A in HCC are summarized in Table 2.
In conclusion, these findings underscore the multi-faceted influence of m6A modification in the regulation of lncRNA functions, thereby controlling the progression of HCC. The m6A–lncRNA axis provides a connection between RNA modification and the complex networks of non-coding RNAs, potentially offering some insights into HCC pathogenesis, suggesting possibilities for new therapeutic approaches.

6. LncRNAs as Serum Biomarkers in Liver Cancer

Beylerli et al. highlighted that lncRNAs are secreted by tumor cells into human biological fluids, forming stable circulating lncRNAs resistant to RNA degradation. Aberrant expression of these lncRNAs has been observed in cancer patients [126]. Thus, for HCC diagnosis and prognosis, lncRNAs are increasingly being recognized as a potent alternative to traditional biomarkers such as alpha-fetoprotein (AFP). The efficacy of AFP as an early detector for HCC has been debated due to concerns regarding its sensitivity and specificity [127,128,129,130,131,132]. LncRNAs offer enhanced sensitivity and specificity, potentially addressing the limitations posed by traditional markers such as AFP [133,134,135]. Consequently, there is a growing demand for new diagnostic markers that could replace AFP, and lncRNAs could serve as one such alternative. Therefore, the expression levels of these lncRNAs could serve as key indicators of disease progression and prognosis.
Notably, lncRNAs such as MVIH, X91348, and HOTTIP have shown potential as prognostic markers in HCC [136,137,138]. The high expression of MVIH, associated with microvascular invasion, is a known independent risk factor for recurrence-free survival and overall survival in HCC patients [136]. Similarly, the low expression of lncRNA X91348 in HCC patients relative to healthy individuals was associated with increased overall survival [137]. Moreover, elevated HOTTIP levels were linked to increased tumor recurrence and decreased survival rates in HCC patients following liver transplantation. Conversely, a decrease in HOTTIP expression correlated with more favorable patient outcomes. Hence, HOTTIP could serve as a significant prognostic marker and potential therapeutic target for HCC [138].
With the progression towards minimally invasive and non-invasive diagnostic techniques, circulating lncRNAs in serum are being extensively studied. For instance, high serum levels of lncRNA-ATB were associated with overall survival, progression-free survival, tumor size, TNM stage, C-reactive protein levels, T stage, and portal vein thrombosis, highlighting their potential as serum biomarkers in HCC patients [139].
Exosomal lncRNAs present multiple benefits when considered as biomarkers. Protected from degradation by RNases within exosomes, these lncRNAs remained stable and detectable in various body fluids, making them a potential non-invasive diagnostic tool [135,140,141,142,143,144]. Moreover, given the tissue- or disease-specific nature of many lncRNAs, the detection of specific exosomal lncRNAs might indicate distinct cancer types. For instance, a signature composed of two lncRNAs, PVT1 and uc002mbe.2, demonstrated satisfactory sensitivity and specificity values for distinguishing liver cancer patients from healthy individuals, thus underscoring their potential as specific biomarkers for HCC [145]. Moreover, the lncRNA LINC00853 not only possessed potential diagnostic value, but it also showed prognostic relevance in HCC. Importantly, increased expression of LINC00853 is associated with lower survival rates in patients with stage II HCC according to the modified Union for International Cancer Control (mUICC II). This underlines the potential of LINC00853 as a liver-cancer-specific marker, providing an avenue for both disease identification and assessment of its progression [134].
Furthermore, several studies identified lncRNAs, such as UCA1 and WRAP53, as promising biomarkers in HCC diagnosis when used in conjunction with AFP [146,147]. Another group also identified LINC00152, RP11-160H22.5, and XLOC014172 as new biomarkers for HCC. These lncRNAs, in combination with the conventional marker AFP, were found to improve the diagnostic accuracy for HCC, indicating their potential for enhancing HCC diagnosis [148].
Within the context of chemotherapy resistance, certain lncRNAs, such as CAHM, were identified as key predictive markers. Utilizing machine learning algorithms, CAHM was characterized as a central lncRNA, with elevated expression in sorafenib-resistant cell lines, highlighting its prospective role as a biomarker for chemotherapy resistance [149]. The names and expression tendencies of lncRNAs with diagnostic potential for HCC are summarized in Table 3.
In summary, lncRNAs have emerged as promising biomarkers for HCC, with their roles in diagnostic and prognostic precision medicine becoming evident. Nonetheless, the literature lacks comparative studies between established biomarkers like PIVKA-II and new lncRNA biomarkers [160]. To validate lncRNAs’ clinical significance, extensive multicentric studies are essential. Additionally, standardizing methodologies for detecting circulating lncRNAs is vital to ensure consistent results. Despite challenges, exploring lncRNAs as HCC biomarkers promises to enhance diagnosis, prognosis, and targeted treatments for this malignancy.

7. Summary and Future Perspectives

The discovery and subsequent study of lncRNAs significantly reshaped our understanding of genomic regulation. These non-coding transcripts, although initially dismissed as incidental transcriptional byproducts, have since been revealed as vital players in gene expression regulation and various biological processes [161,162]. This is particularly true for HCC, where lncRNA dysregulation is closely linked to disease pathogenesis [31].
Fortunately, the environment and numerous techniques for lncRNA research are steadily improving, and there is a wealth of web-based tools and publicly available data to facilitate the study of lncRNAs. These resources have expanded our ability to explore the multifaceted roles of lncRNAs and the mechanisms underlying their regulation (Table 4).
Such advancements in bioinformatics technology have further facilitated our ability to uncover the roles of lncRNAs in HCC, from influencing disease pathogenesis to acting as potential diagnostic markers. One such example is the exploration of upstream regulators like the m6A modification, unveiling lncRNA regulation and adding complexity to our understanding of their role in HCC. While current research predominantly revolves around m6A and its relationship with lncRNA, there are various RNA modifications similar to m6A, such as 5-methylcytosine (m5C), N7-methylguanosine (m7G), and 3-methylcytidine (m3C) [177]. The roles and biological functions of these diverse RNA modifications are not yet well-understood. Given this, they present promising research topics in relation to cancer etiology.
In future studies, research on lncRNAs holds promising potential to open new avenues for therapeutic intervention. Despite initial uncertainties and controversies, the role of lncRNAs in cellular function regulation and their implications in HCC have underscored their importance in biomedical research. Subsequent research endeavors should persist in elucidating the complexities of lncRNA function and dysregulation, deepening our comprehension of HCC, and establishing the foundation for novel diagnostic and therapeutic approaches. The challenge will be to translate this expanding knowledge into clinical applications, advancing lncRNA research from the bench to the bedside.

Author Contributions

Conceptualization, J.W.E. and H.S.K.; investigation, H.S.K.; writing—original draft preparation, J.W.E. and H.S.K.; writing—review and editing, J.Y.C. and J.-Y.J.; visualization, J.W.E.; supervision, H.S.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by grants from the Korea Health Technology R&D Project through the Korea Health Industry Development Institute, funded by the Ministry of Health and Welfare, Republic of Korea (HR21C1003) and the Bio and Medical Technology Development Program of the National Research Foundation (RS-2023-00210847, NRF-2022R1A2C2092422, and NRF-2022R1A2C1092155) funded by the Korean government (Ministry of Science and ICT).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Acknowledgments

This review article discusses the utilization of biospecimens and data derived from the Biobank of Ajou University Hospital, a member institution of the Korea Biobank Network. The illustration in this review was drawn with Biorender (www.biorender.com, accessed on 18 August 2023). In addition, we thank all the members of the MOAGEN (Daejeon, Republic of Korea) for the bioinformatic guidance.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Eddy, S.R. Non–coding RNA genes and the modern RNA world. Nat. Rev. Genet. 2001, 2, 919–929. [Google Scholar] [CrossRef]
  2. Zhu, Z.; Ma, Y.; Li, Y.; Li, P.; Cheng, Z.; Li, H.; Zhang, L.; Tang, Z. The comprehensive detection of miRNA, lncRNA, and circRNA in regulation of mouse melanocyte and skin development. Biol. Res. 2020, 53, 4. [Google Scholar] [CrossRef] [PubMed]
  3. Bahn, J.H.; Zhang, Q.; Li, F.; Chan, T.M.; Lin, X.; Kim, Y.; Wong, D.T.; Xiao, X. The landscape of microRNA, Piwi-interacting RNA, and circular RNA in human saliva. Clin. Chem. 2015, 61, 221–230. [Google Scholar] [CrossRef] [PubMed]
  4. Penny, G.D.; Kay, G.F.; Sheardown, S.A.; Rastan, S.; Brockdorff, N. Requirement for Xist in X chromosome inactivation. Nature 1996, 379, 131–137. [Google Scholar] [CrossRef] [PubMed]
  5. Consortium, E.P. An integrated encyclopedia of DNA elements in the human genome. Nature 2012, 489, 57–74. [Google Scholar] [CrossRef]
  6. Huarte, M. The emerging role of lncRNAs in cancer. Nat. Med. 2015, 21, 1253–1261. [Google Scholar] [CrossRef]
  7. Luo, L.; Zhen, Y.; Peng, D.; Wei, C.; Zhang, X.; Liu, X.; Han, L.; Zhang, Z. The role of N6-methyladenosine-modified non-coding RNAs in the pathological process of human cancer. Cell Death Discov. 2022, 8, 325. [Google Scholar] [CrossRef]
  8. Yang, C.; Hu, Y.; Zhou, B.; Bao, Y.; Li, Z.; Gong, C.; Yang, H.; Wang, S.; Xiao, Y. The role of m6A modification in physiology and disease. Cell Death Dis. 2020, 11, 960. [Google Scholar] [CrossRef]
  9. Sivasudhan, E.; Blake, N.; Lu, Z.L.; Meng, J.; Rong, R. Dynamics of m6A RNA Methylome on the Hallmarks of Hepatocellular Carcinoma. Front. Cell Dev. Biol. 2021, 9, 642443. [Google Scholar] [CrossRef]
  10. Han, P.; Chang, C.P. Long non-coding RNA and chromatin remodeling. RNA Biol. 2015, 12, 1094–1098. [Google Scholar] [CrossRef]
  11. Yoon, J.H.; Abdelmohsen, K.; Gorospe, M. Posttranscriptional gene regulation by long noncoding RNA. J. Mol. Biol. 2013, 425, 3723–3730. [Google Scholar] [CrossRef] [PubMed]
  12. Zhang, C.; Wang, D.; Hao, Y.; Wu, S.; Luo, J.; Xue, Y.; Wang, D.; Li, G.; Liu, L.; Shao, C.; et al. LncRNA CCTT-mediated RNA-DNA and RNA-protein interactions facilitate the recruitment of CENP-C to centromeric DNA during kinetochore assembly. Mol. Cell 2022, 82, 4018–4032.e4019. [Google Scholar] [CrossRef] [PubMed]
  13. Rinn, J.L.; Kertesz, M.; Wang, J.K.; Sqazzo, S.L.; Xu, X.; Brugmann, S.A.; Goodnough, L.H.; Hemls, J.A.; Farnham, P.J.; Segal, E.; et al. Functional demarcation of active and silent chromatin domains in human HOX loci by noncoding RNAs. Cell 2007, 129, 1311–1323. [Google Scholar] [CrossRef] [PubMed]
  14. Perez, C.A.G.; Adachi, S.; Nong, Q.D.; Adhitama, N.; Matsuura, T.; Natsume, T.; Wada, T.; Kato, Y.; Watanabe, H. Sense-overlapping lncRNA as a decoy of translational repressor protein for dimorphic gene expression. PLoS Genet. 2021, 17, e1009683. [Google Scholar] [CrossRef]
  15. 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]
  16. Kino, T.; Hurt, D.E.; Ichijo, T.; Nader, N.; Chrousos, G.P. Noncoding RNA gas5 is a growth arrest- and starvation-associated repressor of the glucocorticoid receptor. Sci. Signal 2010, 3, ra8. [Google Scholar] [CrossRef]
  17. Arab, K.; Park, Y.J.; Lindroth, A.M.; Schafer, A.; Oakes, C.; Weichenhan, D.; Lukanova, A.; Lundin, E.; Risch, A.; Meister, M.; et al. Long noncoding RNA TARID directs demethylation and activation of the tumor suppressor TCF21 via GADD45A. Mol. Cell 2014, 55, 604–614. [Google Scholar] [CrossRef]
  18. Jiang, L.; Shao, C.; Wu, Q.J.; Chen, G.; Zhou, J.; Yang, B.; Li, H.; Gou, L.T.; Zhang, Y.; Wang, Y.; et al. NEAT1 scaffolds RNA-binding proteins and the Microprocessor to globally enhance pri-miRNA processing. Nat. Struct. Mol. Biol. 2017, 24, 816–824. [Google Scholar] [CrossRef]
  19. Kim, T.K.; Hemberg, M.; Gray, J.M. Enhancer RNAs: A class of long noncoding RNAs synthesized at enhancers. Cold Spring Harb. Perspect. Biol. 2015, 7, a018622. [Google Scholar] [CrossRef]
  20. Miao, Y.; Ajami, N.E.; Huang, T.S.; Lin, F.M.; Lou, C.H.; Wang, Y.T.; Li, S.; Kang, J.; Munkacsi, H.; Maurya, M.R.; et al. Enhancer-associated long non-coding RNA LEENE regulates endothelial nitric oxide synthase and endothelial function. Nat. Commun. 2018, 9, 292. [Google Scholar] [CrossRef]
  21. Lou, M.M.; Tang, X.Q.; Wang, G.M.; He, J.; Luo, F.; Guan, M.F.; Wang, F.; Zou, H.; Wang, J.Y.; Zhang, Q.; et al. Long noncoding RNA BS-DRL1 modulates the DNA damage response and genome stability by interacting with HMGB1 in neurons. Nat. Commun. 2021, 12, 4075. [Google Scholar] [CrossRef]
  22. Flores-Concha, M.; Onate, A.A. Long Non-coding RNAs in the Regulation of the Immune Response and Trained Immunity. Front. Genet. 2020, 11, 718. [Google Scholar] [CrossRef]
  23. Sellitto, A.; Pecoraro, G.; Giurato, G.; Nassa, G.; Rizzo, F.; Saggese, P.; Martinez, C.A.; Scafoglio, C.; Tarallo, R. Regulation of Metabolic Reprogramming by Long Non-Coding RNAs in Cancer. Cancers 2021, 13, 3485. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, F.; Wang, Q.; Liu, B.; Mei, L.; Ma, S.; Wang, S.; Wang, R.; Zhang, Y.; Niu, C.; Xiong, Z.; et al. The long noncoding RNA Synage regulates synapse stability and neuronal function in the cerebellum. Cell Death Differ. 2021, 28, 2634–2650. [Google Scholar] [CrossRef] [PubMed]
  25. Deng, S.; Wang, J.; Zhang, L.; Li, J.; Jin, Y. LncRNA HOTAIR Promotes Cancer Stem-Like Cells Properties by Sponging miR-34a to Activate the JAK2/STAT3 Pathway in Pancreatic Ductal Adenocarcinoma. Onco Targets Ther. 2021, 14, 1883–1893. [Google Scholar] [CrossRef]
  26. Gao, L.; Wang, X.; Guo, S.; Xiao, L.; Liang, C.; Wang, Z.; Li, Y.; Liu, Y.; Yao, R.; Liu, Y.; et al. LncRNA HOTAIR functions as a competing endogenous RNA to upregulate SIRT1 by sponging miR-34a in diabetic cardiomyopathy. J. Cell Physiol. 2019, 234, 4944–4958. [Google Scholar] [CrossRef] [PubMed]
  27. Shao, T.; Hu, Y.; Tang, W.; Shen, H.; Yu, Z.; Gu, J. The long noncoding RNA HOTAIR serves as a microRNA-34a-5p sponge to reduce nucleus pulposus cell apoptosis via a NOTCH1-mediated mechanism. Gene 2019, 715, 144029. [Google Scholar] [CrossRef] [PubMed]
  28. Tay, Y.; Kats, L.; Salmena, L.; Weiss, D.; Tan, S.M.; Ala, U.; Karreth, F.; Poliseno, L.; Provero, P.; Di Cunto, F.; et al. Coding-independent regulation of the tumor suppressor PTEN by competing endogenous mRNAs. Cell 2011, 147, 344–357. [Google Scholar] [CrossRef]
  29. Karreth, F.A.; Tay, Y.; Perna, D.; Ala, U.; Tan, S.M.; Rust, A.G.; DeNicola, G.; Webster, K.A.; Weiss, D.; Perez-Mancera, P.A.; et al. In vivo identification of tumor- suppressive PTEN ceRNAs in an oncogenic BRAF-induced mouse model of melanoma. Cell 2011, 147, 382–395. [Google Scholar] [CrossRef]
  30. Gu, X.; Li, M.; Jin, Y.; Liu, D.; Wei, F. Identification and integrated analysis of differentially expressed lncRNAs and circRNAs reveal the potential ceRNA networks during PDLSC osteogenic differentiation. BMC Genet. 2017, 18, 100. [Google Scholar] [CrossRef]
  31. Ghafouri-Fard, S.; Gholipour, M.; Hussen, B.M.; Taheri, M. The Impact of Long Non-Coding RNAs in the Pathogenesis of Hepatocellular Carcinoma. Front. Oncol. 2021, 11, 649107. [Google Scholar] [CrossRef]
  32. Abbastabar, M.; Sarfi, M.; Golestani, A.; Khalili, E. lncRNA involvement in hepatocellular carcinoma metastasis and prognosis. EXCLI J. 2018, 17, 900–913. [Google Scholar] [CrossRef] [PubMed]
  33. Ishibashi, M.; Kogo, R.; Shibata, K.; Sawada, G.; Takahashi, Y.; Kurashige, J.; Akiyoshi, S.; Sasaki, S.; Iwaya, T.; Sudo, T.; et al. Clinical significance of the expression of long non-coding RNA HOTAIR in primary hepatocellular carcinoma. Oncol. Rep. 2013, 29, 946–950. [Google Scholar] [CrossRef] [PubMed]
  34. Fu, W.M.; Zhu, X.; Wang, W.M.; Lu, Y.F.; Hu, B.G.; Wang, H.; Liang, W.C.; Wang, S.S.; Ko, C.H.; Waye, M.M.; et al. Hotair mediates hepatocarcinogenesis through suppressing miRNA-218 expression and activating P14 and P16 signaling. J. Hepatol. 2015, 63, 886–895. [Google Scholar] [CrossRef] [PubMed]
  35. Topel, H.; Bagirsakci, E.; Comez, D.; Bagci, G.; Cakan-Akdogan, G.; Atabey, N. lncRNA HOTAIR overexpression induced downregulation of c-Met signaling promotes hybrid epithelial/mesenchymal phenotype in hepatocellular carcinoma cells. Cell Commun. Signal 2020, 18, 110. [Google Scholar] [CrossRef]
  36. Yang, Z.; Zhou, L.; Wu, L.M.; Lai, M.C.; Xie, H.Y.; Zhang, F.; Zheng, S.S. Overexpression of long non-coding RNA HOTAIR predicts tumor recurrence in hepatocellular carcinoma patients following liver transplantation. Ann. Surg. Oncol. 2011, 18, 1243–1250. [Google Scholar] [CrossRef]
  37. Hajjari, M.; Salavaty, A. HOTAIR: An oncogenic long non-coding RNA in different cancers. Cancer Biol. Med. 2015, 12, 1. [Google Scholar] [CrossRef]
  38. Fang, S.; Shen, Y.; Chen, B.; Wu, Y.; Jia, L.; Li, Y.; Zhu, Y.; Yan, Y.; Li, M.; Chen, R.; et al. H3K27me3 induces multidrug resistance in small cell lung cancer by affecting HOXA1 DNA methylation via regulation of the lncRNA HOTAIR. Ann. Transl. Med. 2018, 6, 440. [Google Scholar] [CrossRef]
  39. Zhang, H.; Diab, A.; Fan, H.; Mani, S.K.; Hullinger, R.; Merle, P.; Andrisani, O. PLK1 and HOTAIR Accelerate Proteasomal Degradation of SUZ12 and ZNF198 during Hepatitis B Virus-Induced Liver Carcinogenesis. Cancer Res. 2015, 75, 2363–2374. [Google Scholar] [CrossRef]
  40. Zhang, H.; Xing, Z.; Mani, S.K.; Bancel, B.; Durantel, D.; Zoulim, F.; Tran, E.J.; Merle, P.; Andrisani, O. RNA helicase DEAD box protein 5 regulates Polycomb repressive complex 2/Hox transcript antisense intergenic RNA function in hepatitis B virus infection and hepatocarcinogenesis. Hepatology 2016, 64, 1033–1048. [Google Scholar] [CrossRef]
  41. Clemson, C.M.; Hutchinson, J.N.; Sara, S.A.; Ensminger, A.W.; Fox, A.H.; Chess, A.; Lawrence, J.B. An architectural role for a nuclear noncoding RNA: NEAT1 RNA is essential for the structure of paraspeckles. Mol. Cell 2009, 33, 717–726. [Google Scholar] [CrossRef]
  42. Liu, Z.; Chang, Q.; Yang, F.; Liu, B.; Yao, H.W.; Bai, Z.G.; Pu, C.S.; Ma, X.M.; Yang, Y.; Wang, T.T.; et al. Long non-coding RNA NEAT1 overexpression is associated with unfavorable prognosis in patients with hepatocellular carcinoma after hepatectomy: A Chinese population-based study. Eur. J. Surg. Oncol. 2017, 43, 1697–1703. [Google Scholar] [CrossRef] [PubMed]
  43. Ning, L.; Li, Z.; Wei, D.; Chen, H.; Yang, C. LncRNA, NEAT1 is a prognosis biomarker and regulates cancer progression via epithelial-mesenchymal transition in clear cell renal cell carcinoma. Cancer Biomark. 2017, 19, 75–83. [Google Scholar] [CrossRef] [PubMed]
  44. Chen, T.; Wang, H.; Yang, P.; He, Z.Y. Prognostic role of long noncoding RNA NEAT1 in various carcinomas: A meta-analysis. Onco Targets Ther. 2017, 10, 993–1000. [Google Scholar] [CrossRef] [PubMed]
  45. Xu, Y.; Cao, Z.; Ding, Y.; Li, Z.; Xiang, X.; Lai, R.; Sheng, Z.; Liu, Y.; Cai, W.; Hu, R.; et al. Long Non-coding RNA NEAT1 Alleviates Acute-on-Chronic Liver Failure Through Blocking TRAF6 Mediated Inflammatory Response. Front. Physiol. 2019, 10, 1503. [Google Scholar] [CrossRef] [PubMed]
  46. Bu, F.T.; Wang, A.; Zhu, Y.; You, H.M.; Zhang, Y.F.; Meng, X.M.; Huang, C.; Li, J. LncRNA NEAT1: Shedding light on mechanisms and opportunities in liver diseases. Liver Int. 2020, 40, 2612–2626. [Google Scholar] [CrossRef]
  47. Mang, Y.; Li, L.; Ran, J.; Zhang, S.; Liu, J.; Li, L.; Chen, Y.; Liu, J.; Gao, Y.; Ren, G. Long noncoding RNA NEAT1 promotes cell proliferation and invasion by regulating hnRNP A2 expression in hepatocellular carcinoma cells. Onco Targets Ther. 2017, 10, 1003–1016. [Google Scholar] [CrossRef]
  48. Zheng, X.; Zhang, Y.; Liu, Y.; Fang, L.; Li, L.; Sun, J.; Pan, Z.; Xin, W.; Huang, P. HIF-2alpha activated lncRNA NEAT1 promotes hepatocellular carcinoma cell invasion and metastasis by affecting the epithelial-mesenchymal transition. J. Cell Biochem. 2018, 119, 3247–3256. [Google Scholar] [CrossRef]
  49. Fang, L.; Sun, J.; Pan, Z.; Song, Y.; Zhong, L.; Zhang, Y.; Liu, Y.; Zheng, X.; Huang, P. Long non-coding RNA NEAT1 promotes hepatocellular carcinoma cell proliferation through the regulation of miR-129-5p-VCP-IkappaB. Am. J. Physiol. Gastrointest. Liver Physiol. 2017, 313, G150–G156. [Google Scholar] [CrossRef]
  50. Wang, Z.; Zou, Q.; Song, M.; Chen, J. NEAT1 promotes cell proliferation and invasion in hepatocellular carcinoma by negative regulating miR-613 expression. Biomed. Pharmacother. 2017, 94, 612–618. [Google Scholar] [CrossRef]
  51. Zhang, Y.; Luo, M.; Cui, X.; O’Connell, D.; Yang, Y. Long noncoding RNA NEAT1 promotes ferroptosis by modulating the miR-362-3p/MIOX axis as a ceRNA. Cell Death Differ. 2022, 29, 1850–1863. [Google Scholar] [CrossRef]
  52. Li, J.; Cao, F.; Yin, H.L.; Huang, Z.J.; Lin, Z.T.; Mao, N.; Sun, B.; Wang, G. Ferroptosis: Past, present and future. Cell Death Dis. 2020, 11, 88. [Google Scholar] [CrossRef]
  53. Ru, Y.; Chen, X.J.; Guo, W.Z.; Gao, S.G.; Qi, Y.J.; Chen, P.; Feng, X.S.; Zhang, S.J. NEAT1_2-SFPQ axis mediates cisplatin resistance in liver cancer cells in vitro. Onco Targets Ther. 2018, 11, 5695–5702. [Google Scholar] [CrossRef] [PubMed]
  54. Chen, S.; Xia, X. Long noncoding RNA NEAT1 suppresses sorafenib sensitivity of hepatocellular carcinoma cells via regulating miR-335-c-Met. J. Cell Physiol. 2019, 234, 14999–15009. [Google Scholar] [CrossRef] [PubMed]
  55. Wang, J.; Liu, X.; Wu, H.; Ni, P.; Gu, Z.; Qiao, Y.; Chen, N.; Sun, F.; Fan, Q. CREB up-regulates long non-coding RNA, HULC expression through interaction with microRNA-372 in liver cancer. Nucleic Acids Res. 2010, 38, 5366–5383. [Google Scholar] [CrossRef]
  56. Li, D.; Liu, X.; Zhou, J.; Hu, J.; Zhang, D.; Liu, J.; Qiao, Y.; Zhan, Q. Long noncoding RNA HULC modulates the phosphorylation of YB-1 through serving as a scaffold of extracellular signal-regulated kinase and YB-1 to enhance hepatocarcinogenesis. Hepatology 2017, 65, 1612–1627. [Google Scholar] [CrossRef] [PubMed]
  57. Ma, Y.; Huang, D.; Yang, F.; Tian, M.; Wang, Y.; Shen, D.; Wang, Q.; Chen, Q.; Zhang, L. Long Noncoding RNA Highly Upregulated in Liver Cancer Regulates the Tumor Necrosis Factor-alpha-Induced Apoptosis in Human Vascular Endothelial Cells. DNA Cell Biol. 2016, 35, 296–300. [Google Scholar] [CrossRef]
  58. Du, Y.; Kong, G.; You, X.; Zhang, S.; Zhang, T.; Gao, Y.; Ye, L.; Zhang, X. Elevation of highly up-regulated in liver cancer (HULC) by hepatitis B virus X protein promotes hepatoma cell proliferation via down-regulating p18. J. Biol. Chem. 2012, 287, 26302–26311. [Google Scholar] [CrossRef] [PubMed]
  59. Xie, H.; Ma, H.; Zhou, D. Plasma HULC as a promising novel biomarker for the detection of hepatocellular carcinoma. Biomed. Res. Int. 2013, 2013, 136106. [Google Scholar] [CrossRef]
  60. Lu, J.; Guo, J.; Liu, J.; Mao, X.; Xu, K. Long Non-coding RNA MALAT1: A Key Player in Liver Diseases. Front. Med. 2021, 8, 734643. [Google Scholar] [CrossRef]
  61. 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. Tumour Biol. 2017, 39, 1010428317718135. [Google Scholar] [CrossRef] [PubMed]
  62. Peng, J.; Wu, H.J.; Zhang, H.F.; Fang, S.Q.; Zeng, R. miR-143-3p inhibits proliferation and invasion of hepatocellular carcinoma cells by regulating its target gene FGF1. Clin. Transl. Oncol. 2021, 23, 468–480. [Google Scholar] [CrossRef] [PubMed]
  63. Yao, W.; Liu, J.; Huang, D. MiR-200a inhibits cell proliferation and EMT by down-regulating the ASPH expression levels and affecting ERK and PI3K/Akt pathways in human hepatoma cells. Am. J. Transl. Res. 2018, 10, 1117–1130. [Google Scholar] [PubMed]
  64. Cui, R.J.; Fan, J.L.; Lin, Y.C.; Pan, Y.J.; Liu, C.; Wan, J.H.; Wang, W.; Jiang, Z.Y.; Zheng, X.L.; Tang, J.B.; et al. miR-124-3p availability is antagonized by LncRNA-MALAT1 for Slug-induced tumor metastasis in hepatocellular carcinoma. Cancer Med. 2019, 8, 6358–6369. [Google Scholar] [CrossRef]
  65. Liu, D.; Zhu, Y.; Pang, J.; Weng, X.; Feng, X.; Guo, Y. Knockdown of long non-coding RNA MALAT1 inhibits growth and motility of human hepatoma cells via modulation of miR-195. J. Cell Biochem. 2018, 119, 1368–1380. [Google Scholar] [CrossRef]
  66. Chen, S.; Wang, G.; Tao, K.; Cai, K.; Wu, K.; Ye, L.; Bai, J.; Yin, Y.; Wang, J.; Shuai, X.; et al. Long noncoding RNA metastasis-associated lung adenocarcinoma transcript 1 cooperates with enhancer of zeste homolog 2 to promote hepatocellular carcinoma development by modulating the microRNA-22/Snail family transcriptional repressor 1 axis. Cancer Sci. 2020, 111, 1582–1595. [Google Scholar] [CrossRef]
  67. Li, S.; Xu, H.; Yu, Y.; He, J.; Wang, Z.; Xu, Y.; Wang, C.; Zhang, H.; Zhang, R.; Zhang, 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, 42431–42446. [Google Scholar] [CrossRef]
  68. Xin, X.; Wu, M.; Meng, Q.; Wang, C.; Lu, Y.; Yang, Y.; Li, X.; Zheng, Q.; Pu, H.; Gui, X.; et al. Long noncoding RNA HULC accelerates liver cancer by inhibiting PTEN via autophagy cooperation to miR15a. Mol. Cancer 2018, 17, 94. [Google Scholar] [CrossRef]
  69. Guo, Y.; Bai, M.; Lin, L.; Huang, J.; An, Y.; Liang, L.; Liu, Y.; Huang, W. LncRNA DLEU2 aggravates the progression of hepatocellular carcinoma through binding to EZH2. Biomed. Pharmacother. 2019, 118, 109272. [Google Scholar] [CrossRef]
  70. Li, B.; Li, A.; You, Z.; Xu, J.; Zhu, S. Epigenetic silencing of CDKN1A and CDKN2B by SNHG1 promotes the cell cycle, migration and epithelial-mesenchymal transition progression of hepatocellular carcinoma. Cell Death Dis. 2020, 11, 823. [Google Scholar] [CrossRef]
  71. Pan, J.; Hu, Y.; Yuan, C.; Wu, Y.; Zhu, X. lncRNA NEAT1 promotes the proliferation and metastasis of hepatocellular carcinoma by regulating the FOXP3/PKM2 axis. Front. Oncol. 2022, 12, 928022. [Google Scholar] [CrossRef] [PubMed]
  72. Dai, Q.; Deng, J.; Zhou, J.; Wang, Z.; Yuan, X.-F.; Pan, S.; Zhang, H.-b. Long non-coding RNA TUG1 promotes cell progression in hepatocellular carcinoma via regulating miR-216b-5p/DLX2 axis. Cancer Cell Int. 2020, 20, 8. [Google Scholar] [CrossRef] [PubMed]
  73. Wang, H.; Ke, J.; Guo, Q.; Barnabo Nampoukime, K.P.; Yang, P.; Ma, K. Long non-coding RNA CRNDE promotes the proliferation, migration and invasion of hepatocellular carcinoma cells through miR-217/MAPK1 axis. J. Cell Mol. Med. 2018, 22, 5862–5876. [Google Scholar] [CrossRef] [PubMed]
  74. Chen, T.; Liu, R.; Niu, Y.; Mo, H.; Wang, H.; Lu, Y.; Wang, L.; Sun, L.; Wang, Y.; Tu, K.; et al. HIF-1alpha-activated long non-coding RNA KDM4A-AS1 promotes hepatocellular carcinoma progression via the miR-411-5p/KPNA2/AKT pathway. Cell Death Dis. 2021, 12, 1152. [Google Scholar] [CrossRef]
  75. Wang, H.Z.; Liu, L.; Xu, Y.; Zhang, G.Y.; Wang, Y.Y. LncRNA UCA1 Affects the Cell Proliferation, Migration, Invasion and Apoptosis of Hepatic Carcinoma Cells by Targeting MicroRNA-193a-3p. Cancer Manag. Res. 2020, 12, 10897–10907. [Google Scholar] [CrossRef]
  76. Ma, J.; Li, T.; Han, X.; Yuan, H. Knockdown of LncRNA ANRIL suppresses cell proliferation, metastasis, and invasion via regulating miR-122-5p expression in hepatocellular carcinoma. J. Cancer Res. Clin. Oncol. 2018, 144, 205–214. [Google Scholar] [CrossRef]
  77. Huang, D.; Bi, C.; Zhao, Q.; Ding, X.; Bian, C.; Wang, H.; Wang, T.; Liu, H. Knockdown long non-coding RNA ANRIL inhibits proliferation, migration and invasion of HepG2 cells by down-regulation of miR-191. BMC Cancer 2018, 18, 919. [Google Scholar] [CrossRef]
  78. Li, Y.; Chen, G.; Yan, Y.; Fan, Q. CASC15 promotes epithelial to mesenchymal transition and facilitates malignancy of hepatocellular carcinoma cells by increasing TWIST1 gene expression via miR-33a-5p sponging. Eur. J. Pharmacol. 2019, 860, 172589. [Google Scholar] [CrossRef]
  79. Duan, R.; Li, C.; Wang, F.; Han, F.; Zhu, L. The Long Noncoding RNA ZFAS1 Potentiates the Development of Hepatocellular Carcinoma via the microRNA-624/MDK/ERK/JNK/P38 Signaling Pathway. Onco Targets Ther. 2020, 13, 4431–4444. [Google Scholar] [CrossRef]
  80. Dou, C.; Sun, L.; Jin, X.; Han, M.; Zhang, B.; Jiang, X.; LV, J.; Li, T. Long non-coding RNA CARLo-5 promotes tumor progression in hepatocellular carcinoma via suppressing miR-200b expression. Oncotarget 2017, 8, 70172–70182. [Google Scholar] [CrossRef]
  81. Xu, J.H.; Chang, W.H.; Fu, H.W.; Shu, W.Q.; Yuan, T.; Chen, P. Upregulated long non-coding RNA LOC90784 promotes cell proliferation and invasion and is associated with poor clinical features in HCC. Biochem. Biophys. Res. Commun. 2017, 490, 920–926. [Google Scholar] [CrossRef] [PubMed]
  82. Li, L.; Han, T.; Liu, K.; Lei, C.; Wnag, Z.; Shi, G. LncRNA H19 promotes the development of hepatitis B related hepatocellular carcinoma through regulating microRNA-22 via EMT pathway. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 5392–5401. [Google Scholar] [PubMed]
  83. LV, J.; Yu, Y.; Li, S.; Luo, L.; Wang, Q. Aflatoxin B1 Promotes Cell Growth and Invasion in Hepatocellular Carcinoma HepG2 Cells through H19 and E2F1. Asian Pac. J. Cancer Prev. 2014, 15, 2565–2570. [Google Scholar] [CrossRef] [PubMed]
  84. Wang, Y.; Zeng, J.; Chen, W.; Fan, J.; Hylemon, P.B.; Zhou, H. Long Noncoding RNA H19: A Novel Oncogene in Liver Cancer. Noncoding RNA 2023, 9, 19. [Google Scholar] [CrossRef]
  85. Wang, Y.; Hu, Y.; Wu, G.; Yang, Y.; Tang, Y.; Zhang, W.; Wang, K.; Liu, Y.; Wang, X.; Li, T. Long noncoding RNA PCAT-14 induces proliferation and invasion by hepatocellular carcinoma cells by inducing methylation of miR-372. Oncotarget 2017, 8, 34429–34441. [Google Scholar] [CrossRef]
  86. Zhang, H.; Liu, Y.; Wang, W.; Liu, F.; Wang, W.; Su, C.; Zhu, H.; Liao, Z.; Zhang, B.; Chen, X. ALKBH5-mediated m6A modification of lincRNA LINC02551 enhances the stability of DDX24 to promote hepatocellular carcinoma growth and metastasis. Cell Death Dis. 2022, 13, 926. [Google Scholar] [CrossRef]
  87. Tao, H.; Zhang, Y.; Yan, T.; Li, J.; Liu, J.; Xiong, Y.; Zhu, J.; Huang, Z.; Wang, P.; Liang, H.; et al. Identification of an EMT-related lncRNA signature and LINC01116 as an immune-related oncogene in hepatocellular carcinoma. Aging 2022, 14, 1473–1491. [Google Scholar] [CrossRef]
  88. Liu, Y.; Wang, D.; Li, Y.; Yan, S.; Dang, H.; Yue, H.; Ling, J.; Chen, F.; Zhao, Y.; Gou, L.; et al. Long noncoding RNA CCAT2 promotes hepatocellular carcinoma proliferation and metastasis through up-regulation of NDRG1. Exp. Cell Res. 2019, 379, 19–29. [Google Scholar] [CrossRef]
  89. Su, R.; Zhang, H.; Zhang, L.; Khan, A.R.; Zhang, X.; Wang, R.; Shao, C.; Wei, X.; Xu, X. Systemic analysis identifying PVT1/DUSP13 axis for microvascular invasion in hepatocellular carcinoma. Cancer Med. 2023, 12, 8937–8955. [Google Scholar] [CrossRef]
  90. Luo, Z.; Cao, P. Long noncoding RNA PVT1 promotes hepatoblastoma cell proliferation through activating STAT3. Cancer Manag. Res. 2019, 11, 8517–8527. [Google Scholar] [CrossRef]
  91. Wang, X.L.; Shi, M.; Xiang, T.; Bu, Y.Z. Long noncoding RNA DGCR5 represses hepatocellular carcinoma progression by inactivating Wnt signaling pathway. J. Cell Biochem. 2019, 120, 275–282. [Google Scholar] [CrossRef] [PubMed]
  92. Sun, Y.; Cao, F.; Qu, L.; Wang, Z.; Liu, X. MEG3 promotes liver cancer by activating PI3K/AKT pathway through regulating AP1G1. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 1459–1467. [Google Scholar]
  93. Wei, Q.; Liu, G.; Huang, Z.; Huang, Y.; Huang, L.; Huang, Z.; Wu, X.; Wei, H.; Pu, J. LncRNA MEG3 Inhibits Tumor Progression by Modulating Macrophage Phenotypic Polarization via miR-145-5p/DAB2 Axis in Hepatocellular Carcinoma. J. Hepatocell. Carcinoma 2023, 10, 1019–1035. [Google Scholar] [CrossRef] [PubMed]
  94. Wu, J.; Pang, R.; Li, M.; Chen, B.; Huang, J.; Zhu, Y. m6A-Induced LncRNA MEG3 Suppresses the Proliferation, Migration and Invasion of Hepatocellular Carcinoma Cell Through miR-544b/BTG2 Signaling. Onco Targets Ther. 2021, 14, 3745–3755. [Google Scholar] [CrossRef]
  95. Qian, G.; Jin, X.; Zhang, L. LncRNA FENDRR Upregulation Promotes Hepatic Carcinoma Cells Apoptosis by Targeting miR-362-5p Via NPR3 and p38-MAPK Pathway. Cancer Biother. Radiopharm. 2020, 35, 629–639. [Google Scholar] [CrossRef] [PubMed]
  96. 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] [PubMed]
  97. Wang, C.; Ke, S.; Li, M.; Lin, C.; Liu, X.; Pan, Q. Downregulation of LncRNA GAS5 promotes liver cancer proliferation and drug resistance by decreasing PTEN expression. Mol. Genet. Genom. 2020, 295, 251–260. [Google Scholar] [CrossRef]
  98. Wang, X.; Li, F.Y.; Zhao, W.; Gao, Z.K.; Shen, B.; Xu, H.; Cui, Y.F. Long non-coding RNA GAS5 overexpression inhibits M2-like polarization of tumour-associated macrophages in SMCC-7721 cells by promoting PTEN expression. Int. J. Exp. Pathol. 2020, 101, 215–222. [Google Scholar] [CrossRef]
  99. Dai, Y.Z.; Liu, Y.D.; Li, J.; Chen, M.T.; Huang, M.; Wang, F.; Yang, Q.S.; Yuan, J.H.; Sun, S.H. METTL16 promotes hepatocellular carcinoma progression through downregulating RAB11B-AS1 in an m6A-dependent manner. Cell Mol. Biol. Lett. 2022, 27, 41. [Google Scholar] [CrossRef]
  100. Park, E.G.; Pyo, S.J.; Cui, Y.; Yoon, S.H.; Nam, J.W. Tumor immune microenvironment lncRNAs. Brief. Bioinform. 2022, 23, bbab504. [Google Scholar] [CrossRef]
  101. Jiang, P.; Li, X. Regulatory Mechanism of lncRNAs in M1/M2 Macrophages Polarization in the Diseases of Different Etiology. Front. Immunol. 2022, 13, 835932. [Google Scholar] [CrossRef]
  102. Hou, Z.H.; Xu, X.W.; Fu, X.Y.; Zhou, L.D.; Liu, S.P.; Tan, D.M. Long non-coding RNA MALAT1 promotes angiogenesis and immunosuppressive properties of HCC cells by sponging miR-140. Am. J. Physiol. Cell Physiol. 2020, 318, C649–C663. [Google Scholar] [CrossRef] [PubMed]
  103. Jiang, R.; Tang, J.; Chen, Y.; Deng, L.; Ji, J.; Xie, Y.; Wang, K.; Jia, W.; Chu, W.M.; Sun, B. The long noncoding RNA lnc-EGFR stimulates T-regulatory cells differentiation thus promoting hepatocellular carcinoma immune evasion. Nat. Commun. 2017, 8, 15129. [Google Scholar] [CrossRef] [PubMed]
  104. Wang, Y.; Yang, L.; Dong, X.; Yang, X.; Zhang, X.; Liu, Z.; Zhao, X.; Wen, T. Overexpression of NNT-AS1 Activates TGF-beta Signaling to Decrease Tumor CD4 Lymphocyte Infiltration in Hepatocellular Carcinoma. Biomed. Res. Int. 2020, 2020, 8216541. [Google Scholar] [CrossRef]
  105. Peng, L.; Chen, Y.; Ou, Q.; Wang, X.; Tang, N. LncRNA MIAT correlates with immune infiltrates and drug reactions in hepatocellular carcinoma. Int. Immunopharmacol. 2020, 89, 107071. [Google Scholar] [CrossRef] [PubMed]
  106. Zheng, Y.; Nie, P.; Peng, D.; He, Z.; Liu, M.; Xie, Y.; Miao, Y.; Zuo, Z.; Ren, J. m6AVar: A database of functional variants involved in m6A modification. Nucleic Acids Res. 2018, 46, D139–D145. [Google Scholar] [CrossRef] [PubMed]
  107. Luo, X.; Li, H.; Liang, J.; Zhao, Q.; Xie, Y.; Ren, J.; Zuo, Z. RMVar: An updated database of functional variants involved in RNA modifications. Nucleic Acids Res. 2021, 49, D1405–D1412. [Google Scholar] [CrossRef] [PubMed]
  108. Alarcon, C.R.; Goodarzi, H.; Lee, H.; Liu, X.; Tavazoie, S.; Tavazoie, S.F. HNRNPA2B1 Is a Mediator of m6A-Dependent Nuclear RNA Processing Events. Cell 2015, 162, 1299–1308. [Google Scholar] [CrossRef]
  109. Liu, J.; Yue, Y.; Han, D.; Wang, X.; Fu, Y.; Zhang, L.; Jia, G.; Yu, M.; Lu, Z.; Deng, X.; et al. A METTL3-METTL14 complex mediates mammalian nuclear RNA N6-adenosine methylation. Nat. Chem. Biol. 2014, 10, 93–95. [Google Scholar] [CrossRef]
  110. Ping, X.L.; Sun, B.F.; Wang, L.; Xiao, W.; Yang, X.; Wang, W.J.; Adhikari, S.; Shi, Y.; Lv, Y.; Chen, Y.S.; et al. Mammalian WTAP is a regulatory subunit of the RNA N6-methyladenosine methyltransferase. Cell Res. 2014, 24, 177–189. [Google Scholar] [CrossRef]
  111. Schwartz, S.; Mumbach, M.R.; Jovanovic, M.; Wang, T.; Maciag, K.; Bushkin, G.G.; Mertins, P.; Ter-Ovanesyan, D.; Habib, N.; Cacchiarelli, D.; et al. Perturbation of m6A writers reveals two distinct classes of mRNA methylation at internal and 5′ sites. Cell Rep. 2014, 8, 284–296. [Google Scholar] [CrossRef] [PubMed]
  112. Patil, D.P.; Chen, C.K.; Pickering, B.F.; Chow, A.; Jackson, C.; Guttman, M.; Jaffrey, S.R. m6A RNA methylation promotes XIST-mediated transcriptional repression. Nature 2016, 537, 369–373. [Google Scholar] [CrossRef] [PubMed]
  113. Jia, G.; Fu, Y.; Zhao, X.; Dai, Q.; Zheng, G.; Yang, Y.; Yi, C.; Lindahl, T.; Pan, T.; Yang, Y.G.; et al. N6-methyladenosine in nuclear RNA is a major substrate of the obesity-associated FTO. Nat. Chem. Biol. 2011, 7, 885–887. [Google Scholar] [CrossRef]
  114. Zheng, G.; Dahl, J.A.; Niu, Y.; Fedorcsak, P.; Huang, C.M.; Li, C.J.; Vagbo, C.B.; Shi, Y.; Wang, W.L.; Song, S.H.; et al. ALKBH5 is a mammalian RNA demethylase that impacts RNA metabolism and mouse fertility. Mol. Cell 2013, 49, 18–29. [Google Scholar] [CrossRef]
  115. Xu, C.; Liu, K.; Ahmed, H.; Loppnau, P.; Schapira, M.; Min, J. Structural Basis for the Discriminative Recognition of N6-Methyladenosine RNA by the Human YT521-B Homology Domain Family of Proteins. J. Biol. Chem. 2015, 290, 24902–24913. [Google Scholar] [CrossRef]
  116. Huang, H.; Weng, H.; Sun, W.; Qin, X.; Shi, H.; Wu, H.; Zhao, B.S.; Mesquita, A.; Liu, C.; Yuan, C.L.; et al. Recognition of RNA N6-methyladenosine by IGF2BP proteins enhances mRNA stability and translation. Nat. Cell Biol. 2018, 20, 285–295. [Google Scholar] [CrossRef] [PubMed]
  117. Du, H.; Zhao, Y.; He, J.; Zhang, Y.; Xi, H.; Liu, M.; Ma, J.; Wu, L. YTHDF2 destabilizes m6A-containing RNA through direct recruitment of the CCR4-NOT deadenylase complex. Nat. Commun. 2016, 7, 12626. [Google Scholar] [CrossRef]
  118. Zuo, X.; Chen, Z.; Gao, W.; Zhang, Y.; Wang, J.; Wang, J.; Cao, M.; Cai, J.; Wu, J.; Wang, X. M6A-mediated upregulation of LINC00958 increases lipogenesis and acts as a nanotherapeutic target in hepatocellular carcinoma. J. Hematol. Oncol. 2020, 13, 5. [Google Scholar] [CrossRef]
  119. Li, L.; Xie, R.; Lu, G. Identification of m6A methyltransferase-related lncRNA signature for predicting immunotherapy and prognosis in patients with hepatocellular carcinoma. Biosci. Rep. 2021, 41, BSR20210760. [Google Scholar] [CrossRef]
  120. Peng, L.; Pan, B.; Zhang, X.; Wang, Z.; Qiu, J.; Wang, X.; Tang, N. Lipopolysaccharide facilitates immune escape of hepatocellular carcinoma cells via m6A modification of lncRNA MIR155HG to upregulate PD-L1 expression. Cell Biol. Toxicol. 2022, 38, 1159–1173. [Google Scholar] [CrossRef]
  121. Liu, J.; Zhang, N.; Zeng, J.; Wang, T.; Shen, Y.; Ma, C.; Yang, M. N6-methyladenosine-modified lncRNA ARHGAP5-AS1 stabilises CSDE1 and coordinates oncogenic RNA regulons in hepatocellular carcinoma. Clin. Transl. Med. 2022, 12, e1107. [Google Scholar] [CrossRef] [PubMed]
  122. Ye, Y.; Wang, M.; Wang, G.; Mai, Z.; Zhou, B.; Han, Y.; Zhuang, J.; Xia, W. lncRNA miR4458HG modulates hepatocellular carcinoma progression by activating m6A-dependent glycolysis and promoting the polarization of tumor-associated macrophages. Cell Mol. Life Sci. 2023, 80, 99. [Google Scholar] [CrossRef] [PubMed]
  123. Hammerle, M.; Gutschner, T.; Uckelmann, H.; Ozgur, S.; Fiskin, E.; Gross, M.; Skawran, B.; Geffers, R.; Longerich, T.; Breuhahn, K.; et al. Posttranscriptional destabilization of the liver-specific long noncoding RNA HULC by the IGF2 mRNA-binding protein 1 (IGF2BP1). Hepatology 2013, 58, 1703–1712. [Google Scholar] [CrossRef] [PubMed]
  124. Zeng, F.; Lin, J.; Xie, X.; Xie, Y.; Zhang, J.; Xu, D.; He, X.; Liu, F.; Xie, B. LncRNA SLC7A11-AS1 promotes the progression of hepatocellular carcinoma by mediating KLF9 ubiquitination. Neoplasma 2023, 70, 361–374. [Google Scholar] [CrossRef]
  125. Zhang, Q.; Wei, T.; Yan, L.; Zhu, S.; Jin, W.; Bai, Y.; Zeng, Y.; Zhang, X.; Yin, Z.; Yang, J.; et al. Hypoxia-Responsive lncRNA AC115619 Encodes a Micropeptide That Suppresses m6A Modifications and Hepatocellular Carcinoma Progression. Cancer Res. 2023, 83, 2496–2512. [Google Scholar] [CrossRef]
  126. Beylerli, O.; Gareev, I.; Sufianov, A.; Ilyasova, T.; Guang, Y. Long noncoding RNAs as promising biomarkers in cancer. Noncoding RNA Res. 2022, 7, 66–70. [Google Scholar] [CrossRef] [PubMed]
  127. Taketa, K.; Okada, S.; Win, N.; Hlaing, N.; Win, K.-M. Evaluation of Tumor Markers for the Detection of Hepatocellular Carcinoma in Yangon General Hospital, Myanmar. Acta Med. Okayama. 2002, 56, 317–320. [Google Scholar]
  128. Khien, V.-V.; Mao, H.; Chinh, T.; Bang, M.; Lac, B.-V.; Hop, T.-V.; Tuan, N.-A.; Don, L.; Taketa, K.; Satomura, S. Clinical evaluation of lentil lectin-reactive alpha-fetoprotein-L3 in histology-proven hepatocellular carcinoma. Int. J. Biol. Markers 2001, 16, 105–111. [Google Scholar] [CrossRef]
  129. Hippo, Y.; Watanabe, K.; Watanabe, A.; Midorikawa, Y.; Yamamoto, S.; Ihara, S.; Tokita, S.; Iwanari, H.; Ito, Y.; Nakano, K.; et al. Identification of Soluble NH2-Terminal Fragment of Glypican-3 as a Serological Marker for Early-Stage Hepatocellular Carcinoma. Cancer Res. 2004, 64, 2418–2423. [Google Scholar] [CrossRef]
  130. Miura, N.; Maeda, Y.; Kanbe, Y.; Yazama, H.; Takeda, Y.; Sato, R.; Tsukamoto, T.; Sato, E.; Marunoto, A.; Harada, T.; et al. Serum HumanTelomerase ReverseTranscriptase Messenger RNA as a NovelTumor Marker for Hepatocellular Carcinoma. Clin. Cancer Res. 2005, 11, 3205–3209. [Google Scholar] [CrossRef]
  131. Kim, H.S.; Yoon, J.H.; Baek, G.O.; Yoon, M.G.; Han, J.E.; Cho, H.J.; Kim, S.S.; Jeong, J.Y.; Cheong, J.Y.; Eun, J.W. Tumor Endothelial Cells-Associated Integrin Alpha-6 as a Promising Biomarker for Early Detection and Prognosis of Hepatocellular Carcinoma. Cancers 2023, 15, 4156. [Google Scholar] [CrossRef] [PubMed]
  132. Shen, Q.; Eun, J.W.; Lee, K.; Kim, H.S.; Yang, H.D.; Kim, S.Y.; Lee, E.K.; Kim, T.; Kang, K.; Kim, S.; et al. Barrier to autointegration factor 1, procollagen-lysine, 2-oxoglutarate 5-dioxygenase 3, and splicing factor 3b subunit 4 as early-stage cancer decision markers and drivers of hepatocellular carcinoma. Hepatology 2018, 67, 1360–1377. [Google Scholar] [CrossRef] [PubMed]
  133. Huang, J.; Zheng, Y.; Xiao, X.; Liu, C.; Lin, J.; Zheng, S.; Yang, B.; Ou, Q. A Circulating Long Noncoding RNA Panel Serves as a Diagnostic Marker for Hepatocellular Carcinoma. Dis. Markers 2020, 2020, 5417598. [Google Scholar] [CrossRef] [PubMed]
  134. Kim, S.S.; Baek, G.O.; Ahn, H.R.; Sung, S.; Seo, C.W.; Cho, H.J.; Nam, S.W.; Cheong, J.Y.; Eun, J.W. Serum small extracellular vesicle-derived LINC00853 as a novel diagnostic marker for early hepatocellular carcinoma. Mol. Oncol. 2020, 14, 2646–2659. [Google Scholar] [CrossRef] [PubMed]
  135. Kim, S.S.; Baek, G.O.; Son, J.A.; Ahn, H.R.; Yoon, M.K.; Cho, H.J.; Yoon, J.H.; Nam, S.W.; Cheong, J.Y.; Eun, J.W. Early detection of hepatocellular carcinoma via liquid biopsy: Panel of small extracellular vesicle-derived long noncoding RNAs identified as markers. Mol. Oncol. 2021, 15, 2715–2731. [Google Scholar] [CrossRef]
  136. Yuan, S.X.; Yang, F.; Yang, Y.; Tao, Q.F.; Zhang, J.; Huang, G.; Yang, Y.; Wang, R.Y.; Yang, S.; Huo, X.S.; 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]
  137. Zeng, Z.; Dong, J.; Li, Y.; Dong, Z.; Liu, Z.; Huang, J.; Wang, Y.; Zhen, Y.; Lu, Y. The expression level and clinical significance of lncRNA X91348 in hepatocellular carcinoma. Artif. Cells Nanomed. Biotechnol. 2019, 47, 3067–3071. [Google Scholar] [CrossRef]
  138. Wu, L.; Yang, Z.; Zhang, J.; Xie, H.; Zhou, L.; Zheng, S. Long noncoding RNA HOTTIP expression predicts tumor recurrence in hepatocellular carcinoma patients following liver transplantation. Hepatobiliary Surg. Nutr. 2018, 7, 429–439. [Google Scholar] [CrossRef]
  139. Lee, Y.R.; Kim, G.; Tak, W.Y.; Jang, S.Y.; Kweon, Y.O.; Park, J.G.; Lee, H.W.; Han, Y.S.; Chun, J.M.; Park, S.Y.; et al. Circulating exosomal noncoding RNAs as prognostic biomarkers in human hepatocellular carcinoma. Int. J. Cancer 2019, 144, 1444–1452. [Google Scholar] [CrossRef]
  140. Lin, L.Y.; Yang, L.; Zeng, Q.; Wang, L.; Chen, M.L.; Zhao, Z.H.; Ye, G.D.; Luo, Q.C.; Lv, P.Y.; Guo, Q.W.; et al. Tumor-originated exosomal lncUEGC1 as a circulating biomarker for early-stage gastric cancer. Mol. Cancer 2018, 17, 84. [Google Scholar] [CrossRef]
  141. Vosough, P.; Khatami, S.H.; Hashemloo, A.; Tajbakhsh, A.; Karimi-Fard, F.; Taghvimi, S.; Taheri-Anganeh, M.; Soltani Fard, E.; Savardashtaki, A.; Movahedpour, A. Exosomal lncRNAs in gastrointestinal cancer. Clin. Chim. Acta 2023, 540, 117216. [Google Scholar] [CrossRef] [PubMed]
  142. Dragomir, M.; Chen, B.; Calin, G.A. Exosomal lncRNAs as new players in cell-to-cell communication. Transl. Cancer Res. 2018, 7, S243–S252. [Google Scholar] [CrossRef] [PubMed]
  143. Li, Q.; Shao, Y.; Zhang, X.; Zheng, T.; Miao, M.; Qin, L.; Wang, B.; Ye, G.; Xiao, B.; Guo, J. Plasma long noncoding RNA protected by exosomes as a potential stable biomarker for gastric cancer. Tumour Biol. 2015, 36, 2007–2012. [Google Scholar] [CrossRef]
  144. Lee, S.A.; Yoo, T.H. Therapeutic application of extracellular vesicles for various kidney diseases: A brief review. BMB Rep. 2022, 55, 3–10. [Google Scholar] [CrossRef] [PubMed]
  145. 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]
  146. 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. J. Int. Med. Res. 2018, 46, 348–356. [Google Scholar] [CrossRef]
  147. Abdelmoety, A.A.; Elhassafy, M.Y.; Omar Said, R.S.; Elsheaita, A.; Mahmoud, M.M. The role of UCA1 and WRAP53 in diagnosis of hepatocellular carcinoma: A single-center case-control study. Clin. Exp. Hepatol. 2023, 9, 129–137. [Google Scholar] [CrossRef]
  148. 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]
  149. Yin, Q.; Huang, X.; Yang, Q.; Lin, S.; Song, Q.; Fan, W.; Li, W.; Li, Z.; Gao, L. LncRNA model predicts liver cancer drug resistance and validate in vitro experiments. Front. Cell Dev. Biol. 2023, 11, 1174183. [Google Scholar] [CrossRef]
  150. Guo, D.; Li, Y.; Chen, Y.; Zhang, D.; Wang, X.; Lu, G.; Ren, M.; Lu, X.; He, S. DANCR promotes HCC progression and regulates EMT by sponging miR-27a-3p via ROCK1/LIMK1/COFILIN1 pathway. Cell Prolif. 2019, 52, e12628. [Google Scholar] [CrossRef]
  151. Xu, X.; Gu, J.; Ding, X.; Ge, G.; Zang, X.; Ji, R.; Shao, M.; Mao, Z.; Zhang, Y.; Zhang, J.; et al. LINC00978 promotes the progression of hepatocellular carcinoma by regulating EZH2-mediated silencing of p21 and E-cadherin expression. Cell Death Dis. 2019, 10, 752. [Google Scholar] [CrossRef] [PubMed]
  152. Zeng, B.; Lin, Z.; Ye, H.; Cheng, D.; Zhang, G.; Zhou, J.; Huang, Z.; Wang, M.; Cai, C.; Zeng, J.; et al. Upregulation of LncDQ is Associated with Poor Prognosis and Promotes Tumor Progression via Epigenetic Regulation of the EMT Pathway in HCC. Cell Physiol. Biochem. 2018, 46, 1122–1133. [Google Scholar] [CrossRef] [PubMed]
  153. Ma, W.; Chen, X.; Wu, X.; Li, J.; Mei, C.; Jing, W.; Teng, L.; Tu, H.; Jiang, X.; Wang, G.; et al. Long noncoding RNA SPRY4-IT1 promotes proliferation and metastasis of hepatocellular carcinoma via mediating TNF signaling pathway. J. Cell Physiol. 2020, 235, 7849–7862. [Google Scholar] [CrossRef]
  154. Wang, X.; Zhang, W.; Tang, J.; Huang, R.; Li, J.; Xu, D.; Xie, Y.; Jiang, R.; Deng, L.; Zhang, X.; et al. LINC01225 promotes occurrence and metastasis of hepatocellular carcinoma in an epidermal growth factor receptor-dependent pathway. Cell Death Dis. 2016, 7, e2130. [Google Scholar] [CrossRef]
  155. Wang, D.; Xing, N.; Yang, T.; Liu, J.; Zhao, H.; He, J.; Ai, Y.; Yang, J. Exosomal lncRNA H19 promotes the progression of hepatocellular carcinoma treated with Propofol via miR-520a-3p/LIMK1 axis. Cancer Med. 2020, 9, 7218–7230. [Google Scholar] [CrossRef] [PubMed]
  156. Motawi, T.M.K.; El-Maraghy, S.A.; Sabry, D.; Mehana, N.A. The expression of long non coding RNA genes is associated with expression with polymorphisms of HULC rs7763881 and MALAT1 rs619586 in hepatocellular carcinoma and HBV Egyptian patients. J. Cell Biochem. 2019, 120, 14645–14656. [Google Scholar] [CrossRef]
  157. Yao, Z.; Jia, C.; Tai, Y.; Liang, H.; Zhong, Z.; Xiong, Z.; Deng, M.; Zhang, Q. Serum exosomal long noncoding RNAs lnc-FAM72D-3 and lnc-EPC1-4 as diagnostic biomarkers for hepatocellular carcinoma. Aging 2020, 12, 11843–11863. [Google Scholar] [CrossRef]
  158. Lin, X.Q.; Huang, Z.M.; Chen, X.; Wu, F.; Wu, W. XIST Induced by JPX Suppresses Hepatocellular Carcinoma by Sponging miR-155-5p. Yonsei Med. J. 2018, 59, 816–826. [Google Scholar] [CrossRef]
  159. Wang, Y.G.; Liu, J.; Shi, M.; Chen, F.X. LncRNA DGCR5 represses the development of hepatocellular carcinoma by targeting the miR-346/KLF14 axis. J. Cell Physiol. 2018, 234, 572–580. [Google Scholar] [CrossRef]
  160. Pote, N.; Cauchy, F.; Albuquerque, M.; Voitot, H.; Belghiti, J.; Castera, L.; Puy, H.; Bedossa, P.; Paradis, V. Performance of PIVKA-II for early hepatocellular carcinoma diagnosis and prediction of microvascular invasion. J. Hepatol. 2015, 62, 848–854. [Google Scholar] [CrossRef]
  161. Statello, L.; Guo, C.J.; Chen, L.L.; Huarte, M. Gene regulation by long non-coding RNAs and its biological functions. Nat. Rev. Mol. Cell Biol. 2021, 22, 96–118. [Google Scholar] [CrossRef] [PubMed]
  162. Fu, X.D. Non-coding RNA: A new frontier in regulatory biology. Natl. Sci. Rev. 2014, 1, 190–204. [Google Scholar] [CrossRef] [PubMed]
  163. The, R.C.; Petrov, A.I.; Kay, S.J.E.; Kalvari, I.; Howe, K.L.; Gray, K.A.; Bruford, E.A.; Kersey, P.J.; Cochrane, G.; Finn, R.D.; et al. RNAcentral: A comprehensive database of non-coding RNA sequences. Nucleic Acids Res. 2017, 45, D128–D134. [Google Scholar] [CrossRef]
  164. Karagkouni, D.; Paraskevopoulou, M.D.; Tastsoglou, S.; Skoufos, G.; Karavangeli, A.; Pierros, V.; Zacharopoulou, E.; Hatzigeorgiou, A.G. DIANA-LncBase v3: Indexing experimentally supported miRNA targets on non-coding transcripts. Nucleic Acids Res. 2020, 48, D101–D110. [Google Scholar] [CrossRef] [PubMed]
  165. Hofacker, I.L.; Stadler, P.F. Memory efficient folding algorithms for circular RNA secondary structures. Bioinformatics 2006, 22, 1172–1176. [Google Scholar] [CrossRef]
  166. Chen, J.; Zhang, J.; Gao, Y.; Li, Y.; Feng, C.; Song, C.; Ning, Z.; Zhou, X.; Zhao, J.; Feng, M.; et al. LncSEA: A platform for long non-coding RNA related sets and enrichment analysis. Nucleic Acids Res. 2021, 49, D969–D980. [Google Scholar] [CrossRef]
  167. Li, Z.; Liu, L.; Jiang, S.; Li, Q.; Feng, C.; Du, Q.; Zou, D.; Xiao, J.; Zhang, Z.; Ma, L. LncExpDB: An expression database of human long non-coding RNAs. Nucleic Acids Res. 2021, 49, D962–D968. [Google Scholar] [CrossRef]
  168. Seifuddin, F.; Singh, K.; Suresh, A.; Judy, J.T.; Chen, Y.C.; Chaitankar, V.; Tunc, I.; Ruan, X.; Li, P.; Chen, Y.; et al. lncRNAKB, a knowledgebase of tissue-specific functional annotation and trait association of long noncoding RNA. Sci. Data 2020, 7, 326. [Google Scholar] [CrossRef]
  169. Volders, P.J.; Helsens, K.; Wang, X.; Menten, B.; Martens, L.; Gevaert, K.; Vandesompele, J.; Mestdagh, P. LNCipedia: A database for annotated human lncRNA transcript sequences and structures. Nucleic Acids Res. 2013, 41, D246–D251. [Google Scholar] [CrossRef]
  170. Zhang, Y.Y.; Zhang, W.Y.; Xin, X.H.; Du, P.F. dbEssLnc: A manually curated database of human and mouse essential lncRNA genes. Comput. Struct. Biotechnol. J. 2022, 20, 2657–2663. [Google Scholar] [CrossRef]
  171. Li, J.; Ma, W.; Zeng, P.; Wang, J.; Geng, B.; Yang, J.; Cui, Q. LncTar: A tool for predicting the RNA targets of long noncoding RNAs. Brief. Bioinform. 2015, 16, 806–812. [Google Scholar] [CrossRef]
  172. Li, J.; Han, L.; Roebuck, P.; Diao, L.; Liu, L.; Yuan, Y.; Weinstein, J.N.; Liang, H. TANRIC: An Interactive Open Platform to Explore the Function of lncRNAs in Cancer. Cancer Res. 2015, 75, 3728–3737. [Google Scholar] [CrossRef] [PubMed]
  173. Li, Z.; Liu, L.; Feng, C.; Qin, Y.; Xiao, J.; Zhang, Z.; Ma, L. LncBook 2.0: Integrating human long non-coding RNAs with multi-omics annotations. Nucleic Acids Res. 2023, 51, D186–D191. [Google Scholar] [CrossRef]
  174. Mas-Ponte, D.; Carlevaro-Fita, J.; Palumbo, E.; Hermoso Pulido, T.; Guigo, R.; Johnson, R. LncATLAS database for subcellular localization of long noncoding RNAs. RNA 2017, 23, 1080–1087. [Google Scholar] [CrossRef] [PubMed]
  175. Kang, J.; Tang, Q.; He, J.; Li, L.; Yang, N.; Yu, S.; Wang, M.; Zhang, Y.; Lin, J.; Cui, T.; et al. RNAInter v4.0: RNA interactome repository with redefined confidence scoring system and improved accessibility. Nucleic Acids Res. 2022, 50, D326–D332. [Google Scholar] [CrossRef] [PubMed]
  176. Li, J.H.; Liu, S.; Zhou, H.; Qu, L.H.; Yang, J.H. starBase v2.0: Decoding miRNA-ceRNA, miRNA-ncRNA and protein-RNA interaction networks from large-scale CLIP-Seq data. Nucleic Acids Res. 2014, 42, D92–D97. [Google Scholar] [CrossRef]
  177. Cui, L.; Ma, R.; Cai, J.; Guo, C.; Chen, Z.; Yao, L.; Wang, Y.; Fan, R.; Wang, X.; Shi, Y. RNA modifications: Importance in immune cell biology and related diseases. Signal Transduct. Target. Ther. 2022, 7, 334. [Google Scholar] [CrossRef]
Figure 1. Diverse mechanisms of lncRNAs’ functions in cellular regulation. (A) Signal. LncRNAs act as molecular indicators, responding to various cellular stimuli. (B) Decoy. LncRNAs can bind and sequester transcription factors or other proteins, preventing them from interacting with their target genomic loci. (C) Guide. LncRNAs direct chromatin-modifying enzymes to specific genomic regions, enabling targeted epigenetic modifications. (D) Scaffold. LncRNAs facilitate the formation of multi-protein complexes, providing a structural platform for these assemblies. (E) Enhancer RNA. LncRNAs can function as enhancers, looping DNA to bring distant regions into proximity for transcriptional activation. (F) miRNA Sponge. LncRNAs can act as sponges for miRNAs, sequestering them and preventing them from binding to their target mRNAs, thus inhibiting miRNA-mediated gene repression.
Figure 1. Diverse mechanisms of lncRNAs’ functions in cellular regulation. (A) Signal. LncRNAs act as molecular indicators, responding to various cellular stimuli. (B) Decoy. LncRNAs can bind and sequester transcription factors or other proteins, preventing them from interacting with their target genomic loci. (C) Guide. LncRNAs direct chromatin-modifying enzymes to specific genomic regions, enabling targeted epigenetic modifications. (D) Scaffold. LncRNAs facilitate the formation of multi-protein complexes, providing a structural platform for these assemblies. (E) Enhancer RNA. LncRNAs can function as enhancers, looping DNA to bring distant regions into proximity for transcriptional activation. (F) miRNA Sponge. LncRNAs can act as sponges for miRNAs, sequestering them and preventing them from binding to their target mRNAs, thus inhibiting miRNA-mediated gene repression.
Cells 12 02272 g001
Figure 2. A schematic view of m6A modification machinery. The m6A methyltransferase complex, composed of METTL3, METTL14, KIAA1429, RBM15, and WTAP, is responsible for adding the m6A modification. In contrast, demethylases, including FTO and ALKBH5, are depicted removing the m6A marks. This m6A modification plays a pivotal role in determining cellular fate and is involved in the onset of diseases such as cancer.
Figure 2. A schematic view of m6A modification machinery. The m6A methyltransferase complex, composed of METTL3, METTL14, KIAA1429, RBM15, and WTAP, is responsible for adding the m6A modification. In contrast, demethylases, including FTO and ALKBH5, are depicted removing the m6A marks. This m6A modification plays a pivotal role in determining cellular fate and is involved in the onset of diseases such as cancer.
Cells 12 02272 g002
Table 1. LncRNAs with differential expression profiles and their roles in hepatocarcinogenesis.
Table 1. LncRNAs with differential expression profiles and their roles in hepatocarcinogenesis.
ExpressionlncRNAKnown FunctionReference
UpregulatedHULCEMT, metastasis, apoptosis[56,67,68]
MALAT1EMT, metastasis,
apoptosis, cell-cycle arrest
[61,65]
HOTAIREMT, metastasis,
apoptosis, cell-cycle arrest
[34,35,36]
DLEU2Vascular invasion, lymphatic metastasis [69]
SNHG1EMT, cell-cycle regulation,
metastasis, apoptosis
[70]
NEAT1Ferroptosis, metastasis, proliferation, invasion, drug resistance[47,51,53,54,71]
TUG1Metastasis, apoptosis[72]
CRNDEProliferation, migration, invasion[73]
KDM4A-AS1EMT, metastasis[74]
UCA1EMT, cell-cycle regulation, apoptosis[75]
ANRILMetastasis, apoptosis[76,77]
CASC15EMT, metastasis,
suppression of apoptosis
[78]
ZFAS1Metastasis[79]
CARLo-5EMT[80]
LOC90784Apoptosis, cell-cycle arrest[81]
H19EMT, metastasis,
suppression of apoptosis
[82,83,84]
PCAT-14cell-cycle arrest[85]
LINC02551EMT, metastasis[86]
LINC01116EMT, cell-cycle regulation, metastasis, immune-cell infiltration[87]
CCAT2Proliferation, migration, invasion[88]
PVT1Microvascular invasion, proliferation[89,90]
DownregulatedDGCR5Proliferation, migration, invasion[91]
MEG3Inhibition of metastasis, angioinvasion and proliferation by cell-cycle regulation[92,93,94]
FENDRRApoptosis, Treg-mediated immune escape[95,96]
GAS5Suppression of proliferation, drug resistance and M2 macrophage polarization[97,98]
RAB11B-AS1Apoptosis[99]
Table 2. m6A modified lncRNAs in HCC.
Table 2. m6A modified lncRNAs in HCC.
ExpressionlncRNAm6A Binding PartnerReference
UpregulatedHULCIGF2BP1[123]
LINC00958METTL3[118]
LINC02362-[119]
SNHG20
SNHG6
MIR4458HGIGF2BP2[122]
LINC02551ALKBH5[86]
MIR155HGMETTL14[120]
SLC7A11-AS1METTL3[124]
ARHGAP5-AS1METTL14
IGF2BP2
[121]
DownregulatedRAB11B-AS1METTL16[99]
MEG3METTL3[94]
AC115619WTAP[125]
Table 3. Potential lncRNAs as diagnostic biomarkers for HCC detection.
Table 3. Potential lncRNAs as diagnostic biomarkers for HCC detection.
ExpressionlncRNABlood DetectionReference
UpregulatedMVIHYes[136]
HOTTIPYes[135,138]
ATBYes[139]
DLEU2Yes[135]
MALAT1Yes
SNHG1Yes
PVT1Yes[145]
uc002mbe.2Yes
LINC00853Yes[134]
UCA1Yes[146,147]
WRAP53Yes[147]
LINC00152Yes[148]
RP11-160H22.5Yes
XLOC014172Yes
DANCR-[150]
LINC00978Yes[151]
LncDQYes[152]
SPRY4-IT1-[153]
UBE2CP3Yes[133]
LINC01225Yes[154]
H19Yes[155]
uc003wbdYes[156]
CAHM-[149]
FAM72D-3Yes[157]
EPC1-4Yes
DownregulatedJPX-[158]
XIST-
DGCR5-[159]
X91348Yes[137]
Table 4. Web-based tools for comprehensive lncRNA research and analysis.
Table 4. Web-based tools for comprehensive lncRNA research and analysis.
NameWebsiteDescriptionReference
RNACentralhttps://rnacentral.org (accessed on 10 September 2023)A public platform offering access to a wide collection of non-coding RNA sequences from various organisms and RNA types[163]
LncBasehttps://diana.e-ce.uth.gr/lncbasev3 (accessed on 10 September 2023)A database cataloging 500,000 verified miRNA–lncRNA interactions across 243 cell types[164]
RNAfoldhttp://rna.tbi.univie.ac.at/cgi-bin/RNAWebSuite/RNAfold.cgi (accessed on 10 September 2023)A tool used for predicting the secondary structure of RNA sequences[165]
LncSEAhttp://bio.liclab.net/LncSEA/index.php (accessed on 10 September 2023)A platform for lncRNA-related sets and enrichment analysis[166]
LncExpDBhttps://ngdc.cncb.ac.cn/lncexpdb (accessed on 10 September 2023)Expression database of human lncRNAs[167]
lncRNAKBhttps://bio.tools/lncrnakb (accessed on 10 September 2023)A knowledgebase of tissue-specific functional annotation and trait association of lncRNA[168]
LNCipediahttp://www.lncipedia.org (accessed on 10 September 2023)A database for annotated human lncRNA transcript sequences and structures[169]
dbEssLnchttps://esslnc.pufengdu.org (accessed on 10 September 2023)A manually curated database of human and mouse essential lncRNA genes[170]
LncTarhttp://www.cuilab.cn/lnctar (accessed on 10 September 2023)A tool for predicting the RNA targets of lncRNAs.[171]
TANRIChttps://bioinformatics.mdanderson.org/public-software/tanric (accessed on 10 September 2023)TANRIC webapp provides analysis of lncRNA in cancer, highlighting potential therapeutic targets and biomarkers[172]
LncBookhttps://ngdc.cncb.ac.cn/lncbook (accessed on 10 September 2023)A comprehensive database of human lncRNAs, offering annotations for understanding their roles in diseases and biological contexts.[173]
lncATLAShttps://lncatlas.crg.eu (accessed on 10 September 2023)A database showing subcellular locations of GENCODE-annotated lncRNAs, using RCI values[174]
RNAInterhttp://www.rnainter.org (accessed on 10 September 2023)A database with a scoring system to rate the confidence of RNA-associated interactions based on experimental evidence and tissue/cell types[175]
ENCORIhttps://rnasysu.com/encori (accessed on 10 September 2023)A platform for studying RNA interactions, integrating diverse data and enabling pan-cancer analysis[176]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Eun, J.W.; Cheong, J.Y.; Jeong, J.-Y.; Kim, H.S. A New Understanding of Long Non-Coding RNA in Hepatocellular Carcinoma—From m6A Modification to Blood Biomarkers. Cells 2023, 12, 2272. https://doi.org/10.3390/cells12182272

AMA Style

Eun JW, Cheong JY, Jeong J-Y, Kim HS. A New Understanding of Long Non-Coding RNA in Hepatocellular Carcinoma—From m6A Modification to Blood Biomarkers. Cells. 2023; 12(18):2272. https://doi.org/10.3390/cells12182272

Chicago/Turabian Style

Eun, Jung Woo, Jae Youn Cheong, Jee-Yeong Jeong, and Hyung Seok Kim. 2023. "A New Understanding of Long Non-Coding RNA in Hepatocellular Carcinoma—From m6A Modification to Blood Biomarkers" Cells 12, no. 18: 2272. https://doi.org/10.3390/cells12182272

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