Next Article in Journal
Drug Repurposing Approach to Identify Candidate Drug Molecules for Hepatocellular Carcinoma
Previous Article in Journal
Data-Driven Modelling of Substituted Pyrimidine and Uracil-Based Derivatives Validated with Newly Synthesized and Antiproliferative Evaluated Compounds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

MicroRNAs in Hepatocellular Carcinoma Pathogenesis: Insights into Mechanisms and Therapeutic Opportunities

by
Khadijeh Mahboobnia
1,2,
Dianne J. Beveridge
1,2,
George C. Yeoh
1,3,
Tasnuva D. Kabir
1,2 and
Peter J. Leedman
1,2,*
1
Laboratory for Cancer Medicine, Harry Perkins Institute of Medical Research, QEII Medical Centre, Perth, WA 6009, Australia
2
Centre for Medical Research, The University of Western Australia, Perth, WA 6009, Australia
3
School of Molecular Sciences, The University of Western Australia, Perth, WA 6009, Australia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(17), 9393; https://doi.org/10.3390/ijms25179393
Submission received: 21 July 2024 / Revised: 18 August 2024 / Accepted: 20 August 2024 / Published: 29 August 2024
(This article belongs to the Section Molecular Oncology)

Abstract

:
Hepatocellular carcinoma (HCC) presents a significant global health burden, with alarming statistics revealing its rising incidence and high mortality rates. Despite advances in medical care, HCC treatment remains challenging due to late-stage diagnosis, limited effective therapeutic options, tumor heterogeneity, and drug resistance. MicroRNAs (miRNAs) have attracted substantial attention as key regulators of HCC pathogenesis. These small non-coding RNA molecules play pivotal roles in modulating gene expression, implicated in various cellular processes relevant to cancer development. Understanding the intricate network of miRNA-mediated molecular pathways in HCC is essential for unraveling the complex mechanisms underlying hepatocarcinogenesis and developing novel therapeutic approaches. This manuscript aims to provide a comprehensive review of recent experimental and clinical discoveries regarding the complex role of miRNAs in influencing the key hallmarks of HCC, as well as their promising clinical utility as potential therapeutic targets.

1. Introduction

1.1. Hepatocellular Carcinoma

Hepatocellular carcinoma (HCC) is a highly heterogeneous disease, accounting for 75–85% of human liver cancers, and is the third leading cause of cancer-related deaths worldwide [1,2]. Its incidence has been steadily increasing globally over the past three decades [1]. In Australia, there was a remarkable 378% increase in HCC cases between 1982 and 2015 [3]. The main causes of HCC are chronic infection with viral hepatitis (hepatitis B virus (HBV) and hepatitis C virus (HCV)), heavy alcohol consumption, and non-alcoholic fatty liver disease (NAFLD) [4]. As the prevalence of viral hepatitis has been reduced due to the global coverage of hepatitis B vaccination and the successful antiviral treatment for HCV, it is expected that NAFLD, which is associated with metabolic syndrome, type 2 diabetes, and central obesity, will be the major risk factor for HCC in the future [1,5,6]. These etiological factors disrupt crucial signaling cascades within hepatocytes, initiate tumor formation and heterogeneity, and compromise the subsequent treatment efficacy [7].
Three distinct transcriptomic signatures labelled S1, S2, and S3 have been identified in HCC, and their clinical severity being worse in S1 [8]. The S1 subtype is characterized by abnormal activation of the Wnt signaling pathway and overexpression of transforming growth factor beta 1 (TGFβ1), which correlates with an aggressive epithelial-to-mesenchymal (EMT) phenotype. S2 tumors have activated Myc and Akt pathways, and the upregulation of stemness markers such as α-fetoprotein (AFP) and epithelial cell adhesion molecule (EPCAM). In contrast, S3 tumors are smaller, more differentiated, and express genes associated with normal hepatocyte function, including glycolipid metabolism-associated genes and tumor suppressor genes p21 and p53, resulting in a less aggressive clinical profile [8].
Key pathways involved in HCC pathogenesis, including the receptor tyrosine kinase (RTK), fibroblast growth factor (FGF), PI3K/Akt and MAPK/ERK, WNT, Hedgehog, Notch, JAK/STAT, and ubiquitin-proteasome pathways, are dysregulated in liver cancer due to genetic and epigenetic factors [9]. Inflammatory hepatic lesions upregulate and activate various transcription factors and gene regulators that orchestrate these molecular cascades, driving different hallmarks of HCC, including uncontrolled cell growth, evasion of cell death signals, tumor microenvironment remodelling, immune escape, metabolic reprogramming, invasion, and metastasis [10].

1.2. microRNAs

Recent studies have significantly advanced our understanding of non-coding RNAs and their critical roles in numerous cellular functions [11]. MicroRNAs (miRNAs/miRs), which are endogenously transcribed RNA molecules of approximately 22 nucleotides, have emerged as key players in various biological processes. miRNAs are encoded by specific genes located in intergenic sequences or within intronic regions [12]. These regions are transcribed by RNA polymerase II or III to form a primary miRNA transcript (pri-miRNA) that is over 1000 nucleotides in length [13,14]. The pri-miRNA is then processed in the nucleus by a microprocessor complex, involving Drosha and DGCR8 proteins, to generate a precursor miRNA (pre-miRNA) of approximately 85 nucleotides [15,16]. This pre-miRNA is then exported to the cytoplasm via the exportin-5 and Ran-GTP protein complexes. In the cytoplasm, the Dicer enzyme and the TRBP/PACT protein complex further cleave and process the pre-miRNA to generate a mature 20–22 nucleotide miRNA duplex consisting of passenger and guide strands [16]. The guide strand, with lower thermodynamic stability at the 5′ end, is favored for integration into the Argonaute 2 (AGO 2) protein, resulting in the formation of an RNA-induced silencing complex [17] [15,16]. RISC targets are subjected to either translational repression or mRNA degradation, primarily through the formation of base pairs between the miRNA’s 5′ seed sequence and the 3′ untranslated region (UTR) of the target mRNA (Figure 1) [18,19].
The dysregulation of miRNAs is a common feature in various cancers, where they can act as either oncogenes (oncomiRs) or tumor suppressors, depending on their expression levels and the context of the cancer type [20]. MiRNAs influence several fundamental biological processes, including cell proliferation, differentiation, apoptosis, and angiogenesis, all of which are critical in cancer development and progression [20,21]. The dysregulation of miRNA expression can occur through various mechanisms, such as gene amplification, deletion, mutation, and epigenetic changes, which can lead to the aberrant expression of miRNAs and contribute to tumorigenesis [21,22]. Numerous studies have highlighted the significant impact that altered miRNA expression has on HCC pathogenesis, metastasis, and chemoresistance [23,24,25,26].
This review delves into the latest literature on the impact of miRNAs on oncogenic pathways in HCC and their implications for treatment outcomes. Furthermore, we offer insights into the current progress in the development of miRNA-based therapies for HCC treatment.

2. Role of miRNAs in Hepatocellular Carcinoma Cell Survival

Growth signaling autonomy, a hallmark of cancer, allows tumor cells to sustain proliferation independently of external growth factors. This autonomy is crucial for cancer progression, where it is associated with enhanced stemness, proliferation, and resistance to therapies [27,28]. In HCC, miRNAs are pivotal in regulating signaling pathways that contribute to cell survival and tumorigenesis. miRNA dysregulation characterized by an overexpression of oncomiRs and a reduction in tumor suppressor miRNAs leads to an excessive proliferation of growth signals and a diminished response to anti-growth and pro-apoptotic signals, ultimately permitting uncontrolled cell division [29].

2.1. miRNAs in Regulating Hepatocellular Carcinoma Cell Cycle and Proliferation

The tightly controlled regulation of the cell cycle involves essential components such as cyclin-dependent kinases (CDKs) like cyclin-D-dependent CDK4/CDK6 and cyclin-E-dependent CDK2. CDK inhibitors (CDKis) from the INK4 and Cip/Kip families, including p21Cip1, p27Kip1, and p57Kip2, also play crucial roles. Various signaling pathways such as RB-E2F, Hedgehog and Wnt, Ras-Raf-MEK-ERK, Hippo, and c-Myc are integral to this regulatory network [30]. When the cell cycle machinery malfunctions, leading to uncontrolled cellular proliferation, it becomes a significant feature of cancer [27]. miRNAs contribute to regulating these critical cell cycle pathways, affecting cellular processes [31].
In HCC, numerous tumor suppressor miRNAs inhibit cell cycle entry, including miR-9, miR-424-5p, miR-621, miR-125b-5p, the miR-29 family, and miR-450b-3p, which target HMGA2, E2F7, CAPRIN1, TXNRD1, RPS15A, and PGK1, respectively, leading to cell cycle arrest in the G0/G1 phase [32,33,34,35,36,37,38].
Overexpression of nucleolar and spindle associated protein 1 (NUSAP1), which is involved in mitosis, spindle assembly, and chromosome attachment [39], has been observed in HCC cells [40,41]. miR-193a-5p has been shown to induce G1 phase cell cycle arrest by targeting NUSAP1 in liver cancer cells, which in turn downregulates Cyclin E1, Cyclin D1, Cyclin B1, and Cyclin A2 and induces p21 expression [42].
Additionally, several miRNAs obstruct cell cycle progression in HCC by regulating the G1 to S phase transition (G1/S checkpoint). For example, miR-214 targets Wnt3a [43] and MELK [44] to decelerate HCC cell proliferation and induce G1 phase arrest. Ectopic overexpression of miR-0308-3p in HCC cells results in cell cycle blockade in the G1/S phase by downregulating CDK6 and Cyclin D1 [45]. Additionally, miR-217 reduced liver cancer cell proliferation and halted G1/S transition by targeting EZH2, Cyclin-D1 [46], MTDH [47], and KLF5 [48].
Recent reports have highlighted the involvement of let-7-5p, miR-31-5p, and miR-3613-3p in regulating the later stages of the cell cycle, specifically by suppressing G2/M transition [49,50,51]. In contrast to cell cycle inhibitory miRNAs, pro-proliferative miRNAs, such as miR-494 [52], miR-191 [53], miR-3682-3p [54], miR-10b [55], and miR-221-3p [56] facilitate tumor growth and cell cycle progression in HCC (Figure 2).

2.2. Regulatory Role of miRNAs in Hepatocellular Carcinoma Cell Death Pathways

Multiple stress conditions can activate various death signaling pathways, including necrosis, apoptosis, necroptosis, mitoptosis, ferroptosis, pyroptosis, and autophagy [57]. Apoptosis, necroptosis (programmed necrosis), and autophagy have been extensively studied in the context of liver cancer, and dysregulation of these death signaling pathways play a crucial role in drug resistance and treatment failure in HCC.

2.2.1. Apoptosis Related miRNAs

Apoptosis, a programmed cell death process, is facilitated by a family of proteases known as caspases (cysteinyl aspartate-specific proteases) [17]. This type of cell death can be initiated by intrinsic signals, such as genotoxic stress, or extrinsic signals, such as the binding of ligands to cell surface death receptors (DRs), including TNF-R1, CD95 (Fas), TRAIL-R1 (DR4), TRAIL-R2 (DR5), DR3, and DR6 [58].
miRNAs modulate apoptotic events by acting as pro- or anti-apoptotic miRNAs. For example, members of the BCL-2 family of genes, the chief mediators of apoptosis, are regulated by multiple miRNAs including miR-378 [59], miR-9-5p [60], miR-448 [61], and miR-133b [62].

2.2.1.1. Pro-Apoptotic miRNAs in Hepatocellular Carcinoma

miRNA profiling of HCC tissues revealed downregulation of pro-apoptotic miRNAs in cancerous cells [63]. Corroborating these observations, studies have shown that treatment of HCC cells with miR-133b mimics, a pro-apoptotic miRNA, induces apoptotic cell death by enhancing the activities of caspase-3/-8 and increasing the Bax/Bcl-2 protein expression ratio by regulating the EGFR/PI3K/Akt/mTOR axis [62]. Similarly, miR-22-3p promotes apoptosis by inhibiting the Akt/PI3K pathway via direct suppression of AKT2 protein expression [64]. Furthermore, Wang et al. recently reported that miR-206 suppresses c-MET expression, thereby reducing the malignant behavior of HCC cells and stimulating apoptosis [65].
Importantly, the JAK1/STAT3 signaling pathway has emerged as an important survival mechanism in HCC, promoting damage repair, cellular renewal, and rejuvenation within the cirrhotic microenvironment, thereby exerting an anti-apoptotic effect on liver cancer cells [66]. miR-26a has been found to be pro-apoptotic in HCC cells by directly targeting JAK1 to modulate its expression, leading to increased apoptosis [67]. Moreover, the JNK and MAPK pathways are closely associated with the stress response and apoptosis [68,69,70], and notably MAP3K2 is upregulated in many cancers [71,72,73]. Recent evidence suggests that miR-302a exerts antiproliferative and pro-apoptotic effects in HCC cells by targeting both MAP3K2 and PBX3 (Figure 3a) [74].

2.2.1.2. Anti-Apoptotic miRNAs in Hepatocellular Carcinoma

Anti-apoptotic miRNAs are aberrantly overexpressed during tumorigenesis [63]. For example, miR-9-5p promotes HCC progression via inhibition of KLF4, thereby activating AKT/mTOR signaling, resulting in increased expression of the anti-apoptotic protein Bcl-2, and reduced expression of the pro-apoptotic protein Bax [60]. Furthermore, miR-33a facilitates HCC cell survival by inhibiting PPARα [75], a pro-apoptotic factor involved in the degradation of the Bcl-2 protein [76,77]. In HCC with underlying cirrhosis, miR-3682-3p was found to impair apoptosis and enhance cell survival by suppressing PHLDA1 expression and subsequent downregulation of Fas [54].
In viral hepatitis, activation of toll-like receptor 3 (TLR3) induces apoptosis of infected hepatocytes by promoting NF-kB transcription and activating caspase 8 [78,79]. miR-155 is often upregulated in HCC, contributing to the downregulation of TLR3, which is associated with apoptosis evasion and poor prognosis in these cancer cells [80]. The gold (Au) tagged antimiR-155 nanocomplexes (NCs) triggers TLR3-dependent apoptosis in HCC cells. These antimiR-155 NCs effectively silence miR-155 in HCC cells, inhibiting proliferation and migration while inducing apoptosis through TLR3 signaling [80].
miR-106b exhibits an anti-apoptotic role in HCC by negatively regulating DR4 [81]. The depletion of miR-106b increased sensitivity to TNF-related apoptosis-inducing ligand (TRAIL) treatment by upregulating DR4 expression and facilitating TRAIL-DR4 mediated cell death (Figure 3b) [81].

2.2.2. Necroptosis Related miRNAs

Necroptosis (programmed necrosis) is a regulated form of lytic cell death that is triggered by various environmental stressors such as chemical and mechanical stress, osmotic shock, toxins, and viral and bacterial products [82,83]. A significant amount of experimental evidence has demonstrated that necroptosis contributes to tumor progression and metastasis by recruiting inflammatory cells, thereby promoting angiogenesis, proliferation, and invasiveness of cancer cells [84].
The role of miRNAs in regulating necroptosis in HCC has been underexplored in scientific literature. A study by Visalli et al. marks a significant step in this direction by identifying three specific miRNAs, namely miR-371-5p, miR-373, and miR-543, which exhibited abnormal overexpression in HCC tumor tissues and were found to drive necroptosis by inhibiting Casp-8 [85].

2.2.3. Autophagy Related miRNAs

Autophagic cell death [86] is a regulated process in which cellular organelles and proteins are targeted for lysosomal degradation and categorized as type II cell death [87]. In tumorigenesis, ACD has dual context-dependent functional roles. ACD suppresses tumor initiation and malignant transformation [87] but also promotes the survival of tumor cells, metastatic progression, and drug resistance by providing substrates for cellular metabolism [88,89]. Autophagic flux involves the formation of autophagosomes, cargo degradation, and eventual recycling of breakdown products. The ULK1 complex, PI3K complex, ATG12-ATG5-ATG16L1, and LC3-PE, and lysosomal proteins such as LAMPs and cathepsins are involved in autophagosome maturation, elongation and closure, and cargo degradation, respectively [90,91]. Comprehending the functions of these proteins and their coordinated actions is crucial for understanding the molecular intricacies of autophagic flux in cancer cells.

2.2.3.1. Pro-Autophagic miRNAs in Hepatocellular Carcinoma

As a key regulator of the autophagy pathway, AMP-activated protein kinase (AMPK) plays a role in various stages of autophagy, including initiation, autophagosome formation, and fusion with lysosomes [92]. Recent research suggests that miR-519d inhibits cell proliferation and promotes apoptosis and autophagy by targeting Rab10 expression and activating the AMP signaling pathway [93]. Experiments showed that restoring miR-519d in vivo suppressed tumor growth, with upregulation of autophagy-related genes e.g., Beclin1, Atg5, p53, and pro-apoptotic Bax, and downregulation of Rab10, mTOR, and Bcl-2 [93]. Additionally, miR-185 has been shown to have a tumor-suppressive effect in HCC cells by inducing autophagy through regulation of the Akt1 pathway [94].

2.2.3.2. Anti-Autophagic miRNAs in Hepatocellular Carcinoma

In HCC cells, miR-181a-5p has been shown to attenuates autophagic flux by targeting Atg7 [95]. Additionally, Mig-6, a cytoplasmic protein that negatively regulates EGFR signaling [96], promotes apoptosis and inhibits the autophagic pathway through the upregulation of miR-193a-3p. Mechanistically, miR-193a-3p inhibits TGF-β2 protein expression, leading to decreased autophagy [97]. Another miRNA, miR-7, directly interferes with Atg5 and inhibits HCC metastasis [98]. Similarly, miR-30a reduces the expression of Beclin 1 and Atg5, resulting in autophagy inhibition and improved resistance to anoikis in HCC cells [99]. Furthermore, miR-26a/b has been identified as a direct inhibitor of ULK1 in HCC, impeding autophagic flux at an early stage and enhancing the sensitivity of cancer cells to chemotherapy [100].
Table 1 provides a comprehensive overview of miRNAs involved in the regulation of cell survival and proliferation in HCC.

3. miRNAs Regulatory Role in Tumor Cell Stemness

Cancer stem cells (CSCs) are a subset of cancerous cells within a tumor mass that have the ability to self-renew and generate new tumors, leading to relapse, metastasis, and resistance to radiation and chemotherapy [145]. Hepatic progenitor cells, hepatoblasts, and adult hepatocytes are potential sources of hepatic CSCs [146]. Specific antigenic markers (e.g., EpCAM, CD133, CD90, CD44, CD47, CD24, CD13, calcium channel α2δ1 isoform 5, K19, OV6, ABCG2, ALDH, and Hoechst dye efflux) can be used to identify and isolate liver CSCs [147,148,149]. These markers have been found to be expressed in different subsets of liver CSCs and have been used in combination to improve the sensitivity and specificity of CSC detection [150,151].
The stemness characteristics of cancer cells are maintained through various signaling pathways including Wnt/β-catenin, IL-6/STAT3, TGF-β, Notch, HH, Hippo, BMI1, NF-κB, PI3K/Akt/mTOR, and Ras/Raf/MAPK [152]. Other mechanisms contributing to the stemness of hepatic cancer cells include extracellular matrix (ECM) remodelling, epithelial-mesenchymal transition (EMT), hypoxia, epigenetic modifications, and autophagy [152]. miRNAs have been found to target important regulatory genes in these signaling pathways to acquire stem-like characteristics.

3.1. miRNAs Inhibiting Hepatocellular Carcinoma Stemness

Upregulation of the E2F family of transcription factors in CSCs plays a significant role in their self-renewal capacity, proliferation, aggressiveness, and resistance to chemotherapy and radiotherapy [153]. miRNA-302a/d inhibits the expression of E2F7 in liver CSCs, leading to attenuation of the AKT/β-catenin/CCND1 signaling pathway, repression of cell proliferation, and reduction in stemness characteristics [119].
Analysis of human HCC samples showed that CBX4 is abnormally upregulated in a subset of CSCs, marked by CD44+ CD133+ expression [154]. The miR-6838-5p/CBX4 axis affects liver CSC function via the ERK pathway with CBX4 as the target of miR-6838-5p [154]. In clinical HCC tissues, low miR-589-5p and high CD90 expression are correlated with vascular invasion and recurrence, and miR-589-5p negatively regulates CD90+ liver CSCs by suppressing MAP3K8 expression [155]. miR-148a and miR-148b suppress HCC progression and stemness, respectively, by targeting ACVR1 and NRP1 [156,157]. Additionally, Shi et al. reported that miR-296-5p mitigated EMT and stemness characteristics of HCC cells by modulating the Brg1/Sall4 axis and blocking NRG1/ERBB2/ERBB3/RAS/MAPK/Fra-2 signaling [158,159].

3.2. miRNAs Promoting Hepatocellular Carcinoma Stemness

Numerous miRNAs have been shown to play a role in promoting stem-like properties in HCC cells. Studies have demonstrated that the Oct4/miR-1246 axis can upregulate Wnt/β-catenin signaling in hepatic CD133+CSCs by targeting AXIN2 and GSK3β, leading to the accumulation of β-catenin [160]. Similarly, miR-5188 enhances the stemness phenotype by directly targeting FOXO1 and activating the Wnt/β-catenin cascade and EMT [161]. Another significant oncomiR in HCC is miR-106b-5p, which contributes to the stemness and aggressiveness of tumors by blocking PTEN expression and activating the PI3K/AKT pathway [162]. Table 2 lists the HCC stemness-regulating miRNAs and their underlying mechanisms.

4. miRNAs Regulatory Role in Hepatocellular Carcinoma Metastasis

The majority of cancer-related deaths (~90%) are attributable to metastasis rather than primary tumor growth [169]. The metastatic process involves several steps, including local tissue invasion, intravasation, survival in circulation, extravasation, and secondary site colonization [169,170]. During local invasion and metastatic dissemination, cancer cells undergo EMT, a process by which epithelial cells acquire a mesenchymal phenotype [27]. Dysregulation of miRNAs in HCC can affect key signaling pathways including TGF-β/Smad, MAPK, NF-kB, JAK/STAT, Hedgehog, Wnt/β-catenin, and Hippo-YAP transcriptional co-activators with TAZ, which are involved in EMT induction, and cancer progression (Figure 4) [171,172].
In HCC, invasion and metastasis occur at both intrahepatic and extrahepatic sites, indicating a highly aggressive tumor phenotype [173,174].

4.1. Anti-Metastatic miRNAs in Hepatocellular Carcinoma

Several miRNAs have been shown to have anti-metastatic effects in HCC, including miR-211-5p, miR-130a-3p, miR-193a-5p, miR-30a, miR-124, and miR-15a-3p [42,99,175,176,177,178,179].
Long-chain acyl-CoA synthetase 4 (ACSL4) is upregulated in HCC [180,181] and contributes to disease progression and poor prognosis by stabilizing c-Myc through the ERK/FBW7 axis [182]. miR-211-5p attenuates aggressive HCC features by directly regulating ACSL4 expression [175].
Recent findings have indicated that miR-130a-3p expression is reduced in HCC tumor tissues, and its restoration can inhibit cell proliferation, migration, and invasion by targeting the androgen receptor (overexpressed in approximately 37% of HCC cases [183]) and the subsequent decrease in β-catenin and Slug expression [176].
Research has linked irregular gene variants resulting from incorrect splicing to the promotion of cancer growth [184,185]. SF3B4, a subunit of the spliceosome complex, is highly dysregulated in HCC and serves as an early diagnostic biomarker [186]. miR-133b suppresses SF3B4, affecting downstream molecules, such as KLF4, KIP1, and SNAI2, potentially inhibiting tumor metastasis [126].
Focal adhesion kinase (FAK) plays a critical role in cell migration, metastasis [187], and angiogenesis [188,189], as a downstream mediator of angiogenic growth factor receptors. miR-7 inhibits the metastasis and invasion of HCC cells by modulating Snail-1, Slug, EGFR, TYRO3, and MMP-9. Additionally, it downregulates FAK expression, leading to downstream suppression of the Akt pathway [190]. Eukaryotic translation initiation factor 4A3 (eIF4A3) regulates mRNA splicing [191] and promotes tumor growth in HCC and other carcinomas [192,193,194]. miR-2113 inhibits cell migration and EMT by reducing the interaction between eIF4A3 and WDR66, a positive regulator of EMT, through downregulating WDR66 expression [195].

4.2. Pro-Metastatic miRNAs (metastamiRs) in Hepatocellular Carcinoma

Literature has consistently emphasized the essential role of miRNAs in various stages of tumor progression, such as transforming the tumor microenvironment and extracellular matrix, stimulating neoangiogenesis, and facilitating tumor cell invasion, metastasis, and colonization. For example, miR-18a [196,197], miR-25 [198], miR-106b-5p [199], miR-183-5p [138], and miR-376c-3p [200] are noteworthy miRNAs that act as pro-metastatic miRNAs in HCC. These miRNAs target genes involved in cancer cell migration and invasion, including Bcl2L10, KLF4, Fbxw7, FOG2, PDCD4, and ARID2.
Moreover, miR-93-5p promoted liver cancer cell metastasis by activating the MAP3K2/p38-JNK/p21 signaling pathway [201]. ChIP analysis of HepG2 HCC cells revealed a positive feedback loop between miR-93-5p, MAP3K2, and c-Jun, where c-Jun targets the miR-93-5p promoter to enhance its transcription [201]. Table 3 classifies metastasis-related miRNAs in HCC.

5. Exosomal miRNAs in Hepatocellular Carcinoma Progression

The complex process of cancer cell metastasis requires malignant cells to evade death signals and nullify immune responses to facilitate local invasion and colonization at secondary sites [264]. In solid cancers, tumor microenvironment (TME)-embedded cancer cells represent a complex and dynamic stroma that consists of fibroblasts, endothelial cells, mesenchymal stem cells, adipocytes, and immune cells (e.g., macrophages, neutrophils, dendritic cells, and lymphocytes). Moreover, blood and lymph vessels as well as non-cellular components, including cytokines, extracellular vesicles (EVs), and extracellular matrix (ECM), are among the other TME components [265]. It is well established that an intercellular communication network exists between a primary tumor and its stroma, driven by various growth factors, cytokines, chemokines, and EVs to sustain cancer cell survival and promote invasion and metastasis [266].
Exosomes are EVs which develop from the endosomal plasma membrane, are 30–150 nm in size and serve as cargo carriers that transport molecules, including DNA, mRNA, miRNAs, proteins, and lipids [267]. Accordingly, exosome-delivered miRNAs can trigger specific signaling pathways in recipient cells to modulate processes involved in tumor development and invasion, such as EMT, angiogenesis, stemness, chemoresistance, and immune responses [268].

5.1. EMT Related Exosomal miRNAs in Hepatocellular Carcinoma

Exosomal miRNAs have emerged as key players in the progression of HCC [269]. For instance, highly metastatic HCC cells secrete miR-92a-3p in exosomes and promote EMT and metastasis of neighbouring less metastatic cancer cells by suppressing PTEN and activating the Akt/Snail signaling pathway [244].
Hypoxia is a common condition in solid tumors, including HCC, where there is upregulation of specific hypoxia-inducible factors (HIFs), leading to cancer progression and therapeutic resistance [270,271]. Tian et al. reported that HIF binding to the promoter regions of exosomal miRNAs miR-21 and miR-10b induces a metastatic phenotype in HCC by upregulating vimentin and reducing the expression of PTEN and E-cadherin [272]. Another study found that, under hypoxic conditions, HCC cells secrete exosomes rich in miR-1273f, which augments proliferation, migration, invasion, and EMT phenotypes in recipient HCC cells by direct inhibition of LHX6 [260].
However, certain exosomal miRNAs act as tumor suppressors, inhibiting cancer development and progression. For example, exosomal miR-125b has been shown to exert anti-metastatic effects by inhibiting MMPs, targeting SMAD2, and disrupting the TGF-β1/SMAD signaling pathway, which collectively leads to reduced EMT [209]. Notably, exosomal miR-125b is a potential early biomarker for detecting extrahepatic metastasis in HCC, given that lower serum exosomal miR-125b levels are associated with an increased risk of metastasis in cancer patients [209].

5.2. Exosomal miRNAs Involved in Non-EMT Related Hepatocellular Carcinoma Progression and Metastsis

Emerging data suggest that metastasis is a multi-directional process where circulating tumor cells (CTCs) can settle in both distant organs and the primary tumor itself, a phenomenon known as “tumor self-seeding” [273]. Tumor-derived inflammatory cytokines, such as interleukin (IL)-6 and IL-8 attract CTCs [274,275]. Exosomal miR-25-5p secreted from primary HCC cells stimulates the trans-endothelial mobility and tumor self-seeding potency of CTCs by downregulating the LRRC7 gene [258]. Another study found that exosome-derived miR-25-5p released from a highly metastatic HCC cell line (CSQT-2) increases the aggressiveness of cancer cells by inhibiting SIK1 [257], thereby activating the Wnt/β-catenin signaling pathway. Moreover, loss of liver glycine N-methyltransferase (GNMT) induces hepatic steatosis and disease progression in HCC [276]. Studies have shown that exosomal miR-224 drives tumor progression and invasiveness by directly targeting GNMT, and the serum expression of exosomal miR-224 is higher in HCC patients than in healthy individuals [261].

6. miRNAs in Tumor Microenvironment Remodelling

Several studies have demonstrated the importance of miRNAs in shaping the tumor microenvironment (TME) and promoting cancer progression (Figure 5) [277]. The following sections summarize the latest research on the role of miRNAs in the TME and provide insights into HCC management.

6.1. miRNAs Regulating Cancer-Associated Fibroblasts

Cancer-associated fibroblasts (CAFs) are activated fibroblasts with high expression of alpha-smooth muscle actin (αSMA) and fibroblast activation protein (FAP) and are located near neoplastic lesions [278,279]. CAFs secrete factors such as TGF-β, hepatocyte growth factor (HGF), stromal cell-derived factor 1 (SDF-1), and IL-1β to mediate immune suppression, extracellular matrix remodelling, maintenance of tumor stemness, angiogenesis, and chemoresistance [280].
In a hypoxic environment, tumor cells release miR-4508 within exosomes, which in turn activates pulmonary fibroblasts to yield a permissive pre-metastatic niche in the lungs of mice by targeting RFX1 [281]. Furthermore, knockdown of RFX1 induces an activated fibroblast phenotype in the pre-metastatic niche via its effect on the p38 MAPK-NF-kB signaling pathway [281]. In addition, miR-4508 is also reported to be elevated in plasma exosomes of patients with HCC after trans-arterial chemoembolization (TACE). Similarly, Fang et al. found a correlation between high serum levels of exosomal miR-1247-3p and lung metastasis in HCC patients [263]. Specifically, researchers found that in the metastatic lung niche, HCC cells secrete exosomal miR-1247-3p, which leads to the conversion of normal fibroblasts into CAFs through direct inhibition of B4GALT3 and activation of β1-integrin/NF-κB signaling in fibroblasts [263].
Zhang et al. found that miR-320a expression was reduced in CAF-derived exosomes compared to normal fibroblasts, and ectopic miR-320a expression impaired the migration, invasion, and EMT phenotype of cancer cells via suppression of the MAPK/ERK1/2-CDK2 axis and direct inhibition of PBX3 [177]. Moreover, reduced miR-150-3p levels have also been reported in CAF-derived exosomes, which upon transfer to HCC cells, inhibit their proliferation and metastatic capabilities [282]. Clinically, poor miR-150-3p expression in HCC tissues is a significant risk factor for recurrence [282].
Another study reported that HCC cell-derived exosomal miR-21 directly targets PTEN, leading to the activation of hepatic stellate cells (HSCs) into CAFs by triggering PDK1/AKT signaling and lipogenesis [259]. Consistently, high serum levels of exosomal miRNA-21 were associated with higher CAF activation and tumor vasculature density in patients with HCC [259]. The expression of miR-101-3p and miR-490-3p have been reported to be downregulated in HCC CAFs [283]. TGFBR1 is a common target gene of these miRNAs, and its expression is positively correlated with the infiltration of myeloid-derived suppressor cells, regulatory T cells, and M2 macrophages, which create an immunosuppressive TME and facilitate HCC progression [283].

6.2. miRNAs in Regulating Tumor-Associated Macrophages

Tumor-associated macrophages (TAMs) are an M2-polarized subtype of macrophages that are considered key factors in the development of an immunosuppressive TME that supports HCC progression and metastasis [284]. Therapeutic strategies targeting TAMs in animal models of HCC have efficiently attenuated tumor growth [284,285]. Several studies have highlighted the important roles of specific exosomal miRNAs in modulating TAM infiltration and influencing HCC aggressiveness and metastasis (Figure 5).
Nakano et al. demonstrated that in patients with post-transplant HCC recurrence, high exosomal miR-4669 levels contribute to an immunosuppressive TME by inducing M2 macrophage polarization, which enhances tumor aggressiveness [286]. Similarly, exosomal miR-452-5p secreted by HCC cells was found to promote M2 macrophage polarization by targeting TIMP3, thereby contributing to HCC progression and metastasis [287].
In contrast, the miR-144/miR-451a cluster promotes M1 macrophage polarization and antitumor activity in HCC by targeting HGF and MIF [288].
Furthermore, Chen et al. found that deficiency of miR-125a and miR-125b in TAM-derived exosomes promoted the proliferation, stem cell properties, and metastatic capacity of HCC cells, and re-expression of miR-125a/b in HCC cells suppressed the growth and sphere formation ability of liver cancer cells by targeting CD90 [163].
Importantly, Ke et al. showed that a lack of miR-148b leads to the overexpression of CSF1, which induces TAM infiltration and promotes HCC metastasis through CSF1/CSF1R signaling [289]. Moreover, Zhou et al. demonstrated that a reduction in miR-28-5p levels increases HCC growth and metastasis via IL-34-mediated TAM infiltration [290].

6.3. miRNAs in Regulating Natural Killer Cells, T Cells, and Dendritic Cells

Immune checkpoints (e.g., programmed death protein 1 (PD-1)) play a crucial role in regulating immune responses, and miRNAs can be used to target and modulate the expression of immune checkpoints on the surface of natural killer (NK) cells and T cells or that of their ligands on cancer cells, leading to potential applications in antitumor immunotherapy [291]. PD-1 is a cell surface receptor found on various immune cells that promotes self-tolerance and downregulates the immune response. The PD-1 ligand, PD-L1, is highly expressed in cancer cells and can inhibit the proliferation of immune cells, leading to immune tolerance within the tumor microenvironment [292].
In HCC models, miR-223 was reported to exert an indirect downregulatory effect on PD-1 and PD-L1 through the suppression of the HIF-1α-driven CD39/CD73-adenosine pathway, thereby impeding HCC progression [293]. Wang et al. reported that KDM1A interacts with MEF2D to increase PD-L1 expression in HCC cells, and miR-329-3p targeting of KDM1A inhibits tumor-induced immunosuppression and sensitizes HCC cells to T cell-mediated cytotoxicity, by modulating PD-L1 expression [294].
Sun et al. found that miR-200c and PD-L1 expression were inversely correlated in HBV-induced HCC, as miR-200c directly targets the 3ʹ-UTR of PD-L1 which leads to a reversal of CD8+ T cell exhaustion [295]. However, HBV-induced STAT3 activation triggers SALL4 expression, which in turn suppresses miR-200c transcription, suggesting that the HBV-pSTAT3-SALL4-miR-200c axis regulates PD-L1 expression [295].
Wan et al. recently highlighted the potential of miR-22 as a therapeutic target for modulating immune responses in HCC [296]. miR-22 delivery via adeno-associated virus serotype 8 significantly affected the function of hepatocytes and T cells by silencing HIF-1α and enhancing retinoic acid signaling in both cell types. miR-22 was observed to decrease the abundance of IL17-producing T cells and inhibit IL17 signaling in the liver, thereby promoting the expansion of cytotoxic T cells and reducing the population of regulatory T cells [296].
CXCL10 plays a crucial role in CD8 + T and NK cell trafficking [297]. According to experimental evidence, DPP4 is involved in the N-terminal truncation of CXCL10, which limits T-cell and NK cell migration [298]. Huang et al. reported that the miR-30-5p/Snail/DPP4/CXCL10 axis influences the HCC-immune microenvironment by enhancing the stability of CXCL10 and improving CD8+ T-cell infiltration [299].
Chen et al. showed that CX3CL1 stimulates the chemotactic migration of CX3CR1+ NK cells through STAT3 signaling, and miR-561-5p promotes tumor growth and lung metastasis by suppressing CX3CL1 expression, leading to low NK cell infiltration and poor prognosis in patients with HCC [4].
In a HCC mouse model, miR-1258 inhibited tumor development and metastasis by stimulating TLR7/8 expression and inducing the antitumor activity of dendritic cells (DCs) and NK cells [300].
Xie et al. demonstrated that the expression of MICB in tumor cells was negatively correlated with miR-889 in HCC tissues and that miR-889 upregulation reduced HCC cell susceptibility to NK lysis [301].
Table 4 presents a comprehensive list of various miRNAs involved in the modulation of the tumor microenvironment in HCC.

6.4. miRNAs Involved in Tumor Angiogenesis Regulation

Tumor neovascularization, also referred to as tumor angiogenesis, is a sophisticated biological process that develops new blood vessels within the tumor microenvironment [314], and serves as an essential step for cancer progression and metastasis. As a pathological condition, an imbalance between angiogenic stimulators and inhibitors results in robust neovascularization in tumors [315]. Several studies have explored angiogenesis-related miRNAs in various human carcinomas (Figure 5).

6.4.1. Anti-Angiogenic miRNAs in Hepatocellular Carcinoma

The interaction between HCC cells and VECs is fundamental for the construction of tumor blood vessels. Research has identified that miR-199a-3p disrupts this interaction by directly inhibiting VEGFA expression in HCC cells and decreasing VEGFR1 and VEGFR2 expression in VECs [316]. HOXA3, HOXB3, and HOXD3 control angiogenesis, and HOXD3 directly targets the VEGFR promoter region [317]. miR-203a suppresses HCC cell invasion, metastasis, and angiogenesis by inhibiting the VEGFR pathway and targeting HOXD3 expression [318]. The PI3K/AKT pathway and its downstream mediator SGK3 facilitate tumor angiogenesis [319]. Wu et al. showed that miR-144-3p is a critical regulator of this pathway by reducing SGK3 expression and downregulating VEGF2, thereby inhibiting the angiogenic capabilities of HCC cells [320].
The ERG plays an essential role in endothelial differentiation and angiogenesis [321]. Hepatocytes release exosomal miR-200b-3p, which attenuates angiogenesis by blocking endothelial ERG expression [322]. Furthermore, downregulation of miR-3064-5p correlates with high angiogenesis potential in HCC tissues by inhibiting VEGFA and angiogenin while inducing endostatin and MMP12 expression through the FOXA1/CD24/Src pathway [323]. This relationship underscores the intricate balance of pro- and anti-angiogenic factors governed by microRNAs in the tumor microenvironment.

6.4.2. Pro-Angiogenic miRNAs in Hepatocellular Carcinoma

Researchers have discovered several pro-angiogenic miRNAs that play roles in various cancers by specifically regulating the VEGF/HIF-1 pathway [324]. One of the key factors in this process is tumor hypoxia, which activates the hypoxia-inducible factors HIF-1 and HIF-2, which in turn trigger the transcription of genes involved in angiogenesis [325]. For instance, in HCC cells under hypoxic conditions, both cellular and exosomal miR-155 were found to be overexpressed and to have the ability to induce tube formation in human umbilical vein endothelial cells [326]. Clinical investigations have also demonstrated a positive correlation between miR-155, HIF-1α, and VEGF expression in HCC tissue [326].
Conversely, there are genes and factors with anti-angiogenic properties that promote vascular stability. HOXA5 inhibits angiogenesis by increasing the expression of anti-angiogenic factors (e.g., p53) and decreasing the expression of pro-angiogenic factors (e.g., HIF-1α and VEGFR2), thereby promoting vascular stability [327]. Studies have shown that miR-130b-3p is a pro-angiogenic miRNA as it can suppress HOXA5 expression, leading to enhanced angiogenic capacity in HCC cells [328].
Smad4 is another important factor in the regulation of angiogenesis, exerting its anti-angiogenic properties via the suppression of VEGF expression and upregulation of the angiogenesis inhibitor TSP-1 [329]. Lin et al. reported that exosomal miR-210-3p facilitates intercellular communication between HCC and endothelial cells by inhibiting Smad4 and STAT6 in vascular endothelial cells, thereby promoting angiogenesis [330].
Table 5 provides an overview of angiogenesis-related miRNAs in HCC.

7. miRNAs in Hepatocellular Carcinoma Drug Resistance

Tumor heterogeneity and poor survival rates in HCC pose significant clinical challenges. As chronic liver disease progresses to HCC, changes in TME can have a major impact on drug metabolism and response to therapy [335]. Drug resistance remains a major obstacle in HCC management, stemming from multiple factors, including enhanced drug efflux, reduced drug uptake, intracellular detention, high drug metabolism, aberrant apoptotic and autophagic signaling, TME remodelling, and the acquisition of stem cell-like features [336]. Recently, researchers have focused on miRNAs and their roles in the development of chemoresistance in HCC (Figure 6) [337]. This section reviews the current understanding of the role of miRNAs in drug sensitivity in HCC and examines the potential mechanisms and clinical implications.

7.1. miRNAs in Chemotherapy Response

The most common chemotherapy intervention for unresectable HCC is transcatheter arterial chemoembolization (TACE) using doxorubicin, cisplatin, and 5-fluorouracil (5-FU). miRNAs can modulate drug responses through various mechanisms, including the regulation of autophagy, membrane transporters, EMT, CSCs, TME, and metabolic reprogramming (Figure 6) [337].

7.1.1. miRNAs Promoting Hepatocellular Carcinoma Cells Sensitivity to Chemotherapy

In a study involving four distinct HCC cohorts, researchers discovered that the miR-125b/HIF-1α axis plays a critical role in determining HCC cell sensitivity to TACE therapy [338]. This study revealed that miR-125b directly reduces the translation of HIF-1α and disrupts the autocrine HIF-1α/PDGFβ/pAkt/HIF-1α loop by targeting the PDGFβ receptor. Moreover, the loss of miR-125b and increased HIF-1α expression upregulate CD24 and erythropoietin, resulting in the enrichment of doxorubicin-resistant CD24+ cancer stem cell population [338].
Other studies have demonstrated that miR-26a/b and miR-223 suppress doxorubicin-induced autophagic flux by targeting ULK1 and FOXO3a, respectively, resulting in an improved sensitivity to doxorubicin and increased apoptosis [100,339].
Additionally, the HIF-2α-MALAT1-miR-216b axis regulates multidrug resistance in HCC cells by modulating the expression of the autophagosome marker LC3-II and autophagy [340]. MALAT1 is an oncogenic long non-coding RNA whose expression is induced by HIF-2α, which then downregulates miR-216b in HCC cells. MALAT1 siRNA and miR-216b mimics inhibit LC3-II levels and autophagy, while enhancing 5-FU-induced apoptosis [340].
Moreover, miR-361-5p and miR-610 have been shown to enhance cisplatin sensitivity by reducing the expression of MAP3K9 and HDGF in HCC cells [341,342]. The RNA-binding protein MSI1 is upregulated in malignancies and modulates cancer cell proliferation by influencing the Notch, Wnt, and Akt signaling pathways [343]. Interestingly, miR-10a-5p has been found to promote cisplatin sensitivity by downregulating MSI1 and impairing the Akt signaling pathway in HCC [344].

7.1.2. miRNAs Promoting Hepatocellular Carcinoma Cells Resistance to Chemotherapy

Myocyte-specific factor 2C (MEF2C) plays an essential role in regulating cell differentiation, stemness, proliferation, and migration [345,346]. In a study by Kang et al., cells expressing miR-551a displayed resistance to 5-FU induced cell death and exhibited enhanced survival and sphere formation after 5-FU treatment, with these effects attributable to the miR-551a targeting of MEF2C [347].
Furthermore, miR-182 levels were significantly increased in HCC patients undergoing cisplatin-based chemotherapy, and the upregulation of miR-182 enhanced cell viability during cisplatin treatment by targeting TP53INP1 [348].
Table 6 outlines the various miRNAs that play a role in the response of HCC to chemotherapy.

7.2. miRNAs in Targeted Therapy Response

In recent years, there has been a growing interest among researchers in studying the influence of miRNA dysregulation on the efficacy of targeted therapies, such as sorafenib and lenvatinib, in the context of HCC. Sorafenib, a first-line targeted therapy approved for advanced HCC treatment (since 2008) [365], functions by suppressing EGFR 1–3 and PDGFR-II, acting as an anti-angiogenic agent [366]. However, sorafenib only provides a modest improvement in patient survival of approximately three months, primarily owing to the development of resistance [367]. Another targeted therapeutic agent, lenvatinib, was approved by the FDA in 2018 [368]. It is an oral multikinase inhibitor that effectively blocks various receptors, including VEGFRs, FGFRs, RET, C-kit, PDGFR-α, and PDGFR-β, and inhibits downstream signaling pathways involved in tumor angiogenesis and cancer cell proliferation [369].
The mechanisms underlying sorafenib and lenvatinib resistance in HCC are not yet fully understood. An in-depth understanding of miRNAs and their impact on the development of targeted therapy resistance is important for the development of innovative therapeutic strategies for HCC. Targeting these miRNAs or their downstream signaling pathways could potentially restore the sensitivity of targeted therapy and enhance treatment efficacy [370].

7.2.1. miRNAs Improving Sorafenib Response

Sorafenib resistance in hepatocellular carcinoma (HCC) involves a complex interplay of molecular mechanisms. Kabir et al. [371] identified a miR-7/TYRO3 axis that regulates the growth and invasiveness of sorafenib-resistant HCC cells, suggesting a potential therapeutic role for miR-7. Several other miRNAs have been identified as potential enhancers of the efficacy of sorafenib treatment in HCC cells. For example, miR-449a-5p improves sorafenib-induced apoptosis and reduces angiogenesis in HCC cells by downregulating PEA15, PPP1CA, and TUFT1, and modulating the Akt/ERK signaling pathway [372]. Another miRNA, miR-4277, sensitizes HCC cells to sorafenib treatment by targeting CYP3A4 and reducing its metabolism and clearance [373].
Autophagy, a major mechanism underlying acquired sorafenib resistance, can be induced by sorafenib treatment via regulation of various targets [374]. Importantly, restoration of miR-30a-5p by hydroxychloroquine treatment has been found to improve sorafenib response by suppressing autophagy through ATG5 and Beclin-1 [375]. Furthermore, the miR-30a-5p/CLCF1 axis has been implicated in regulating sorafenib resistance by directly targeting CLCF1, modulating PI3K/Akt signaling, and attenuating aerobic glycolysis in sorafenib-resistant HCC cells [376].
miR-204 has also been identified as a regulator of autophagy and the sorafenib response by suppressing ATG3 in HCC [377]. Moreover, ectopic miR-142-3p overexpression sensitizes HCC cells to sorafenib by decreasing sorafenib-induced autophagy and enhancing sorafenib-induced apoptosis by targeting ATG5 and ATG16L1 [378].
miR-375 also plays a pivotal role in reducing the resistance of HCC cells to sorafenib by inhibiting autophagy and exerting anti-angiogenic effects through targeted regulation of SIRT5 and PDGFC [332,379]. Additionally, sorafenib administration has been shown to induce miR-375 expression in hepatoma cells via the transcription factor ASH1. The expression of miR-375 is decreased in sorafenib-resistant cells, but restoring its levels can partially re-sensitize cells to sorafenib, primarily through the degradation of the AEG-1 gene [332].
In a study by Lin et al., liver X receptor-α activation with its agonist (GW3965) was found to induce the transcription of miR-378a-3p, which improves sorafenib response by regulating IGF1R expression and inhibiting ERK/PI3K signaling [380].
The pregnane X receptor (PXR), a member of the nuclear receptor superfamily, regulates genes involved in xenobiotic metabolism and drug resistance, such as CYP450 and ATP-binding cassette transporters [381], and studies have shown that sorafenib therapy triggers PXR expression in HCC cells, resulting in increased drug clearance and reduced sensitivity [382]. However, miR-140-3p and miR-148a can target PXR expression, enhance sorafenib retention in HCC cells, and restore sensitivity [383,384].
Another study identified FOXO3a, a crucial transcription factor in the cellular stress response, as a significant player in miR-124-3p.1 mediated sorafenib efficacy in HCC cells [385]. miR-124-3p.1 enhances sorafenib-induced apoptosis by increasing the nuclear localization of FOXO3a and maintaining its dephosphorylation and acetylation through the targeting of upstream regulators Akt2 and SIRT1 [385]. Additionally, Pei et al. found that PAK5 phosphorylates β-catenin, facilitating its translocation to the nucleus and promoting ABCB1 transcription. Importantly, miR-138-1-3p inhibits PAK5, reduces β-catenin/ABCB1 signaling, and improves sorafenib response [386]. ADAM-17, which is involved in the processing and activation of Notch proteins [387], is targeted by miR-3163, which enhances sorafenib sensitivity by suppressing Notch signaling activation [386]. Moreover, miR-345-5p targets TOP2A mRNA and improves the effects of sorafenib treatment in HCC by inducing apoptosis [388]. In a separate discovery, miR-3689a-3p emerged in CRISPR/Cas9 screening as a key miRNA that increased sorafenib sensitivity in HCC cells. miR-3689a-3p targets CCS, disrupting the copper balance and SOD1 function, ultimately leading to sorafenib-induced HCC cell death [389].

7.2.2. miRNAs Inducing Sorafenib Resistance

Lu et al. discovered that miR-23a-3p is highly expressed in sorafenib-non-responder HCC patients [390], and further proteomic analysis revealed that miR-23a-3p blocks ferroptosis by suppressing the expression of ACSL4, an essential enzyme for ferroptosis [391]. In another study, miR-125b-5p induced sorafenib resistance by promoting Snail-mediated EMT via ATXN1 inhibition [392].
Pollutri et al. found that miR-494 expression was significantly upregulated in HCCs with stem cell-like characteristics in both humans and rats, and high miR-494 levels were linked to a poor sorafenib response [52]. miR-494’s oncogenic function is attributed to its targeting of p27, PUMA, and PTEN genes, as well as its activation of the mTOR pathway in HCC cell lines [52]. Furthermore, miR-494 contributes to metabolic plasticity and survival of HCC cells by promoting glycolysis by targeting G6pc and activating the HIF-1α pathway [393]. Elevated serum levels of miR-494 have been linked to sorafenib resistance in preclinical models and patients with HCC. Furthermore, the use of combination therapy with antagomiR-494 and sorafenib or 2-deoxy-glucose has demonstrated improved anticancer effects in HCC cells [393].
A recent study found that high expression of miR-21-5p and USP24 is associated with cancer progression and drug resistance in HCC by promoting autophagy via USP24-mediated SIRT7 ubiquitination [394]. Interestingly, inhibition of miR-21-5p led to a decrease in SIRT7 ubiquitination, the LC3II/I ratio, the expression of Beclin1, and an increase in p62 levels. These molecular changes collectively contribute to improved sensitivity to sorafenib [394].

7.2.3. miRNAs Improving Lenvatinib Response

Several miRNAs have been identified as regulators of lenvatinib sensitivity in HCC. miR-128-3p/c-Met axis is involved in the mechanisms underlying lenvatinib resistance by regulating Akt and ERK, which are involved in cell cycle progression and apoptosis, respectively [113]. Wang et al. reported that miR-24-3p suppresses the expression of the anti-apoptotic protein BCL2L2, thereby improving HCC cell sensitivity to lenvatinib treatment [395]. Additionally, miR-34a inhibits autophagy by targeting Beclin-1 in HCC cells, which in turn enhances lenvatinib sensitivity [396].

7.2.4. miRNAs Inducing Lenvatinib Resistance

Several miRNAs have been identified as key players in lenvatinib resistance in HCC. miR-183-5p.1 has been shown to be upregulated in liver tumor-initiating cells (T-ICs), and the overexpression of miR-183-5p.1 promotes self-renewal and tumorigenesis by targeting the MUC15/c-MET/PI3K/Akt/SOX2 axis, thereby reducing lenvatinib sensitivity [397]. Another miRNA implicated in reducing lenvatinib efficacy is miR-520c-3p via its targeting of MBD2, leading to increased FGFR4 expression [398].
Additionally, miR-3154 functions as an oncomiR, is elevated in both HCC and liver cancer stem cells, and plays a significant role in HCC progression by targeting HNF4α and promoting self-renewal, proliferation, metastasis, and tumorigenesis [399].Increased miR-3154 expression was observed in lenvatinib-resistant HCC cell lines, and lenvatinib sensitivity was improved following miR-3154 knockdown. Therefore, miR-3154 may serve as a predictive marker for HCC patient response to lenvatinib treatment, as confirmed through cohort and xenograft analyses, thereby providing valuable insights into the potential clinical benefits of targeting miR-3154 in combination with lenvatinib [399]. Table 7 presents an overview of miRNAs that affect the treatment efficacy of sorafenib or lenvatinib in HCC cells.

8. miRNA-Based Therapeutics for Hepatocellular Carcinoma Therapy

miRNA-based therapeutic approaches for HCC treatment offer a dual-pronged strategy: miRNA replacement therapy [419], which introduces tumor-suppressive miRNAs to restore normal function and inhibit tumor growth, and miRNA antagonism [420], which inhibits oncogenic miRNAs to mitigate their adverse effects. These miRNA-targeted strategies are promising because of their ability to simultaneously regulate multiple genes and pathways that are often dysregulated in HCC, offering a more comprehensive approach to cancer therapy than drugs targeting single molecules [421]. To improve in vivo delivery of miRNA-based therapeutics, various strategies can be employed to enhance their stability, targeting efficiency, and cellular uptake while avoiding immunogenic responses. Some promising approaches include nanoparticle-based delivery systems, exosome-mediated delivery, chemical modification, and local delivery (Figure 7) [422].

8.1. Restoring Tumor Suppressive miRNA Function

Numerous strategies have been explored in the pursuit of achieving the goal of restoring tumor repressor miRNAs back to the tumor to reduce tumorigenesis. One such approach involves utilizing synthetic miRNA mimics, which are chemically modified versions of specific miRNAs, and these synthetic molecules can be effectively delivered to cancer cells to replicate the actions of tumor-suppressive miRNAs [422]. For instance, miR-34a, which is downregulated in various cancer cells [423], has been the focus of extensive research. Notably, researchers have developed a liposome-formulated miR-34a mimic called MRX34, which has undergone clinical trials in HCC patients with advanced solid tumors [424]. Although trials were terminated owing to the occurrence of five immune-related serious adverse events involving the death of patients [425], the development of MRX34 demonstrated a feasible approach for miRNA drug discovery using liposomal nanoparticles.
In 2014, TargomiRs were introduced to the field of RNA therapeutics as non-living bacterial nanoparticles that serve as effective drug delivery vehicles in which various molecules, including nucleic acids, can be encapsulated [426]. EGFR-targeted TargomiRs loaded with miR-16 mimics have undergone phase I clinical trials in patients with recurrent malignant pleural mesothelioma and non-small cell lung cancer [86]. To avoid undesirable off-target effects, researchers opted to target EGFR using panitumumab because of its high expression in mesothelioma cells. Encouragingly, a phase I study yielded promising results with no reported adverse effects and has paved the way for a phase II clinical trial [86].
Another promising miRNA of interest is miR-193a-3p, which is known to suppress the growth of various cancer types including HCC [427]. To harness this potential, researchers have developed INT-1B3, a lipid nanoparticle-formulated miR-193a-3p mimic [428]. INT-1B3 upregulates the tumor-suppressive PTEN pathway and downregulates several oncogenic pathways in cancer cells [429]. Preclinical studies have demonstrated the safe and effective delivery of INT-1B3 to tumors in vivo [428]. Currently, INT-1B3 is undergoing a phase 1 clinical trial to evaluate its safety, tolerability, pharmacokinetics, pharmacodynamics, and antitumor activity in patients with various solid cancers [421].
Moreover, miR-122, which is frequently downregulated in HCC, has been extensively studied for its tumor suppressive properties. Multiple studies have demonstrated that restoring miR-122 expression inhibits tumor growth, angiogenesis, and metastasis in HCC models [430,431,432]. Hsu et al. successfully delivered miR-122 to HCC cells using cationic lipid nanoparticles consisting of 2-dioleyloxy-N,N-dimethyl-3-aminopropane (DODMA), egg phosphatidylcholine, cholesterol, and cholesterol-polyethylene glycol (LNP-DP1) [433]. This delivery method significantly reduced the expression of miR-122 target genes.
Furthermore, miR-26a plays a crucial role in liver cancer development. Although it is highly expressed in the adult liver, its expression is significantly decreased in human and mouse liver tumors. When miR-26a is delivered to tumors in animal models of liver cancer using adeno-associated virus (AAV), it suppresses cancer cell growth and promotes apoptosis in vivo, resulting in a significant reduction in tumor growth [434].

8.2. Inhibition of Oncogenic miRNA Function

Numerous studies have highlighted the potential use of anti-miRs in cancer treatment. One specific example is an LNA-based antagomiR designed to target miR-155, namely Cobomarsen (MRG-106), which has been used in clinical trials in patients with various types of lymphoma e.g., mycosis fungoides, cutaneous T-cell lymphoma, diffuse large B-cell lymphoma, and adult T-cell leukemia/lymphoma. The initial phase I trial demonstrated favorable tolerability and efficacy [435]; however, the subsequent phase II trial was prematurely terminated owing to factors associated with business operations and participant considerations [436]. In another study, nanoparticles were used to simultaneously target miR-21 and miR-155 in lymphoma cell lines. They developed polylactic-co-glycolic acid (PLGA) nanoparticles capable of delivering different classes of anti-miRs, including phosphorothioates (PS) and peptide nucleic acids (PNAs). This innovative approach resulted in significant reductions in the levels of miR-21 and miR-155 and their downstream target genes, and a decrease in cancer cell viability, indicating promising therapeutic potential [437].
A study conducted by Lee et al. focused on the development of lactosylated gramicidin-containing lipid nanoparticles (Lac-GLN) for targeted delivery of anti-miR-155 to HCC cells, resulting in increased expression of miR-155 target genes, such as C/EBPβ and FOXP3. This formulation also exhibited preferential accumulation in mouse hepatocytes [438]. Liang et al. devised a nanoplatform called PCACP (PEI-βCD@Ad-CDM-PEG) to facilitate the delivery of miRNAs. This platform was used to create a miRNA cocktail therapy by encapsulating miR-199a/b-3p mimics and antimiR-10b within PCACP, specifically targeting HCC. This therapy effectively inhibits the proliferation of HCC cells and suppresses tumor growth by targeting specific pathways involved in cancer progression, including mTOR, PAK4, RHOC, and EMT. Notably, in patient-derived xenografts (PDXs), the personalized PCACP/miR-cocktail system showed significant tumor suppression and multi-target regulation, indicating potential improvements over conventional therapies [439].
A promising miRNA inhibitor for the treatment of pancreatic cancer is TTX-MC138, which is specifically designed to suppress the overexpression of miR-10b, an oncogenic miRNA [440]. TTX-MC138 utilizes advanced dextran-coated iron-oxide nanoparticles for enhanced stability and targeted delivery to cancer cells. Preclinical studies have yielded promising results, and further investigation is currently underway [421].
Another avenue of research involves the use of exosomes for the targeted delivery of miRNA-based therapies. In a particular study, researchers explored the co-delivery of 5-FU and an miR-21 inhibitor to colorectal cancer cells using exosomes, where there was effective downregulation of miR-21 expression, resulting in cell cycle arrest, reduced proliferation, increased apoptosis, restoration of tumor suppressor gene expression, and reversal of drug resistance [441].
In addition to targeting oncogenic miRNAs, scientists have developed inhibitors, such as miravirsen and RG101, to combat viral infections. These inhibitors disrupt the activity of miR-122, a liver-specific miRNA that is crucial for hepatitis C virus (HCV) replication [442]. Miravirsen, a 15-mer LNA-PS-modified antisense oligonucleotide (ASO), has been used in clinical trials as a targeted therapy for HCV infections [443]. Similarly, RG101, an N-acetylgalactosamine (GalNAc)-conjugated anti-miR-122 oligonucleotide [444], has shown promise in reducing the viral load in patients with chronic HCV and is currently being tested in a Phase II clinical trial [445]. These inhibitors are potential treatment options for patients that do not respond to traditional therapies. Table 8 presents a compilation of miRNAs that have been studied for targeted delivery in both in vivo experiments and clinical trials, highlighting the diversity of the ongoing research in this field.

9. Conclusions and Future Perspectives

HCC represents a significant global health burden owing to its high prevalence and recurrence rates, despite the availability of current therapies. Recent advancements in miRNA sequencing have greatly expanded our understanding of the involvement of miRNAs in hepatocarcinogenesis and drug response. These non-coding RNAs play a crucial role in regulating target genes and signaling pathways, acting as either tumor suppressors or oncogenes, thereby influencing various cancer hallmark traits. Harnessing the potential of miRNAs in HCC therapy offers an innovative approach for molecular cancer treatment. Moreover, serum exosomal miRNAs have emerged as potential biomarkers for monitoring treatment efficacy and improving patient outcome [449]. However, translating miRNA research into clinical application has several critical barriers that need to be addressed. One major challenge is achieving precise modulation of miRNAs in vivo, considering the complex dosage requirements and potential for unforeseen side effects owing to the multitude of gene targets regulated by miRNAs. Additionally, ensuring the stability of miRNA molecules in the bloodstream and their targeted delivery specifically to cancer cells while avoiding adverse effects on normal cells adds further complexity to therapeutic development. To overcome these obstacles, ongoing research has focused on the development of advanced delivery systems such as lipid-based nanoparticles or molecular conjugates to improve the stability and specificity of miRNA-based therapies. Furthermore, identification of key miRNAs that play pivotal roles in liver cancer could refine therapeutic targets, enhance treatment efficacy, and minimize potential side effects.

Author Contributions

Conceptualization, K.M. and T.D.K.; writing, K.M.; review and editing, K.M., T.D.K., D.J.B. and P.J.L.; visualization, K.M.; supervision, P.J.L., G.C.Y. and T.D.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was in part funded by a grant from the Cancer Council NSW RG21-04.

Acknowledgments

We would like to express our sincere gratitude to the funders of our research and to everyone who contributed to the completion of this manuscript, including those who provided insightful feedback.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. 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] [PubMed]
  2. Chan, L.K.; Tsui, Y.M.; Ho, D.W.; Ng, I.O. Cellular heterogeneity and plasticity in liver cancer. Semin. Cancer Biol. 2021, 82, 134–149. [Google Scholar] [CrossRef] [PubMed]
  3. Australian Institute of Health and Welfare. Cancer in Australia 2019 (Cancer Series No. 119; Cat. No. CAN 123); AIHW: Canberra, Australia, 2019. [Google Scholar]
  4. Chen, E.-B.; Zhou, Z.-J.; Xiao, K.; Zhu, G.-Q.; Yang, Y.; Wang, B.; Zhou, S.-L.; Chen, Q.; Yin, D.; Wang, Z. The miR-561-5p/CX3CL1 signaling axis regulates pulmonary metastasis in hepatocellular carcinoma involving CX3CR1+ natural killer cells infiltration. Theranostics 2019, 9, 4779. [Google Scholar] [CrossRef]
  5. Baecker, A.; Liu, X.; La Vecchia, C.; Zhang, Z.-F. Worldwide incident hepatocellular carcinoma cases attributable to major risk factors. Eur. J. Cancer Prev. Off. J. Eur. Cancer Prev. Organ. ECP 2018, 27, 205. [Google Scholar] [CrossRef]
  6. Kanwal, F.; Kramer, J.; Asch, S.M.; Chayanupatkul, M.; Cao, Y.; El-Serag, H.B. Risk of hepatocellular cancer in HCV patients treated with direct-acting antiviral agents. Gastroenterology 2017, 153, 996–1005.e1001. [Google Scholar] [CrossRef]
  7. Tian, Z.; Xu, C.; Yang, P.; Lin, Z.; Wu, W.; Zhang, W.; Ding, J.; Ding, R.; Zhang, X.; Dou, K. Molecular pathogenesis: Connections between viral hepatitis-induced and non-alcoholic steatohepatitis-induced hepatocellular carcinoma. Front. Immunol. 2022, 13, 984728. [Google Scholar] [CrossRef] [PubMed]
  8. Hoshida, Y.; Nijman, S.M.B.; Kobayashi, M.; Chan, J.A.; Brunet, J.-P.; Chiang, D.Y.; Villanueva, A.; Newell, P.; Ikeda, K.; Hashimoto, M. Integrative transcriptome analysis reveals common molecular subclasses of human hepatocellular carcinoma. Cancer Res. 2009, 69, 7385–7392. [Google Scholar] [CrossRef] [PubMed]
  9. Dhanasekaran, R.; Bandoh, S.; Roberts, L.R. Molecular pathogenesis of hepatocellular carcinoma and impact of therapeutic advances. F1000Research 2016, 5, 879. [Google Scholar] [CrossRef]
  10. Dariya, B.; Peela, S.; Nagaraju, G.P.; Nagaraju, G.P.; Peela, S. Chapter 1—Cell origin, biology, and pathophysiology of hepatocellular carcinoma. In Theranostics and Precision Medicine for the Management of Hepatocellular Carcinoma; Academic Press: Cambridge, MA, USA, 2022; pp. 1–9. [Google Scholar] [CrossRef]
  11. 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]
  12. Bartel, D.P. MicroRNAs: Genomics, biogenesis, mechanism, and function. Cell 2004, 116, 281–297. [Google Scholar] [CrossRef]
  13. Borchert, G.M.; Lanier, W.; Davidson, B.L. RNA polymerase III transcribes human microRNAs. Nat. Struct. Mol. Biol. 2006, 13, 1097–1101. [Google Scholar] [CrossRef] [PubMed]
  14. Lee, Y.; Kim, M.; Han, J.; Yeom, K.H.; Lee, S.; Baek, S.H.; Kim, V.N. MicroRNA genes are transcribed by RNA polymerase II. EMBO J. 2004, 23, 4051–4060. [Google Scholar] [CrossRef] [PubMed]
  15. Macfarlane, L.A.; Murphy, P.R. MicroRNA: Biogenesis, Function and Role in Cancer. Curr. Genom. 2010, 11, 537–561. [Google Scholar] [CrossRef] [PubMed]
  16. Lund, E.; Dahlberg, J.E. Substrate selectivity of exportin 5 and Dicer in the biogenesis of microRNAs. Cold Spring Harb. Symp. Quant. Biol. 2006, 71, 59–66. [Google Scholar] [CrossRef]
  17. Pistritto, G.; Trisciuoglio, D.; Ceci, C.; Garufi, A.; D’Orazi, G. Apoptosis as anticancer mechanism: Function and dysfunction of its modulators and targeted therapeutic strategies. Aging 2016, 8, 603. [Google Scholar] [CrossRef]
  18. Bartel, D.P. MicroRNAs: Target recognition and regulatory functions. Cell 2009, 136, 215–233. [Google Scholar] [CrossRef]
  19. Guo, H.; Ingolia, N.T.; Weissman, J.S.; Bartel, D.P. Mammalian microRNAs predominantly act to decrease target mRNA levels. Nature 2010, 466, 835–840. [Google Scholar] [CrossRef] [PubMed]
  20. Di Leva, G.; Garofalo, M.; Croce, C.M. MicroRNAs in cancer. Annu. Rev. Pathol. Mech. Dis. 2014, 9, 287–314. [Google Scholar] [CrossRef]
  21. Peng, Y.; Croce, C.M. The role of MicroRNAs in human cancer. Signal Transduct. Target. Ther. 2016, 1, 15004. [Google Scholar] [CrossRef]
  22. Siddhartha, R.; Garg, M. MicroRNAs in Cancer: Diagnostics and Therapeutics. In Handbook of Oncobiology: From Basic to Clinical Sciences; Springer Nature: Berlin/Heidelberg, Germany, 2024; p. 819. [Google Scholar]
  23. Mizuguchi, Y.; Takizawa, T.; Yoshida, H.; Uchida, E. Dysregulated miRNA in progression of hepatocellular carcinoma: A systematic review. Hepatol. Res. 2016, 46, 391–406. [Google Scholar] [CrossRef]
  24. Wei, L.; Wang, X.; Lv, L.; Liu, J.; Xing, H.; Song, Y.; Xie, M.; Lei, T.; Zhang, N.; Yang, M. The emerging role of microRNAs and long noncoding RNAs in drug resistance of hepatocellular carcinoma. Mol. Cancer 2019, 18, 147. [Google Scholar] [CrossRef]
  25. Ramakrishnan, K.; Vishwakarma, R.; Dev, R.R.; Raju, R.; Rehman, N. Etiologically Significant microRNAs in Hepatitis B Virus-Induced Hepatocellular Carcinoma. OMICS J. Integr. Biol. 2024, 28, 280–290. [Google Scholar] [CrossRef] [PubMed]
  26. Tavakoli Pirzaman, A.; Alishah, A.; Babajani, B.; Ebrahimi, P.; Sheikhi, S.A.; Moosaei, F.; Salarfar, A.; Doostmohamadian, S.; Kazemi, S. The role of microRNAs in hepatocellular cancer: A narrative review focused on tumor microenvironment and drug resistance. Technol. Cancer Res. Treat. 2024, 23, 15330338241239188. [Google Scholar] [CrossRef] [PubMed]
  27. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef]
  28. Sinha, S.; Farfel, A.; Luker, K.E.; Parker, B.A.; Yeung, K.T.; Luker, G.D.; Ghosh, P. Growth signaling autonomy in circulating tumor cells aids metastatic seeding. Proc. Natl. Acad. Sci. USA Nexus 2024, 3, pgae014. [Google Scholar] [CrossRef] [PubMed]
  29. Hwang, H.W.; Mendell, J.T. MicroRNAs in cell proliferation, cell death, and tumorigenesis. Br. J. Cancer 2006, 94, 776–780. [Google Scholar] [CrossRef] [PubMed]
  30. Duronio, R.J.; Xiong, Y. Signaling pathways that control cell proliferation. Cold Spring Harb. Perspect. Biol. 2013, 5, a008904. [Google Scholar] [CrossRef]
  31. Bueno, M.J.; Malumbres, M. MicroRNAs and the cell cycle. Biochim. Biophys. Acta (BBA) Mol. Basis Dis. 2011, 1812, 592–601. [Google Scholar] [CrossRef]
  32. Xu, X.; Zou, H.; Luo, L.; Wang, X.; Wang, G. MicroRNA-9 exerts antitumor effects on hepatocellular carcinoma progression by targeting HMGA2. FEBS Open Bio 2019, 9, 1784–1797. [Google Scholar] [CrossRef]
  33. Zhao, Y.; Zhu, C.; Chang, Q.; Peng, P.; Yang, J.; Liu, C.; Liu, Y.; Chen, X.; Liu, Y.; Cheng, R. MiR-424-5p regulates cell cycle and inhibits proliferation of hepatocellular carcinoma cells by targeting E2F7. PLoS ONE 2020, 15, e0242179. [Google Scholar] [CrossRef]
  34. Zhang, Y.; You, W.; Zhou, H.; Chen, Z.; Han, G.; Zuo, X.; Zhang, L.; Wu, J.; Wang, X. Downregulated miR-621 promotes cell proliferation via targeting CAPRIN1 in hepatocellular carcinoma. Am. J. Cancer Res. 2018, 8, 2116. [Google Scholar]
  35. Hu, B.; Yang, X.B.; Yang, X.; Sang, X.T. LncRNA CYTOR affects the proliferation, cell cycle and apoptosis of hepatocellular carcinoma cells by regulating the miR-125b-5p/KIAA1522 axis. Aging 2020, 13, 2626–2639. [Google Scholar] [CrossRef]
  36. Hua, S.; Quan, Y.; Zhan, M.; Liao, H.; Li, Y.; Lu, L. miR-125b-5p inhibits cell proliferation, migration, and invasion in hepatocellular carcinoma via targeting TXNRD1. Cancer Cell Int. 2019, 19, 203. [Google Scholar] [CrossRef]
  37. Zhang, N.-S.; Dai, G.-L.; Liu, S.-J. MicroRNA-29 family functions as a tumor suppressor by targeting RPS15A and regulating cell cycle in hepatocellular carcinoma. Int. J. Clin. Exp. Pathol. 2017, 10, 8031. [Google Scholar]
  38. Chen, Z.; Zhuang, W.; Wang, Z.; Xiao, W.; Don, W.; Li, X.; Chen, X. MicroRNA-450b-3p inhibits cell growth by targeting phosphoglycerate kinase 1 in hepatocellular carcinoma. J. Cell. Biochem. 2019, 120, 18805–18815. [Google Scholar] [CrossRef] [PubMed]
  39. Ribbeck, K.; Raemaekers, T.; Carmeliet, G.; Mattaj, I.W. A role for NuSAP in linking microtubules to mitotic chromosomes. Curr. Biol. 2007, 17, 230–236. [Google Scholar] [CrossRef]
  40. Hou, S.; Hua, L.; Wang, W.; Li, M.; Xu, L. Nucleolar spindle associated protein 1 (NUSAP1) facilitates proliferation of hepatocellular carcinoma cells. Transl. Cancer Res. 2019, 8, 2113–2120. [Google Scholar] [CrossRef] [PubMed]
  41. Wang, Y.; Ju, L.; Xiao, F.; Liu, H.; Luo, X.; Chen, L.; Lu, Z.; Bian, Z. Downregulation of nucleolar and spindle-associated protein 1 expression suppresses liver cancer cell function. Exp. Ther. Med. 2019, 17, 2969–2978. [Google Scholar] [CrossRef] [PubMed]
  42. Roy, S.; Hooiveld, G.J.; Seehawer, M.; Caruso, S.; Heinzmann, F.; Schneider, A.T.; Frank, A.K.; Cardenas, D.V.; Sonntag, R.; Luedde, M.; et al. microRNA 193a-5p Regulates Levels of Nucleolar- and Spindle-Associated Protein 1 to Suppress Hepatocarcinogenesis. Gastroenterology 2018, 155, 1951–1966.e1926. [Google Scholar] [CrossRef]
  43. Yang, Y.; Zhao, Z.; Hou, N.; Li, Y.; Wang, X.; Wu, F.; Sun, R.; Han, J.; Sun, H.; Song, T. MicroRNA-214 targets Wnt3a to suppress liver cancer cell proliferation. Mol. Med. Rep. 2017, 16, 6920–6927. [Google Scholar] [CrossRef]
  44. Li, Y.; Li, Y.; Chen, Y.; Xie, Q.; Dong, N.; Gao, Y.; Deng, H.; Lu, C.; Wang, S. MicroRNA-214-3p inhibits proliferation and cell cycle progression by targeting MELK in hepatocellular carcinoma and correlates cancer prognosis. Cancer Cell Int. 2017, 17, 102. [Google Scholar] [CrossRef]
  45. Dai, X.; Huang, R.; Hu, S.; Zhou, Y.; Sun, X.; Gui, P.; Yu, Z.; Zhou, P. A novel miR-0308-3p revealed by miRNA-seq of HBV-positive hepatocellular carcinoma suppresses cell proliferation and promotes G1/S arrest by targeting double CDK6/Cyclin D1 genes. Cell Biosci. 2020, 10, 24. [Google Scholar] [CrossRef]
  46. Dai, X.; Chen, C.; Yang, Q.; Xue, J.; Chen, X.; Sun, B.; Luo, F.; Liu, X.; Xiao, T.; Xu, H. Exosomal circRNA_100284 from arsenite-transformed cells, via microRNA-217 regulation of EZH2, is involved in the malignant transformation of human hepatic cells by accelerating the cell cycle and promoting cell proliferation. Cell Death Dis. 2018, 9, 454. [Google Scholar] [CrossRef]
  47. Zhang, M.; Li, M.; Li, N.; Zhang, Z.; Liu, N.; Han, X.; Liu, Q.; Liao, C. miR-217 suppresses proliferation, migration, and invasion promoting apoptosis via targeting MTDH in hepatocellular carcinoma. Oncol. Rep. 2017, 37, 1772–1778. [Google Scholar] [CrossRef] [PubMed]
  48. Gao, W.; Lu, Y.X.; Wang, F.; Sun, J.; Bian, J.X.; Wu, H.Y. miRNA-217 inhibits proliferation of hepatocellular carcinoma cells by regulating KLF5. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 7874–7883. [Google Scholar]
  49. Li, S.; Peng, F.; Ning, Y.; Jiang, P.; Peng, J.; Ding, X.; Zhang, J.; Jiang, T.; Xiang, S. SNHG16 as the miRNA let-7b-5p sponge facilitates the G2/M and epithelial-mesenchymal transition by regulating CDC25B and HMGA2 expression in hepatocellular carcinoma. J. Cell. Biochem. 2020, 121, 2543–2558. [Google Scholar] [CrossRef]
  50. Zhao, G.; Han, C.; Zhang, Z.; Wang, L.; Xu, J. Increased expression of microRNA-31-5p inhibits cell proliferation, migration, and invasion via regulating Sp1 transcription factor in HepG2 hepatocellular carcinoma cell line. Biochem. Biophys. Res. Commun. 2017, 490, 371–377. [Google Scholar] [CrossRef] [PubMed]
  51. Zhang, D.; Liu, E.; Kang, J.; Yang, X.; Liu, H. MiR-3613-3p affects cell proliferation and cell cycle in hepatocellular carcinoma. Oncotarget 2017, 8, 93014. [Google Scholar] [CrossRef]
  52. Pollutri, D.; Patrizi, C.; Marinelli, S.; Giovannini, C.; Trombetta, E.; Giannone, F.A.; Baldassarre, M.; Quarta, S.; Vandewynckel, Y.-P.; Vandierendonck, A. The epigenetically regulated miR-494 associates with stem-cell phenotype and induces sorafenib resistance in hepatocellular carcinoma. Cell Death Dis. 2018, 9, 4. [Google Scholar] [CrossRef] [PubMed]
  53. Tian, F.; Yu, C.; Wu, M.; Wu, X.; Wan, L.; Zhu, X. MicroRNA-191 promotes hepatocellular carcinoma cell proliferation by has_circ_0000204/miR-191/KLF6 axis. Cell Prolif. 2019, 52, e12635. [Google Scholar] [CrossRef]
  54. Yao, B.; Niu, Y.; Li, Y.; Chen, T.; Wei, X.; Liu, Q. High-matrix-stiffness induces promotion of hepatocellular carcinoma proliferation and suppression of apoptosis via miR-3682-3p-PHLDA1-FAS pathway. J. Cancer 2020, 11, 6188–6203. [Google Scholar] [CrossRef] [PubMed]
  55. Zhu, Q.; Gong, L.; Wang, J.; Tu, Q.; Yao, L.; Zhang, J.-R.; Han, X.-J.; Zhu, S.-J.; Wang, S.-M.; Li, Y.-H.; et al. miR-10b exerts oncogenic activity in human hepatocellular carcinoma cells by targeting expression of CUB and sushi multiple domains 1 (CSMD1). BMC Cancer 2016, 16, 806. [Google Scholar] [CrossRef] [PubMed]
  56. Chen, Z.; Xiang, B.; Qi, L.; Zhu, S.; Li, L. miR-221-3p promotes hepatocellular carcinogenesis by downregulating O6-methylguanine-DNA methyltransferase. Cancer Biol. Ther. 2020, 21, 915–926. [Google Scholar] [CrossRef] [PubMed]
  57. Yan, G.; Elbadawi, M.; Efferth, T. Multiple cell death modalities and their key features. World Acad. Sci. J. 2020, 2, 39–48. [Google Scholar] [CrossRef]
  58. Walczak, H. Death receptor-ligand systems in cancer, cell death, and inflammation. Cold Spring Harb. Perspect. Biol. 2013, 5, a008698. [Google Scholar] [CrossRef]
  59. Wang, J.; Li, Y.; Ma, Q.; Huang, J. miR-378 in combination with ultrasonic irradiation and SonoVue microbubbles transfection inhibits hepatoma cell growth. Mol. Med. Rep. 2020, 21, 2493–2501. [Google Scholar] [CrossRef]
  60. Dong, X.; Wang, F.; Xue, Y.; Lin, Z.; Song, W.; Yang, N.; Li, Q. MicroRNA-9-5p downregulates Klf4 and influences the progression of hepatocellular carcinoma via the AKT signaling pathway. Int. J. Mol. Med. 2019, 43, 1417–1429. [Google Scholar] [CrossRef]
  61. Liao, Z.-B.; Tan, X.-L.; Dong, K.-S.; Zhang, H.-W.; Chen, X.-P.; Chu, L.; Zhang, B.-X. miRNA-448 inhibits cell growth by targeting BCL-2 in hepatocellular carcinoma. Dig. Liver Dis. 2019, 51, 703–711. [Google Scholar] [CrossRef]
  62. Wang, X.; Zeng, J.; Wang, L.; Zhang, X.; Liu, Z.; Zhang, H.; Dong, J. Overexpression of microRNA-133b is associated with the increased survival of patients with hepatocellular carcinoma after curative hepatectomy: Involvement of the EGFR/PI3K/Akt/mTOR signaling pathway. Oncol. Rep. 2017, 38, 141–150. [Google Scholar] [CrossRef]
  63. Li, C.; Hashimi, S.M.; Good, D.A.; Cao, S.; Duan, W.; Plummer, P.N.; Mellick, A.S.; Wei, M.Q. Apoptosis and micro RNA aberrations in cancer. Clin. Exp. Pharmacol. Physiol. 2012, 39, 739–746. [Google Scholar] [CrossRef]
  64. Zhou, X.; Wang, X.; Zhou, Y.; Cheng, L.; Zhang, Y.; Zhang, Y. Long Noncoding RNA NEAT1 Promotes Cell Proliferation and Invasion And Suppresses Apoptosis In Hepatocellular Carcinoma By Regulating miRNA-22-3p/akt2 In Vitro And In Vivo. OncoTargets Ther. 2019, 12, 8991–9004. [Google Scholar] [CrossRef]
  65. Wang, Y.; Tai, Q.; Zhang, J.; Kang, J.; Gao, F.; Zhong, F.; Cai, L.; Fang, F.; Gao, Y. MiRNA-206 inhibits hepatocellular carcinoma cell proliferation and migration but promotes apoptosis by modulating cMET expression. Acta Biochim. Biophys. Sin. 2019, 51, 243–253. [Google Scholar] [CrossRef]
  66. Lee, H.; Jeong, A.J.; Ye, S.-K. Highlighted STAT3 as a potential drug target for cancer therapy. BMB Rep. 2019, 52, 415–423. [Google Scholar] [CrossRef] [PubMed]
  67. Yang, L.; Xue, H.; Sun, Y.; Zhang, L.; Xue, F.; Ge, R. CircularRNA-9119 protects hepatocellular carcinoma cells from apoptosis by intercepting miR-26a/JAK1/STAT3 signaling. Cell Death Dis. 2020, 11, 605. [Google Scholar] [CrossRef] [PubMed]
  68. Uhlik, M.T.; Abell, A.N.; Cuevas, B.D.; Nakamura, K.; Johnson, G.L. Wiring diagrams of MAPK regulation by MEKK1, 2, and 3. Biochem. Cell Biol. 2004, 82, 658–663. [Google Scholar] [CrossRef]
  69. Cuevas, B.D.; Abell, A.N.; Johnson, G.L. Role of mitogen-activated protein kinase kinase kinases in signal integration. Oncogene 2007, 26, 3159–3171. [Google Scholar] [CrossRef] [PubMed]
  70. Su, B.; Cheng, J.; Yang, J.; Guo, Z. MEKK2 is required for T-cell receptor signals in JNK activation and interleukin-2 gene expression. J. Biol. Chem. 2001, 276, 14784–14790. [Google Scholar] [CrossRef]
  71. Mirza, A.A.; Kahle, M.P.; Ameka, M.; Campbell, E.M.; Cuevas, B.D. MEKK2 regulates focal adhesion stability and motility in invasive breast cancer cells. Biochim. Biophys. Acta BBA Mol. Cell Res. 2014, 1843, 945–954. [Google Scholar] [CrossRef]
  72. Mazur, P.K.; Reynoird, N.; Khatri, P.; Jansen, P.W.T.C.; Wilkinson, A.W.; Liu, S.; Barbash, O.; Van Aller, G.S.; Huddleston, M.; Dhanak, D. SMYD3 links lysine methylation of MAP3K2 to Ras-driven cancer. Nature 2014, 510, 283–287. [Google Scholar] [CrossRef]
  73. Li, X.; Azhati, B.; Wang, W.; Rexiati, M.; Xing, C.; Wang, Y. Circular RNA UBAP2 promotes the proliferation of prostate cancer cells via the miR-1244/MAP3K2 axis. Oncol. Lett. 2021, 21, 486. [Google Scholar] [CrossRef]
  74. Wang, M.; Lv, G.; Jiang, C.; Xie, S.; Wang, G. miR-302a inhibits human HepG2 and SMMC-7721 cells proliferation and promotes apoptosis by targeting MAP3K2 and PBX3. Sci. Rep. 2019, 9, 2032. [Google Scholar] [CrossRef] [PubMed]
  75. Chang, W.; Zhang, L.; Xian, Y.; Yu, Z. MicroRNA-33a promotes cell proliferation and inhibits apoptosis by targeting PPARα in human hepatocellular carcinoma. Exp. Ther. Med. 2017, 13, 2507–2514. [Google Scholar] [CrossRef] [PubMed]
  76. Gao, J.; Liu, Q.; Xu, Y.; Gong, X.; Zhang, R.; Zhou, C.; Su, Z.; Jin, J.; Shi, H.; Shi, J.; et al. PPARα induces cell apoptosis by destructing Bcl2. Oncotarget 2015, 6, 44635–44642. [Google Scholar] [CrossRef] [PubMed]
  77. Maggiora, M.; Oraldi, M.; Muzio, G.; Canuto, R.A. Involvement of PPARα and PPARγ in apoptosis and proliferation of human hepatocarcinoma HepG2 cells. Cell Biochem. Funct. 2010, 28, 571–577. [Google Scholar] [CrossRef]
  78. Zheng, X.; Li, S.; Yang, H. Roles of Toll-Like Receptor 3 in Human Tumors. Front. Immunol. 2021, 12, 667454. [Google Scholar] [CrossRef]
  79. Khvalevsky, E.; Rivkin, L.; Rachmilewitz, J.; Galun, E.; Giladi, H. TLR3 signaling in a hepatoma cell line is skewed towards apoptosis. J. Cell. Biochem. 2007, 100, 1301–1312. [Google Scholar] [CrossRef] [PubMed]
  80. Yin, L.; Cai, W.; Liang, Y.; Yao, J.; Wang, X.; Shen, J. In situ self-assembly of Au-antimiR-155 nanocomplexes mediates TLR3-dependent apoptosis in hepatocellular carcinoma cells. Aging 2020, 13, 241–261. [Google Scholar] [CrossRef]
  81. Xu, C.; Shi, L.; Chen, W.; Fang, P.; Li, J.; Jin, L.; Pan, Z.; Pan, C. MiR-106b inhibitors sensitize TRAIL-induced apoptosis in hepatocellular carcinoma through increase of death receptor 4. Oncotarget 2017, 8, 41921–41931. [Google Scholar] [CrossRef]
  82. Christofferson, D.E.; Yuan, J. Necroptosis as an alternative form of programmed cell death. Curr. Opin. Cell Biol. 2010, 22, 263–268. [Google Scholar] [CrossRef]
  83. Belizário, J.; Vieira-Cordeiro, L.; Enns, S. Necroptotic Cell Death Signaling and Execution Pathway: Lessons from Knockout Mice. Mediat. Inflamm. 2015, 2015, 128076. [Google Scholar] [CrossRef]
  84. Gong, Y.; Fan, Z.; Luo, G.; Yang, C.; Huang, Q.; Fan, K.; Cheng, H.; Jin, K.; Ni, Q.; Yu, X.; et al. The role of necroptosis in cancer biology and therapy. Mol. Cancer 2019, 18, 100. [Google Scholar] [CrossRef] [PubMed]
  85. Visalli, M.; Bartolotta, M.; Polito, F.; Oteri, R.; Barbera, A.; Arrigo, R.; Di Giorgio, R.M.; Navarra, G.; Aguennouz, M.H. miRNA expression profiling regulates necroptotic cell death in hepatocellular carcinoma. Int. J. Oncol. 2018, 53, 771–780. [Google Scholar] [CrossRef]
  86. van Zandwijk, N.; Pavlakis, N.; Kao, S.C.; Linton, A.; Boyer, M.J.; Clarke, S.; Huynh, Y.; Chrzanowska, A.; Fulham, M.J.; Bailey, D.L.; et al. Safety and activity of microRNA-loaded minicells in patients with recurrent malignant pleural mesothelioma: A first-in-man, phase 1, open-label, dose-escalation study. Lancet Oncol. 2017, 18, 1386–1396. [Google Scholar] [CrossRef]
  87. Morselli, E.; Galluzzi, L.; Kepp, O.; Mariño, G.; Michaud, M.; Vitale, I.; Maiuri, M.C.; Kroemer, G. Oncosuppressive functions of autophagy. Antioxid. Redox Signal. 2011, 14, 2251–2269. [Google Scholar] [CrossRef] [PubMed]
  88. Kenific, C.M.; Thorburn, A.; Debnath, J. Autophagy and metastasis: Another double-edged sword. Curr. Opin. Cell Biol. 2010, 22, 241–245. [Google Scholar] [CrossRef] [PubMed]
  89. Fung, C.; Lock, R.; Gao, S.; Salas, E.; Debnath, J. Induction of autophagy during extracellular matrix detachment promotes cell survival. Mol. Biol. Cell 2008, 19, 797–806. [Google Scholar] [CrossRef] [PubMed]
  90. Frankel, L.B.; Lund, A.H. MicroRNA regulation of autophagy. Carcinogenesis 2012, 33, 2018–2025. [Google Scholar] [CrossRef]
  91. Cao, W.; Li, J.; Yang, K.; Cao, D. An overview of autophagy: Mechanism, regulation and research progress. Bull. Du Cancer 2021, 108, 304–322. [Google Scholar] [CrossRef]
  92. Wang, S.; Li, H.; Yuan, M.; Fan, H.; Cai, Z. Role of AMPK in autophagy. Front. Physiol. 2022, 13, 1015500. [Google Scholar] [CrossRef]
  93. Zhang, Y.J.; Pan, Q.; Yu, Y.; Zhong, X.P. microRNA-519d Induces Autophagy and Apoptosis of Human Hepatocellular Carcinoma Cells Through Activation of the AMPK Signaling Pathway via Rab10. Cancer Manag. Res. 2020, 12, 2589–2602. [Google Scholar] [CrossRef]
  94. Zhou, L.; Liu, S.; Han, M.; Feng, S.; Liang, J.; Li, Z.; Li, Y.; Lu, H.; Liu, T.; Ma, Y. MicroRNA-185 induces potent autophagy via AKT signaling in hepatocellular carcinoma. Tumor Biol. 2017, 39, 1010428317694313. [Google Scholar] [CrossRef]
  95. Guo, J.; Ma, Y.; Peng, X.; Jin, H.; Liu, J. LncRNA CCAT1 promotes autophagy via regulating ATG7 by sponging miR-181 in hepatocellular carcinoma. J. Cell. Biochem. 2019, 120, 17975–17983. [Google Scholar] [CrossRef] [PubMed]
  96. Ying, H.; Zheng, H.; Scott, K.; Wiedemeyer, R.; Yan, H.; Lim, C.; Huang, J.; Dhakal, S.; Ivanova, E.; Xiao, Y. Mig-6 controls EGFR trafficking and suppresses gliomagenesis. Proc. Natl. Acad. Sci. USA 2010, 107, 6912–6917. [Google Scholar] [CrossRef] [PubMed]
  97. Qu, L.; Tian, Y.; Hong, D.; Wang, F.; Li, Z. Mig-6 Inhibits Autophagy in HCC Cell Lines by Modulating miR-193a-3p. Int. J. Med. Sci. 2022, 19, 338. [Google Scholar] [CrossRef] [PubMed]
  98. Yuan, J.; Li, Y.; Liao, J.; Liu, M.; Zhu, L.; Liao, K. MicroRNA-7 inhibits hepatocellular carcinoma cell invasion and metastasis by regulating Atg5-mediated autophagy. Transl. Cancer Res. 2020, 9, 3965–3972. [Google Scholar] [CrossRef]
  99. Fu, X.T.; Shi, Y.H.; Zhou, J.; Peng, Y.F.; Liu, W.R.; Shi, G.M.; Gao, Q.; Wang, X.Y.; Song, K.; Fan, J.; et al. MicroRNA-30a suppresses autophagy-mediated anoikis resistance and metastasis in hepatocellular carcinoma. Cancer Lett. 2018, 412, 108–117. [Google Scholar] [CrossRef] [PubMed]
  100. Jin, F.; Wang, Y.; Li, M.; Zhu, Y.; Liang, H.; Wang, C.; Wang, F.; Zhang, C.-Y.; Zen, K.; Li, L. MiR-26 enhances chemosensitivity and promotes apoptosis of hepatocellular carcinoma cells through inhibiting autophagy. Cell Death Dis. 2018, 8, e2540. [Google Scholar] [CrossRef]
  101. Su, Y.; Yao, L.; An, H.; Liu, J.; Ye, F.; Shen, J.; Ni, Z.; Huang, B.; Lin, J. MicroRNA-204-5p Inhibits Hepatocellular Carcinoma by Targeting the Regulator of G Protein Signaling 20. ACS Pharmacol. Transl. Sci. 2023, 6, 1817–1828. [Google Scholar] [CrossRef]
  102. Ye, R.; Lu, X.; Liu, J.; Duan, Q.; Xiao, J.; Duan, X.; Yue, Z.; Liu, F. CircSOD2 Contributes to Tumor Progression, Immune Evasion and Anti-PD-1 Resistance in Hepatocellular Carcinoma by Targeting miR-497-5p/ANXA11 Axis. Biochem Genet. 2023, 61, 597–614. [Google Scholar] [CrossRef]
  103. Chen, Z.; Du, J.; Yang, C.; Si, G.; Chen, Y. circ-CFH promotes the development of HCC by regulating cell proliferation, apoptosis, migration, invasion, and glycolysis through the miR-377-3p/RNF38 axis. Open Life Sci. 2022, 17, 248–260. [Google Scholar] [CrossRef]
  104. Zhang, T.; Zhang, Y.; Liu, J.; Ma, Y.; Ye, Q.; Yan, X.; Ding, L. MicroRNA-377-3p inhibits hepatocellular carcinoma growth and metastasis through negative regulation of CPT1C-mediated fatty acid oxidation. Cancer Metab. 2022, 10, 2. [Google Scholar] [CrossRef]
  105. Wang, C.; Li, C.; Hao, R. miR-559 Inhibits Proliferation, Autophagy, and Angiogenesis of Hepatocellular Carcinoma Cells by Targeting PARD3. Mediat. Inflamm. 2022, 2022, 3121492. [Google Scholar] [CrossRef]
  106. Zheng, X.S.; Liu, H.J.; Zhang, L.L.; Li, H.; Wang, C.J.; Xin, Y.J.; Hao, R. MiR-559 targets GP73 to suppress proliferation and invasion of hepatocellular carcinoma in vitro. Kaohsiung J. Med. Sci. 2020, 36, 793–798. [Google Scholar] [CrossRef] [PubMed]
  107. Zhang, H.; Liang, H.; Wu, S.; Zhang, Y.; Yu, Z. MicroRNA-638 induces apoptosis and autophagy in human liver cancer cells by targeting enhancer of zeste homolog 2 (EZH2). Environ. Toxicol. Pharmacol. 2021, 82, 103559. [Google Scholar] [CrossRef]
  108. Guan, J.; Liu, Z.; Xiao, M.; Hao, F.; Wang, C.; Chen, Y.; Lu, Y.; Liang, J. MicroRNA-199a-3p inhibits tumorigenesis of hepatocellular carcinoma cells by targeting ZHX1/PUMA signal. Am. J. Transl. Res. 2017, 9, 2457–2465. [Google Scholar] [PubMed]
  109. Fornari, F.; Milazzo, M.; Chieco, P.; Negrini, M.; Calin, G.A.; Grazi, G.L.; Pollutri, D.; Croce, C.M.; Bolondi, L.; Gramantieri, L. MiR-199a-3p regulates mTOR and c-Met to influence the doxorubicin sensitivity of human hepatocarcinoma cells. Cancer Res. 2010, 70, 5184–5193. [Google Scholar] [CrossRef]
  110. Yang, Y.; Yang, Z.; Zhang, R.; Jia, C.; Mao, R.; Mahati, S.; Zhang, Y.; Wu, G.; Sun, Y.n.; Jia, X.y. MiR-27a-3p enhances the cisplatin sensitivity in hepatocellular carcinoma cells through inhibiting PI3K/Akt pathway. Biosci. Rep. 2021, 41, BSR20192007. [Google Scholar] [CrossRef]
  111. Zhang, R.; Guo, C.; Liu, T.; Li, W.; Chen, X. MicroRNA miR-495 regulates the development of Hepatocellular Carcinoma by targeting C1q/tumor necrosis factor-related protein-3 (CTRP3). Bioengineered 2021, 12, 6902–6912. [Google Scholar] [CrossRef] [PubMed]
  112. Yang, X.; Yang, S.; Song, J.; Yang, W.; Ji, Y.; Zhang, F.; Rao, J. Dysregulation of miR-23b-5p promotes cell proliferation via targeting FOXM1 in hepatocellular carcinoma. Cell Death Discov. 2021, 7, 47. [Google Scholar] [CrossRef]
  113. Xu, X.; Jiang, W.; Han, P.; Zhang, J.; Tong, L.; Sun, X. MicroRNA-128-3p Mediates Lenvatinib Resistance of Hepatocellular Carcinoma Cells by Downregulating c-Met. J. Hepatocell. Carcinoma 2022, 9, 113–126. [Google Scholar] [CrossRef]
  114. Hu, C.; Cui, S.; Zheng, J.; Yin, T.; Lv, J.; Long, J.; Zhang, W.; Wang, X.; Sheng, S.; Zhang, H. MiR-875-5p inhibits hepatocellular carcinoma cell proliferation and migration by repressing astrocyte elevated gene-1 (AEG-1) expression. Transl. Cancer Res. 2018, 7, 158. [Google Scholar] [CrossRef]
  115. Zhang, Y.; Wei, Y.; Li, X.; Liang, X.; Wang, L.; Song, J.; Zhang, X.; Zhang, C.; Niu, J.; Zhang, P.; et al. microRNA-874 suppresses tumor proliferation and metastasis in hepatocellular carcinoma by targeting the DOR/EGFR/ERK pathway. Cell Death Dis. 2018, 9, 130. [Google Scholar] [CrossRef]
  116. Qin, X.; Chen, J.; Wu, L.; Liu, Z. MiR-30b-5p acts as a tumor suppressor, repressing cell proliferation and cell cycle in human hepatocellular carcinoma. Biomed. Pharmacother. 2017, 89, 742–750. [Google Scholar] [CrossRef] [PubMed]
  117. Li, L.; Ai, R.; Yuan, X.; Dong, S.; Zhao, D.; Sun, X.; Miao, T.; Guan, W.; Guo, P.; Yu, S.; et al. LINC00886 Facilitates Hepatocellular Carcinoma Tumorigenesis by Sequestering microRNA-409-3p and microRNA-214-5p. J. Hepatocell. Carcinoma 2023, 10, 863–881. [Google Scholar] [CrossRef] [PubMed]
  118. Jiang, T.; Li, M.; Li, Q.; Guo, Z.; Sun, X.; Zhang, X.; Liu, Y.; Yao, W.; Xiao, P. MicroRNA-98-5p Inhibits Cell Proliferation and Induces Cell Apoptosis in Hepatocellular Carcinoma via Targeting IGF2BP1. Oncol. Res. 2017, 25, 1117–1127. [Google Scholar] [CrossRef]
  119. Ma, Y.-S.; Lv, Z.-W.; Yu, F.; Chang, Z.-Y.; Cong, X.-L.; Zhong, X.-M.; Lu, G.-X.; Zhu, J.; Fu, D. MicroRNA-302a/d inhibits the self-renewal capability and cell cycle entry of liver cancer stem cells by targeting the E2F7/AKT axis. J. Exp. Clin. Cancer Res. 2018, 37, 252. [Google Scholar] [CrossRef]
  120. Sun, L.; Wang, L.; Chen, T.; Yao, B.; Wang, Y.; Li, Q.; Yang, W.; Liu, Z. microRNA-1914, which is regulated by lncRNA DUXAP10, inhibits cell proliferation by targeting the GPR39-mediated PI3K/AKT/mTOR pathway in HCC. J. Cell. Mol. Med. 2019, 23, 8292–8304. [Google Scholar] [CrossRef]
  121. Dong, G.; Zhang, S.; Shen, S.; Sun, L.; Wang, X.; Wang, H.; Wu, J.; Liu, T.; Wang, C.; Wang, H.; et al. SPATS2, negatively regulated by miR-145-5p, promotes hepatocellular carcinoma progression through regulating cell cycle. Cell Death Dis. 2020, 11, 837. [Google Scholar] [CrossRef]
  122. Simile, M.M.; Peitta, G.; Tomasi, M.L.; Brozzetti, S.; Feo, C.F.; Porcu, A.; Cigliano, A.; Calvisi, D.F.; Feo, F.; Pascale, R.M. MicroRNA-203 impacts on the growth, aggressiveness and prognosis of hepatocellular carcinoma by targeting MAT2A and MAT2B genes. Oncotarget 2019, 10, 2835. [Google Scholar] [CrossRef]
  123. Hui, L.; Zheng, F.; Bo, Y.; Sen-Lin, M.; Ai-Jun, L.; Wei-Ping, Z.; Yong-Jie, Z.; Lei, Y. MicroRNA let-7b inhibits cell proliferation via upregulation of p21 in hepatocellular carcinoma. Cell Biosci. 2020, 10, 83. [Google Scholar] [CrossRef]
  124. Feng, Y.; Xia, S.; Hui, J.; Xu, Y. Circular RNA circBNC2 facilitates glycolysis and stemness of hepatocellular carcinoma through the miR-217/high mobility group AT-hook 2 (HMGA2) axis. Heliyon 2023, 9, e17120. [Google Scholar] [CrossRef] [PubMed]
  125. Wan, L.; Yuan, X.; Liu, M.; Xue, B. miRNA-223-3p regulates NLRP3 to promote apoptosis and inhibit proliferation of hep3B cells. Exp. Ther. Med. 2018, 15, 2429–2435. [Google Scholar] [CrossRef]
  126. Liu, Z.; Li, W.; Pang, Y.; Zhou, Z.; Liu, S.; Cheng, K.; Qin, Q.; Jia, Y.; Liu, S. SF3B4 is regulated by microRNA-133b and promotes cell proliferation and metastasis in hepatocellular carcinoma. EBioMedicine 2018, 38, 57–68. [Google Scholar] [CrossRef] [PubMed]
  127. Wang, Y.; Wang, Q.; Song, J. Inhibition of autophagy potentiates the proliferation inhibition activity of microRNA-7 in human hepatocellular carcinoma cells. Oncol. Lett. 2017, 14, 3566–3572. [Google Scholar] [CrossRef] [PubMed]
  128. Wang, J.; Chen, J.; Liu, Y.; Zeng, X.; Wei, M.; Wu, S.; Xiong, Q.; Song, F.; Yuan, X.; Xiao, Y. Hepatitis B virus induces autophagy to promote its replication by the axis of miR-192-3p-XIAP through NF kappa B signaling. Hepatology 2019, 69, 974–992. [Google Scholar] [CrossRef]
  129. Ting, G.; Li, X.; Kwon, H.Y.; Ding, T.; Zhang, Z.; Chen, Z.; Li, C.; Liu, Y.; Yang, Y. microRNA-219-5p targets NEK6 to inhibit hepatocellular carcinoma progression. Am. J. Transl. Res. 2020, 12, 7528–7541. [Google Scholar]
  130. Li, L.; Jia, L.; Ding, Y. Upregulation of miR-375 inhibits human liver cancer cell growth by modulating cell proliferation and apoptosis via targeting ErbB2. Oncol. Lett. 2018, 16, 3319–3326. [Google Scholar] [CrossRef]
  131. Wang, D.; Sun, X.; Wei, Y.; Liang, H.; Yuan, M.; Jin, F.; Chen, X.; Liu, Y.; Zhang, C.-Y.; Li, L.; et al. Nuclear miR-122 directly regulates the biogenesis of cell survival oncomiR miR-21 at the posttranscriptional level. Nucleic Acids Res. 2018, 46, 2012–2029. [Google Scholar] [CrossRef]
  132. Li, Y.; Wang, X.; Li, Z.; Liu, B.; Wu, C. MicroRNA-4651 represses hepatocellular carcinoma cell growth and facilitates apoptosis via targeting FOXP4. Biosci. Rep. 2020, 40, BSR20194011. [Google Scholar] [CrossRef]
  133. Liu, Q.; Shi, H.; Yang, J.; Jiang, N. Long non-coding RNA NEAT1 promoted hepatocellular carcinoma cell proliferation and reduced apoptosis through the regulation of Let-7b-IGF-1R Axis. OncoTargets Ther. 2019, 12, 10401. [Google Scholar] [CrossRef]
  134. Komoll, R.M.; Hu, Q.; Olarewaju, O.; von Döhlen, L.; Yuan, Q.; Xie, Y.; Tsay, H.C.; Daon, J.; Qin, R.; Manns, M.P.; et al. MicroRNA-342-3p is a potent tumour suppressor in hepatocellular carcinoma. J. Hepatol. 2021, 74, 122–134. [Google Scholar] [CrossRef] [PubMed]
  135. Sun, J.; Liu, Q.; Jiang, Y.; Cai, Z.; Liu, H.; Zuo, H. Engineered small extracellular vesicles loaded with miR-654-5p promote ferroptosis by targeting HSPB1 to alleviate sorafenib resistance in hepatocellular carcinoma. Cell Death Discov. 2023, 9, 362. [Google Scholar] [CrossRef] [PubMed]
  136. Liu, Y.; Li, J. Circular RNA 0016142 Knockdown Induces Ferroptosis in Hepatocellular Carcinoma Cells via Modulation of the MicroRNA-188-3p/Glutathione Peroxidase 4 Axis. Biochem. Genet. 2023, 62, 333–351. [Google Scholar] [CrossRef]
  137. Xing, K.; Bian, X.; Shi, D.; Dong, S.; Zhou, H.; Xiao, S.; Bai, J.; Wu, W. miR-612 Enhances RSL3-Induced Ferroptosis of Hepatocellular Carcinoma Cells via Mevalonate Pathway. J. Hepatocell. Carcinoma 2023, 10, 2173–2185. [Google Scholar] [CrossRef]
  138. Duan, X.; Li, W.; Hu, P.; Jiang, B.; Yang, J.; Zhou, L.; Mao, X.; Tian, B. MicroRNA-183-5p contributes to malignant progression through targeting PDCD4 in human hepatocellular carcinoma. Biosci. Rep 2020, 40, BSR20201761. [Google Scholar] [CrossRef]
  139. Hu, Z.; Li, L.; Li, M.; Zhang, X.; Zhang, Y.; Ran, J.; Li, L. miR-21-5p Inhibits Ferroptosis in Hepatocellular Carcinoma Cells by Regulating the AKT/mTOR Signaling Pathway through MELK. J. Immunol. Res. 2023, 2023, 8929525. [Google Scholar] [CrossRef]
  140. Shen, H.; Li, H.; Zhou, J. Circular RNA hsa_circ_0032683 inhibits the progression of hepatocellular carcinoma by sponging microRNA-338-5p. Bioengineered 2022, 13, 2321–2335. [Google Scholar] [CrossRef] [PubMed]
  141. Cheng, N.; Wu, J.; Yin, M.; Xu, J.; Wang, Y.; Chen, X.; Nie, Z.; Yin, J. LncRNA CASC11 promotes cancer cell proliferation in hepatocellular carcinoma by inhibiting miRNA-188-5p. Biosci. Rep. 2019, 39, BSR20190251. [Google Scholar] [CrossRef]
  142. Chen, Y.L.; Xu, Q.P.; Guo, F.; Guan, W.H. MicroRNA-302d downregulates TGFBR2 expression and promotes hepatocellular carcinoma growth and invasion. Exp. Ther. Med. 2017, 13, 681–687. [Google Scholar] [CrossRef]
  143. Wang, Q.; Yang, X.; Zhou, X.; Wu, B.; Zhu, D.; Jia, W.; Chu, J.; Wang, J.; Wu, J.; Kong, L. MiR-3174 promotes proliferation and inhibits apoptosis by targeting FOXO1 in hepatocellular carcinoma. Biochem. Biophys. Res. Commun. 2020, 526, 889–897. [Google Scholar] [CrossRef]
  144. Li, Y.; Wu, B.; Sun, R.; Zhao, M.; Li, N. miR-93-5p knockdown repressed hepatocellular carcinoma progression via increasing ERBB4 and TETs-dependent DNA demethylation. Autoimmunity 2021, 54, 547–560. [Google Scholar] [CrossRef] [PubMed]
  145. Walcher, L.; Kistenmacher, A.-K.; Suo, H.; Kitte, R.; Dluczek, S.; Strauß, A.; Blaudszun, A.-R.; Yevsa, T.; Fricke, S.; Kossatz-Boehlert, U. Cancer Stem Cells—Origins and Biomarkers: Perspectives for Targeted Personalized Therapies. Front. Immunol. 2020, 11, 1280. [Google Scholar] [CrossRef]
  146. Holczbauer, Á.; Factor, V.M.; Andersen, J.B.; Marquardt, J.U.; Kleiner, D.E.; Raggi, C.; Kitade, M.; Seo, D.; Akita, H.; Durkin, M.E. Modeling pathogenesis of primary liver cancer in lineage-specific mouse cell types. Gastroenterology 2013, 145, 221–231. [Google Scholar] [CrossRef] [PubMed]
  147. Li-Li, L.; Da, F.; Yu-Shui, M.; Xizhong, S. The power and the promise of liver cancer stem cell markers. Stem Cells Dev. 2011, 20, 2023–2030. [Google Scholar] [CrossRef]
  148. Min, W.; Juan, X.; Jianxin, J.; Renyi, Q. CD133 and ALDH may be the molecular markers of cholangiocarcinoma stem cells. Int. J. Cancer 2011, 128, 1996–1997. [Google Scholar] [CrossRef]
  149. Shimaa, E.; Raja, C.; Mohammed, Z. Ultrasensitive Label-free Electrochemical Immunosensors for Multiple Cell Surface Biomarkers on Liver Cancer Stem Cells. Electroanalysis 2017, 29, 1994–2000. [Google Scholar] [CrossRef]
  150. Li, H. Efficacy analysis of combined detection of 5 Serological Tumor markers including MIF and PIVKA-II for early diagnosis of Primary Hepatic Cancer. Pak. J. Med. Sci. 2021, 37, 1456–1460. [Google Scholar] [CrossRef]
  151. Msy Rulan, A.; Indriyani; Ahmad Azmi, N.; Isabella Kurnia, L. Specific Markers for Hepatic Progenitor Cells. OnLine J. Biol. Sci. 2017, 17, 187–192. [Google Scholar] [CrossRef]
  152. Lv, D.; Chen, L.; Du, L.; Zhou, L.; Tang, H. Emerging Regulatory Mechanisms Involved in Liver Cancer Stem Cell Properties in Hepatocellular Carcinoma. Front. Cell Dev. Biol. 2021, 9, 691410. [Google Scholar] [CrossRef]
  153. Xie, D.; Pei, Q.; Li, J.; Wan, X.; Ye, T. Emerging Role of E2F Family in Cancer Stem Cells. Front. Oncol. 2021, 11, 723137. [Google Scholar] [CrossRef]
  154. Dou, Z.; Lu, F.; Hu, J.; Wang, H.; Li, B.; Li, X. MicroRNA-6838-5p suppresses the self-renewal and metastasis of human liver cancer stem cells through downregulating CBX4 expression and inactivating ERK signaling. Biol. Chem. 2023, 404, 29–39. [Google Scholar] [CrossRef]
  155. Zhang, X.; Jiang, P.; Shuai, L.; Chen, K.; Li, Z.; Zhang, Y.; Jiang, Y.; Li, X. miR-589-5p inhibits MAP3K8 and suppresses CD90+ cancer stem cells in hepatocellular carcinoma. J. Exp. Clin. Cancer Res. 2016, 35, 176. [Google Scholar] [CrossRef] [PubMed]
  156. Li, L.; Liu, Y.; Guo, Y.; Liu, B.; Zhao, Y.; Li, P.; Song, F.; Zheng, H.; Yu, J.; Song, T.; et al. Regulatory MiR-148a-ACVR1/BMP circuit defines a cancer stem cell-like aggressive subtype of hepatocellular carcinoma. Hepatology 2015, 61, 574–584. [Google Scholar] [CrossRef] [PubMed]
  157. Liu, Q.; Xu, Y.; Wei, S.; Gao, W.; Chen, L.; Zhou, T.; Wang, Z.; Ying, M.; Zheng, Q. miRNA-148b suppresses hepatic cancer stem cell by targeting neuropilin-1. Biosci. Rep. 2015, 35, e00229. [Google Scholar] [CrossRef] [PubMed]
  158. Shi, D.M.; Shi, X.L.; Xing, K.L.; Zhou, H.X.; Lu, L.L.; Wu, W.Z. miR-296-5p suppresses stem cell potency of hepatocellular carcinoma cells via regulating Brg1/Sall4 axis. Cell. Signal. 2020, 72, 109650. [Google Scholar] [CrossRef] [PubMed]
  159. Shi, D.-M.; Li, L.-X.; Bian, X.-Y.; Shi, X.-J.; Lu, L.-L.; Zhou, H.-X.; Pan, T.-J.; Zhou, J.; Fan, J.; Wu, W.-Z. miR-296-5p suppresses EMT of hepatocellular carcinoma via attenuating NRG1/ERBB2/ERBB3 signaling. J. Exp. Clin. Cancer Res. 2018, 37, 294. [Google Scholar] [CrossRef]
  160. Chai, S.; Ng, K.Y.; Tong, M.; Lau, E.Y.; Lee, T.K.; Chan, K.W.; Yuan, Y.F.; Cheung, T.T.; Cheung, S.T.; Wang, X.Q.; et al. Octamer 4/microRNA-1246 signaling axis drives Wnt/β-catenin activation in liver cancer stem cells. Hepatology 2016, 64, 2062–2076. [Google Scholar] [CrossRef]
  161. Lin, X.; Zuo, S.; Luo, R.; Li, Y.; Yu, G.; Zou, Y.; Zhou, Y.; Liu, Z.; Liu, Y.; Hu, Y.; et al. HBX-induced miR-5188 impairs FOXO1 to stimulate β-catenin nuclear translocation and promotes tumor stemness in hepatocellular carcinoma. Theranostics 2019, 9, 7583–7598. [Google Scholar] [CrossRef]
  162. Shi, D.-M.; Bian, X.-Y.; Qin, C.-D.; Wu, W.-Z. miR-106b-5p promotes stem cell-like properties of hepatocellular carcinoma cells by targeting PTEN via PI3K/Akt pathway. OncoTargets Ther. 2018, 11, 571–585. [Google Scholar] [CrossRef]
  163. Wang, Y.; Wang, B.; Xiao, S.; Li, Y.; Chen, Q. miR-125a/b inhibits tumor-associated macrophages mediated in cancer stem cells of hepatocellular carcinoma by targeting CD90. J. Cell. Biochem. 2018, 120, 3046–3055. [Google Scholar] [CrossRef]
  164. Li, B.; Liu, D.; Yang, P.; Li, H.-Y.; Wang, D. miR-613 inhibits liver cancer stem cell expansion by regulating SOX9 pathway. Gene 2019, 707, 78–85. [Google Scholar] [CrossRef] [PubMed]
  165. Seol, H.S.; Akiyama, Y.; Lee, S.-E.; Shimada, S.; Jang, S.J. Loss of miR-100 and miR-125b results in cancer stem cell properties through IGF2 upregulation in hepatocellular carcinoma. Sci. Rep. 2020, 10, 21412. [Google Scholar] [CrossRef] [PubMed]
  166. Li, R.; Xu, T.; Wang, H.; Wu, N.; Liu, F.; Jia, X.; Mi, J.; Lv, J.; Gao, H. Dysregulation of the miR-325–3p/DPAGT1 axis supports HBV-positive HCC chemoresistance. Biochem. Biophys. Res. Commun. 2019, 519, 358–365. [Google Scholar] [CrossRef] [PubMed]
  167. Guo, J.C.; Yang, Y.J.; Zhang, J.Q.; Guo, M.; Xiang, L.; Yu, S.F.; Ping, H.; Zhuo, L. microRNA-448 inhibits stemness maintenance and self-renewal of hepatocellular carcinoma stem cells through the MAGEA6-mediated AMPK signaling pathway. J. Cell. Physiol. 2019, 234, 23461–23474. [Google Scholar] [CrossRef]
  168. Zhong, F.; Wang, Y. YY1-regulated lncRNA SOCS2-AS1 suppresses HCC cell stemness and progression via miR-454-3p/CPEB1. Biochem. Biophys. Res. Commun. 2023, 679, 98–109. [Google Scholar] [CrossRef]
  169. Chaffer, C.L.; Weinberg, R.A. A perspective on cancer cell metastasis. Science 2011, 331, 1559–1564. [Google Scholar] [CrossRef]
  170. Geiger, T.R.; Peeper, D.S. Metastasis mechanisms. Biochim. Biophys. Acta 2009, 1796, 293–308. [Google Scholar] [CrossRef]
  171. Loh, C.-Y.; Chai, J.Y.; Tang, T.F.; Wong, W.F.; Sethi, G.; Shanmugam, M.K.; Chong, P.P.; Looi, C.Y. The E-Cadherin and N-Cadherin Switch in Epithelial-to-Mesenchymal Transition: Signaling, Therapeutic Implications, and Challenges. Cells 2019, 8, 1118. [Google Scholar] [CrossRef]
  172. Ruiz-Manriquez, L.M.; Carrasco-Morales, O.; Sanchez, Z.E.; Osorio-Perez, S.M.; Estrada-Meza, C.; Pathak, S.; Banerjee, A.; Bandyopadhyay, A.; Duttaroy, A.K.; Paul, S. MicroRNA-mediated regulation of key signaling pathways in hepatocellular carcinoma: A mechanistic insight. Front. Genet. 2022, 13, 910733. [Google Scholar] [CrossRef]
  173. Mitsunobu, M.; Toyosaka, A.; Oriyama, T.; Okamoto, E.; Nakao, N. Intrahepatic metastases in hepatocellular carcinoma: The role of the portal vein as an efferent vessel. Clin. Exp. Metastasis 1996, 14, 520–529. [Google Scholar] [CrossRef]
  174. Katyal, S.; Oliver, J.H., 3rd; Peterson, M.S.; Ferris, J.V.; Carr, B.S.; Baron, R.L. Extrahepatic metastases of hepatocellular carcinoma. Radiology 2000, 216, 698–703. [Google Scholar] [CrossRef] [PubMed]
  175. Qin, X.; Zhang, J.; Lin, Y.; Sun, X.-M.; Zhang, J.-N.; Cheng, Z.-Q. Identification of MiR-211-5p as a tumor suppressor by targeting ACSL4 in Hepatocellular Carcinoma. J. Transl. Med. 2020, 18, 326. [Google Scholar] [CrossRef] [PubMed]
  176. Qin, Z.; Liu, X.; Li, Z.; Wang, G.; Feng, Z.; Liu, Y.; Yang, H.; Tan, C.; Zhang, Z.; Li, K. LncRNA LINC00667 aggravates the progression of hepatocellular carcinoma by regulating androgen receptor expression as a miRNA-130a-3p sponge. Cell Death Discov. 2021, 7, 387. [Google Scholar] [CrossRef]
  177. Cai, Q.Q.; Dong, Y.W.; Wang, R.; Qi, B.; Guo, J.X.; Pan, J.; Liu, Y.Y.; Zhang, C.Y.; Wu, X.Z. MiR-124 inhibits the migration and invasion of human hepatocellular carcinoma cells by suppressing integrin αV expression. Sci. Rep. 2017, 7, 40733. [Google Scholar] [CrossRef]
  178. Zhao, B.; Lu, Y.; Cao, X.; Zhu, W.; Kong, L.; Ji, H.; Zhang, F.; Lin, X.; Guan, Q.; Ou, K. MiRNA-124 inhibits the proliferation, migration and invasion of cancer cell in hepatocellular carcinoma by downregulating lncRNA-UCA1. OncoTargets Ther. 2019, 12, 4509. [Google Scholar] [CrossRef] [PubMed]
  179. Jiang, C.; He, Z.; Hu, X.; Ma, P. MiRNA-15a-3p inhibits the metastasis of hepatocellular carcinoma by interacting with HMOX1. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 12694–12700. [Google Scholar]
  180. Sung, Y.K.; Hwang, S.Y.; Park, M.K.; Bae, H.I.; Kim, W.H.; Kim, J.C.; Kim, M. Fatty acid-CoA ligase 4 is overexpressed in human hepatocellular carcinoma. Cancer Sci. 2003, 94, 421–424. [Google Scholar] [CrossRef]
  181. Sung, Y.K.; Park, M.K.; Hong, S.H.; Hwang, S.Y.; Kwack, M.H.; Kim, J.C.; Kim, M.K. Regulation of cell growth by fatty acid-CoA ligase 4 in human hepatocellular carcinoma cells. Exp. Mol. Med. 2007, 39, 477–482. [Google Scholar] [CrossRef] [PubMed]
  182. Chen, J.; Ding, C.; Chen, Y.; Hu, W.; Lu, Y.; Wu, W.; Zhang, Y.; Yang, B.; Wu, H.; Peng, C.; et al. ACSL4 promotes hepatocellular carcinoma progression via c-Myc stability mediated by ERK/FBW7/c-Myc axis. Oncogenesis 2020, 9, 42. [Google Scholar] [CrossRef]
  183. Nagasue, N.; Kohno, H.; Chang, Y.C.; Hayashi, T.; Utsumi, Y.; Nakamura, T.; Yukaya, H. Androgen and estrogen receptors in hepatocellular carcinoma and the surrounding liver in women. Cancer 1989, 63, 112–116. [Google Scholar] [CrossRef]
  184. Grosso, A.R.; Martins, S.; Carmo-Fonseca, M. The emerging role of splicing factors in cancer. EMBO Rep. 2008, 9, 1087–1093. [Google Scholar] [CrossRef] [PubMed]
  185. Kim, E.; Goren, A.; Ast, G. Insights into the connection between cancer and alternative splicing. Trends Genet. 2008, 24, 7–10. [Google Scholar] [CrossRef]
  186. Shen, Q.; Nam, S.W. SF3B4 as an early-stage diagnostic marker and driver of hepatocellular carcinoma. BMB Rep. 2018, 51, 57–58. [Google Scholar] [CrossRef] [PubMed]
  187. Zhao, X.; Guan, J.-L. Focal adhesion kinase and its signaling pathways in cell migration and angiogenesis. Adv. Drug Deliv. Rev. 2011, 63, 610–615. [Google Scholar] [CrossRef] [PubMed]
  188. Qi, J.H.; Claesson-Welsh, L. VEGF-induced activation of phosphoinositide 3-kinase is dependent on focal adhesion kinase. Exp. Cell Res. 2001, 263, 173–182. [Google Scholar] [CrossRef]
  189. Kim, I.; Kim, H.G.; Moon, S.O.; Chae, S.W.; So, J.N.; Koh, K.N.; Ahn, B.C.; Koh, G.Y. Angiopoietin-1 induces endothelial cell sprouting through the activation of focal adhesion kinase and plasmin secretion. Circ. Res. 2000, 86, 952–959. [Google Scholar] [CrossRef]
  190. Liu, L.; Yang, X.; Li, N.F.; Lin, L.; Luo, H. Circ_0015756 promotes proliferation, invasion and migration by microRNA-7-dependent inhibition of FAK in hepatocellular carcinoma. Cell Cycle 2019, 18, 2939–2953. [Google Scholar] [CrossRef] [PubMed]
  191. Chan, C.C.; Dostie, J.; Diem, M.D.; Feng, W.; Mann, M.; Rappsilber, J.; Dreyfuss, G. eIF4A3 is a novel component of the exon junction complex. Rna 2004, 10, 200–209. [Google Scholar] [CrossRef]
  192. Xia, Q.; Kong, X.-T.; Zhang, G.-A.; Hou, X.-J.; Qiang, H.; Zhong, R.-Q. Proteomics-based identification of DEAD-box protein 48 as a novel autoantigen, a prospective serum marker for pancreatic cancer. Biochem. Biophys. Res. Commun. 2005, 330, 526–532. [Google Scholar] [CrossRef]
  193. Lin, Y.; Liang, R.; Mao, Y.; Ye, J.; Mai, R.; Gao, X.; Liu, Z.; Wainwright, T.; Li, Q.; Luo, M. Comprehensive analysis of biological networks and the eukaryotic initiation factor 4A-3 gene as pivotal in hepatocellular carcinoma. J. Cell. Biochem. 2020, 121, 4094–4107. [Google Scholar] [CrossRef]
  194. Zhu, Y.; Ren, C.; Yang, L. Effect of eukaryotic translation initiation factor 4A3 in malignant tumors. Oncol. Lett. 2021, 21, 358. [Google Scholar] [CrossRef]
  195. Zhang, L.; Chen, Y.; Bao, C.; Zhang, X.; Li, H. Eukaryotic initiation Factor 4AIII facilitates hepatocellular carcinoma cell proliferation, migration, and epithelial-mesenchymal transition process via antagonistically binding to WD repeat domain 66 with miRNA-2113. J. Cell. Physiol. 2020, 235, 8199–8209. [Google Scholar] [CrossRef] [PubMed]
  196. Wang, X.; Lu, J.; Cao, J.; Ma, B.; Gao, C.; Qi, F. MicroRNA-18a promotes hepatocellular carcinoma proliferation, migration, and invasion by targeting Bcl2L10. OncoTargets Ther. 2018, 11, 7919–7934. [Google Scholar] [CrossRef] [PubMed]
  197. Liu, L.; Cai, X.; Liu, E.; Tian, X.; Tian, C. MicroRNA-18a promotes proliferation and metastasis in hepatocellular carcinoma via targeting KLF4. Oncotarget 2017, 8, 68263–68269. [Google Scholar] [CrossRef]
  198. El-Mezayen, H.; Yamamura, K.; Yusa, T.; Nakao, Y.; Uemura, N.; Kitamura, F.; Itoyama, R.; Yamao, T.; Higashi, T.; Hayashi, H.; et al. MicroRNA-25 Exerts an Oncogenic Function by Regulating the Ubiquitin Ligase Fbxw7 in Hepatocellular Carcinoma. Ann. Surg. Oncol. 2021, 28, 7973–7982. [Google Scholar] [CrossRef]
  199. Yu, L.-X.; Zhang, B.-L.; Yang, M.-Y.; Liu, H.; Xiao, C.-H.; Zhang, S.-G.; Liu, R. MicroRNA-106b-5p promotes hepatocellular carcinoma development via modulating FOG2. OncoTargets Ther. 2019, 12, 5639–5647. [Google Scholar] [CrossRef] [PubMed]
  200. Wang, Y.; Chang, W.; Chang, W.; Chang, X.; Zhai, S.; Pan, G.; Dang, S. MicroRNA-376c-3p Facilitates Human Hepatocellular Carcinoma Progression via Repressing AT-Rich Interaction Domain 2. J. Cancer 2018, 9, 4187–4196. [Google Scholar] [CrossRef]
  201. Shi, X.; Liu, T.T.; Yu, X.N.; Balakrishnan, A.; Zhu, H.R.; Guo, H.Y.; Zhang, G.C.; Bilegsaikhan, E.; Sun, J.L.; Song, G.Q.; et al. microRNA-93-5p promotes hepatocellular carcinoma progression via a microRNA-93-5p/MAP3K2/c-Jun positive feedback circuit. Oncogene 2020, 39, 5768–5781. [Google Scholar] [CrossRef]
  202. Lv, G.; Wu, M.; Wang, M.; Jiang, X.; Du, J.; Zhang, K.; Li, D.; Ma, N.; Peng, Y.; Wang, L.; et al. miR-320a regulates high mobility group box 1 expression and inhibits invasion and metastasis in hepatocellular carcinoma. Liver Int. 2017, 37, 1354–1364. [Google Scholar] [CrossRef]
  203. Xie, F.; Yuan, Y.; Xie, L.; Ran, P.; Xiang, X.; Huang, Q.; Qi, G.; Guo, X.; Xiao, C.; Zheng, S. miRNA-320a inhibits tumor proliferation and invasion by targeting c-Myc in human hepatocellular carcinoma. OncoTargets Ther. 2017, 10, 885–894. [Google Scholar] [CrossRef]
  204. Che, J.; Su, Z.; Yang, W.; Xu, L.; Li, Y.; Wang, H.; Zhou, W. Tumor-suppressor p53 specifically binds to miR-29c-3p and reduces ADAM12 expression in hepatocellular carcinoma. Dig. Liver Dis. 2023, 55, 412–421. [Google Scholar] [CrossRef]
  205. Wu, Y.-H.; Yu, B.; Zhou, J.-M.; Shen, X.-H.; Chen, W.-X.; Ai, X.; Leng, C.; Liang, B.-Y.; Shao, Y.-J. MicroRNA-188-5p inhibits hepatocellular carcinoma proliferation and migration by targeting forkhead box N2. BMC Cancer 2023, 23, 511. [Google Scholar] [CrossRef] [PubMed]
  206. He, C.; Liu, Z.; Jin, L.; Zhang, F.; Peng, X.; Xiao, Y.; Wang, X.; Lyu, Q.; Cai, X. lncRNA TUG1-mediated Mir-142-3p downregulation contributes to metastasis and the epithelial-to-mesenchymal transition of hepatocellular carcinoma by targeting ZEB1. Cell. Physiol. Biochem. 2018, 48, 1928–1941. [Google Scholar] [CrossRef]
  207. Yu, Q.; Xiang, L.; Yin, L.; Liu, X.; Yang, D.; Zhou, J. Loss-of-function of miR-142 by hypermethylation promotes TGF-β-mediated tumour growth and metastasis in hepatocellular carcinoma. Cell Prolif. 2017, 50, e12384. [Google Scholar] [CrossRef] [PubMed]
  208. Fu, Y.; Sun, L.-Q.; Huang, Y.; Quan, J.; Hu, X.; Tang, D.; Kang, R.; Li, N.; Fan, X.-G. miR-142-3p Inhibits the Metastasis of Hepatocellular Carcinoma Cells by Regulating HMGB1 Gene Expression. Curr. Mol. Med. 2018, 18, 135–141. [Google Scholar] [CrossRef]
  209. Kim, H.S.; Kim, J.S.; Park, N.R.; Nam, H.; Sung, P.S.; Bae, S.H.; Choi, J.Y.; Yoon, S.K.; Hur, W.; Jang, J.W. Exosomal miR-125b Exerts Anti-Metastatic Properties and Predicts Early Metastasis of Hepatocellular Carcinoma. Front. Oncol. 2021, 11, 637247. [Google Scholar] [CrossRef]
  210. Liu, X.; Wang, H.; Yang, M.; Hou, Y.; Chen, Y.; Bie, P. Exosomal miR-29b from cancer-associated fibroblasts inhibits the migration and invasion of hepatocellular carcinoma cells. Transl. Cancer Res. 2020, 9, 2576. [Google Scholar] [CrossRef] [PubMed]
  211. Shen, D.; Zhao, H.Y.; Gu, A.D.; Wu, Y.W.; Weng, Y.H.; Li, S.J.; Song, J.Y.; Gu, X.F.; Qiu, J.; Zhao, W. miRNA-10a-5p inhibits cell metastasis in hepatocellular carcinoma via targeting SKA1. Kaohsiung J. Med. Sci. 2021, 37, 784–794. [Google Scholar] [CrossRef]
  212. Deng, L.; Tang, J.; Yang, H.; Cheng, C.; Lu, S.; Jiang, R.; Sun, B. MTA1 modulated by miR-30e contributes to epithelial-to-mesenchymal transition in hepatocellular carcinoma through an ErbB2-dependent pathway. Oncogene 2017, 36, 3976–3985. [Google Scholar] [CrossRef]
  213. Ma, L.; Tao, C.; Zhang, Y. MicroRNA-517c Functions as a Tumor Suppressor in Hepatocellular Carcinoma via Downregulation of KPNA2 and Inhibition of PI3K/AKT Pathway. J. Healthc. Eng. 2022, 2022, 7026174. [Google Scholar] [CrossRef]
  214. Liu, Y.; Xiao, X.; Wang, J.; Wang, Y.; Yu, Y. Silencing CircEIF3I/miR-526b-5p Axis Epigenetically Targets HGF/c-Met Signal to Hinder the Malignant Growth, Metastasis and Angiogenesis of Hepatocellular Carcinoma. Biochem. Genet. 2023, 61, 48–68. [Google Scholar] [CrossRef] [PubMed]
  215. Wang, Y.; Li, C.-f.; Sun, L.-b.; Li, Y.-c. microRNA-4270-5p inhibits cancer cell proliferation and metastasis in hepatocellular carcinoma by targeting SATB2. Hum. Cell 2020, 33, 1155–1164. [Google Scholar] [CrossRef]
  216. Majid, A.; Wang, J.; Nawaz, M.; Abdul, S.; Ayesha, M.; Guo, C.; Liu, Q.; Liu, S.; Sun, M.-Z. miR-124-3p suppresses the invasiveness and metastasis of hepatocarcinoma cells via targeting CRKL. Front. Mol. Biosci. 2020, 7, 223. [Google Scholar] [CrossRef] [PubMed]
  217. Han, B.; Huang, J.; Yang, Z.; Zhang, J.; Wang, X.; Xu, N.; Meng, H.; Wu, J.; Huang, Q.; Yang, X. miR-449a is related to short-term recurrence of hepatocellular carcinoma and inhibits migration and invasion by targeting notch1. OncoTargets Ther. 2019, 12, 10975. [Google Scholar] [CrossRef]
  218. Huang, Z.; Wen, J.; Yu, J.; Liao, J.; Liu, S.; Cai, N.; Liang, H.; Chen, X.; Ding, Z.; Zhang, B. MicroRNA-148a-3p inhibits progression of hepatocelluar carcimoma by repressing SMAD2 expression in an Ago2 dependent manner. J. Exp. Clin. Cancer Res. 2020, 39, 150. [Google Scholar] [CrossRef] [PubMed]
  219. Cao, N.; Mu, L.; Yang, W.; Liu, L.; Liang, L.; Zhang, H. MicroRNA-298 represses hepatocellular carcinoma progression by inhibiting CTNND1-mediated Wnt/β-catenin signaling. Biomed. Pharmacother. 2018, 106, 483–490. [Google Scholar] [CrossRef]
  220. Chang, J.-H.; Xu, B.-W.; Shen, D.; Zhao, W.; Wang, Y.; Liu, J.-L.; Meng, G.-X.; Li, G.-Z.; Zhang, Z.-L. BRF2 is mediated by microRNA-409-3p and promotes invasion and metastasis of HCC through the Wnt/β-catenin pathway. Cancer Cell Int. 2023, 23, 46. [Google Scholar] [CrossRef]
  221. Jia, C.; Tang, D.; Sun, C.; Yao, L.; Li, F.; Hu, Y.; Zhang, X.; Wu, D. MicroRNA-466 inhibits the aggressive behaviors of hepatocellular carcinoma by directly targeting metadherin. Oncol. Rep. 2018, 40, 3890–3898. [Google Scholar] [CrossRef]
  222. Yang, G.; Guo, S.; Liu, H.T.; Yang, G. MiR-138-5p predicts negative prognosis and exhibits suppressive activities in hepatocellular carcinoma HCC by targeting FOXC1. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 8788–8800. [Google Scholar]
  223. Hu, Y.; Yang, Z.; Bao, D.; Ni, J.-S.; Lou, J. miR-455-5p suppresses hepatocellular carcinoma cell growth and invasion via IGF-1R/AKT/GLUT1 pathway by targeting IGF-1R. Pathol. Res. Pract. 2019, 215, 152674. [Google Scholar] [CrossRef] [PubMed]
  224. Yu, Y.; You, S.; Fan, R.; Shan, X. UCK2 regulated by miR-139-3p regulates the progression of hepatocellular carcinoma cells. Future Oncol. 2022, 18, 979–990. [Google Scholar] [CrossRef]
  225. Ahmed, F.; Onwumeh-Okwundu, J.; Yukselen, Z.; Endaya Coronel, M.-K.; Zaidi, M.; Guntipalli, P.; Garimella, V.; Gudapati, S.; Mezidor, M.D.; Andrews, K.; et al. Atezolizumab plus bevacizumab versus sorafenib or atezolizumab alone for unresectable hepatocellular carcinoma: A systematic review. World J. Gastrointest. Oncol. 2021, 13, 1813–1832. [Google Scholar] [CrossRef]
  226. Li, L.-M.; Chen, C.; Ran, R.-X.; Huang, J.-T.; Sun, H.-L.; Zeng, C.; Zhang, Z.; Zhang, W.; Liu, S.-M. Corrigendum: Loss of TARBP2 drives the progression of hepatocellular carcinoma via miR-145-SERPINE1 axis. Front. Oncol. 2021, 11, 620912. [Google Scholar] [CrossRef]
  227. Ni, J.-S.; Zheng, H.; Ou, Y.-L.; Tao, Y.-P.; Wang, Z.-G.; Song, L.-H.; Yan, H.-L.; Zhou, W.-P. miR-515–5p suppresses HCC migration and invasion via targeting IL6/JAK/STAT3 pathway. Surg. Oncol. 2020, 34, 113–120. [Google Scholar] [CrossRef] [PubMed]
  228. Li, Y.; Zhou, T.; Cheng, X.; Li, D.; Zhao, M.; Zheng, W.V. microRNA-378a-3p regulates the progression of hepatocellular carcinoma by regulating PD-L1 and STAT3. Bioengineered 2022, 13, 4730–4743. [Google Scholar] [CrossRef]
  229. Zhan, Y.; Zheng, N.; Teng, F.; Bao, L.; Liu, F.; Zhang, M.; Guo, M.; Guo, W.; Ding, G.; Wang, Q. MiR-199a/b-5p inhibits hepatocellular carcinoma progression by post-transcriptionally suppressing ROCK1. Oncotarget 2017, 8, 67169. [Google Scholar] [CrossRef] [PubMed]
  230. Zou, H.; Yang, L. miR-378a-5p improved the prognosis and suppressed the progression of hepatocellular carcinoma by targeting the VEGF pathway. Transl. Cancer Res. 2020, 9, 1558–1566. [Google Scholar] [CrossRef] [PubMed]
  231. Ye, Y.; Zhuang, J.; Wang, G.; He, S.; Zhang, S.; Wang, G.; Ni, J.; Wang, J.; Xia, W. MicroRNA-495 suppresses cell proliferation and invasion of hepatocellular carcinoma by directly targeting insulin-like growth factor receptor-1. Exp. Ther. Med. 2018, 15, 1150–1158. [Google Scholar] [PubMed]
  232. Liu, B.; Tian, Y.; Chen, M.; Shen, H.; Xia, J.; Nan, J.; Yan, T.; Wang, Y.; Shi, L.; Shen, B. CircUBAP2 promotes MMP9-mediated oncogenic effect via sponging miR-194-3p in hepatocellular carcinoma. Front. Cell Dev. Biol. 2021, 9, 1634. [Google Scholar] [CrossRef]
  233. Lin, Y.; Liu, J.; Huang, Y.; Liu, D.; Zhang, G.; Kan, H. microRNA-489 plays an anti-metastatic role in human hepatocellular carcinoma by targeting matrix metalloproteinase-7. Transl. Oncol. 2017, 10, 211–220. [Google Scholar] [CrossRef]
  234. Hong, H.; Sui, C.; Qian, T.; Xu, X.; Zhu, X.; Fei, Q.; Yang, J.; Xu, M. Long noncoding RNA LINC00460 conduces to tumor growth and metastasis of hepatocellular carcinoma through miR-342-3p-dependent AGR2 up-regulation. Aging 2020, 12, 10544–10555. [Google Scholar] [CrossRef] [PubMed]
  235. Liu, Q.; Dai, S.-J.; Dong, L.; Li, H. Long noncoding RNA RP11-909N17. 2 promotes proliferation, invasion, and migration of hepatocellular carcinoma by regulating microRNA-767-3p. Biochem. Cell Biol. 2020, 98, 709–718. [Google Scholar] [CrossRef]
  236. He, H.; Wang, Y.; Ye, P.; Yi, D.; Cheng, Y.; Tang, H.; Zhu, Z.; Wang, X.; Jin, S. Long noncoding RNA ZFPM2-AS1 acts as a miRNA sponge and promotes cell invasion through regulation of miR-139/GDF10 in hepatocellular carcinoma. J. Exp. Clin. Cancer Res. 2020, 39, 159. [Google Scholar] [CrossRef]
  237. Sandbothe, M.; Buurman, R.; Reich, N.; Greiwe, L.; Vajen, B.; Gürlevik, E.; Schäffer, V.; Eilers, M.; Kühnel, F.; Vaquero, A.; et al. The microRNA-449 family inhibits TGF-β-mediated liver cancer cell migration by targeting SOX4. J. Hepatol. 2017, 66, 1012–1021. [Google Scholar] [CrossRef]
  238. Li, Q.; Li, S.; Wu, Y.; Gao, F. miRNA-708 functions as a tumour suppressor in hepatocellular carcinoma by targeting SMAD3. Oncol. Lett. 2017, 14, 2552–2558. [Google Scholar] [CrossRef] [PubMed]
  239. Chen, Y.; Guo, Y.; Li, Y.; Yang, J.; Liu, J.; Wu, Q.; Wang, R. miR-300 regulates tumor proliferation and metastasis by targeting lymphoid enhancer-binding factor 1 in hepatocellular carcinoma. Int. J. Oncol. 2019, 54, 1282–1294. [Google Scholar] [CrossRef] [PubMed]
  240. Li, L.; Sun, P.; Zhang, C.; Li, Z.; Zhou, W. MiR-98 suppresses the effects of tumor-associated macrophages on promoting migration and invasion of hepatocellular carcinoma cells by regulating IL-10. Biochimie 2018, 150, 23–30. [Google Scholar] [CrossRef]
  241. Zhang, J.-J.; Chen, J.-T.; Hua, L.; Yao, K.-H.; Wang, C.-Y. miR-98 inhibits hepatocellular carcinoma cell proliferation via targeting EZH2 and suppressing Wnt/β-catenin signaling pathway. Biomed. Pharmacother. 2017, 85, 472–478. [Google Scholar] [CrossRef]
  242. Zhang, Q.; Huang, F.; Yao, Y.; Wang, J.; Wei, J.; Wu, Q.; Xiang, S.; Xu, L. Interaction of transforming growth factor-β-Smads/microRNA-362-3p/CD82 mediated by M2 macrophages promotes the process of epithelial-mesenchymal transition in hepatocellular carcinoma cells. Cancer Sci. 2019, 110, 2507–2519. [Google Scholar] [CrossRef]
  243. Wang, J.; Zhu, Y.; Ai, X.; Wan, H.; Jia, W.; Chu, J.; Xu, B.; Kong, X.; Kong, L. Long noncoding RNA 02027 inhibits proliferation, migration and invasion of hepatocellular carcinoma via miR-625-3p/PDLIM5 pathway. J. Gene Med. 2023, 25, e3485. [Google Scholar] [CrossRef]
  244. Yang, B.; Feng, X.; Liu, H.; Tong, R.; Wu, J.; Li, C.; Yu, H.; Chen, Y.; Cheng, Q.; Chen, J.; et al. High-metastatic cancer cells derived exosomal miR92a-3p promotes epithelial-mesenchymal transition and metastasis of low-metastatic cancer cells by regulating PTEN/Akt pathway in hepatocellular carcinoma. Oncogene 2020, 39, 6529–6543. [Google Scholar] [CrossRef] [PubMed]
  245. Wang, L.; Cui, M.; Qu, F.; Cheng, D.; Yu, J.; Tang, Z.; Cheng, L.; Wei, Y.; Wu, X.; Liu, X. MiR-92a-3p promotes the malignant progression of hepatocellular carcinoma by mediating the PI3K/AKT/mTOR signaling pathway. Curr. Pharm. Des. 2021, 27, 3244–3250. [Google Scholar] [CrossRef] [PubMed]
  246. 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]
  247. Guo, L.; Gao, R.; Gan, J.; Zhu, Y.; Ma, J.; Lv, P.; Zhang, Y.; Li, S.; Tang, H. Downregulation of TNFRSF19 and RAB43 by a novel miRNA, miR-HCC3, promotes proliferation and epithelial–mesenchymal transition in hepatocellular carcinoma cells. Biochem. Biophys. Res. Commun. 2020, 525, 425–432. [Google Scholar] [CrossRef] [PubMed]
  248. Hu, Z.; Wang, P.; Lin, J.; Zheng, X.; Yang, F.; Zhang, G.; Chen, D.; Xie, J.; Gao, Z.; Peng, L. MicroRNA-197 promotes metastasis of hepatocellular carcinoma by activating Wnt/β-catenin signaling. Cell. Physiol. Biochem. 2018, 51, 470–486. [Google Scholar]
  249. Hong, Y.; Ye, M.; Wang, F.; Fang, J.; Wang, C.; Luo, J.; Liu, J.; Liu, J.; Liu, L.; Zhao, Q. MiR-21-3p promotes hepatocellular carcinoma progression via SMAD7/YAP1 regulation. Front. Oncol. 2021, 11, 642030. [Google Scholar] [CrossRef]
  250. Du, W.; Zhang, X.; Wan, Z. miR-3691-5p promotes hepatocellular carcinoma cell migration and invasion through activating PI3K/Akt signaling by targeting PTEN. OncoTargets Ther. 2019, 12, 4897–4906. [Google Scholar] [CrossRef]
  251. Qian, Y.; Chen, H.; Chen, L.; Ge, C.; Zhu, D.; Zhou, D. Suppression of hepatocellular carcinoma progression by long noncoding RNA apolipoprotein C1 pseudogene via the regulation of the microRNA-106b-PTEN axis. Transl. Cancer Res. 2023, 12, 3752–3763. [Google Scholar] [CrossRef]
  252. Cao, M.-Q.; You, A.B.; Zhu, X.-D.; Zhang, W.; Zhang, Y.-Y.; Zhang, S.-Z.; Zhang, K.-w.; Cai, H.; Shi, W.-K.; Li, X.-L. miR-182-5p promotes hepatocellular carcinoma progression by repressing FOXO3a. J. Hematol. Oncol. 2018, 11, 12. [Google Scholar] [CrossRef]
  253. Xian, Y.; Wang, L.; Yao, B.; Yang, W.; Mo, H.; Zhang, L.; Tu, K. MicroRNA-769-5p contributes to the proliferation, migration and invasion of hepatocellular carcinoma cells by attenuating RYBP. Biomed. Pharmacother. 2019, 118, 109343. [Google Scholar] [CrossRef]
  254. Han, S.; Wang, L.; Sun, L.; Wang, Y.; Yao, B.; Chen, T.; Liu, R.; Liu, Q. MicroRNA-1251-5p promotes tumor growth and metastasis of hepatocellular carcinoma by targeting AKAP12. Biomed. Pharmacother. 2020, 122, 109754. [Google Scholar] [CrossRef] [PubMed]
  255. Tang, S.; Chen, Y.; Feng, S.; Yi, T.; Liu, X.; Li, Q.; Liu, Z.; Zhu, C.; Hu, J.; Yu, X. MiR-483-5p promotes IGF-II transcription and is associated with poor prognosis of hepatocellular carcinoma. Oncotarget 2017, 8, 99871. [Google Scholar] [CrossRef]
  256. Liu, Y.-M.; Cao, Y.; Zhao, P.-S.; Wu, L.-Y.; Lu, Y.-M.; Wang, Y.-L.; Zhao, J.-F.; Liu, X.-G. CircCCNB1 silencing acting as a miR-106b-5p sponge inhibited GPM6A expression to promote HCC progression by enhancing DYNC1I1 expression and activating the AKT/ERK signaling pathway. Int. J. Biol. Sci. 2022, 18, 637. [Google Scholar] [CrossRef]
  257. Fu, X.; Tang, Y.; Wu, W.; Ouyang, Y.; Tan, D.; Huang, Y. Exosomal microRNA-25 released from cancer cells targets SIK1 to promote hepatocellular carcinoma tumorigenesis. Dig. Liver Dis. 2021, 54, 954–963. [Google Scholar] [CrossRef]
  258. Liu, H.; Chen, W.; Zhi, X.; Chen, E.-J.; Wei, T.; Zhang, J.; Shen, J.; Hu, L.-Q.; Zhao, B.; Feng, X.-H.; et al. Tumor-derived exosomes promote tumor self-seeding in hepatocellular carcinoma by transferring miRNA-25-5p to enhance cell motility. Oncogene 2018, 37, 4964–4978. [Google Scholar] [CrossRef]
  259. Zhou, Y.; Ren, H.; Dai, B.; Li, J.; Shang, L.; Huang, J.; Shi, X. Hepatocellular carcinoma-derived exosomal miRNA-21 contributes to tumor progression by converting hepatocyte stellate cells to cancer-associated fibroblasts. J. Exp. Clin. Cancer Res. 2018, 37, 324. [Google Scholar] [CrossRef]
  260. Yu, Y.; Min, Z.; Zhou, Z.; Linhong, M.; Tao, R.; Yan, L.; Song, H. Hypoxia-induced exosomes promote hepatocellular carcinoma proliferation and metastasis via miR-1273f transfer. Exp. Cell Res. 2019, 385, 111649. [Google Scholar] [CrossRef] [PubMed]
  261. Cui, Y.; Xu, H.-F.; Liu, M.-Y.; Xu, Y.-J.; He, J.-C.; Zhou, Y.; Cang, S.-D. Mechanism of exosomal microRNA-224 in development of hepatocellular carcinoma and its diagnostic and prognostic value. World J. Gastroenterol. 2019, 25, 1890–1898. [Google Scholar] [CrossRef]
  262. Yang, Y.; Mao, F.; Guo, L.; Shi, J.; Wu, M.; Cheng, S.; Guo, W. Tumor cells derived-extracellular vesicles transfer miR-3129 to promote hepatocellular carcinoma metastasis by targeting TXNIP. Dig. Liver Dis. 2021, 53, 474–485. [Google Scholar] [CrossRef]
  263. Fang, T.; Lv, H.; Lv, G.; Li, T.; Wang, C.; Han, Q.; Yu, L.; Su, B.; Guo, L.; Huang, S. Tumor-derived exosomal miR-1247-3p induces cancer-associated fibroblast activation to foster lung metastasis of liver cancer. Nat. Commun. 2018, 9, 191. [Google Scholar] [CrossRef]
  264. Zhuyan, J.; Chen, M.; Zhu, T.; Bao, X.; Zhen, T.; Xing, K.; Wang, Q.; Zhu, S. Critical steps to tumor metastasis: Alterations of tumor microenvironment and extracellular matrix in the formation of pre-metastatic and metastatic niche. Cell Biosci. 2020, 10, 89. [Google Scholar] [CrossRef] [PubMed]
  265. Anderson, N.M.; Simon, M.C. The tumor microenvironment. Curr. Biol. 2020, 30, R921–R925. [Google Scholar] [CrossRef] [PubMed]
  266. Brassart-Pasco, S.; Brézillon, S.; Brassart, B.; Ramont, L.; Oudart, J.-B.; Monboisse, J.C. Tumor Microenvironment: Extracellular Matrix Alterations Influence Tumor Progression. Front. Oncol. 2020, 10, 397. [Google Scholar] [CrossRef]
  267. Yu, S.; Cao, H.; Shen, B.; Feng, J. Tumor-derived exosomes in cancer progression and treatment failure. Oncotarget 2015, 6, 37151–37168. [Google Scholar] [CrossRef] [PubMed]
  268. Li, C.; Zhou, T.; Chen, J.; Li, R.; Chen, H.; Luo, S.; Chen, D.; Cai, C.; Li, W. The role of Exosomal miRNAs in cancer. J. Transl. Med. 2022, 20, 6. [Google Scholar] [CrossRef]
  269. Li, S.; Yao, J.; Xie, M.; Liu, Y.; Zheng, M. Exosomal miRNAs in hepatocellular carcinoma development and clinical responses. J. Hematol. Oncol. 2018, 11, 54. [Google Scholar] [CrossRef]
  270. Semenza, G.L. Hypoxia-inducible factors: Mediators of cancer progression and targets for cancer therapy. Trends Pharmacol. Sci. 2012, 33, 207–214. [Google Scholar] [CrossRef] [PubMed]
  271. Zeng, Z.; Lu, Q.; Liu, Y.; Zhao, J.; Zhang, Q.; Hu, L.; Shi, Z.; Tu, Y.; Xiao, Z.; Xu, Q.; et al. Effect of the Hypoxia Inducible Factor on Sorafenib Resistance of Hepatocellular Carcinoma. Front. Oncol. 2021, 11, 641522. [Google Scholar] [CrossRef] [PubMed]
  272. Tian, X.P.; Wang, C.Y.; Jin, X.H.; Li, M.; Wang, F.W.; Huang, W.J.; Yun, J.P.; Xu, R.H.; Cai, Q.Q.; Xie, D. Acidic Microenvironment Up-Regulates Exosomal miR-21 and miR-10b in Early-Stage Hepatocellular Carcinoma to Promote Cancer Cell Proliferation and Metastasis. Theranostics 2019, 9, 1965–1979. [Google Scholar] [CrossRef]
  273. Kim, M.-Y.; Oskarsson, T.; Acharyya, S.; Nguyen, D.X.; Zhang, X.H.F.; Norton, L.; Massagué, J. Tumor self-seeding by circulating cancer cells. Cell 2009, 139, 1315–1326. [Google Scholar] [CrossRef]
  274. Comen, E.; Norton, L.; Massagué, J. Clinical implications of cancer self-seeding. Nat. Rev. Clin. Oncol. 2011, 8, 369–377. [Google Scholar] [CrossRef]
  275. Jayatilaka, H.; Phillip, J.M. Targeting metastasis through the inhibition of interleukin 6 and 8. Future Med. 2019, 8, BMT20. [Google Scholar] [CrossRef]
  276. Hughey, C.C.; James, F.D.; Wang, Z.; Goelzer, M.; Wasserman, D.H. Dysregulated transmethylation leading to hepatocellular carcinoma compromises redox homeostasis and glucose formation. Mol. Metab. 2019, 23, 1–13. [Google Scholar] [CrossRef] [PubMed]
  277. Pan, Z.; Tian, Y.; Niu, G.; Cao, C. Role of microRNAs in remodeling the tumor microenvironment. Int. J. Oncol. 2020, 56, 407–416. [Google Scholar] [CrossRef] [PubMed]
  278. Qin, X.; Guo, H.; Wang, X.; Zhu, X.; Yan, M.; Wang, X.; Xu, Q.; Shi, J.; Lu, E.; Chen, W. Exosomal miR-196a derived from cancer-associated fibroblasts confers cisplatin resistance in head and neck cancer through targeting CDKN1B and ING5. Genome Biol. 2019, 20, 12. [Google Scholar] [CrossRef]
  279. Qin, X.; Yan, M.; Zhang, J.; Wang, X.; Shen, Z.; Lv, Z.; Li, Z.; Wei, W.; Chen, W. TGFβ3-mediated induction of Periostin facilitates head and neck cancer growth and is associated with metastasis. Sci. Rep. 2016, 6, 20587. [Google Scholar] [CrossRef]
  280. Linares, J.; Marín-Jiménez, J.A.; Badia-Ramentol, J.; Calon, A. Determinants and Functions of CAFs Secretome During Cancer Progression and Therapy. Front. Cell Dev. Biol. 2021, 8, 621070. [Google Scholar] [CrossRef] [PubMed]
  281. Jia, W.; Liang, S.; Lin, W.; Li, S.; Yuan, J.; Jin, M.; Nie, S.; Zhang, Y.; Zhai, X.; Zhou, L.; et al. Hypoxia-induced exosomes facilitate lung pre-metastatic niche formation in hepatocellular carcinoma through the miR-4508-RFX1-IL17A-p38 MAPK-NF-κB pathway. Int. J. Biol. Sci. 2023, 19, 4744–4762. [Google Scholar] [CrossRef]
  282. Yugawa, K.; Yoshizumi, T.; Mano, Y.; Itoh, S.; Harada, N.; Ikegami, T.; Kohashi, K.; Oda, Y.; Mori, M. Cancer-associated fibroblasts promote hepatocellular carcinoma progression through downregulation of exosomal miR-150-3p. Eur. J. Surg. Oncol. 2021, 47, 384–393. [Google Scholar] [CrossRef]
  283. Eun, J.W.; Ahn, H.R.; Baek, G.O.; Yoon, M.G.; Son, J.A.; Weon, J.H.; Yoon, J.H.; Kim, H.S.; Han, J.E.; Kim, S.S.; et al. Aberrantly Expressed MicroRNAs in Cancer-Associated Fibroblasts and Their Target Oncogenic Signatures in Hepatocellular Carcinoma. Int. J. Mol. Sci. 2023, 24, 4272. [Google Scholar] [CrossRef] [PubMed]
  284. Li, Z.; Wu, T.; Zheng, B.; Chen, L. Individualized precision treatment: Targeting TAM in HCC. Cancer Lett. 2019, 458, 86–91. [Google Scholar] [CrossRef]
  285. Zhu, Y.; Yang, J.; Xu, D.; Gao, X.-M.; Zhang, Z.; Hsu, J.L.; Li, C.-W.; Lim, S.-O.; Sheng, Y.-Y.; Zhang, Y. Disruption of tumour-associated macrophage trafficking by the osteopontin-induced colony-stimulating factor-1 signalling sensitises hepatocellular carcinoma to anti-PD-L1 blockade. Gut 2019, 68, 1653–1666. [Google Scholar] [CrossRef] [PubMed]
  286. Nakano, T.; Chen, C.-L.; Chen, I.-H.; Tseng, H.-P.; Chiang, K.-C.; Lai, C.-Y.; Hsu, L.-W.; Goto, S.; Lin, C.-C.; Cheng, Y.-F. Overexpression of miR-4669 Enhances Tumor Aggressiveness and Generates an Immunosuppressive Tumor Microenvironment in Hepatocellular Carcinoma: Its Clinical Value as a Predictive Biomarker. Int. J. Mol. Sci. 2023, 24, 7908. [Google Scholar] [CrossRef] [PubMed]
  287. Hu, Z.; Chen, J.; Zhao, Y.; Yan, C.; Wang, Y.; Zhu, J.; Li, L. Exosomal miR-452-5p Induce M2 Macrophage Polarization to Accelerate Hepatocellular Carcinoma Progression by Targeting TIMP3. J. Immunol. Res. 2022, 2022, 1032106. [Google Scholar] [CrossRef]
  288. Zhao, J.; Li, H.; Zhao, S.; Wang, E.; Zhu, J.; Feng, D.; Zhu, Y.; Dou, W.; Fan, Q.; Hu, J. Epigenetic silencing of miR-144/451a cluster contributes to HCC progression via paracrine HGF/MIF-mediated TAM remodeling. Mol. Cancer 2021, 20, 46. [Google Scholar] [CrossRef]
  289. Ke, M.; Zhang, Z.; Cong, L.; Zhao, S.; Li, Y.; Wang, X.; Lv, Y.; Zhu, Y.; Dong, J. MicroRNA-148b-colony-stimulating factor-1 signaling-induced tumor-associated macrophage infiltration promotes hepatocellular carcinoma metastasis. Biomed. Pharmacother. 2019, 120, 109523. [Google Scholar] [CrossRef]
  290. Zhou, S.L.; Hu, Z.Q.; Zhou, Z.J.; Dai, Z.; Wang, Z.; Cao, Y.; Fan, J.; Huang, X.W.; Zhou, J. miR-28-5p-IL-34-macrophage feedback loop modulates hepatocellular carcinoma metastasis. Hepatology 2016, 63, 1560–1575. [Google Scholar] [CrossRef]
  291. Xing, Y.; Wang, Z.; Lu, Z.; Xia, J.; Xie, Z.; Jiao, M.; Liu, R.; Chu, Y. MicroRNAs: Immune modulators in cancer immunotherapy. Immunother. Adv. 2021, 1, ltab006. [Google Scholar] [CrossRef]
  292. Han, Y.; Liu, D.; Li, L. PD-1/PD-L1 pathway: Current researches in cancer. Am. J. Cancer Res. 2020, 10, 727. [Google Scholar]
  293. Fu, Y.; Mackowiak, B.; Feng, D.; Lu, H.; Guan, Y.; Lehner, T.; Pan, H.; Wang, X.W.; He, Y.; Gao, B. MicroRNA-223 attenuates hepatocarcinogenesis by blocking hypoxia-driven angiogenesis and immunosuppression. Gut 2023, 72, 1942–1958. [Google Scholar] [CrossRef]
  294. Wang, Y.; Cao, K. KDM1A Promotes Immunosuppression in Hepatocellular Carcinoma by Regulating PD-L1 through Demethylating MEF2D. J. Immunol. Res. 2021, 2021, 9965099. [Google Scholar] [CrossRef]
  295. Sun, C.; Lan, P.; Han, Q.; Huang, M.; Zhang, Z.; Xu, G.; Song, J.; Wang, J.; Wei, H.; Zhang, J. Oncofetal gene SALL4 reactivation by hepatitis B virus counteracts miR-200c in PD-L1-induced T cell exhaustion. Nat. Commun. 2018, 9, 1241. [Google Scholar] [CrossRef] [PubMed]
  296. Hu, Y.; Setayesh, T.; Vaziri, F.; Wu, X.; Hwang, S.T.; Chen, X.; Wan, Y.-J.Y. miR-22 gene therapy treats HCC by promoting anti-tumor immunity and enhancing metabolism. Mol. Ther. 2023, 31, 1829–1845. [Google Scholar] [CrossRef] [PubMed]
  297. Liu, M.; Guo, S.; Stiles, J.K. The emerging role of CXCL10 in cancer (Review). Oncol. Lett. 2011, 2, 583–589. [Google Scholar] [CrossRef] [PubMed]
  298. Decalf, J.; Tarbell, K.V.; Casrouge, A.; Price, J.D.; Linder, G.; Mottez, E.; Sultanik, P.; Mallet, V.; Pol, S.; Duffy, D.; et al. Inhibition of DPP4 activity in humans establishes its in vivo role in CXCL10 post-translational modification: Prospective placebo-controlled clinical studies. EMBO Mol. Med. 2016, 8, 679–683. [Google Scholar] [CrossRef] [PubMed]
  299. Huang, X.-Y.; Zhang, P.-F.; Wei, C.-Y.; Peng, R.; Lu, J.-C.; Gao, C.; Cai, J.-B.; Yang, X.; Fan, J.; Ke, A.-W. Circular RNA circMET drives immunosuppression and anti-PD1 therapy resistance in hepatocellular carcinoma via the miR-30-5p/snail/DPP4 axis. Mol. Cancer 2020, 19, 92. [Google Scholar] [CrossRef]
  300. Hai, Y.; Hong, Y.; Yang, Y. miR-1258 Enhances the Anti-Tumor Effect of Liver Cancer Natural Killer (NK) Cells by Stimulating Toll-Liker Receptor (TLR) 7/8 to Promote Natural Killer (NK)-Dendritic Cell (DC) Interaction. J. Biomater. Tissue Eng. 2022, 12, 1241–1246. [Google Scholar] [CrossRef]
  301. Xie, H.; Zhang, Q.; Zhou, H.; Zhou, J.; Zhang, J.; Jiang, Y.; Wang, J.; Meng, X.; Zeng, L.; Jiang, X. microRNA-889 is downregulated by histone deacetylase inhibitors and confers resistance to natural killer cytotoxicity in hepatocellular carcinoma cells. Cytotechnology 2018, 70, 513–521. [Google Scholar] [CrossRef]
  302. Zhang, Z.; Li, X.; Sun, W.; Yue, S.; Yang, J.; Li, J.; Ma, B.; Wang, J.; Yang, X.; Pu, M. Loss of exosomal miR-320a from cancer-associated fibroblasts contributes to HCC proliferation and metastasis. Cancer Lett. 2017, 397, 33–42. [Google Scholar] [CrossRef]
  303. Chai, Z.-T.; Zhu, X.-D.; Ao, J.-Y.; Wang, W.-Q.; Gao, D.-M.; Kong, J.; Zhang, N.; Zhang, Y.-Y.; Ye, B.-G.; Ma, D.-N.; et al. microRNA-26a suppresses recruitment of macrophages by down-regulating macrophage colony-stimulating factor expression through the PI3K/Akt pathway in hepatocellular carcinoma. J. Hematol. Oncol. 2015, 8, 56. [Google Scholar] [CrossRef]
  304. Wang, L.; Yi, X.; Xiao, X.; Zheng, Q.; Ma, L.; Li, B. miR-628-5p from M1 polarized macrophages hinders m6A modification of circFUT8 to suppress hepatocellular carcinoma progression. Cell. Mol. Biol. Lett. 2022, 27, 106. [Google Scholar] [CrossRef]
  305. Chen, L.; Gibbons, D.L.; Goswami, S.; Cortez, M.A.; Ahn, Y.H.; Byers, L.A.; Zhang, X.; Yi, X.; Dwyer, D.; Lin, W.; et al. Metastasis is regulated via microRNA-200/ZEB1 axis control of tumor cell PD-L1 expression and intratumoral immunosuppression. Nat. Commun. 2015, 5, 5241. [Google Scholar] [CrossRef] [PubMed]
  306. Xu, G.; Zhang, P.; Liang, H.; Xu, Y.; Shen, J.; Wang, W.; Li, M.; Huang, J.; Ni, C.; Zhang, X.; et al. Circular RNA hsa_circ_0003288 induces EMT and invasion by regulating hsa_circ_0003288/miR-145/PD-L1 axis in hepatocellular carcinoma. Cancer Cell Int. 2021, 21, 212. [Google Scholar] [CrossRef] [PubMed]
  307. Fan, F.; Chen, K.; Lu, X.; Li, A.; Liu, C.; Wu, B. Dual targeting of PD-L1 and PD-L2 by PCED1B-AS1 via sponging hsa-miR-194-5p induces immunosuppression in hepatocellular carcinoma. Hepatol. Int. 2021, 15, 444–458. [Google Scholar] [CrossRef] [PubMed]
  308. Lin, Y.; Liu, S.; Su, L.; Su, Q.; Lin, J.; Huang, X.; Wang, C. miR-570 Inhibits Proliferation, Angiogenesis, and Immune Escape of Hepatocellular Carcinoma. Cancer Biother. Radiopharm. 2018, 33, 252–257. [Google Scholar] [CrossRef] [PubMed]
  309. Huang, F.; Wang, B.; Zeng, J.; Sang, S.; Lei, J.; Lu, Y. MicroRNA-374b inhibits liver cancer progression via down regulating programmed cell death-1 expression on cytokine-induced killer cells. Oncol. Lett. 2018, 15, 4797–4804. [Google Scholar] [CrossRef]
  310. Xu, Y.; Luan, G.; Liu, F.; Zhang, Y.; Li, Z.; Liu, Z.; Yang, T. Exosomal miR-200b-3p induce macrophage polarization by regulating transcriptional repressor ZEB1 in hepatocellular carcinoma. Hepatol. Int. 2023, 17, 889–903. [Google Scholar] [CrossRef]
  311. Liu, G.; Ouyang, X.; Sun, Y.; Xiao, Y.; You, B.; Gao, Y.; Yeh, S.; Li, Y.; Chang, C. The miR-92a-2-5p in exosomes from macrophages increases liver cancer cells invasion via altering the AR/PHLPP/p-AKT/β-catenin signaling. Cell Death Differ. 2020, 27, 3258–3272. [Google Scholar] [CrossRef]
  312. Li, W.; Xin, X.; Li, X.; Geng, J.; Sun, Y. Exosomes secreted by M2 macrophages promote cancer stemness of hepatocellular carcinoma via the miR-27a-3p/TXNIP pathways. Int. Immunopharmacol. 2021, 101, 107585. [Google Scholar] [CrossRef]
  313. Tian, B.; Zhou, L.; Wang, J.; Yang, P. miR-660-5p-loaded M2 macrophages-derived exosomes augment hepatocellular carcinoma development through regulating KLF3. Int. Immunopharmacol. 2021, 101, 108157. [Google Scholar] [CrossRef]
  314. Herbert, S.P.; Stainier, D.Y.R. Molecular control of endothelial cell behaviour during blood vessel morphogenesis. Nat. Rev. Mol. Cell Biol. 2011, 12, 551–564. [Google Scholar] [CrossRef] [PubMed]
  315. Hanahan, D.; Folkman, J. Patterns and emerging mechanisms of the angiogenic switch during tumorigenesis. Cell 1996, 86, 353–364. [Google Scholar] [CrossRef]
  316. Ghosh, A.; Dasgupta, D.; Ghosh, A.; Roychoudhury, S.; Kumar, D.; Gorain, M.; Butti, R.; Datta, S.; Agarwal, S.; Gupta, S. MiRNA199a-3p suppresses tumor growth, migration, invasion and angiogenesis in hepatocellular carcinoma by targeting VEGFA, VEGFR1, VEGFR2, HGF and MMP2. Cell Death Dis. 2017, 8, e2706. [Google Scholar] [CrossRef] [PubMed]
  317. Contarelli, S.; Fedele, V.; Melisi, D. HOX Genes Family and Cancer: A Novel Role for Homeobox B9 in the Resistance to Anti-Angiogenic Therapies. Cancers 2020, 12, 3299. [Google Scholar] [CrossRef] [PubMed]
  318. Wang, L.; Tong, D.; Guo, Q.; Wang, X.; Wu, F.; Li, Q.; Yang, J.; Zhao, L.; Qin, Y.; Liu, Y. HOXD3 targeted by miR-203a suppresses cell metastasis and angiogenesis through VEGFR in human hepatocellular carcinoma cells. Sci. Rep. 2018, 8, 2431. [Google Scholar] [CrossRef] [PubMed]
  319. Hou, M.; Lai, Y.; He, S.; He, W.; Shen, H.; Ke, Z. SGK3 (CISK) may induce tumor angiogenesis (Hypothesis). Oncol. Lett. 2015, 10, 23–26. [Google Scholar] [CrossRef]
  320. Wu, M.; Huang, C.; Huang, X.; Liang, R.; Feng, Y.; Luo, X. MicroRNA-144-3p suppresses tumor growth and angiogenesis by targeting SGK3 in hepatocellular carcinoma. Oncol. Rep. 2017, 38, 2173–2181. [Google Scholar] [CrossRef]
  321. Shah, A.V.; Birdsey, G.M.; Randi, A.M. Regulation of endothelial homeostasis, vascular development and angiogenesis by the transcription factor ERG. Vasc. Pharmacol. 2016, 86, 3–13. [Google Scholar] [CrossRef] [PubMed]
  322. Moh-Moh-Aung, A.; Fujisawa, M.; Ito, S.; Katayama, H.; Ohara, T.; Ota, Y.; Yoshimura, T.; Matsukawa, A. Decreased miR-200b-3p in cancer cells leads to angiogenesis in HCC by enhancing endothelial ERG expression. Sci. Rep. 2020, 10, 10418. [Google Scholar] [CrossRef]
  323. Zhang, P.; Ha, M.; Li, L.; Huang, X.; Liu, C. MicroRNA-3064-5p sponged by MALAT1 suppresses angiogenesis in human hepatocellular carcinoma by targeting the FOXA1/CD24/Src pathway. FASEB J. 2020, 34, 66–81. [Google Scholar] [CrossRef]
  324. Orso, F.; Quirico, L.; Dettori, D.; Coppo, R.; Virga, F.; Ferreira, L.C.; Paoletti, C.; Baruffaldi, D.; Penna, E.; Taverna, D. Role of miRNAs in tumor and endothelial cell interactions during tumor progression. Semin. Cancer Biol. 2020, 60, 214–224. [Google Scholar] [CrossRef]
  325. Muz, B.; de la Puente, P.; Azab, F.; Azab, A.K. The role of hypoxia in cancer progression, angiogenesis, metastasis, and resistance to therapy. Hypoxia 2015, 3, 83–92. [Google Scholar] [CrossRef]
  326. Matsuura, Y.; Wada, H.; Eguchi, H.; Gotoh, K.; Kobayashi, S.; Kinoshita, M.; Kubo, M.; Hayashi, K.; Iwagami, Y.; Yamada, D. Exosomal miR-155 derived from hepatocellular carcinoma cells under hypoxia promotes angiogenesis in endothelial cells. Dig. Dis. Sci. 2019, 64, 792–802. [Google Scholar] [CrossRef] [PubMed]
  327. Rhoads, K.; Arderiu, G.; Charboneau, A.; Hansen, S.L.; Hoffman, W.; Boudreau, N. A role for Hox A5 in regulating angiogenesis and vascular patterning. Lymphat. Res. Biol. 2005, 3, 240–252. [Google Scholar] [CrossRef]
  328. Liao, Y.; Wang, C.; Yang, Z.; Liu, W.; Yuan, Y.; Li, K.; Zhang, Y.; Wang, Y.; Shi, Y.; Qiu, Y. Dysregulated Sp1/miR-130b-3p/HOXA5 axis contributes to tumor angiogenesis and progression of hepatocellular carcinoma. Theranostics 2020, 10, 5209. [Google Scholar] [CrossRef]
  329. Schwarte-Waldhoff, I.; Schmiegel, W. Smad4 transcriptional pathways and angiogenesis. Int. J. Gastrointest. Cancer 2002, 31, 47–59. [Google Scholar] [CrossRef] [PubMed]
  330. Lin, X.-J.; Fang, J.-H.; Yang, X.-J.; Zhang, C.; Yuan, Y.; Zheng, L.; Zhuang, S.-M. Hepatocellular carcinoma cell-secreted exosomal microRNA-210 promotes angiogenesis in vitro and in vivo. Mol. Ther. Nucleic Acids 2018, 11, 243–252. [Google Scholar] [CrossRef] [PubMed]
  331. Yang, X.; Zhang, X.-F.; Lu, X.; Jia, H.-L.; Liang, L.; Dong, Q.-Z.; Ye, Q.-H.; Qin, L.-X. MicroRNA-26a Suppresses Angiogenesis in Human Hepatocellular Carcinoma by Targeting Hepatocyte Growth Factor-cMet Pathway. Hepatology 2014, 59, 1874–1885. [Google Scholar] [CrossRef]
  332. Li, D.; Wang, T.; Sun, F.-F.; Feng, J.-Q.; Peng, J.-J.; Li, H.; Wang, C.; Wang, D.; Liu, Y.; Bai, Y.-D. MicroRNA-375 represses tumor angiogenesis and reverses resistance to sorafenib in hepatocarcinoma. Cancer Gene Ther. 2021, 28, 126–140. [Google Scholar] [CrossRef]
  333. Liu, Y.; Tang, H.; Zhang, Y.; Wang, Q.; Li, S.; Wang, Z.; Shi, X. Circular RNA hsa_circ_0000519 contributes to angiogenesis and tumor progression in hepatocellular carcinoma through the miR-1296/E2F7 axis. Hum. Cell 2023, 36, 738–751. [Google Scholar] [CrossRef]
  334. Yang, H.D.; Kim, H.S.; Kim, S.Y.; Na, M.J.; Yang, G.; Eun, J.W.; Wang, H.J.; Cheong, J.Y.; Park, W.S.; Nam, S.W. HDAC6 suppresses Let-7i-5p to elicit TSP1/CD47-mediated anti-tumorigenesis and phagocytosis of hepatocellular carcinoma. Hepatology 2019, 70, 1262–1279. [Google Scholar] [CrossRef] [PubMed]
  335. Özkan, A.; Stolley, D.L.; Cressman, E.N.K.; McMillin, M.; DeMorrow, S.; Yankeelov, T.E.; Rylander, M.N. Tumor Microenvironment Alters Chemoresistance of Hepatocellular Carcinoma Through CYP3A4 Metabolic Activity. Front. Oncol. 2021, 11, 662135. [Google Scholar] [CrossRef]
  336. Duan, B.; Huang, C.; Bai, J.; Zhang, Y.L.; Wang, X.; Yang, J.; Li, J. Multidrug Resistance in Hepatocellular Carcinoma. In Hepatocellular Carcinoma; Tirnitz-Parker, J.E.E., Ed.; Codon Publications: Brisbane, Australia, 2019. [Google Scholar] [CrossRef]
  337. Liang, Y.; Liang, Q.; Qiao, L.; Xiao, F. MicroRNAs modulate drug resistance-related mechanisms in hepatocellular carcinoma. Front. Oncol. 2020, 10, 920. [Google Scholar] [CrossRef]
  338. Wei, X.; Zhao, L.; Ren, R.; Ji, F.; Xue, S.; Zhang, J.; Liu, Z.; Ma, Z.; Wang, X.W.; Wong, L. MiR-125b Loss Activated HIF1α/pAKT Loop, Leading to Transarterial Chemoembolization Resistance in Hepatocellular Carcinoma. Hepatology 2021, 73, 1381–1398. [Google Scholar] [CrossRef] [PubMed]
  339. Zhou, Y.; Chen, E.; Tang, Y.; Mao, J.; Shen, J.; Zheng, X.; Xie, S.; Zhang, S.; Wu, Y.; Liu, H. miR-223 overexpression inhibits doxorubicin-induced autophagy by targeting FOXO3a and reverses chemoresistance in hepatocellular carcinoma cells. Cell Death Dis. 2019, 10, 843. [Google Scholar] [CrossRef] [PubMed]
  340. Yuan, P.; Cao, W.; Zang, Q.; Li, G.; Guo, X.; Fan, J. The HIF-2α-MALAT1-miR-216b axis regulates multi-drug resistance of hepatocellular carcinoma cells via modulating autophagy. Biochem. Biophys. Res. Commun. 2016, 478, 1067–1073. [Google Scholar] [CrossRef] [PubMed]
  341. Ren, Q.; Xiao, X.; Leng, X.; Zhang, Q.; Zhou, X.; Ren, Z.; Xiao, H. MicroRNA-361-5p induces hepatocellular carcinoma cell apoptosis and enhances drug sensitivity by targeting MAP3K9. Exp. Ther. Med. 2021, 21, 574. [Google Scholar] [CrossRef] [PubMed]
  342. Xu, Y.; Wang, H.; Gao, W. MiRNA-610 acts as a tumour suppressor to depress the cisplatin resistance in hepatocellular carcinoma through targeted silencing of hepatoma-derived growth factor. Arch. Med. Sci. AMS 2020, 16, 1394. [Google Scholar] [CrossRef] [PubMed]
  343. Forouzanfar, M.; Lachinani, L.; Dormiani, K.; Nasr-Esfahani, M.H.; Gure, A.O.; Ghaedi, K. Intracellular functions of RNA-binding protein, Musashi1, in stem and cancer cells. Stem Cell Res. Ther. 2020, 11, 193. [Google Scholar] [CrossRef]
  344. Gao, J.; Dai, C.; Yu, X.; Yin, X.B.; Zhou, F. Long noncoding RNA LEF1-AS1 acts as a microRNA-10a-5p regulator to enhance MSI1 expression and promote chemoresistance in hepatocellular carcinoma cells through activating AKT signaling pathway. J. Cell. Biochem. 2021, 122, 86–99. [Google Scholar] [CrossRef]
  345. Ignatius, M.S.; Hayes, M.N.; Lobbardi, R.; Chen, E.Y.; McCarthy, K.M.; Sreenivas, P.; Motala, Z.; Durbin, A.D.; Molodtsov, A.; Reeder, S.; et al. The NOTCH1/SNAIL1/MEF2C Pathway Regulates Growth and Self-Renewal in Embryonal Rhabdomyosarcoma. Cell Rep. 2017, 19, 2304–2318. [Google Scholar] [CrossRef]
  346. Li, H.; Radford, J.C.; Ragusa, M.J.; Shea, K.L.; McKercher, S.R.; Zaremba, J.D.; Soussou, W.; Nie, Z.; Kang, Y.J.; Nakanishi, N.; et al. Transcription factor MEF2C influences neural stem/progenitor cell differentiation and maturation in vivo. Proc. Natl. Acad. Sci. USA 2008, 105, 9397–9402. [Google Scholar] [CrossRef] [PubMed]
  347. Kang, H.; Kim, C.; Ji, E.; Ahn, S.; Jung, M.; Hong, Y.; Kim, W.; Lee, E.K. The MicroRNA-551a/MEF2C axis regulates the survival and sphere formation of cancer cells in response to 5-fluorouracil. Mol. Cells 2019, 42, 175. [Google Scholar]
  348. Qin, J.; Luo, M.; Qian, H.; Chen, W. Upregulated miR-182 increases drug resistance in cisplatin-treated HCC cell by regulating TP53INP1. Gene 2014, 538, 342–347. [Google Scholar] [CrossRef] [PubMed]
  349. Fan, Y.-P.; Liao, J.-Z.; Lu, Y.-Q.; Tian, D.-A.; Ye, F.; Zhao, P.-X.; Xiang, G.-Y.; Tang, W.-X.; He, X.-X. MiR-375 and doxorubicin co-delivered by liposomes for combination therapy of hepatocellular carcinoma. Mol. Ther. Nucleic Acids 2017, 7, 181–189. [Google Scholar] [CrossRef] [PubMed]
  350. Yang, T.; Zheng, Z.-M.; Li, X.-N.; Li, Z.-F.; Wang, Y.; Geng, Y.-F.; Bai, L.; Zhang, X.-B. MiR-223 modulates multidrug resistance via downregulation of ABCB1 in hepatocellular carcinoma cells. Exp. Biol. Med. 2013, 238, 1024–1032. [Google Scholar] [CrossRef] [PubMed]
  351. Tu, C.; Chen, W.; Wang, S.; Tan, W.; Guo, J.; Shao, C.; Wang, W. MicroRNA-383 inhibits doxorubicin resistance in hepatocellular carcinoma by targeting eukaryotic translation initiation factor 5A2. J. Cell. Mol. Med. 2019, 23, 7190–7199. [Google Scholar] [CrossRef]
  352. Gao, X.; Jiang, Y.; Li, Y. Inhibitory effect of miR-140-5p on doxorubicin resistance of hepatocellular carcinoma. Exp. Ther. Med. 2021, 21, 507. [Google Scholar] [CrossRef]
  353. Chen, M.; Wu, L.; Tu, J.; Zhao, Z.; Fan, X.; Mao, J.; Weng, Q.; Wu, X.; Huang, L.; Xu, M. miR-590-5p suppresses hepatocellular carcinoma chemoresistance by targeting YAP1 expression. EBioMedicine 2018, 35, 142–154. [Google Scholar] [CrossRef]
  354. Yahya, S.M.M.; Fathy, S.A.; El-Khayat, Z.A.; El-Toukhy, S.E.; Hamed, A.R.; Hegazy, M.G.A.; Nabih, H.K. Possible Role of microRNA-122 in Modulating Multidrug Resistance of Hepatocellular Carcinoma. Indian J. Clin. Biochem. 2018, 33, 21–30. [Google Scholar] [CrossRef]
  355. Pan, C.; Wang, X.; Shi, K.; Zheng, Y.; Li, J.; Chen, Y.; Jin, L.; Pan, Z. MiR-122 Reverses the Doxorubicin-Resistance in Hepatocellular Carcinoma Cells through Regulating the Tumor Metabolism. PLoS ONE 2016, 11, e0152090. [Google Scholar] [CrossRef]
  356. Tian, T.; Fu, X.; Lu, J.; Ruan, Z.; Nan, K.; Yao, Y.; Yang, Y. MicroRNA-760 inhibits doxorubicin resistance in hepatocellular carcinoma through regulating Notch1/Hes1-PTEN/Akt signaling pathway. J. Biochem. Mol. Toxicol. 2018, 32, e22167. [Google Scholar] [CrossRef] [PubMed]
  357. Li, X.; He, J.; Ren, X.; Zhao, H.; Zhao, H. Circ_0003998 enhances doxorubicin resistance in hepatocellular carcinoma by regulating miR-218-5p/EIF5A2 pathway. Diagn. Pathol. 2020, 15, 141. [Google Scholar] [CrossRef] [PubMed]
  358. Chen, Z.; Ma, T.; Huang, C.; Zhang, L.; Lv, X.; Xu, T.; Hu, T.; Li, J. MiR-27a modulates the MDR1/P-glycoprotein expression by inhibiting FZD7/β-catenin pathway in hepatocellular carcinoma cells. Cell Signal. 2013, 25, 2693–2701. [Google Scholar] [CrossRef]
  359. Qin, Y.-F.; Zhou, Z.-Y.; Fu, H.-W.; Lin, H.-M.; Xu, L.-B.; Wu, W.-R.; Liu, C.; Xu, X.-L.; Zhang, R. Hepatitis B Virus Surface Antigen Promotes Stemness of Hepatocellular Carcinoma through Regulating MicroRNA-203a. J. Clin. Transl. Hepatol. 2023, 11, 118–129. [Google Scholar] [CrossRef] [PubMed]
  360. Jiang, J.-X.; Gao, S.; Pan, Y.-Z.; Yu, C.; Sun, C.-Y. Overexpression of microRNA-125b sensitizes human hepatocellular carcinoma cells to 5-fluorouracil through inhibition of glycolysis by targeting hexokinase II. Mol. Med. Rep. 2014, 10, 995–1002. [Google Scholar] [CrossRef]
  361. Zheng, R.-P.; Ma, D.-K.; Li, Z.; Zhang, H.-F. MiR-145 Regulates the Chemoresistance of Hepatic Carcinoma Cells Against 5-Fluorouracil by Targeting Toll-Like Receptor 4. Cancer Manag. Res. 2020, 12, 6165–6175. [Google Scholar] [CrossRef] [PubMed]
  362. Bai, B.; Liu, Y.; Fu, X.-M.; Qing, H.-Y.; Li, G.-K.; Wang, H.-C.; Sun, S.-L. Dysregulation of EZH2/miR-138-5p Axis Contributes to Radiosensitivity in Hepatocellular Carcinoma Cell by Downregulating Hypoxia-Inducible Factor 1 Alpha (HIF-1α). Oxidative Med. Cell. Longev. 2022, 2022, 7608712. [Google Scholar] [CrossRef]
  363. Fu, X.; Liu, M.; Qu, S.; Ma, J.; Zhang, Y.; Shi, T.; Wen, H.; Yang, Y.; Wang, S.; Wang, J. Exosomal microRNA-32-5p induces multidrug resistance in hepatocellular carcinoma via the PI3K/Akt pathway. J. Exp. Clin. Cancer Res. 2018, 37, 52. [Google Scholar] [CrossRef]
  364. Jin, X.; Cai, L.; Wang, C.; Deng, X.; Yi, S.; Lei, Z.; Xiao, Q.; Xu, H.; Luo, H.; Sun, J. CASC2/miR-24/miR-221 modulates the TRAIL resistance of hepatocellular carcinoma cell through caspase-8/caspase-3. Cell Death Dis. 2018, 9, 318. [Google Scholar] [CrossRef]
  365. Simpson, D.; Keating, G.M. Sorafenib. Drugs 2008, 68, 251–258. [Google Scholar] [CrossRef]
  366. Berretta, M.; Rinaldi, L.; Di Benedetto, F.; Lleshi, A.; De Re, V.; Facchini, G.; De Paoli, P.; Di Francia, R. Angiogenesis Inhibitors for the Treatment of Hepatocellular Carcinoma. Front. Pharmacol. 2016, 7, 428. [Google Scholar] [CrossRef] [PubMed]
  367. Llovet, J.M.; Montal, R.; Sia, D.; Finn, R.S. Molecular therapies and precision medicine for hepatocellular carcinoma. Nat. Rev. Clin. Oncol. 2018, 15, 599–616. [Google Scholar] [CrossRef]
  368. Nair, A.; Reece, K.; Donoghue, M.B.; Yuan, W.V.; Rodriguez, L.; Keegan, P.; Pazdur, R. FDA Supplemental Approval Summary: Lenvatinib for the Treatment of Unresectable Hepatocellular Carcinoma. Oncologist 2021, 26, e484–e491. [Google Scholar] [CrossRef]
  369. Capozzi, M.; De Divitiis, C.; Ottaiano, A.; von Arx, C.; Scala, S.; Tatangelo, F.; Delrio, P.; Tafuto, S. Lenvatinib, a molecule with versatile application: From preclinical evidence to future development in anti-cancer treatment. Cancer Manag. Res. 2019, 11, 3847–3860. [Google Scholar] [CrossRef] [PubMed]
  370. Hu, X.; Zhu, H.; Shen, Y.; Zhang, X.; He, X.; Xu, X. The role of non-coding RNAs in the sorafenib resistance of hepatocellular carcinoma. Front. Oncol. 2021, 11, 696705. [Google Scholar] [CrossRef] [PubMed]
  371. Kabir, T.D.; Ganda, C.; Brown, R.M.; Beveridge, D.J.; Richardson, K.L.; Chaturvedi, V.; Candy, P.; Epis, M.; Wintle, L.; Kalinowski, F.; et al. A microRNA-7/growth arrest specific 6/TYRO3 axis regulates the growth and invasiveness of sorafenib-resistant cells in human hepatocellular carcinoma. Hepatology 2018, 67, 216–231. [Google Scholar] [CrossRef]
  372. Reinkens, T.; Stalke, A.; Huge, N.; Vajen, B.; Eilers, M.; Schäffer, V.; Dittrich-Breiholz, O.; Schlegelberger, B.; Illig, T.; Skawran, B. Ago-RIP Sequencing Identifies New MicroRNA-449a-5p Target Genes Increasing Sorafenib Efficacy in Hepatocellular Carcinoma. J. Cancer 2022, 13, 62. [Google Scholar] [CrossRef]
  373. He, X.; Sun, H.; Jiang, Q.; Chai, Y.; Li, X.; Wang, Z.; Zhu, B.; You, S.; Li, B.; Hao, J. Hsa-miR-4277 Decelerates the Metabolism or Clearance of Sorafenib in HCC Cells and Enhances the Sensitivity of HCC Cells to Sorafenib by Targeting Cyp3a4. Front. Oncol. 2021, 11, 735447. [Google Scholar] [CrossRef]
  374. Sun, T.; Liu, H.; Ming, L. Multiple Roles of Autophagy in the Sorafenib Resistance of Hepatocellular Carcinoma. Cell. Physiol. Biochem. 2017, 44, 716–727. [Google Scholar] [CrossRef]
  375. Chen, M.-Y.; Yadav, V.K.; Chu, Y.C.; Ong, J.R.; Huang, T.-Y.; Lee, K.-F.; Lee, K.-H.; Yeh, C.-T.; Lee, W.-H. Hydroxychloroquine (HCQ) modulates autophagy and oxidative DNA damage stress in hepatocellular carcinoma to overcome sorafenib resistance via TLR9/SOD1/hsa-miR-30a-5p/Beclin-1 axis. Cancers 2021, 13, 3227. [Google Scholar] [CrossRef] [PubMed]
  376. Zhang, Z.; Tan, X.; Luo, J.; Yao, H.; Si, Z.; Tong, J.-S. The miR-30a-5p/CLCF1 axis regulates sorafenib resistance and aerobic glycolysis in hepatocellular carcinoma. Cell Death Dis. 2020, 11, 902. [Google Scholar] [CrossRef] [PubMed]
  377. Li, X.; Zhou, Y.; Yang, L.; Ma, Y.; Peng, X.; Yang, S.; Li, H.; Liu, J. LncRNA NEAT1 promotes autophagy via regulating miR-204/ATG3 and enhanced cell resistance to sorafenib in hepatocellular carcinoma. J. Cell. Physiol. 2020, 235, 3402–3413. [Google Scholar] [CrossRef] [PubMed]
  378. Zhang, K.; Chen, J.; Zhou, H.; Chen, Y.; Zhi, Y.; Zhang, B.; Chen, L.; Chu, X.; Wang, R.; Zhang, C. PU. 1/microRNA-142-3p targets ATG5/ATG16L1 to inactivate autophagy and sensitize hepatocellular carcinoma cells to sorafenib. Cell Death Dis. 2018, 9, 312. [Google Scholar] [CrossRef]
  379. Wang, D.; Yang, J. MiR-375 attenuates sorafenib resistance of hepatocellular carcinoma cells by inhibiting cell autophagy. Acta Biochim. Pol. 2023, 70, 239–246. [Google Scholar] [CrossRef]
  380. Lin, Z.; Xia, S.; Liang, Y.; Ji, L.; Pan, Y.; Jiang, S.; Wan, Z.; Tao, L.; Chen, J.; Lin, C. LXR activation potentiates sorafenib sensitivity in HCC by activating microRNA-378a transcription. Theranostics 2020, 10, 8834. [Google Scholar] [CrossRef]
  381. Kliewer, S.A. Nuclear receptor PXR: Discovery of a pharmaceutical anti-target. J. Clin. Investig. 2015, 125, 1388–1389. [Google Scholar] [CrossRef]
  382. Feng, F.; Jiang, Q.; Cao, S.; Cao, Y.; Li, R.; Shen, L.; Zhu, H.; Wang, T.; Sun, L.; Liang, E. Pregnane X receptor mediates sorafenib resistance in advanced hepatocellular carcinoma. Biochim. Biophys. Acta BBA Gen. Subj. 2018, 1862, 1017–1030. [Google Scholar] [CrossRef]
  383. Li, J.; Zhao, J.; Wang, H.; Li, X.; Liu, A.; Qin, Q.; Li, B. MicroRNA-140-3p enhances the sensitivity of hepatocellular carcinoma cells to sorafenib by targeting pregnenolone X receptor. OncoTargets Ther. 2018, 11, 5885. [Google Scholar] [CrossRef]
  384. Li, B.; Feng, F.; Jia, H.; Jiang, Q.; Cao, S.; Wei, L.; Zhang, Y.; Lu, J. Rhamnetin decelerates the elimination and enhances the antitumor effect of the molecular-targeting agent sorafenib in hepatocellular carcinoma cells via the miR-148a/PXR axis. Food Funct. 2021, 12, 2404–2417. [Google Scholar] [CrossRef]
  385. Dong, Z.-B.; Wu, H.-M.; He, Y.-C.; Huang, Z.-T.; Weng, Y.-H.; Li, H.; Liang, C.; Yu, W.-M.; Chen, W. MiRNA-124-3p. 1 sensitizes hepatocellular carcinoma cells to sorafenib by regulating FOXO3a by targeting AKT2 and SIRT1. Cell Death Dis. 2022, 13, 35. [Google Scholar] [CrossRef]
  386. Li, T.-T.; Mou, J.; Pan, Y.-J.; Huo, F.-C.; Du, W.-Q.; Liang, J.; Wang, Y.; Zhang, L.-S.; Pei, D.-S. MicroRNA-138-1-3p sensitizes sorafenib to hepatocellular carcinoma by targeting PAK5 mediated β-catenin/ABCB1 signaling pathway. J. Biomed. Sci. 2021, 28, 56. [Google Scholar] [CrossRef] [PubMed]
  387. Bozkulak, E.C.; Weinmaster, G. Selective use of ADAM10 and ADAM17 in activation of Notch1 signaling. Mol. Cell. Biol. 2009, 29, 5679–5695. [Google Scholar] [CrossRef] [PubMed]
  388. Ruan, Y.; Chen, T.; Zheng, L.; Cai, J.; Zhao, H.; Wang, Y.; Tao, L.; Xu, J.; Ji, L.; Cai, X. cDCBLD2 mediates sorafenib resistance in hepatocellular carcinoma by sponging miR-345-5p binding to the TOP2A coding sequence. Int. J. Biol. Sci. 2023, 19, 4608–4626. [Google Scholar] [CrossRef] [PubMed]
  389. Lu, Y.; Chan, Y.-T.; Wu, J.; Feng, Z.; Yuan, H.; Li, Q.; Xing, T.; Xu, L.; Zhang, C.; Tan, H.-Y.; et al. CRISPR/Cas9 screens unravel miR-3689a-3p regulating sorafenib resistance in hepatocellular carcinoma via suppressing CCS/SOD1-dependent mitochondrial oxidative stress. Drug Resist. Updat. 2023, 71, 318. [Google Scholar] [CrossRef]
  390. Lu, Y.; Chan, Y.-T.; Tan, H.-Y.; Zhang, C.; Guo, W.; Xu, Y.; Sharma, R.; Chen, Z.-S.; Zheng, Y.-C.; Wang, N. Epigenetic regulation of ferroptosis via ETS1/miR-23a-3p/ACSL4 axis mediates sorafenib resistance in human hepatocellular carcinoma. J. Exp. Clin. Cancer Res. 2022, 41, 3. [Google Scholar] [CrossRef] [PubMed]
  391. Doll, S.; Proneth, B.; Tyurina, Y.Y.; Panzilius, E.; Kobayashi, S.; Ingold, I.; Irmler, M.; Beckers, J.; Aichler, M.; Walch, A.; et al. ACSL4 dictates ferroptosis sensitivity by shaping cellular lipid composition. Nat. Chem. Biol. 2017, 13, 91–98. [Google Scholar] [CrossRef]
  392. Hirao, A.; Sato, Y.; Tanaka, H.; Nishida, K.; Tomonari, T.; Hirata, M.; Bando, M.; Kida, Y.; Tanaka, T.; Kawaguchi, T. MiR-125b-5p Is Involved in Sorafenib Resistance through Ataxin-1-Mediated Epithelial-Mesenchymal Transition in Hepatocellular Carcinoma. Cancers 2021, 13, 4917. [Google Scholar] [CrossRef]
  393. Bergamini, C.; Leoni, I.; Rizzardi, N.; Melli, M.; Galvani, G.; Coada, C.A.; Giovannini, C.; Monti, E.; Liparulo, I.; Valenti, F.; et al. MiR-494 induces metabolic changes through G6pc targeting and modulates sorafenib response in hepatocellular carcinoma. J. Exp. Clin. Cancer Res. 2023, 42, 145. [Google Scholar] [CrossRef]
  394. Hu, Z.; Zhao, Y.; Mang, Y.; Zhu, J.; Yu, L.; Li, L.; Ran, J. MiR-21-5p promotes sorafenib resistance and hepatocellular carcinoma progression by regulating SIRT7 ubiquitination through USP24. Life Sci. 2023, 15, 121773. [Google Scholar] [CrossRef]
  395. Yu, T.; Yu, J.; Lu, L.; Zhang, Y.; Zhou, Y.; Zhou, Y.; Huang, F.; Sun, L.; Guo, Z.; Hou, G.; et al. MT1JP-mediated miR-24-3p/BCL2L2 axis promotes Lenvatinib resistance in hepatocellular carcinoma cells by inhibiting apoptosis. Cell. Oncol. 2021, 44, 821–834. [Google Scholar] [CrossRef]
  396. Gu, D.; Tong, M.; Wang, J.; Zhang, B.; Liu, J.; Song, G.; Zhu, B. Overexpression of the lncRNA HOTAIRM1 promotes lenvatinib resistance by downregulating miR-34a and activating autophagy in hepatocellular carcinoma. Discov. Oncol. 2023, 14, 66. [Google Scholar] [CrossRef] [PubMed]
  397. Han, T.; Zheng, H.; Zhang, J.; Yang, P.; Li, H.; Cheng, Z.; Xiang, D.; Wang, R. Downregulation of MUC15 by miR-183-5p.1 promotes liver tumor-initiating cells properties and tumorigenesis via regulating c-MET/PI3K/AKT/SOX2 axis. Cell Death Dis. 2022, 13, 200. [Google Scholar] [CrossRef] [PubMed]
  398. Liu, D.; Liu, W.; Chen, X.; Yin, J.; Ma, L.; Liu, M.; Zhou, X.; Xian, L.; Li, P.; Tan, X.; et al. circKCNN2 suppresses the recurrence of hepatocellular carcinoma at least partially via regulating miR-520c-3p/methyl-DNA-binding domain protein 2 axis. Clin. Transl. Med. 2022, 12, e662. [Google Scholar] [CrossRef]
  399. Wei, Y.; Wei, L.; Han, T.; Ding, S. miR-3154 promotes hepatocellular carcinoma progression via suppressing HNF4α. Carcinogenesis 2022, 43, 1002–1014. [Google Scholar] [CrossRef]
  400. Wang, G.; Zhao, W.; Wang, H.; Qiu, G.; Jiang, Z.; Wei, G.; Li, X. Exosomal MiR-744 inhibits proliferation and sorafenib chemoresistance in hepatocellular carcinoma by targeting PAX2. Med. Sci. Monit. Int. Med. J. Exp. Clin. Res. 2019, 25, 7209. [Google Scholar] [CrossRef]
  401. Feng, Y.; Jiang, W.; Zhao, W.; Lu, Z.; Gu, Y.; Dong, Y. miR-124 regulates liver cancer stem cells expansion and sorafenib resistance. Exp. Cell Res. 2020, 394, 112162. [Google Scholar] [CrossRef] [PubMed]
  402. Xu, C.; Sun, W.; Liu, J.; Pu, H.; Li, Y. Circ_RBM23 knockdown suppresses chemoresistance, proliferation, migration and invasion of sorafenib-resistant HCC cells through miR-338–3p/RAB1B axis. Pathol. Res. Pract. 2023, 245, 154435. [Google Scholar] [CrossRef] [PubMed]
  403. Xu, Y.; Huang, J.; Ma, L.; Shan, J.; Shen, J.; Yang, Z.; Liu, L.; Luo, Y.; Yao, C.; Qian, C. MicroRNA-122 confers sorafenib resistance to hepatocellular carcinoma cells by targeting IGF-1R to regulate RAS/RAF/ERK signaling pathways. Cancer Lett. 2016, 371, 171–181. [Google Scholar] [CrossRef]
  404. Turato, C.; Fornari, F.; Pollutri, D.; Fassan, M.; Quarta, S.; Villano, G.; Ruvoletto, M.; Bolondi, L.; Gramantieri, L.; Pontisso, P. MiR-122 targets SerpinB3 and is involved in sorafenib resistance in hepatocellular carcinoma. J. Clin. Med. 2019, 8, 171. [Google Scholar] [CrossRef]
  405. Chen, B.W.; Zhou, Y.; Wei, T.; Wen, L.; Zhang, Y.b.; Shen, S.c.; Zhang, J.; Ma, T.; Chen, W.; Ni, L. lncRNA-POIR promotes epithelial–mesenchymal transition and suppresses sorafenib sensitivity simultaneously in hepatocellular carcinoma by sponging miR-182-5p. J. Cell. Biochem. 2021, 122, 130–142. [Google Scholar] [CrossRef] [PubMed]
  406. Zhang, P.F.; Wang, F.; Wu, J.; Wu, Y.; Huang, W.; Liu, D.; Huang, X.Y.; Zhang, X.M.; Ke, A.W. LncRNA SNHG3 induces EMT and sorafenib resistance by modulating the miR-128/CD151 pathway in hepatocellular carcinoma. J. Cell. Physiol. 2019, 234, 2788–2794. [Google Scholar] [CrossRef]
  407. Qin, X.; Wang, S. LncASAP1-IT1 promotes hepatocellular carcinoma progression through the regulation of the miR-1294/TGFBR1 pathway in vitro and in vivo. J. Gastrointest. Oncol. 2023, 14, 1451–1461. [Google Scholar] [CrossRef] [PubMed]
  408. Zhou, Y.; Huang, Y.; Dai, T.; Hua, Z.; Xu, J.; Lin, Y.; Han, L.; Yue, X.; Ho, L.; Lu, J. LncRNA TTN-AS1 intensifies sorafenib resistance in hepatocellular carcinoma by sponging miR-16-5p and upregulation of cyclin E1. Biomed. Pharmacother. 2021, 133, 111030. [Google Scholar] [CrossRef]
  409. Zhang, M.; Zhang, H.; Hong, H.; Zhang, Z. MiR-374b re-sensitizes hepatocellular carcinoma cells to sorafenib therapy by antagonizing PKM2 mediated glycolysis pathway. Am. J. Cancer Res. 2019, 9, 765–778. [Google Scholar]
  410. Yang, S.; Wang, M.; Yang, L.; Li, Y.; Ma, Y.; Peng, X.; Li, X.; Li, B.; Jin, H.; Li, H. MicroRNA-375 Targets ATG14 to Inhibit Autophagy and Sensitize Hepatocellular Carcinoma Cells to Sorafenib. OncoTargets Ther. 2020, 13, 3557–3570. [Google Scholar] [CrossRef]
  411. Ji, L.; Lin, Z.; Wan, Z.; Xia, S.; Jiang, S.; Cen, D.; Cai, L.; Xu, J.; Cai, X. miR-486-3p mediates hepatocellular carcinoma sorafenib resistance by targeting FGFR4 and EGFR. Cell Death Dis. 2020, 11, 250. [Google Scholar] [CrossRef] [PubMed]
  412. Yang, B.; Wang, C.; Xie, H.; Wang, Y.; Huang, J.; Rong, Y.; Zhang, H.; Kong, H.; Yang, Y.; Lu, Y. MicroRNA-3163 targets ADAM-17 and enhances the sensitivity of hepatocellular carcinoma cells to molecular targeted agents. Cell Death Dis. 2019, 10, 784. [Google Scholar] [CrossRef] [PubMed]
  413. Niture, S.; Gadi, S.; Qi, Q.; Gyamfi, M.A.; Varghese, R.S.; Rios-Colon, L.; Chimeh, U.; Vandana; Ressom, H.W.; Kumar, D. MicroRNA-483-5p Inhibits Hepatocellular Carcinoma Cell Proliferation, Cell Steatosis, and Fibrosis by Targeting PPARα and TIMP2. Cancers 2023, 15, 1715. [Google Scholar] [CrossRef]
  414. Song, M.; Ma, L.; Shen, C.; Liu, W.; Zhang, P.; Bi, R.; Zhao, C. FGD5-AS1/miR-5590-3p/PINK1 induces Lenvatinib resistance in hepatocellular carcinoma. Cell. Signal. 2023, 111, 110828. [Google Scholar] [CrossRef]
  415. Tan, W.; Lin, Z.; Chen, X.; Li, W.; Zhu, S.; Wei, Y.; Huo, L.; Chen, Y.; Shang, C. miR-126-3p contributes to sorafenib resistance in hepatocellular carcinoma via downregulating SPRED1. Ann. Transl. Med. 2021, 9, 38. [Google Scholar] [CrossRef] [PubMed]
  416. Huang, Y.; Zhang, J.; Li, H.; Peng, H.; Gu, M.; Wang, H. miR-96 regulates liver tumor-initiating cells expansion by targeting TP53INP1 and predicts Sorafenib resistance. J. Cancer 2020, 11, 6545–6555. [Google Scholar] [CrossRef] [PubMed]
  417. He, H.; Zhou, J.; Cheng, F.; Li, H.; Quan, Y. MiR-3677-3p promotes development and sorafenib resistance of hepatitis B-related hepatocellular carcinoma by inhibiting FOXM1 ubiquitination. Hum. Cell 2023, 36, 1773–1789. [Google Scholar] [CrossRef] [PubMed]
  418. Tang, X.; Yang, W.; Shu, Z.; Shen, X.; Zhang, W.; Cen, C.; Cao, L.; Zhang, M.; Zheng, S.; Yu, J. MicroRNA-223 promotes hepatocellular carcinoma cell resistance to sorafenib by targeting FBW7. Oncol. Rep. 2019, 41, 1231–1237. [Google Scholar] [CrossRef]
  419. Elhefnawi, M.; Salah, Z.; Soliman, B. The promise of miRNA replacement therapy for hepatocellular carcinoma. Curr. Gene Ther. 2019, 19, 290–304. [Google Scholar] [CrossRef] [PubMed]
  420. Moshiri, F.; Callegari, E.; D’Abundo, L.; Corrà, F.; Lupini, L.; Sabbioni, S.; Negrini, M. Inhibiting the oncogenic mir-221 by microRNA sponge: Toward microRNA-based therapeutics for hepatocellular carcinoma. Gastroenterol. Hepatol. Bed Bench 2014, 7, 43. [Google Scholar]
  421. Kim, T.; Croce, C.M. MicroRNA: Trends in clinical trials of cancer diagnosis and therapy strategies. Exp. Mol. Med. 2023, 55, 1314–1321. [Google Scholar] [CrossRef]
  422. Reda El Sayed, S.; Cristante, J.; Guyon, L.; Denis, J.; Chabre, O.; Cherradi, N. MicroRNA Therapeutics in Cancer: Current Advances and Challenges. Cancers 2021, 13, 2680. [Google Scholar] [CrossRef]
  423. Kalfert, D.; Ludvikova, M.; Pesta, M.; Ludvik, J.; Dostalova, L.; Kholová, I. Multifunctional Roles of miR-34a in Cancer: A Review with the Emphasis on Head and Neck Squamous Cell Carcinoma and Thyroid Cancer with Clinical Implications. Diagnostics 2020, 10, 563. [Google Scholar] [CrossRef]
  424. Bader, A.G. miR-34—A microRNA replacement therapy is headed to the clinic. Front. Genet. 2012, 3, 120. [Google Scholar] [CrossRef]
  425. Beg, M.S.; Brenner, A.J.; Sachdev, J.; Borad, M.; Kang, Y.K.; Stoudemire, J.; Smith, S.; Bader, A.G.; Kim, S.; Hong, D.S. Phase I study of MRX34, a liposomal miR-34a mimic, administered twice weekly in patients with advanced solid tumors. Investig. New Drugs 2017, 35, 180–188. [Google Scholar] [CrossRef]
  426. Jivrajani, M.; Nivsarkar, M. Ligand-targeted bacterial minicells: Futuristic nano-sized drug delivery system for the efficient and cost effective delivery of shRNA to cancer cells. Nanomedicine 2016, 12, 2485–2498. [Google Scholar] [CrossRef] [PubMed]
  427. Khordadmehr, M.; Shahbazi, R.; Sadreddini, S.; Baradaran, B. miR-193: A new weapon against cancer. J. Cell. Physiol. 2019, 234, 16861–16872. [Google Scholar] [CrossRef] [PubMed]
  428. Telford, B.J.; Yahyanejad, S.; de Gunst, T.; den Boer, H.C.; Vos, R.M.; Stegink, M.; van den Bosch, M.T.; Alemdehy, M.F.; van Pinxteren, L.A.; Schaapveld, R.Q. Multi-modal effects of 1B3, a novel synthetic miR-193a-3p mimic, support strong potential for therapeutic intervention in oncology. Oncotarget 2021, 12, 422. [Google Scholar] [CrossRef] [PubMed]
  429. van den Bosch, M.T.; Yahyanejad, S.; Alemdehy, M.F.; Telford, B.J.; de Gunst, T.; den Boer, H.C.; Vos, R.M.; Stegink, M.; van Pinxteren, L.A.; Schaapveld, R.Q. Transcriptome-wide analysis reveals insight into tumor suppressor functions of 1B3, a novel synthetic miR-193a-3p mimic. Mol. Ther. Nucleic Acids 2021, 23, 1161–1171. [Google Scholar] [CrossRef]
  430. Hsu, S.H.; Wang, B.; Kota, J.; Yu, J.; Costinean, S.; Kutay, H.; Yu, L.; Bai, S.; La Perle, K.; Chivukula, R.R.; et al. Essential metabolic, anti-inflammatory, and anti-tumorigenic functions of miR-122 in liver. J. Clin. Investig. 2012, 122, 2871–2883. [Google Scholar] [CrossRef]
  431. Bai, S.; Nasser, M.W.; Wang, B.; Hsu, S.H.; Datta, J.; Kutay, H.; Yadav, A.; Nuovo, G.; Kumar, P.; Ghoshal, K. MicroRNA-122 inhibits tumorigenic properties of hepatocellular carcinoma cells and sensitizes these cells to sorafenib. J. Biol. Chem. 2009, 284, 32015–32027. [Google Scholar] [CrossRef] [PubMed]
  432. Tsai, W.C.; Hsu, P.W.; Lai, T.C.; Chau, G.Y.; Lin, C.W.; Chen, C.M.; Lin, C.D.; Liao, Y.L.; Wang, J.L.; Chau, Y.P.; et al. MicroRNA-122, a tumor suppressor microRNA that regulates intrahepatic metastasis of hepatocellular carcinoma. Hepatology 2009, 49, 1571–1582. [Google Scholar] [CrossRef]
  433. Hsu, S.H.; Yu, B.; Wang, X.; Lu, Y.; Schmidt, C.R.; Lee, R.J.; Lee, L.J.; Jacob, S.T.; Ghoshal, K. Cationic lipid nanoparticles for therapeutic delivery of siRNA and miRNA to murine liver tumor. Nanomedicine 2013, 9, 1169–1180. [Google Scholar] [CrossRef]
  434. Kota, J.; Chivukula, R.R.; O’Donnell, K.A.; Wentzel, E.A.; Montgomery, C.L.; Hwang, H.W.; Chang, T.C.; Vivekanandan, P.; Torbenson, M.; Clark, K.R.; et al. Therapeutic microRNA delivery suppresses tumorigenesis in a murine liver cancer model. Cell 2009, 137, 1005–1017. [Google Scholar] [CrossRef]
  435. Querfeld, C.; Foss, F.M.; Kim, Y.H.; Pinter-Brown, L.; William, B.M.; Porcu, P.; Pacheco, T.; Haverkos, B.M.; DeSimone, J.; Guitart, J.; et al. Phase 1 Trial of Cobomarsen, an Inhibitor of Mir-155, in Cutaneous T Cell Lymphoma. Blood 2018, 132, 2903. [Google Scholar] [CrossRef]
  436. James, A.; Ruckman, J.; Pestano, L.; Hopkins, R.; Rodgers, R.; Marshall, W.; Rubin, P.; Escolar, D. SOLAR: A phase 2, global, randomized, active comparator study to investigate the efficacy and safety of cobomarsen in subjects with mycosis fungoides (MF). Hematol. Oncol. 2019, 37, 562–563. [Google Scholar] [CrossRef]
  437. Dhuri, K.; Pradeep, S.P.; Shi, J.; Anastasiadou, E.; Slack, F.J.; Gupta, A.; Zhong, X.B.; Bahal, R. Simultaneous Targeting of Multiple oncomiRs with Phosphorothioate or PNA-Based Anti-miRs in Lymphoma Cell Lines. Pharm. Res. 2022, 39, 2709–2720. [Google Scholar] [CrossRef]
  438. Zhang, M.; Zhou, X.; Wang, B.; Yung, B.C.; Lee, L.J.; Ghoshal, K.; Lee, R.J. Lactosylated gramicidin-based lipid nanoparticles (Lac-GLN) for targeted delivery of anti-miR-155 to hepatocellular carcinoma. J. Control Release 2013, 168, 251–261. [Google Scholar] [CrossRef] [PubMed]
  439. Shao, S.; Hu, Q.; Wu, W.; Wang, M.; Huang, J.; Zhao, X.; Tang, G.; Liang, T. Tumor-triggered personalized microRNA cocktail therapy for hepatocellular carcinoma. Biomater. Sci. 2020, 8, 6579–6591. [Google Scholar] [CrossRef] [PubMed]
  440. Sheedy, P.; Medarova, Z. The fundamental role of miR-10b in metastatic cancer. Am. J. Cancer Res. 2018, 8, 1674–1688. [Google Scholar] [PubMed]
  441. Liang, G.; Zhu, Y.; Ali, D.J.; Tian, T.; Xu, H.; Si, K.; Sun, B.; Chen, B.; Xiao, Z. Engineered exosomes for targeted co-delivery of miR-21 inhibitor and chemotherapeutics to reverse drug resistance in colon cancer. J. Nanobiotechnol. 2020, 18, 10. [Google Scholar] [CrossRef]
  442. Iacomino, G.; Siani, A. Role of microRNAs in obesity and obesity-related diseases. Genes Nutr. 2017, 12, 23. [Google Scholar] [CrossRef]
  443. Gebert, L.F.; Rebhan, M.A.; Crivelli, S.E.; Denzler, R.; Stoffel, M.; Hall, J. Miravirsen (SPC3649) can inhibit the biogenesis of miR-122. Nucleic Acids Res. 2014, 42, 609–621. [Google Scholar] [CrossRef]
  444. van der Ree, M.H.; de Vree, J.M.; Stelma, F.; Willemse, S.; van der Valk, M.; Rietdijk, S.; Molenkamp, R.; Schinkel, J.; Van Nuenen, A.C.; Beuers, U. Safety, tolerability, and antiviral effect of RG-101 in patients with chronic hepatitis C: A phase 1B, double-blind, randomised controlled trial. Lancet 2017, 389, 709–717. [Google Scholar] [CrossRef]
  445. Deng, Y.; Campbell, F.; Han, K.; Theodore, D.; Deeg, M.; Huang, M.; Hamatake, R.; Lahiri, S.; Chen, S.; Horvath, G.; et al. Randomized clinical trials towards a single-visit cure for chronic hepatitis C: Oral GSK2878175 and injectable RG-101 in chronic hepatitis C patients and long-acting injectable GSK2878175 in healthy participants. J. Viral Hepat. 2020, 27, 699–708. [Google Scholar] [CrossRef]
  446. Hong, D.S.; Kang, Y.-K.; Borad, M.; Sachdev, J.; Ejadi, S.; Lim, H.Y.; Brenner, A.J.; Park, K.; Lee, J.-L.; Kim, T.-Y. Phase 1 study of MRX34, a liposomal miR-34a mimic, in patients with advanced solid tumours. Br. J. Cancer 2020, 122, 1630–1637. [Google Scholar] [CrossRef] [PubMed]
  447. Xu, Y.; Liu, N.; Wei, Y.; Zhou, D.; Lin, R.; Wang, X.; Shi, B. Anticancer effects of miR-124 delivered by BM-MSC derived exosomes on cell proliferation, epithelial mesenchymal transition, and chemotherapy sensitivity of pancreatic cancer cells. Aging 2020, 12, 19660–19676. [Google Scholar] [CrossRef] [PubMed]
  448. Ghosh, S.; Lazarus, D.; Robertson, N.; Crosier, E.; Liu, P.; Medarova, Z. Abstract 548: The microRNA-10b targeted therapeutic, TTX-MC138, is effective in preclinical pancreatic adenocarcinoma. Cancer Res. 2023, 83, 548. [Google Scholar] [CrossRef]
  449. Preethi, K.A.; Selvakumar, S.C.; Ross, K.; Jayaraman, S.; Tusubira, D.; Sekar, D. Liquid biopsy: Exosomal microRNAs as novel diagnostic and prognostic biomarkers in cancer. Mol. Cancer 2022, 21, 54. [Google Scholar] [CrossRef]
Figure 1. Schematic of miRNA biogenesis, illustrating the steps from transcription to the formation of functional miRNAs within the RNA-induced silencing complex [17] for mRNA targeting and gene expression regulation. *** The guide strand in the RISC complex is retained, while the passenger strand is removed and subsequently degraded. Abbreviations: DGCR8—DiGeorge Syndrome Critical Region 8; PACT—Protein Activator of the Interferon-Induced Protein Kinase; TRBP—TAR RNA-binding protein; DICER—Endoribonuclease Dicer; AGO2—Argonaute-2; RISC—RNA-induced silencing complex. This figure was sourced from BioRender.
Figure 1. Schematic of miRNA biogenesis, illustrating the steps from transcription to the formation of functional miRNAs within the RNA-induced silencing complex [17] for mRNA targeting and gene expression regulation. *** The guide strand in the RISC complex is retained, while the passenger strand is removed and subsequently degraded. Abbreviations: DGCR8—DiGeorge Syndrome Critical Region 8; PACT—Protein Activator of the Interferon-Induced Protein Kinase; TRBP—TAR RNA-binding protein; DICER—Endoribonuclease Dicer; AGO2—Argonaute-2; RISC—RNA-induced silencing complex. This figure was sourced from BioRender.
Ijms 25 09393 g001
Figure 2. miRNAs regulate different stages of cell cycle progression in HCC, shedding light on the intricate roles of these molecules in modulating cell division and proliferation. Abbreviations: INK4—inhibitors of cyclin-dependent kinases; CIP/KIP—cyclin-dependent kinase (CDK) Inhibitory Proteins; CDC25C—Cell Division Cycle 25 C; CDC25A—Cell Division Cycle 25 A. This figure was sourced from BioRender.
Figure 2. miRNAs regulate different stages of cell cycle progression in HCC, shedding light on the intricate roles of these molecules in modulating cell division and proliferation. Abbreviations: INK4—inhibitors of cyclin-dependent kinases; CIP/KIP—cyclin-dependent kinase (CDK) Inhibitory Proteins; CDC25C—Cell Division Cycle 25 C; CDC25A—Cell Division Cycle 25 A. This figure was sourced from BioRender.
Ijms 25 09393 g002
Figure 3. miRNAs play a pivotal role in maintaining the delicate balance between cell survival and death by modulating the activity of key pathways such as PI3K/AKT/mTOR, MAPK, TGF-β, Wnt/β-catenin, and JAK/STAT. Pro-apoptotic miRNAs (a): Some miRNAs act as pro-apoptotic regulators targeting anti-apoptotic genes in signaling pathways, promoting programmed cell death. Anti-apoptotic miRNAs (b): Conversely, certain miRNAs serve as anti-apoptotic regulators by targeting pro-apoptotic genes, inhibiting apoptosis and promoting cell survival. Abbreviations: CKIα—Casein Kinase I alpha; GSK-3β—Glycogen Synthase Kinase 3 beta; TCF—T Cell Factor; LEF—Lymphoid Enhancer-Binding Factor; PDCD4—Programmed Cell Death Protein 4; FADD—Fas-Associated protein with Death Domain; PHLDA1—Pleckstrin Homology-like Domain, Family A, Member 1; Smurf2—SMAD-specific E3 ubiquitin protein ligase 2; IRAK—Interleukin-1 Receptor-Associated Kinase; TAK1—Transforming growth factor Beta-Activated Kinase 1. This figure was sourced from BioRender.
Figure 3. miRNAs play a pivotal role in maintaining the delicate balance between cell survival and death by modulating the activity of key pathways such as PI3K/AKT/mTOR, MAPK, TGF-β, Wnt/β-catenin, and JAK/STAT. Pro-apoptotic miRNAs (a): Some miRNAs act as pro-apoptotic regulators targeting anti-apoptotic genes in signaling pathways, promoting programmed cell death. Anti-apoptotic miRNAs (b): Conversely, certain miRNAs serve as anti-apoptotic regulators by targeting pro-apoptotic genes, inhibiting apoptosis and promoting cell survival. Abbreviations: CKIα—Casein Kinase I alpha; GSK-3β—Glycogen Synthase Kinase 3 beta; TCF—T Cell Factor; LEF—Lymphoid Enhancer-Binding Factor; PDCD4—Programmed Cell Death Protein 4; FADD—Fas-Associated protein with Death Domain; PHLDA1—Pleckstrin Homology-like Domain, Family A, Member 1; Smurf2—SMAD-specific E3 ubiquitin protein ligase 2; IRAK—Interleukin-1 Receptor-Associated Kinase; TAK1—Transforming growth factor Beta-Activated Kinase 1. This figure was sourced from BioRender.
Ijms 25 09393 g003aIjms 25 09393 g003b
Figure 4. The schematic representation highlights how miRNAs involved in the metastatic process through modulation of key molecular pathways such as TGFβ, PI3K/AKT, Wnt/β-catenin, JNK/STAT, and Hippo/YAP, which are vital for cancer cell invasion and migration. Abbreviations: SERPINE1—Serpin Family E Member 1; B4GALT3—Beta-1,4 Galactosyltransferase 3; NCOR2—Nuclear Receptor Corepressor 2; NKD1—Naked Cuticle Homolog 1; DKK2—Dickkopf WNT Signaling Pathway Inhibitor 2; SIK1—Salt Inducible Kinase 1; CRKL—Crk-Like protein; NRG1—Neuregulin 1; MTA1—Metastasis-Associated Protein 1; BRF2—RNA Polymerase III Transcription Initiation Factor Subunit. This figure was sourced from BioRender.
Figure 4. The schematic representation highlights how miRNAs involved in the metastatic process through modulation of key molecular pathways such as TGFβ, PI3K/AKT, Wnt/β-catenin, JNK/STAT, and Hippo/YAP, which are vital for cancer cell invasion and migration. Abbreviations: SERPINE1—Serpin Family E Member 1; B4GALT3—Beta-1,4 Galactosyltransferase 3; NCOR2—Nuclear Receptor Corepressor 2; NKD1—Naked Cuticle Homolog 1; DKK2—Dickkopf WNT Signaling Pathway Inhibitor 2; SIK1—Salt Inducible Kinase 1; CRKL—Crk-Like protein; NRG1—Neuregulin 1; MTA1—Metastasis-Associated Protein 1; BRF2—RNA Polymerase III Transcription Initiation Factor Subunit. This figure was sourced from BioRender.
Ijms 25 09393 g004
Figure 5. miRNAs influence complex interactions between cancer cells and their environment, modulating processes like fibroblast activation, EMT, angiogenesis, and immune evasion. Abbreviations: TGFβ—Transforming Growth Factor beta; VEGF—Vascular Endothelial Growth Factor; CSF—Colony Stimulating Factor; NK—Natural Killer; DC—Dendritic Cell. This figure was sourced from BioRender.
Figure 5. miRNAs influence complex interactions between cancer cells and their environment, modulating processes like fibroblast activation, EMT, angiogenesis, and immune evasion. Abbreviations: TGFβ—Transforming Growth Factor beta; VEGF—Vascular Endothelial Growth Factor; CSF—Colony Stimulating Factor; NK—Natural Killer; DC—Dendritic Cell. This figure was sourced from BioRender.
Ijms 25 09393 g005aIjms 25 09393 g005b
Figure 6. miRNAs play multifaceted roles in orchestrating chemoresistance in HCC, influencing various cellular processes like proliferation, EMT, autophagy, apoptosis, ferroptosis, metabolism, stemness, ABC transporter activity, hypoxia, and ROS regulation. Abbreviations: CSC—Cancer Stem Cells; ABC—ATP-Binding Cassette transporter; CYPs—Cytochrome P450 enzymes; G6Pase—Glucose-6-phosphatase. This figure was sourced from BioRender.
Figure 6. miRNAs play multifaceted roles in orchestrating chemoresistance in HCC, influencing various cellular processes like proliferation, EMT, autophagy, apoptosis, ferroptosis, metabolism, stemness, ABC transporter activity, hypoxia, and ROS regulation. Abbreviations: CSC—Cancer Stem Cells; ABC—ATP-Binding Cassette transporter; CYPs—Cytochrome P450 enzymes; G6Pase—Glucose-6-phosphatase. This figure was sourced from BioRender.
Ijms 25 09393 g006
Figure 7. Various tested microRNA delivery systems include liposomes, lipid nanoparticles, ligand-targeted micelles, magnetic nanoparticles, adeno-associated viruses, and exosomes for efficient and targeted delivery to specific tissues. Abbreviation: LNA-PS—Locked Nucleic Acid-Phosphorothioate; ASO—Antisense Oligonucleotide; GalNAc—N-Acetylgalactosamine; PLGA—Poly(lactic-co-glycolic acid). This figure was sourced from BioRender.
Figure 7. Various tested microRNA delivery systems include liposomes, lipid nanoparticles, ligand-targeted micelles, magnetic nanoparticles, adeno-associated viruses, and exosomes for efficient and targeted delivery to specific tissues. Abbreviation: LNA-PS—Locked Nucleic Acid-Phosphorothioate; ASO—Antisense Oligonucleotide; GalNAc—N-Acetylgalactosamine; PLGA—Poly(lactic-co-glycolic acid). This figure was sourced from BioRender.
Ijms 25 09393 g007
Table 1. miRNAs regulating hepatocellular carcinoma cell survival.
Table 1. miRNAs regulating hepatocellular carcinoma cell survival.
miRNAExpression in LiverTarget GenesPathwayCellular ProcessRefs.
Tumor suppressor miRNAs
miR-204-5pDownRGS20N/AProliferation (−), Apoptosis (+), Cell cycle arrest (G0/G1)[101]
miR-497-5pDownANXA11N/AProliferation (−), Apoptosis (+), Cell cycle arrest (G0/G1)[102]
miR-377-3pDownRNF38
CPT1C
CPT1C-mediated fatty acid oxidationProliferation (−), Apoptosis (+)[103,104]
miR-559DownPARD3
GP73
N/AProliferation (−), Autophagy (−)[105,106]
miR-638DownEZH2N/AProliferation (−), Apoptosis (+), Autophagy (+)[107]
miR-199a-3pDownmTOR, c-Met
upregulation of ZHX1 and PUMA
mTOR pathway, ZHX1/PUMA signalingProliferation (−), Apoptosis (+), Cell cycle arrest (G0/G1)[108,109]
miR-27a-3pDownN/API3K/Akt signalingProliferation (−), Apoptosis (+), Cell cycle arrest (G0/G1)[110]
miR-9DownHMGA2N/AProliferation (−), Cell cycle arrest (G0/G1)[32]
miR-185DownRHEB, RICTOR, and AKT1AKT1 pathwayCell cycle arrest (G0/G1), Apoptosis (+), Autophagy (+)[94]
miR-424-5pDownE2F7N/AProliferation (−), Cell cycle arrest (G0/G1)[33]
miR-193a-5pDownNUSAP1N/AProliferation (−), Cell cycle arrest (G0/G1), Apoptosis (+)[42]
miR-621DownCAPRIN1CAPRIN1/CCND2/c-MYC axisProliferation (−), Cell cycle arrest (G0/G1)[34]
miR-125b-5pDownKIAA1522cyclinD1, CDK6, cyclin E and CDK2, and p21Proliferation (−), Cell cycle arrest (G0/G1),
Apoptosis (+)
[35]
miR-29 family (miR-29a, b, c)DownRPS15Acyclin A and cyclin D1, p21Proliferation (−), Cell cycle arrest (G0/G1),
Apoptosis (+)
[37]
miR-450b-3pDownPGK1AKT signalingProliferation (−), Cell cycle arrest (G0/G1)[38]
miR-495DownCTRP3N/AProliferation (−), Cell cycle arrest (G0/G1)[111]
miR-23b-5pDownFOXM1c-MYC and cyclin D1 axisProliferation (−), Cell cycle arrest (G0/G1)[112]
miR-128-3pDownc-Metc-Met signalingProliferation (−), Cell cycle arrest (G0/G1), Apoptosis (+)[113]
miR-875-5pDownAEG-1N/AProliferation (−), Cell cycle arrest (G1/S)[114]
miR-0308-3pDownCDK6/Cyclin D1N/AProliferation (−), Cell cycle arrest (G1/S)[45]
miR-378DownCyclin D1, Bcl-2, Akt, β-catenin and SurvivinN/AProliferation (−), Cell cycle arrest (G1/S), Apoptosis (+)[59]
miR-874DownDORDOR/EGFR/ERK pathwayProliferation (−), Cell cycle arrest (G1/S)[115]
miR-30b-5pDownDNMT3A, USP37N/AProliferation (−), Cell cycle arrest (G1/S)[116]
miR-214-3pDownMELKN/AProliferation (−), Cell cycle arrest (G1/S), Apoptosis (+)[44]
miR-214DownWnt3aWnt/β-catenin pathwayProliferation (−), Cell cycle arrest (G1/S)[43]
miR-214-5pDownE2F2NF-κB pathwayProliferation (−), Apoptosis (+)[117]
miR-409-3pDownRAB10NF-κB pathwayProliferation (−), Apoptosis (+)[117]
miR-98-5pDownIGF2BP1N/AProliferation (−), Cell cycle arrest (G1/S), Apoptosis (+)[118]
miR-217DownMTDH, EZH2, cyclin-D1, KLF5N/AProliferation (−), Cell cycle arrest (G1/S), Apoptosis (+)[46,47,48]
miR-302a/dDownE2F7AKT/βcatenin/CCND1 pathway AKT1-cyclin D1 pathwayProliferation (−), Cell cycle arrest (G1/S), Apoptosis (+)[119]
miR-1914DownGPR39PI3K/AKT/mTOR pathwayProliferation (−), Cell cycle arrest (G1/S), Apoptosis (+)[120]
miR-206DowncMET, CCND1, and CDK6cMet signalingProliferation (−), Cell cycle arrest (G1/S), Apoptosis (+)[65]
miR-22-3pDownAKT2AKT/PI3K pathwayProliferation (−), Cell cycle arrest (G1/S), Apoptosis (+)[64]
miR-145-5pDownSPATS2, p21 and p27N/AProliferation (−), Cell cycle arrest (G1/S), Apoptosis (+)[121]
miR-203DownMAT2A, MAT2B, NRasAKT/PI3K pathway
RAS/MAPK signaling
Proliferation (−), Cell cycle arrest (S/G2), Apoptosis (+)[122]
let-7b-5pDownp21, CDC25B, HMGA2CDC25B/CDK1 axisProliferation (−)
Cell cycle arrest (G2/M)
[49,123]
miR-3613-3pDownBIRC5, CDK1, NUF2, ZWINT, and SPC24N/AProliferation (−)
Cell cycle arrest (G2/M)
[51]
miR-31-5pDownSP1SP1/cyclin D1Proliferation (−)
Cell cycle arrest (G2/M)
[50]
miR-217DownHMGA2N/AProliferation (−)
Cell cycle arrest (G2/M)
[124]
miR-448DownBCL-2N/AProliferation (−), Apoptosis (+)[61]
miR-223DownNLRP3, Rab1NLRP3 inflammasome pathway, mTOR pathwayProliferation (−),
Apoptosis (+)
[125]
miR-133bDownEGFR, SF3B4EGFR/PI3K/Akt/mTOR pathway
SF3B4/KLF4/KIP1/SNAI2 axis
Proliferation (−)
Apoptosis (+)
[62,126]
miR-181a-5pDownATG7N/AProliferation (−), Autophagy (−)[95]
miR-7DownmTOR, ATG5AutophagyProliferation (−), Autophagy (+)[98,127]
miR-192-3pDownXIAPNF-κB signalingApoptosis (+), Autophagy (−)[128]
miR-519dDownRab10AMPK Signaling PathwayApoptosis (+), Autophagy (+)[93]
miR-219-5pDownNEK6β-catenin/c-Myc pathwayProliferation (−)[129]
miR-375DownErbB2N/AProliferation (−)
Apoptosis (+)
[130]
miR-122DownmiR-21miR-21–PDCD4 pathwayProliferation (−)
Apoptosis (+)
[131]
miR-4651DownFOXP4N/AProliferation (−)
Apoptosis (+)
[132]
Let-7bDownIGF-1RN/AProliferation (−)
Apoptosis (+)
[133]
miR-26aDownJAK1JAK1-STAT3 signalingProliferation (−)
Apoptosis (+)
[67]
miR-26a/bDownULK1N/AApoptosis (+), Autophagy (−)[100]
miR-342-3pDownMCT1N/AProliferation (−)
Apoptosis (+), Necrosis (+)
[134]
miR-654-5pDownHSPB1N/AProliferation (−),
Ferroptosis (+)
[135]
miR-188-3pDownGPX4N/AProliferation (−),
Ferroptosis (+)
[136]
miR-612DownCoQ10HADHA-mediated MVA pathwayProliferation (−),
Ferroptosis (+)
[137]
Oncogenic miRNAs
miR-183-5pUpPDCD4N/AProliferation (+), Apoptosis (−)[138]
miR-494Upp27, PUMA, and PTENmTOR pathwayProliferation (+), Cell cycle (induces G1/S transition), Apoptosis (−)[52]
miR-3682-3pUpPHLDA1PHLDA1-Fas pathwayProliferation (+), Cell cycle (induces G1/S transition), Apoptosis (−)[54]
miR-191UpKLF6N/AProliferation (+), Cell cycle (induces G1/S transition)[53]
miR-10bUpCSMD1N/AProliferation (+), Cell cycle (induces G1/S transition), Apoptosis (−)[55]
miR-221-3pUpMGMTN/AProliferation (+), Cell cycle (induces G2/M transition), Apoptosis (−)[56]
miR-21-5pUpMELKAKT/mTOR signalingProliferation (+), Ferroptosis (−)[139]
miR-338-5pUpRTN4N/AProliferation (+), Apoptosis (−)[140]
miR-9-5pUpKlf4AKT/mTOR signalingProliferation (+), Apoptosis (−)[141]
miR-302dUpTGFBR2TGF-β signalingProliferation (+), Apoptosis (−)[142]
miR-371-5p, miR-373 and miR-543UpCasp-8N/ANecrosis (+)[85]
miR-106bUpDR4TRAIL pathwayProliferation (+), Apoptosis (−)[81]
miR-3174UpFOXO1N/AProliferation (+), Apoptosis (−)[143]
miR-33aUpPPARαN/AProliferation (+), Apoptosis (−)[75]
miR-155UpTLR3TLR3-NF-kB pathwayProliferation (+), Apoptosis (−)[80]
miR-93-5pUpERBB4TETs dependent DNA demethylationProliferation (+), Apoptosis (−)[144]
Note: (−) inhibiting effect; (+) inducing effect.
Table 2. miRNAs regulating stemness of hepatocellular carcinoma cells.
Table 2. miRNAs regulating stemness of hepatocellular carcinoma cells.
miRNATarget GenesPathwayStemnessRef.
Tumor suppressor miRNAs
miR-6838-5pCBX4ERK signaling(−)[154]
miR-148bNRP1ACVR1-BMP-Wnt axis(−)[157]
miR-148aACVR1Brg1/Sall4 axis(−)[156]
miR-296-5pBrg1AKT/βcatenin/CCND1 signaling(−)[158]
miR-302a/dE2F7N/A(−)[119]
miR-589-5pMAP3K8SOX9 signaling(−)[155]
miR-125a/bCD90N/A(−)[163]
miR-613SOX9PI3K/AKT/mTOR pathway(−)[164]
miR-100 and miR-125bIGF2hexosamine pathway(−)[165]
miR-325-3pDPAGT1N/A(−)[166]
miR-217HMGA2N/A(−)[124]
miR-448MAGEAAMPK signaling(−)[167]
Oncogenic miRNAs
miR-106b-5pPTENPTEN/PI3K/AKT pathway(+)[162]
miR-1246AXIN2, GSK3βWnt/β-catenin pathway(+)[160]
miR-5188FOXO1β-catenin/Wnt signaling(+)[161]
miR-454-3pCPEB1N/A(+)[168]
Note: (−) inhibiting effect; (+) inducing effect.
Table 3. miRNAs regulating hepatocellular carcinoma metastasis.
Table 3. miRNAs regulating hepatocellular carcinoma metastasis.
miRNATarget GenesPathwayCellular ProcessRefs.
Anti-metastatic miRNAs
miR-320aHMGB1, c-MycN/AN/A[202,203]
miR-29c-3pADAM12N/AN/A[204]
miR-188-5pFOXN2N/AN/A[205]
miR-142-3pZEB1, TGF-β1, HMGB1TGF-β pathwayEMT, Angiogenesis[206,207,208]
Exosomal miR-125bSMAD2TGFβ1/SMAD signalingEMT[209]
Exosomal miR-29bDNMT3bN/AEMT[210]
miR-10a-5pSKA1N/AEMT[211]
miR-2113WDR66N/AEMT[195]
miR-30eMTA1MTA1/ErbB2 axisEMT[212]
miR-517cKPNA2PI3K/AKT pathwayEMT[213]
miR-130a-3pARN/AEMT, Angiogenesis[176]
miR-526b-5pHGFHGF/c-Met pathwayEMT, Angiogenesis[214]
miR-4270-5pSATB2N/AEMT[215]
miR-124-3pCRKL, Sp1RAF/MEK/ERK1/2 pathwaysEMT[177,216]
miR-875-5pAEG1/MTDHN/AEMT[114]
miR-296-5pNRG1NRG1/ERBB2/ERBB3/RAS/MAPK/Fra-2 signalingEMT[159]
miR-449aNotch1Notch pathwayEMT[217]
miR-30aBeclin 1, Atg5Autophagy pathwayAutophagy[99]
miR-7FAKAkt pathwayN/A[190]
miR-148a-3pSMAD2TGF-β signalingN/A[218]
miR-298CTNND1Wnt/β-catenin signalingN/A[219]
miR-409-3pBRF2Wnt/β-catenin signalingN/A[220]
miR-466MTDH/AEG-1N/AN/A[221]
miR-138-5pFOXC1N/AN/A[222]
miR-455-5pIGF-1RIGF-1R-AKT-GLUT1 axisN/A[223]
miR-139-3pUCK2N/AN/A[224]
miR--219-5pNEK6β-catenin/c-Myc pathwayN/A[225]
miR-145SERPINE1HIF-1 signaling pathwayN/A[226]
miR-515–5pIL6IL-6/JNK/STAT3 signalingN/A[227]
miR-378a-3pPD-L1STAT3 signalingImmune response[228]
miR-199a/b-5pROCK1ROCK1/MLC and PI3K/Akt signalingN/A[229]
miR-378a-5pVEGFVEGF pathwayAngiogenesis[230]
miR-495IGF1RN/AN/A[231]
miR-194-3pMMP9N/AECM degradation[232]
miR-489MMP7N/AECM degradation[233]
miR-211-5pACSL4N/AN/A[175]
miR-342-3pAGR2N/AApoptosis, cell cycle, ECM degradation[234]
miR-767-3pSMIM7N/AN/A[235]
miR-139GDF10N/AN/A[236]
miR-15a-3pHMOX1N/AN/A[179]
miR-449 familySOX4TGF-β pathwayN/A[237]
miR-708SMAD3TGF-β pathwayN/A[238]
miR-300LEF-1N/AN/A[239]
miR-98EZH2, IL-10Wnt/β-catenin pathwayN/A[240,241]
Metastatic miRNAs
miR-362-3pCD82TGF-β signalingEMT[242]
miR-625-3pPDLIM5N/AEMT[243]
miR-92a-3pPTENAkt/Snail pathway, PI3K/AKT/mTOR SignalingEMT[244,245]
miR-93-5pdirectly upregulates MAP3K2, inhibits ERBB4 and TETsMAP3K2/
p38-JNK/p21 signaling pathway
N/A[144,201]
miR-10a-5pNCOR2Wnt/β-catenin pathwayEMT[246]
miR-HCC3TNFRSF19, RAB43N/AEMT[247]
miR-197Axin-2, NKD1, DKK2Wnt/β-catenin pathwayEMT[248]
miR-21-3pSMAD7SMAD7/YAP1 axisEMT[249]
miR-3691-5pPTENPI3K/Akt signalingN/A[250]
miR-106bPTENN/AN/A[251]
miR-18aKLF4, Bcl2L10N/AN/A[196,197]
miR-106b-5pFOG2N/AN/A[199]
miR-376c-3pARID2N/AN/A[200]
miR-182-5pFOXO3aAKT/FOXO3a pathway, Wnt/β-catenin signalingN/A[252]
miR-769-5pRYBPN/AN/A[253]
miR-1251-5pAKAP12N/AN/A[254]
miR-483-5pPositive regulator of IGF-IIN/AN/A[255]
miR-106b-5pGPM6AAKT/ERK signalingN/A[256]
Exosomal miR-25-5pSIK1 Fbxw7, LRRC7Wnt/β-catenin signalingN/A[198,257,258]
Exosomal miR-21PTENPDK1/AKT signallingN/A[259]
Exosomal miR-1273fLHX6Wnt/β-catenin signalingEMT[260]
Exosomal miR-92a-3pPTENAkt/Snail signalingN/A[244]
Exosomal miR-224GNMTN/AN/A[261]
Extracellular vesicle (EV)-miR-3129TXNIPN/AN/A[262]
Exosomal miR-1247-3pB4GALT3β1-integrin/NF-κB signalingN/A[263]
Table 4. miRNAs regulating tumor microenvironment in hepatocellular carcinoma.
Table 4. miRNAs regulating tumor microenvironment in hepatocellular carcinoma.
miRNASource of Cell TypeTarget GenePathwayFunctionRef.
Tumor suppressor miRNAs
Exosomal miR-320aCAFsPBX3MAPK pathwayInhibits HCC proliferation and metastasis[302]
Exosomal miR-150-3pCAFsN/AN/AInhibits HCC invasion[282]
miR-101-3p/miR-490-3pCAFsTGFBR1N/AInduce infiltration of myeloid-derived suppressor cells, regulatory T cells, and M2 macrophages[283]
miR-26aHCC cellsM-CSFPI3K/Akt pathwayInhibits macrophage recruitment and M2
polarization
[303]
Exosomal miR-628-5pM1 macrophagesMETTL14circFUT8/miR-552-3p/CHMP4B pathwayInhibits HCC progression[304]
miR-144/miR-451aHCC cellsHGF, MIFN/AInduce macrophage M1 polarization[288]
Exosomal miR-125a/bM2 macrophagesCD90N/AInhibit cell proliferation and stem cell properties[163]
miR-148bHCC cellsCSF1CSF1/CSF1R pathwayInhibits TAM infiltration, HCC growth and metastasis[289]
miR-28-5pHCC cellsIL-34TGFβ1 signalingInhibits TAM infiltration, HCC growth and metastasis[290]
miR-200HCC cellsPD-L1N/AInduces CD8+ T cells viability, inhibits metastasis[305]
miR-145HCC cellsPD-L1PI3K/AKT signalingInhibits tumor growth, EMT, and metastasis[306]
miR-194-5pHCC cellsPD-L1 and PD-L2N/AReduces cytotoxic T cells apoptosis[307]
miR-223HCC cellsHIF1αCD39/CD73-adenosine pathwayInhibits infiltration of PD-1+ T cells and PD-L1+ macrophages, inhibits angiogenesis[293]
miR-329-3pHCC cellsKDM1AN/AInhibits the expression of PD-L1 in HCC cells via increasing MEF2D methylation, inhibits tumor growth[294]
miR-200cHBV+ HCC cellsCD274HBV-pSTAT3-SALL4-miR-200c-PD-L1 axisInhibits HBV-mediated PD-L1 expression and CD8+ T cell exhaustion[295]
miR-22HCC cellsHIF1αRetinoic acid signalingInhibits IL17 signaling, expands cytotoxic T cells and reduces Treg[296]
miR-30-5pHCC cellsSnailSnail-DDP4- CXCL10 axisInduces CD8+ T cell infiltration[299]
miR-570HCC cellsCD31 and VEGFN/AIncreases CD8+IFN-γ+ T cells, induces apoptosis, inhibits angiogenesis[308]
miR-374bCIK cellsPD-1N/AInduces cytotoxicity of cytokine-induced killer cells[309]
miR-1258NK cells and DC cellsN/AN/AStimulates TLR7/8 expression, activates NKs and promotes DCs maturation, inhibits tumor growth and metastasis[300]
Oncogenic miRNAs
Exosomal miR-1247-3pHCC cellsB4GALT3β1-integrin–NF-κB signalingInduces CAFs activation and lung metastasis[263]
Exosomal miR-4508 HCC cellsRFX1RFX1-IL17A-p38 MAPK-NF-κB pathwayActivates lung fibroblasts and induces lung metastasis[281]
Exosomal miR-21HCC cellsPTENPDK1/AKT signalingConverts normal HSCs to CAFs, induces angiogenesis[259]
Exosomal miR-200b-3pHCC cellsZEB1JAK/STAT signalingInduces macrophage M2 polarization[310]
Exosomal miR-4669HCC cellsN/AN/AInduces M2 macrophage polarization, migration ability, and sorafenib resistance[286]
Exosoma miR-92a-2-5pM2 macrophagesARAR/PHLPP/p-AKT/β-catenin signalingInduces HCC cells invasion[311]
Exosomal miR-27a-3pM2 macrophagesTXNIPN/AInduces HCC cells stemness, proliferation, drug resistance, migration, invasion, and tumorigenicity [312]
Exosomal miR-660-5pM2 macrophagesKLF3N/AInduces EMT[313]
Exosomal miR-452-5pHCC cellsTIMP3N/AInduces M2 macrophage polarization,[287]
miR-889HCC cellsMICBN/AReduces NK cell-mediated cytotoxicity[301]
miR-561-5pHCC cellsCX3CL1STAT3 signalingReduces CX3CR1+ NK cell infiltration, induces tumor growth and lung metastasis[4]
Table 5. miRNAs regulating hepatocellular carcinoma angiogenesis.
Table 5. miRNAs regulating hepatocellular carcinoma angiogenesis.
miRNATarget GenesPathwayRef.
Anti-angiogenic miRNAs
miR-26aHGFHGF/c-Met pathway[331]
miR-526b-5pHGFHGF/c-Met pathway[214]
miR-200bTranscription factor ERGN/A[322]
miR-203aHOXD3N/A[318]
miR-144-3pSGK3N/A[320]
miR-375PDGFC, AEG-1N/A[332]
miR-3064-5pFOXA1FOXA1/CD24/Src pathway[323]
miR-199a-3pMMP2, HGF, VEGFA, VEGFR1N/A[316]
miR-378a-5pVEGFVEGF pathway[230]
miR-223HIF1αCD39/CD73-adenosine pathway[293]
miR-1296E2F7N/A[333]
Pro-angiogenic miRNAs
Let-7i-5pTSP1N/A[334]
miR-210-3pSMAD4, STAT6N/A[330]
miR-130b-3pHOXA5N/A[328]
Table 6. miRNAs regulating chemotherapy resistance in hepatocellular carcinoma.
Table 6. miRNAs regulating chemotherapy resistance in hepatocellular carcinoma.
miRNATarget GenesDrug ResponseCellular ProcessPathwayRef.
miRNAs improving drug sensitivity
miR-375MDR1, AEG1,
YAP1, and ATG7
doxorubicinProliferation, autophagyN/A[349]
miR-223ABCB1doxorubicinN/AN/A[350]
miR-125bHIF1A, YBX1, PDGFRBdoxorubicinstemnessHIF1α/PDGFβ/pAKT[338]
miR-383EIF5A2doxorubicinProliferation, apoptosisN/A[351]
miR-140-5pPIN1doxorubicinProliferationN/A[352]
miR-590-5pYAP1doxorubicinProliferation, stemnessHippo signaling[353]
miR-122ABCB1 and ABCF2doxorubicinCell cycleN/A[354]
miR-122PKM2doxorubicinGlycolysisN/A[355]
miR-760Notch1doxorubicinProliferation, apoptosisNotch1/Hes1-PTEN/Akt Signaling[356]
miR-26a/bULK1doxorubicinAutophagy, apoptosisN/A[100]
miR-218-5pEIF5A2doxorubicinN/AN/A[357]
miR-325-3pDPAGT1doxorubicinN/AHexosamine pathway[166]
miR-223FOXO3adoxorubicinAutophagyN/A[339]
miR 361 5pMAP3K9cisplatinApoptosisN/A[341]
miR-610HDGFcisplatinProliferation and apoptosisN/A[342]
miR-10a-5pMSI1cisplatinProliferation and apoptosisAKT signaling[344]
miR-27a-3pABCB15-fluorouracil
cisplatin
Proliferation and apoptosisPI3K/Akt pathway[110,358]
miR-203aBMI15-fluorouracilProliferation, stemnessN/A[359]
miR-125bHK II5-fluorouracilGlycolysisN/A[360]
miR-216bMALAT15-fluorouracilAutophagyN/A[340]
miR-145TLR45-fluorouracilApoptosisN/A[361]
miR-138-5pHIF-1αRadiosensitivityN/AMigration/invasion, EMT[362]
miRNAs inducing drug resistance
miR-182TP53INP1cisplatinViabilityN/A[348]
miR-551aMEF2C5-fluorouracilViability and sphere formationN/A[347]
Exosomal miR-32-5pPTEN5-fluorouracilAngiogenesis, EMTPI3K/Akt pathway[363]
miR-24 and miR-221caspase 8/3TRAILAngiogenesisN/A[364]
Table 7. miRNAs regulating targeted therapy resistance in hepatocellular carcinoma.
Table 7. miRNAs regulating targeted therapy resistance in hepatocellular carcinoma.
miRNATarget GenesTargeted
Therapy Agent
Cellular ProcessPathwayRefs.
miRNAs improving drug sensitivity
Exosomal miR-744PAX2SorafenibN/AN/A[400]
miR-3689a-3pCCSSorafenibmitochondrial oxidative stress, ApoptosisCCS/SOD1 axis[389]
miR-124CAV1SorafenibStemnessN/A[401]
miR-338-3pRAB1BSorafenibApoptosis, invasionN/A[402]
miR-122IGF-1RSorafenibApoptosisRAS/RAF/ERK
signaling
[403]
miR-345-5pTOP2ASorafenibApoptosisN/A[388]
miR-122SerpinB3SorafenibApoptosisN/A[404]
miR-138-1-3pPAK5SorafenibApoptosisβ-catenin/ABCB1 signaling[386]
miR-654-5pHSPB1SorafenibFerroptosisN/A[135]
miR-182-5pN/ASorafenibEMTN/A[405]
miR-128CD151SorafenibEMTN/A[406]
miR-1294TGFβR1SorafenibEMTN/A[407]
miR-16-5pcyclin E1SorafenibCell cyclePTEN/Akt signaling[408]
miR-449a-5pPEA15, PPP1CA, TUFT1SorafenibN/AAKT and ERK signaling[372]
miR-4277CYP3A4SorafenibDrug metabolismN/A[373]
miR-374bhnRNPA1SorafenibAerobic glycolysisPKM2[409]
miR-30a-5pATG5, Beclin-1, CLCF1SorafenibAutophagy, aerobic glycolysisPI3K/AKT signaling[375,376]
miR-204ATG3SorafenibAutophagyN/A[377]
miR-378a-3pIGF1RSorafenibN/AERK/PI3K signaling[380]
miR-140-3pPXRSorafenibDrug clearanceN/A[383]
miR-375PDGFC, AEG-1, ATG14SorafenibAngiogenesis, autophagyN/A[332,410]
miR-124-3p.1AKT2, SIRT1SorafenibN/AFOXO3a pathway[385]
miR-142-3pATG5, ATG16L1SorafenibAutophagyN/A[378]
miR-148aPXRSorafenibDrug clearanceN/A[384]
miR-486-3pFGFR4, EGFRSorafenibN/AN/A[411]
miR-3163ADAM-17SorafenibN/ANotch signaling[412]
miR-483-5pPPARα, TIMP2SorafenibApoptosis, steatosis, fibrosisNotch signaling[413]
miR-3163ADAM17SorafenibN/ANotch signaling[412]
miR-5590-3pPINK1LenvatinibApoptosisN/A[414]
miR-128-3pc-MetLenvatinibApoptosis, cell cyclec-Met pathway[113]
miR-24-3pBCL2L2LenvatinibApoptosisN/A[395]
miR-34aBeclin-1LenvatinibAutophagyN/A[396]
miRNAs inducing drug resistance
miR-21-5pUSP24SorafenibAutophagyUSP24-SIRT7 axis[394]
miR-23a-3pACSL4SorafenibFerroptosisN/A[390]
miR-125b-5pAtaxin-1SorafenibEMTN/A[392]
miR-494p27, PUMA, PTEN, G6pcSorafenibCell cycle, Survival, invasion, stemness, glycogenolysis, gluconeogenesismTOR pathway[52]
miR-126-3pSPRED1SorafenibN/AERK signaling pathway[415]
miR-96TP53INP1Sorafenibpromotes liver T-ICs expansionN/A[416]
miR-3677-3pFBXO31SorafenibProliferation, invasion, stemness, apoptosisN/A[417]
miR-4669SIRT1SorafenibImmunosuppressive TMEN/A[286]
miR-223FBW7SorafenibN/AN/A[418]
miR-183-5p.1MUC15 LenvatinibProliferation, apoptosis, stemnessc-MET/PI3K/AKT/SOX2 signaling[397]
miR-520c-3pMBD2LenvatinibProliferation, cell cycleGF19/FGFR4/FRS2 signaling[398]
miR-3154HNF4αLenvatinibProliferation, apoptosis, stemnessN/A[399]
Table 8. miRNAs as therapeutics for cancer treatment.
Table 8. miRNAs as therapeutics for cancer treatment.
Therapeutic
Molecule
Target miRNADiseasePhaseDelivery PlatformRoute of
Administration
Ref.
miRNA replacement therapy
MRX34miR-34aAdvanced solid tumorsPhase ILiposomalIV[446]
TargomiRmiR-16MPM, NSCLCPhase IEGFR targeting minicellIV[86]
INT-1B3miR-193a-3pAdvanced solid tumorsPhase ILNPIV[428]
N/AmiR-124Pancreatic cancerPreclinicalExosomeSC[447]
N/AmiR-122HCCPreclinicalLNPIV[433]
N/AmiR-26aHCCPreclinicalMSCV-derived retroviral constructIV[434]
N/AmiR-22HCCPreclinicalAAVIV[402]
miRNA inhibition therapy
MRG-106 (cobomarsen)miR-155CTCL, CLL, DLBCL, ATLLPhase IILNAIT, SC, IV[436]
MiravirsenmiR-122Chronic HCVPhase IILNA-PS-modified ASOSC[443]
RG-101miR-122Chronic HCVPhase IIGalNAc-ASOSC[445]
TTX-MC138miR-10bPACPreclinicalNPIV[448]
N/AmiR-155HCCPreclinicalLNPIV[438]
N/AmiR-21CRCPreclinicalExosomeIV[441]
N/AmiR-21 and miR-155LymphomaN/ANP containing PS and PNAN/A[437]
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

Mahboobnia, K.; Beveridge, D.J.; Yeoh, G.C.; Kabir, T.D.; Leedman, P.J. MicroRNAs in Hepatocellular Carcinoma Pathogenesis: Insights into Mechanisms and Therapeutic Opportunities. Int. J. Mol. Sci. 2024, 25, 9393. https://doi.org/10.3390/ijms25179393

AMA Style

Mahboobnia K, Beveridge DJ, Yeoh GC, Kabir TD, Leedman PJ. MicroRNAs in Hepatocellular Carcinoma Pathogenesis: Insights into Mechanisms and Therapeutic Opportunities. International Journal of Molecular Sciences. 2024; 25(17):9393. https://doi.org/10.3390/ijms25179393

Chicago/Turabian Style

Mahboobnia, Khadijeh, Dianne J. Beveridge, George C. Yeoh, Tasnuva D. Kabir, and Peter J. Leedman. 2024. "MicroRNAs in Hepatocellular Carcinoma Pathogenesis: Insights into Mechanisms and Therapeutic Opportunities" International Journal of Molecular Sciences 25, no. 17: 9393. https://doi.org/10.3390/ijms25179393

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