Next Article in Journal
Effects of Donor Ages and Propagation Methods on Seedling Growth of Platycladus orientalis (L.) Franco in Winter
Next Article in Special Issue
Circulating microRNA Biomarker for Detecting Breast Cancer in High-Risk Benign Breast Tumors
Previous Article in Journal
Magnoflorine from Berberis vulgaris Roots—Impact on Hippocampal Neurons in Mice after Short-Term Exposure
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

An Insight into the Arising Role of MicroRNAs in Hepatocellular Carcinoma: Future Diagnostic and Therapeutic Approaches

by
Evangelos Koustas
1,2,*,
Eleni-Myrto Trifylli
2,
Panagiotis Sarantis
1,
Nikolaos Papadopoulos
3,
Konstantinos Papanikolopoulos
2,
Georgios Aloizos
2,
Christos Damaskos
4,5,
Nikolaos Garmpis
6,7,
Anna Garmpi
7,
Dimitris Matthaios
8 and
Michalis V. Karamouzis
1
1
Department of Biological Chemistry, Medical School, National and Kapodistrian University of Athens, 75, M. Asias Street, 11527 Athens, Greece
2
First Department of Internal Medicine, 417 Army Equity Fund Hospital, 11521 Athens, Greece
3
Second Department of Internal Medicine, 401 General Army Hospital of Athens, 11525 Athens, Greece
4
‘N.S. Christeas’ Laboratory of Experimental Surgery and Surgical Research, Medical School, National and Kapodistrian University of Athens, 11527 Athens, Greece
5
Renal Transplantation Unit, ‘Laiko’ General Hospital, 11527 Athens, Greece
6
Second Department of Propaedeutic Surgery, ‘Laiko’ General Hospital, Medical School, National and Kapodistrian University of Athens, 11527 Athens, Greece
7
First Department of Pathology, Medical School, National and Kapodistrian University of Athens, 11527 Athens, Greece
8
Oncology Department, General Hospital of Rhodes, 85100 Rhodes, Greece
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(8), 7168; https://doi.org/10.3390/ijms24087168
Submission received: 5 February 2023 / Revised: 3 April 2023 / Accepted: 6 April 2023 / Published: 12 April 2023

Abstract

:
Hepatocellular carcinoma (HCC) constitutes a frequent highly malignant form of primary liver cancer and is the third cause of death attributable to malignancy. Despite the improvement in the therapeutic strategies with the exploration of novel pharmacological agents, the survival rate for HCC is still low. Shedding light on the multiplex genetic and epigenetic background of HCC, such as on the emerging role of microRNAs, is considered quite promising for the diagnosis and the prediction of this malignancy, as well as for combatting drug resistance. MicroRNAs (miRNAs) constitute small noncoding RNA sequences, which play a key role in the regulation of several signaling and metabolic pathways, as well as of pivotal cellular functions such as autophagy, apoptosis, and cell proliferation. It is also demonstrated that miRNAs are significantly implicated in carcinogenesis, either acting as tumor suppressors or oncomiRs, while aberrations in their expression levels are closely associated with tumor growth and progression, as well as with local invasion and metastatic dissemination. The arising role of miRNAs in HCC is in the spotlight of the current scientific research, aiming at the development of novel therapeutic perspectives. In this review, we will shed light on the emerging role of miRNAs in HCC.

1. Introduction

Hepatocellular carcinoma (HCC) is considered one of the most frequently diagnosed malignancies on a global level. Based on recent epidemiological data, it is considered the third cause of death related to cancer, accounting for 830,000 deaths annually [1]. Despite the novel therapeutic approaches, HCC’s five-year survival rates still remain considerably low (≈18%) [2]. Meanwhile, there is a sex disparity in HCC, with increased incidence rates in males in comparison to females, which are mainly attributed to the molecular background of HCC [3]. More particularly, females express a liver-associated autosomal gene, the so-called CYP39A1, which plays a key role in the suppression of hepatic carcinogenesis and the development of HCC [3].
Starting by mentioning the most common risk factors of HCC development, such as viral hepatitis B and C and alcohol abuse, it has to be underlined that HCC is attributed ever more to non-viral factors such as non-alcoholic fatty liver disease [4].
Despite the novel therapeutic strategies such as immunotherapy, radiofrequency combined with immunomodulation and other local ablative approaches, as well as transplantation and novel techniques in the surgical management of this malignancy, survival is still relatively low. However, a better understanding of the molecular basis of HCC is considered pivotal for the optimal management of this malignancy [5,6]. The role of miRNAs is in the spotlight of the current scientific research for several malignancies, including HCC.
MicroRNAs (miRNAs) are short non-coding RNA sequences, composed of 19–25 nucleotides with a pivotal regulatory role for the cellular functional state, including autophagy, apoptosis, proliferation, and apoptosis, as well as metabolic pathways and signaling [7,8]. They are considered a double-edged sword for carcinogenesis as they play a binary role either as tumor suppressors or promoters, given their oncogenic behavior [9].
Aberrations in the expression levels of these regulatory molecules are closely implicated in tumorigenesis, as well as in tumor progression and dissemination, due to the disruption of fundamental cell functions [10].
In this review, we will shed light on the emerging role of miRNAs in hepatocellular cancer, as well as their manipulation for future therapeutic and diagnostic perspectives.

2. A Glance through the Canonical MiRNA Biogenesis

MicroRNAs (miRNAs) play a fundamental role in the regulation of genetic information, several signaling metabolic pathways, hematopoiesis, as well as programmed cell death and autophagy, while it is demonstrated that miRNAs are closely associated even with the microbiome [11].
The mechanism of miRNAs biogenesis is closely related to the DNA transcription of specific miRNAs coding genes, under the enzymatic action of RNA polymerase II. This procedure takes place inside the nucleus, by which primary miRNA (pri-miRNA) arises [12]. Then, the aforementioned miRNA molecule is cleaved under the action of the ribonuclease complex, the so-called DiGeorge Syndrome Critical Region 8 (DGCR8)–Drosha (ribonuclease III) [13]. The latter procedure gives rise to another miRNA molecule, the precursor miRNA (pre-miRNA), which is shorter than the pri-miRNA (≥1000 nucleotides), has a sequence of up to one hundred nucleotides [14].
Afterwards, the pre-miRNA is relocated to the cytoplasm via the action of Exportin 5 and RAS-related nuclear protein–guanosine-5′-triphosphate (Ran-GTP), where it is further cleaved by the Dicer-TRBP complex, with the former being an RNase III endonuclease that removes its terminal loop, and the latter an RNA-binding protein [15].
After the removal of the pre-miRNA terminal loop, the miRNA is presented as a short Duplex strand, which is further loaded into an RNA-induced silencing complex (RISC) complex. This complex is composed of Argonaute RISC Catalytic Component 2 (Ago2) that separates one of the strands, based on its target specificity [16]. More particularly, the so-called leading strand (the mature miRNA strand) is merged with the RISC complex, whereas the other is called the passenger strand and is further degraded [17,18]. The miRNA-RISC (miRISC) targets a partially complementary mRNA strand, binding on its 3′untranslated region (3′UTR), while the miRNA sequence that is complementary is called the seed sequence [14,19,20].
Furthermore, the attachment of the miRNA strand to the partially complementary mRNA sequence leads to the repression of the mRNA translation, as well as its silencing or degradation [21].
Last but not least, there is another pathway of biogenesis, the so-called non-canonical, which does not depend either on the action of the Drosha–DGCR8 or –DICER complexes [22], and there are also different seed regions from the canonical. We demonstrate a schematic representation of canonical miRNA biogenesis in Figure 1.

3. The Emerging Role of MiRNAs in HCC

There are several aberrations occurring in malignant cells, including modifications in the interactions between miRNAs and their targets, while modifications in their expression levels are closely implicated in tumor progression and dissemination, due to the disruption of fundamental cell functions [23]. It is widely demonstrated that the miRNA expression profiling of malignant tumors, the so-called miRNA signature, is closely related to the stage of the tumor, the tumor growth and progression, as well as with diagnosis, prognosis, and anti-neoplastic treatment response [24].
Several studies demonstrate the emerging role of miRNAs in HCC, being closely associated with tumor proliferative and invasive behavior, metastatic dissemination, as well as with the sensitivity of cancer cells to treatment [25].

3.1. The Oncogenic MiRNAs in HCC

There are several aberrations in the expression levels of miRNA that promote HCC progression. Some of the miRNAs are upregulated and facilitate tumor proliferation, progression, neoangiogenesis, as well as local invasion distant dissemination and drug resistance [26]. In this section, we will discuss the role of several miRNAs that act as tumor promoters, the so-called oncomiRs [27]. Some miRNAs are found to be upregulated or downregulated and they are closely associated with the proliferation of malignant hepatocytes [28].
More particularly, the miR-92a level is increased in HCC tissue and is closely associated with tumor proliferation and invasion via targeting FOXA2 [29], while miR-203a-3p.1 promotes metastatic dissemination via targeting IL24, which suppresses HCC metastasis [30]. MiR-4417 is another oncogenic miRNA that inhibits apoptosis and promotes tumor proliferation in HCC via its regulatory effect on Pyruvate Kinase Muscle 2 (PKM2) Phosphorylation and Tripartite Motif-Containing 35 (TRIM35) [31]. Additionally, miR-21 is usually found to be elevated in HCC, especially in NAFLD-associated cases [32]. This miRNA targets several pathways, including STAT3 and PI3K/AKT pathways, as well as TGF-β/SMADs and the cell cycle [32,33]. The implication of miR-21 in the cell cycle of hepatocytes is reflected by its inhibitory effect on hepatocellular carcinoma, down-regulated 1 (HEPN1), which otherwise suppresses the proliferation of HepG2 cells [33].
Meanwhile, miR-18a constitutes an oncogenic miRNA that enhances HCC cell proliferation and tumor migratory behavior via targeting Bcl2l10 [34], whereas the upregulation of miR-155-5p and miR-331-3p is correlated with HCC progression via downregulating the PI3K/Akt pathway through PTEN inhibition and E2F1 suppression, respectively [35,36].
Likewise, the upregulation of miR-765 and miR-302d causes HCC proliferation and migration via suppressing INPP4B and TGFBR2 expression, respectively [37,38]. Migration and metastatic dissemination are also correlated with an upregulated miR-487a level via interacting with SPRED2 and PIK3R1 [39], while an increased miR-454 level exerts a significant oncogenic effect on tumor-initiating cells (T-ICs), leading to recurrent HCC, metastatic dissemination, and resistance to therapeutic modalities [40].
Moreover, miR-346 induces tumor growth and progression by targeting breast cancer metastasis suppressor 1 (BRMS1) [41], while miR-3910 also promotes progression and migration of the disease via targeting MST1 and inducing the Hippo-YAP signaling pathway [42].
On top of that, miR-873 induces HCC growth and metastatic dissemination via the Warburg effect, which is mediated by the AKT/mTOR pathway [43]. A list of several oncogenic miRNAs and their implication in HCC is demonstrated in Table 1.

3.2. MiRNAs as Tumor Suppressors for HCC

Recent studies have revealed that the loss or the decrease in tumor-suppressive miRNA expression levels could potentially lead to oncogenesis. More particularly, miR-199a-5p expression is closely associated with glycometabolism in hepatocytes, while in the case of hepatocarcinogenesis, it reprograms the glycolysis of the malignant cells via binding to 3′ UTR hexokinase 2(HK2) mRNA [44]. In addition, it has to be underlined that tumor-suppressing miR-199a is one of the most downregulated miRNAs in HCC specimens, while the enhancement of its expression levels could be notably beneficial as it can induce the impairment of tumor growth via interfering with aerobic glycolysis, which is crucial for their proliferation [45]. However, it has recently been shown that miR-199a/b-3p has also a tumor-inhibiting effect via interfering with the MAPK/ERK pathway by suppressing the p21-activated kinase 4 (PAK4) [46].
Moreover, miRNA-148-3p targets SMAD2, resulting in the downregulation of the latter and tumor suppression [47], while miR-195 also suppresses HCC development via interacting with the G1-S cell cycle regulatory genes such as CDC25A, CDK6, CDK4, as well as CCNE1. Based on the aforementioned, miR-195 can significantly alter the cell cycle of malignant hepatocytes [48].
Furthermore, miR-296-5p is another tumor-suppressing miRNA for HCC tissues which suppresses the HCC stem cell lines and the Neuregulin-ERBB signaling pathway, as well as limits the EMT phenomenon [49,50]. Additionally, miR-206 constitutes a pivotal suppressing miRNA, as it could totally inhibit HCC development in mice, when it is delivered in the malignant hepatic tissue [51]. The above phenomenon is attributed to the significant targets of this miRNA, such as the cyclin-dependent kinase (CDK) enzymes CDK6 and CCND1 that regulate the progression of the cell cycle and C-MET signaling pathway via targeting cMET protein [51].
Meanwhile, it has been demonstrated that miR-29a suppresses metastatic dissemination and invasion via targeting VEGFA, LOXL2, and LOX [52]. However, the expression of miR-29a is lowered in HCC tissue, and its enhancement could be potentially used as a therapeutic strategy [53]. Additionally, miR-766-3p could be also characterized as tumor-suppressive miRNA in HCC as it is closely associated with the Wnt/β-catenin signaling pathway, which constitutes a pivotal pathway for the physiological differentiation and multiplication of hepatocytes [54]. Similarly, miR-148b also serves as tumor-suppressing miRNA in HCC via targeting the WNT1 Wnt family member 1 protein that is involved in tumor growth [55], while miR-193a-5p significantly inhibits HCC formation and progression, via targeting the nucleolar- and spindle-associated proteins [56], which are involved in genome stability regulation and cell multiplication [57].
Additionally, miR-30e-3p influences the HCC phenotype via targeting mdm2 and activating p53 [58,59], while miR-1249 upregulation is closely associated with the upregulation of the Hedgehog pathway via suppressing PTCH1 expression [60].
Finally, miR-122 is also considered a tumor suppressor for HCC, while it is quite plentiful liver-specific miRNA that has a pivotal role in lipid metabolism [61]. This miRNA is commonly downregulated in HCC tissues, whereas it has been reported that the knockdown of miR-122 leads to steatohepatitis and hepatocarcinogenesis [62]. The aforementioned phenomenon is mainly attributed to the indirect activation of p53 via targeting its regulator, the so-called mouse double minute 2 homolog (mdm2) [63]. In Table 2, we outline a list of tumor-suppressive miRNAs in HCC.

3.3. The Interplay between MiRNAs and the Major HCC Predisposing Diseases

Viral hepatitis C and B constitute major risk factors that eventually lead to HCC development [64]. The deregulation of several cell functions, such as cell cycle, signaling, and metabolic pathways, as well as immunity, is mainly attributed to the interplay between miRNA expression levels and the presence of the virus [65]. More particularly, it is reported that HCV replication is modulated by several miRNAs such as miR-141 that amplify HCV replication by reducing the expression of deleted liver cancer 1 (DLC-1), which is a tumor suppressor gene [66]. Another miRNA that alters the anti-viral immunity and enhances HCV replication is miR-122, which constitutes a significant druggable target [67]. An example of the manipulation of miR-122 as a therapeutic target is the application of Miravirsen, which is an anti-miR agent in patients with HCV-genotype 1 (phase2 trial, NCT01200420) [68]. However, there are other miRNAs that have an anti-HCV effect, such as miR-199a, miR-27a, miR-196, as well as let-7b, miR-431, miR-27a, miR-29, and miR-448 [69]. Medical treatment with Interferon-B (INF)-B utilizes the anti-viral effect of the aforementioned miRNAs and increases their expression levels, leading to the suppression of viral replication [70]. Additionally, the expression levels of several genes are either upregulated or suppressed by a wide variety of miRNAs such as CREB1, stearoyl-CoA desaturase (SCD), and PPARG, respectively. Moreover, HBV also induces the deregulation of miRNA expression levels, which eventually leads to the development of HCC [71].
More particularly, miR-122 is not only implicated in HCV but also in HBV replication, presenting high expression levels [72], whereas miR-155, that is upregulated by HBV X protein (HBx), suppresses the replication of HBV via the blockade of the CCAAT/enhancer-binding protein [73], as well as via the activation of the IFN gene expression and the IFN signaling pathway, which enhances cellular resistance against HBV virus [74,75]. Meanwhile, HBV transcription is upregulated via miR-34 downregulation, which is attributed to HBx protein [76].
In addition, miR-122 is also closely implicated in the progression of either alcoholic or non-alcoholic fatty liver diseases (ALD or NAFLD). In both cases, miR-122 plays a key role in the progression of the diseases via altering the lipid metabolic pathway and inducing inflammation [77,78], while there are several miR-122-targeted genes that are closely associated with fatty acid (FA), cholesterol, and triglyceride metabolism, such as diacylglycerol O-acyltransferase 1 (Dgat1) and 1-acyl-sn-glycerol-3-phosphate acyltransferase alpha (Agpat1) genes, which are implicated in endogenous triglyceride synthesis [79]. Meanwhile, it has been demonstrated that the loss of miR-122 in animal models (mice) induces the brisk development of inflammation via the over-secretion of monocyte chemoattractant protein-1 and pro-inflammatory cytokines (TNF-a, IL6) [80], whereas the downregulation of miR-122 is also related with hepatocarcinogenesis and fibrosis progression in non-alcoholic steatohepatitis (NASH) [81].
Moreover, miR-21 downregulation is associated with the progression of the disease from NASH to HCC, as was demonstrated in animal and human HCC models [82]. The aforementioned phenomenon is attributed to the disruption of several signaling pathways, such as STAT3, AKT/PKB, and TGF-b [83].
Furthermore, the suppression and loss of let-7 expression, which is related to alcohol abuse, promotes ALD progression and eventually carcinogenesis, due to the disruption of Lin-28 homolog B (LIN28B) gene expression [84].
Similarly, there are several other miRNAs that are closely implicated in hepatocarcinogenesis and the fibrotic injury of the hepatic parenchyma, such as miR-192, miR-16, miR-199a, miR-15, and miR-497 [85].
Meanwhile, fibrosis is also closely associated with the action of several miRNAs such as miR200, human miR-29, and 34 families, miR-199, that are responsible for the modification of extracellular matrix (ECM), as well as with the activation of hepatic stellate cells (HSCs) [86,87]. It is worth mentioning in the case of the inflamed hepatocytes in NAFLD that there is an increased production of extracellular vesicles that contain several cargoes, such as miR-128-3p and MiR-192, which are closely implicated in the development and progression of the fibrotic injury via interacting with the HSCs, which further produce several fibrogenic proteins [88].

3.4. MiRNAs Signature in Premalignant State, HCC Staging, and Prognosis

As was aforementioned, miR-122 normally constitutes a suppressor of hepatic inflammation, while its loss leads to inflammation, fibrosis, and the formation of HCC. However, it is reported that miR-122 expression levels are deregulated not only in malignant but also in premalignant lesions, a phenomenon that usually has a background of chronic viral hepatitis and cirrhosis [86]. These premalignant lesions are characterized by nodularity, and they are subcategorized based on their histological and cytological appearance into (i) macroscopic dysplastic nodules (DNs) and (ii) microscopic dysplastic foci (DF). The former category of DNs is further subclassified based on the grade of atypia into high-grade and low-grade dysplastic nodules (HGDNs, LGDNs). The LGDNs are characterized by reduced expression levels of miR-199b, as well as miR-145, during their advance into small (≤2 cm) HCC, whereas HGDNs that are characterized by a higher tendency of malignant transformation present middling levels of miR-224 compared to LGDNs and HCCs (≤2 cm) that do not present and firmly express, respectively [87].
A wide amount of HCC cases (25%) are attributed to viral hepatitis C; for this reason, there are several bioinformatics analyses that intend to determine non-invasive biomarkers, such as miRNAs for identifying which HCV patients will eventually develop HCC. Some of the circulatory miRNAs that are increased in patients with viral hepatitis C and HCC include miR-3607, miR-215, miR-142, miR-199a, miR-150, miR-224, miR-183, miR-150, miR-424, and miR-217 compared to healthy donors. Meanwhile, there is a notable increase in miR-217 and miR-183 and a decrease in miR-3607 and miR-142 in bioptic HCC specimens [88].
Furthermore, there are several metastasis-related miRNAs demonstrated in animal models such as miR-331-3p, miR-487a, miR-29a, miR-1247-3p, miR-425p, and miR-219-4p that promote metastatic dissemination, and they are associated with a worrisome prognosis [89].
Last but not least, tumor-suppressing miR-199-5p, which is significantly downregulated in HCC, is closely associated with a poor prognosis, large-sized HCC [90], as well as with intravascular tumor thrombus and more advanced tumor–node–metastasis TNM stage [91].

3.5. MiRNAs as Predictive and Prognostic Biomarkers for HCC

It is widely demonstrated that HCC prognosis still remains unfavorable, a phenomenon that is mostly attributed to the late diagnostic time, when the disease is already disseminated. There are several conventional HCC biomarkers which are currently used in screening, although their value is limited in clinical practice [92]. The utilization of miRNAs as predictive and prognostic biomarkers in HCC has been in the spotlight in recent years. More particularly, some of the miRNAs that are closely related to a poor prognosis for HCC patients are the following: miR-1468↑, miR-32-5p↑, miR-940↓, miR-221↑, miR-137↓ and miR-296-5p↓, as well as miR-92a↑, miR-638↓, miR-122↓, miR-487a↑, and miR-148↑, with the majority of the aforementioned being associated with viral hepatitis B and C [93,94].
The expression levels of miR-32-5p are found up-regulated and closely associated with multi drug-resistant HCC tumors and the downregulation of PTEN [95]. The low expression of phosphatase and tensin homolog (PTEN) is attributed to the activation of the PI3K/AKT signaling pathway, a phenomenon that leads to tumor progress, drug resistance, epithelial–mesenchymal transition (EMT), as well as neoangiogenesis [95,96]. Meanwhile, upregulated miR-221 targets the PHD finger protein 2 (PHF2) gene, a phenomenon that induces its downregulation, as well as tumor dissemination and worrisome survival rates in HCC patients [97].
Moreover, miR-940 downregulated levels are closely associated with poor overall survival; however, when its expression levels are increased, there is increased apoptosis and reduced growth of the malignant hepatocytes [98]. Similarly, miR-137, miR-296, and miR-638 are also downregulated in HCC cases. More specifically, the former constitutes a significant prognostic biomarker for HCC, being closely implicated with the HCC microenvironment, while it targets the afamin (AFM) gene, leading to tumor invasive behavior and metastatic dissemination [99]. MiR-296-5p is found significantly reduced in HCC tissues compared with the normal tissues in the tumor vicinity, with its normal/increased levels are related to a favorable prognosis, the suppression of EMT phenomenon [50] via downregulating the Nrg1/Erbb signaling pathway, and it exerts a suppressive effect on the stem cell potential of HCC cell lines [50].
Meanwhile, the downregulation of miR-296 is also closely associated with an unfavorable prognosis in HCC patients, whereas its upregulation induces tumor growth suppression and apoptosis, as well as cell cycle downregulation, via targeting the fibroblast growth factor receptor 1 (FGFR1), which plays a key role in cell cycle progression and tumor proliferation [100].
Furthermore, miR-638 is also found to be reduced in the serum of HCC patients, being a novel prognostic factor for HCC [101]. Patients who underwent hepatectomy and had increased levels of miR-638 presented a more frequent development of distant or local metastasis [102]. Additionally, attenuation of its expression levels leads to local invasion and EMT via interacting with the sex determining region Y -box 2 (SOX2) [103]. Nevertheless, it is also related to a favorable prognosis and tumor suppression via inhibiting cyclin D1 [102,104].
Additionally, miR-4258, miR-638, miR-3648, and miR-663a are noticeably up-regulated in HUVECs under the exposure to HuH-7M-derived exosomes, which are closely related to increased vascular permeability and reduced ZO-1 and VE-cadherin expression levels in pre-metastatic stages, while they are secreted by highly aggressive and invasive HCC cell lines [102].
Further, the levels of miR-1246, miR-92a, miR-122, miR-487a, and miR-148a are also upregulated in HCC patients. MiR-1246 and miR-497 are closely associated with tumor staging, diagnosis, and prognosis, while miR-1246 via targeting the activator of the transcription and developmental regulator (AUTS2) gene leads to reduced survival in HCC [105]. MiR-92a is implicated in HCC progression and invasive behavior by interacting with FOXA2 and repressing it [106].
Finally, miR 487a is implicated in HBV-related HCC cases by promoting tumor growth [107]. Meanwhile, miR-148 constitutes a significant biomarker not only for predicting the HCC reoccurrence, but also in HCC screening, especially in cases in which the broadly used alpha fetoprotein (AFP) is low or normal [108], whereas microRNA-148a-3p suppresses HCC progression by downregulating SMAD2 mRNA via integrating with Ago2 [47,109,110].
Lastly, the absence of miR-122 is closely associated with viral hepatitis C infection-related HCC and drug resistance via interacting with several genes and signaling pathways, such as zinc finger protein (Snail)1/2, BCL9, CREB1, and the Wnt signaling pathway, respectively [111].

3.6. MiRNAs as Diagnostic or Monitoring Tools

It is demonstrated that salivary miRNAs could be potentially utilized as diagnostic biomarkers in HCC as they are considered highly specific and sensitive [112]. In addition, there are systematic analyses of several miRNAs (>700 miRNAs), which demonstrate a constant trend of overexpressed miRNAs, such as miR-148a-3p, miR-122-5p, miR-885-5p, and miR-125b-5p, as well as miR-100-5p, miR-1974, and miR-365a-3p [113].
Furthermore, miR-34a-5p is considered a promising biomarker for the correlation of HCC with cirrhosis [113], while it has to be underlined that based on bioinformatics analyses, there is an ethnical disparity between Asia, America, and Europe, with the latter having 5–37% upregulated miRNA expression levels. Finally, miR-199a, miR-182-5p, miR-422a, and miR-1269a have been demonstrated as the most auspicious diagnostic biomarkers for HCC [114].
Moreover, it is reported that miR-199a-5p could be potentially used as a monitoring marker as it is closely associated with glucose uptake by cancer cells. The results of miR-199a-5p replacement can be accessed via [18F]-FDG PET-CT performance, which demonstrates the active uptake of glucose by malignant tumor cells [115].
However, there is another promising HCC diagnostic, prognostic, and predictive tool, such as EVs, which are in the spotlight of current research and will be further discussed in Section 4. In Table 3, we demonstrate a summary of the functional role of the aforementioned miRNAs in HCC management either as prognostic, predictive, or diagnostic biomarkers.

4. The Interplay between EVs and MiRNAs in HCC

EVs constitute nanostructures that are composed of a bilayered lipid membrane, enclosing a high variety of nucleic acids including miRNAs, long non-coding RNAs (lncRNAs), autophagosomes, mitochondrial DNA, protein, and lipid molecules. They are characterized by great heterogeneity, which is reflected by the fact that they can have a wide variety of sizes, cargoes, and origins [117]. They are subclassified into exosomes, microvesicles, and apoptotic bodies based on their diameter, which is 40–150 nm, 150 nm–1000 nm, and over 1000 nm, respectively. EVs play a pivotal role in cross-talk intercellular communication between the parental cell and the recipients, a phenomenon that is mediated via the uptake of EVs cargoes [118].
Focusing on their functional role in HCC, it is reported that there is a great production of small extracellular vesicles (sEVs) (30–130 nm) from malignant hepatocytes, allowing cross-talk communication between HCC cells, in order to enhance cell proliferation, migration, and invasion, as well as metastatic dissemination [119]. There are several exosomal miRNAs that can be potentially isolated from serum, plasma, and urine, which can be utilized as diagnostic tools and prognostic or predictive biomarkers for HCC [120,121,122,123,124].
In Table 4 and Table 5, we demonstrate some examples of extracellular vesicle (EV)-contained miRNA cargoes, their biological role, and their utilization in HCC management.

5. The Interplay between MiRNAs and Autophagy in HCC

Autophagy constitutes a strictly orchestrated homeostatic pathway that reassures the ideal conditions for cell survival under stress, such as under conditions of inadequate nutrients and oxygen, as well as under the augmentation of non-functioning cytoplasmic organelles, which are eventually degraded [139,148]. There are five distinct phases on the autophagy pathway, including the (i) initiation step, (ii) the nucleation of the phagophore, (iii) the phagophore elongation step, (iv) the formation of autophagolysosome, and the (v) cargo degradation step [149,150]. MiRNAs are closely implicated in autophagy regulation, orchestrating each phase of the pathway from the initiation step to the last step of cargo degradation [151]. The emerging role of miRNAs in autophagy orchestration is in the spotlight of studies as produces opens new opportunities for the retrieval of new therapeutic modalities and strategies for the optimal management of HCC. More particularly, several stimuli that induce cellular stress initiate the autophagy pathway, but also some miRNAs that are closely related to autophagy regulation, so-called autophagomiRs [152]. The initiation step is started via the inhibition of the mammalian target of rapamycin (mTOR) and the Unc-51-like kinase1 complex (ULK1) activation, with the former being targeted by several miRNAs such as miR-199a, miR-7, miR-144, miR-7, as well as miR-100 and miR-338-3p [153]. After the engulfment of the cargo, the autophagophore is nucleated, which is achieved via the phosphorylation of class III PI3K by ULK1, followed by Beclin1-PI3K complex formation [154]. Several miRNAs target the components of this step, including the ULK1 suppressive miRNAs such as miR-17-5p, miR-26a-5p, miR-106b, and miR-372. Meanwhile, BECN1 is targeted by miR-181, miR-17/17-5p, and miR-376a/b, which also suppresses the pathway [151,152,155,156], while this is also targeted by activating miRNAs such as miR-221, and it is targeted under chemotherapy and a lack of nutrients by the miR-30 family, miR-9, miR-409-3p, as well as miR-376, miR-199-5p, and miR-20a, respectively [151,152,157,158].
Moreover, Beclin1-PI3K is targeted by activating miR-29b and other suppressive miRNAs such as miR-519a, miR-181a, miR-125a, and miR-374a [151,152,159]. Meanwhile, the elongation of phagophore, including the formation of the autophagosome, is regulated by activating miR-21-3p, miR-9a-5p, miR-155, and miR-20 [151,152,160], while miR-374, miR-519a, miR-520, miR-23b-3p, miR-204, miR-181a, miR-7, and miR-142-3p suppress the elongation step [151,152,160,161].
Focusing on HCC, there are several reports that demonstrate the interplay of autophagy and miRNAs [162]. As it was previously referred to, there are several autophagy-related miRNAs that are either downregulated and upregulated in HCC, such as miR-7 [163], miR-559 [164], miR-101 [165], miR-142-3p [166], miR-181a [167], miR-519d [168], and miR-25 [138], respectively. The deregulation of the expression levels of autophagomiRs is closely related to tumor progression, as well as drug resistance [169].
More particularly interaction between ATG14 (member of class III-PI3K complex) and miR-375 is related to drug resistance (targeted or chemotherapy), including sorafenib, while it is demonstrated that the action of miR-375 on ATG14 leads to the sensitization of HCC cells to sorafenib, via suppressing the cytoprotective effect of autophagy for cancer cells [170].
Similarly, it is also demonstrated that miR-23b-3p modulates the autophagy-induced drug resistance of sorafenib in the HCC HepG2 cell line, while it not only targets ATG12 but also GLS1 (glutaminase), which is related to the high exogenous glutamine that is associated with progressed disease [171].
Moreover, sorafenib resistance in HCC has been also related to miR-25 upregulation, which is closely associated with advanced-stage, lymphatic dissemination and autophagy induction via targeting FBXW7 protein [138,172].
In addition, HCC resistance to sorafenib has been also related to autophagy-related miR-423-5p, which could be potentially utilized as a predictive biomarker for HCC response to sorafenib [173].
Furthermore, the expression levels of miR-519d have been identified as notably increased in HCC, while it induces apoptosis and the autophagy pathway of HCC cells via activating AMPK signaling by binding in the 3ʹ-untranslated region (UTR) of the Ras-related protein 10 (Rab10). More particularly, the overexpressed miR-519d binds to Rab10 3′-UTR, suppressing its expression and the activation (phosphorylation) of AMPK and mTOR [168].
Moreover, it is demonstrated that miR-513b-5p targets the PIK3R3 gene and alters its expression during tumor proliferation and progression in HCC, constituting a potential druggable target [174].
Another miRNA that is downregulated in HCC is miR-559, which normally targets the PARD3 gene. It is shown that the expression level of the aforementioned gene is increased in HCC, a phenomenon that is closely associated with tumor proliferation and progression. Studies have demonstrated that the enhancement of miR-559 expression or the silencing (knockdown) of PARD3 limits the proliferation of HCC cells via autophagy inhibition, as well as inhibits neoangiogenesis via decreasing angiopoietin 2 and vascular endothelial growth factor (VEGF) expression levels [164].
Furthermore, it is demonstrated in hepatoma xenografts (in vivo) that miR-375 suppresses hypoxia-induced autophagy and resensitizes malignant cells to hypoxia, resulting in tumor suppression [175]. Additionally, autophagy-related miR-7 is found to be downregulated in HCC, while it constitutes a tumor suppressor for several malignancies, including HCC. However, the enhancement of miR-7 expression levels limits tumor proliferation via binding on mTOR, while the suppression of autophagy can potentially intensify the anti-proliferative activity of miR-7 in HCC cells [163].
Last but not least, it has to be underlined that autophagy also regulates miRNA expression levels via the selective degradation of several miRNAs, such as miR-224, which constitutes an oncogenic miRNA. More particularly, HBV-associated HCC cases present decreased autophagy activity, resulting in the aggregation of oncogenic miR-224 [176]. We demonstrate in Table 6 several autophagomiRs that are implicated in HCC.

6. The Interplay between MiRNAs and Microbiome in HCC

The significance of the human microbiome has been relatively underestimated in recent years; however, there are various studies have been published recently regarding the gut microbiome and its implication in HCC. It is demonstrated that hepatic and gut functions are closely interacting via the gut–liver axis, through the portal circulation. Under microbial dysbiosis, the multilayered gut barrier is disrupted and permeable, leading to the leakage of microbial products, the so-called microbiota-associated molecular patterns (MAMPs) that can interact with liver parenchyma and several hepatic functions, such as the production of bile acids (BAs) [177]. The modification of BAs synthesis is closely related to hepatobiliary malignancy via the deregulation of several cell functions [178,179,180].
More particularly, the baseline gut microbiome is comprised of Firmicutes, Verrucomicrobia, and Actinobacteria phyla, whereas under dysbiosis, there is an overgrowth of Proteobacteria and Bacteroidetes phyla [181]. It has also been demonstrated that the gut microbiome presents increased levels of lipopolysaccharides (LPS)–Gram-negative bacteria in HCC [182]. HCC miRNA signature includes several miRNAs that are closely associated with hepatic pathophysiological mechanisms, such as miR-21 and miR-666, with the former being positively and negatively correlated with Bacteroides acidifaciens and Firmicutes, respectively, while the latter is also negatively correlated with Firmicutes [183].
The interplay between miRNAs and microbiota is in the spotlight of several types of research for the identification of novel therapeutic targets, as well as diagnostic and prognostic markers. Meanwhile, several tumor-suppressive miRNAs have been demonstrated in female animal models, such as miR-122 and miR-26a/a-1. The expression of the aforementioned tumor suppressive miRNAs has been closely related to farnesoid X receptor (FXR) overexpression, which constitutes a BA receptor, implying gender disparity in HCC risk [183,184].

7. Future Therapeutic Opportunities

The utilization of miRNAs and their targets has entered the spotlight of recent research for the development of novel therapeutic strategies. Additionally, new strategies have been demonstrated by taking advantage of the interplay between miRNA and autophagy, as well as the gut microbiome [185]. MiRNAs play a binary role in hepatic carcinogenesis, and their expression levels can be either inhibited or enhanced for achieving tumor suppression [186].
An example of utilizing miRNA targets is the attempt of enhancing the tumor-suppressing effect of miR-122, which targets ADAM10, IGF1R, as well as cyclin G1 and ADAM17. Meanwhile, the suppression of the aforementioned targets could also enhance the expression levels of miR-122 [62].
Additionally, the downregulated levels of miR-296 are notably related to a worrisome prognosis in HCC [50], while the enhancement of its expression levels could limit tumor growth, induce apoptosis, and regulates the cell cycle via interacting with the fibroblast growth factor receptor 1 (FGFR1) gene [100].
Moreover, tumor-suppressing miR-199, which constitutes one of the most downregulated in HCC, can be potentially utilized as a therapeutic target [187]. More particularly, miR-199a-5p is closely related to the regulation of glycometabolism, while under increased levels of expression, it significantly reduces glycolysis products adenosine triphosphate (ATP) and G6P, resulting in the limitation of HCC cells via lowering glucose uptake by malignant hepatocytes [187]. Additionally, another therapeutic modality that is proposed is the utilization of miR-122 injections during trans-arterial chemo-embolization (TACE), which is based on the increased tendency of hepatocytes to uptake short RNA sequences [188].
Furthermore, the silencing of miR-92a could significantly suppress tumor invasion and growth [29], while the enhancement of miR-29a expression levels could also put a break on metastatic dissemination via targeting VEGFA, LOXL2, and LOX [52]. Similarly, the inhibition of oncogenic miR-203a-3p.1, which targets IL24, could also impede HCC metastasis [30].
Moreover, the enhancement of miR-940 expression levels could induce apoptosis and suppress HCC growth, while the same effect was also demonstrated after the inhibition of the estrogen-related receptor gamma (ESRRG), an miR-940 target [98].
Meanwhile, taking advantage of autophagomiRs and their targets on the autophagy pathway, there are several novel therapeutic strategies that can optimize HCC management. More particularly, HCC resistance to sorafenib can be overpassed via targeting miR-25, which is upregulated in HCC via targeting the FBXW7 protein [138,172], while the enhancement of miR-375 action can potentially sensitize the malignant hepatocytes to sorafenib via inhibiting the cytoprotective autophagy pathway in malignant cells [170]. Likewise, the enhancement of miR-559 expression levels could potentially limit HCC growth and drug resistance, as well as neoangiogenesis, via suppressing the expression of PARD3 [164].
In addition, there are miRNAs mimics and inhibitors, such as the miR-221 inhibitor and miR-122 mimic, that demonstrated favorable effects as they significantly reduce proliferation and pro-inflammatory markers, as well as neoangiogenesis [189]. Meanwhile, there is a phase I clinical trial (NCT01829971) for MRX34 which constitutes an miR-34a mimic that is evaluated in advanced solid tumors, including HCC [190]. It was reported that oral dexamethasone as a pre-treatment and daily intravenous MRX34 (daily administration for 5 days with 2 weeks off (21 days in total)) had severe side effects, even resulting in death, thus it was terminated, while the phase II study of MRX34 included a 70 mg/m2 dose for HCC [190,191].
Meanwhile, the utilization of gold nanomaterials as miRNA therapeutic agents can potentially restore the miRNA biological function and drug resistance, as they can release miRNA inhibitors or mimics, as well as nanoconjugates, meaning they can carry therapeutic agents that can stimulate endogenous miRNAs [192]. An example of this therapeutic modulation in HCC is the utilization of gold nanoparticles (AuNPs) that transfer miR-326 mimic, which promote its overexpression and the inhibition of the PDK1/AKT/c-myc axis, leading to invasion–migration inhibition, increased apoptosis, and EMT modification [193].
Last but not least, the off-label utilization of amiodarone, which constitutes a widely used antiarrhythmic treatment, could be potentially effective as a suppressor agent for HCC, via inducing autophagy-induced miR-224 degradation [194].

8. Conclusions

Hepatocellular carcinoma constitutes a highly aggressive and deadly form of primary liver cancer. Despite all the novel therapeutic modalities, the survival rate for this malignancy remains low. Shedding light on miRNA profiling and its interaction with autophagy, as well as gut microbiome, could potentially open new therapeutic horizons for the development of novel druggable targets. Non-coding miRNAs constitute a strong weapon for the diagnosis and treatment and potentially the monitoring and staging of HCC. Finally, the manipulation of the miRNA signature of HCC tumors could potentially overpass drug resistance, and it can be utilized for the identification and treatment of metastatic HCC tumors. However, further research is considered crucial for the discovery of novel diagnostic and therapeutic perspectives.

Author Contributions

All authors participated in the review. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rumgay, H.; Arnold, M.; Ferlay, J.; Lesi, O.; Cabasag, C.J.; Vignat, J.; Laversanne, M.; McGlynn, K.A.; Soerjomataram, I. Global burden of primary liver cancer in 2020 and predictions to 2040. J. Hepatol. 2022, 77, 1598–1606. [Google Scholar] [CrossRef] [PubMed]
  2. Vogel, A.; Meyer, T.; Sapisochin, G.; Salem, R.; Saborowski, A. Hepatocellular carcinoma. Lancet 2022, 400, 1345–1362. [Google Scholar] [CrossRef]
  3. Kotsiliti, E. Understanding HCC sex disparities. Nat. Rev. Gastroenterol. Hepatol. 2022, 19, 147. [Google Scholar] [CrossRef] [PubMed]
  4. Liu, Y.; Liu, L. Changes in the Epidemiology of Hepatocellular Carcinoma in Asia. Cancers 2022, 14, 4473. [Google Scholar] [CrossRef]
  5. Vitale, A.; Scarpa, M.; Caputo, D.; Spoletini, G.; Giovinazzo, F. (Eds.) Novel Insights into the Treatment of HCC and Liver Tumors; Frontiers Media SA: Lausanne, Switzerland, 2022. [Google Scholar]
  6. Da Costa, A.C.; Sodergren, M.; Jayant, K.; Cruz, F.S.; Spalding, D.; Pai, M.; Habib, N. Radiofrequency combined with immunomodulation for hepatocellular carcinoma: State of the art and innovations. World J. Gastroenterol. 2020, 26, 2040–2048. [Google Scholar] [CrossRef]
  7. Koustas, E.; Trifylli, E.-M.; Sarantis, P.; Papadopoulos, N.; Papanikolopoulos, K.; Aloizos, G.; Damaskos, C.; Garmpis, N.; Garmpi, A.; Karamouzis, M.V. The Emerging Role of MicroRNAs and Autophagy Mechanism in Pancreatic Cancer Progression: Future Therapeutic Approaches. Genes 2022, 13, 1868. [Google Scholar] [CrossRef] [PubMed]
  8. 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]
  9. Peng, Y.; Croce, C.M. The role of MicroRNAs in human cancer. Signal Transduct. Target. Ther. 2016, 1, 15004. [Google Scholar] [CrossRef] [Green Version]
  10. Miroshnichenko, S.; Patutina, O. Enhanced Inhibition of Tumorigenesis Using Combinations of miRNA-Targeted Therapeutics. Front. Pharmacol. 2019, 10, 488. [Google Scholar] [CrossRef]
  11. Rani, V.; Sengar, R.S. Biogenesis and mechanisms of microRNA-mediated gene regulation. Biotechnol. Bioeng. 2022, 119, 685–692. [Google Scholar] [CrossRef]
  12. 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] [Green Version]
  13. Lee, D.; Shin, C. Emerging roles of DROSHA beyond primary microRNA processing. RNA Biol. 2018, 15, 186–193. [Google Scholar] [CrossRef] [Green Version]
  14. O’Brien, J.; Hayder, H.; Zayed, Y.; Peng, C. Overview of MicroRNA Biogenesis, Mechanisms of Actions, and Circulation. Front. Endocrinol. 2018, 9, 402. [Google Scholar] [CrossRef] [Green Version]
  15. Paturi, S.; Deshmukh, M.V. A Glimpse of “Dicer Biology” through the Structural and Functional Perspective. Front. Mol. Biosci. 2021, 8, 643657. [Google Scholar] [CrossRef] [PubMed]
  16. Doench, J.G.; Sharp, P.A. Specificity of microRNA target selection in translational repression. Genes Dev. 2004, 18, 504–511. [Google Scholar] [CrossRef] [Green Version]
  17. Medley, J.C.; Panzade, G.; Zinovyeva, A.Y. microRNA strand selection: Unwinding the rules. Wiley Interdiscip. Rev. RNA 2021, 12, e1627. [Google Scholar] [CrossRef]
  18. AGO2 Argonaute RISC Catalytic Component 2 [Homo Sapiens (Human)]-Gene-NCBI. Nih.gov. Available online: https://www.ncbi.nlm.nih.gov/gene/27161 (accessed on 5 March 2023).
  19. Jungers, C.F.; Djuranovic, S. Modulation of miRISC-Mediated Gene Silencing in Eukaryotes. Front. Mol. Biosci. 2022, 9, 832916. [Google Scholar] [CrossRef] [PubMed]
  20. Ergin, K.; Çetinkaya, R. Regulation of MicroRNAs. In Methods in Molecular Biology; Springer: New York, NY, USA, 2022; pp. 1–32. [Google Scholar]
  21. Naeli, P.; Winter, T.; Hackett, A.P.; Alboushi, L.; Jafarnejad, S.M. The intricate balance between microRNA-induced mRNA decay and translational repression. FEBS J. 2022. [Google Scholar] [CrossRef]
  22. Stavast, C.; Erkeland, S. The Non-Canonical Aspects of MicroRNAs: Many Roads to Gene Regulation. Cells 2019, 8, 1465. [Google Scholar] [CrossRef] [Green Version]
  23. Smolarz, B.; Durczyński, A.; Romanowicz, H.; Szyłło, K.; Hogendorf, P. miRNAs in Cancer (Review of Literature). Int. J. Mol. Sci. 2022, 23, 2805. [Google Scholar] [CrossRef] [PubMed]
  24. Hussen, B.M.; Hidayat, H.J.; Salihi, A.; Sabir, D.K.; Taheri, M.; Ghafouri-Fard, S. MicroRNA: A signature for cancer progression. Biomed. Pharmacother. 2021, 138, 111528. [Google Scholar] [CrossRef]
  25. Sarkar, F.H.; Li, Y.; Wang, Z.; Kong, D.; Ali, S. Implication of microRNAs in drug resistance for designing novel cancer therapy. Drug Resist. Updates 2010, 13, 57–66. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Chen, E.; Xu, X.; Liu, R.; Liu, T. Small but Heavy Role: MicroRNAs in Hepatocellular Carcinoma Progression. BioMed Res. Int. 2018, 2018, 6784607. [Google Scholar] [CrossRef] [PubMed]
  27. Otmani, K.; Rouas, R.; Lewalle, P. OncomiRs as noncoding RNAs having functions in cancer: Their role in immune suppression and clinical implications. Front. Immunol. 2022, 13, 913951. [Google Scholar] [CrossRef]
  28. Oura, K.; Morishita, A.; Masaki, T. Molecular and Functional Roles of MicroRNAs in the Progression of Hepatocellular Carcinoma—A Review. Int. J. Mol. Sci. 2020, 21, 8362. [Google Scholar] [CrossRef]
  29. Wang, L.; Wu, J.; Xie, C. miR-92a promotes hepatocellular carcinoma cells proliferation and invasion by FOXA2 targeting. Iran. J. Basic Med. Sci. 2017, 20, 783–790. [Google Scholar] [CrossRef] [PubMed]
  30. Huo, W.; Du, M.; Pan, X.; Zhu, X.; Gao, Y.; Li, Z. miR-203a-3p.1 targets IL-24 to modulate hepatocellular carcinoma cell growth and metastasis. FEBS Open Bio 2017, 7, 1085–1091. [Google Scholar] [CrossRef] [PubMed]
  31. Song, L.; Zhang, W.; Chang, Z.; Pan, Y.; Zong, H.; Fan, Q.; Wang, L. miR-4417 Targets Tripartite Motif-Containing 35 (TRIM35) and Regulates Pyruvate Kinase Muscle 2 (PKM2) Phosphorylation to Promote Proliferation and Suppress Apoptosis in Hepatocellular Carcinoma Cells. Med. Sci. Monit. 2017, 23, 1741–1750. [Google Scholar] [CrossRef] [Green Version]
  32. Lai, C.-Y.; Yeh, K.-Y.; Lin, C.-Y.; Hsieh, Y.-W.; Lai, H.-H.; Chen, J.-R.; Hsu, C.-C.; Her, G.M. MicroRNA-21 Plays Multiple Oncometabolic Roles in the Process of NAFLD-Related Hepatocellular Carcinoma via PI3K/AKT, TGF-β, and STAT3 Signaling. Cancers 2021, 13, 940. [Google Scholar] [CrossRef] [PubMed]
  33. Hu, S.; Tao, R.; Wang, S.; Wang, C.; Zhao, X.; Zhao, H.; Li, L.; Zhu, S.; He, Y.; Jiang, X.; et al. MicroRNA-21 promotes cell proliferation in human hepatocellular carcinoma partly by targeting HEPN1. Tumor Biol. 2015, 36, 5467–5472. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, X.; Lu, J.; Cao, J.; Ma, B.; Gao, C.; Qi, F. MicroRNA-18a promotes hepatocellular carcinoma proliferation, migration, and invasion by targeting Bcl2L10. Onco Targets Ther. 2018, 11, 7919–7934. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Fu, X.; Wen, H.; Jing, L.; Yang, Y.; Wang, W.; Liang, X.; Nan, K.; Yao, Y.; Tian, T. MicroRNA-155-5p promotes hepatocellular carcinoma progression by suppressing PTEN through the PI3K/Akt pathway. Cancer Sci. 2017, 108, 620–631. [Google Scholar] [CrossRef] [Green Version]
  36. Jin, W.; Zhong, N.; Wang, L.; Yu, J.; Yin, F.; Zhang, K. MiR-331-3p Inhibition of the Hepatocellular Carcinoma (HCC) Bel-7402 Cell Line by Down-Regulation of E2F1. J. Nanosci. Nanotechnol. 2019, 19, 5476–5482. [Google Scholar] [CrossRef] [PubMed]
  37. Xie, B.-H.; He, X.; Hua, R.-X.; Zhang, B.; Tan, G.-S.; Xiong, S.-Q.; Liu, L.-S.; Chen, W.; Yang, J.-Y.; Wang, X.-N.; et al. Mir-765 promotes cell proliferation by downregulating INPP4B expression in human hepatocellular carcinoma. Cancer Biomark. 2016, 16, 405–413. [Google Scholar] [CrossRef] [PubMed]
  38. 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] [Green Version]
  39. Chang, R.-M.; Xiao, S.; Lei, X.; Yang, H.; Fang, F.; Yang, L.-Y. miRNA-487a Promotes Proliferation and Metastasis in Hepatocellular Carcinoma. Clin. Cancer Res. 2017, 23, 2593–2604. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Lei, Z.; Tang, X.; Si, A.; Yang, P.; Wang, L.; Luo, T.; Guo, G.; Zhang, Q.; Cheng, Z. microRNA-454 promotes liver tumor-initiating cell expansion by regulating SOCS6. Exp. Cell Res. 2020, 390, 111955. [Google Scholar] [CrossRef]
  41. Guo, Z.; Li, J.; Sun, J.; Sun, L.; Zhou, Y.; Yu, Z. miR-346 Promotes HCC Progression by Suppressing Breast Cancer Metastasis Suppressor 1 Expression. Oncol. Res. Featur. Preclin. Clin. Cancer Ther. 2018, 26, 1073–1081. [Google Scholar] [CrossRef] [PubMed]
  42. Cheng, L.; Wang, H.; Han, S. MiR-3910 Promotes the Growth and Migration of Cancer Cells in the Progression of Hepatocellular Carcinoma. Dig. Dis. Sci. 2017, 62, 2812–2820. [Google Scholar] [CrossRef]
  43. Zhang, Y.; Zhang, C.; Zhao, Q.; Wei, W.; Dong, Z.; Shao, L.; Li, J.; Wu, W.; Zhang, H.; Huang, H.; et al. The MiR-873/NDFIP1 Axis Promotes Hepatocellular Carcinoma Growth and Metastasis through the AKT/MTOR-Mediated Warburg Effect. Am. J. Cancer Res. 2019, 9, 927–944. [Google Scholar] [PubMed]
  44. Li, B.; He, L.; Zuo, D.; He, W.; Wang, Y.; Zhang, Y.; Liu, W.; Yuan, Y. Mutual Regulation of MiR-199a-5p and HIF-1α Modulates the Warburg Effect in Hepatocellular Carcinoma. J. Cancer 2017, 8, 940–949. [Google Scholar] [CrossRef] [Green Version]
  45. Feng, J.; Li, J.; Wu, L.; Yu, Q.; Ji, J.; Wu, J.; Dai, W.; Guo, C. Emerging roles and the regulation of aerobic glycolysis in hepatocellular carcinoma. J. Exp. Clin. Cancer Res. 2020, 39, 126. [Google Scholar] [CrossRef]
  46. Zeng, B.; Shi, W.; Tan, G. MiR-199a/b-3p inhibits gastric cancer cell proliferation via down-regulating PAK4/MEK/ERK signaling pathway. BMC Cancer 2018, 18, 34. [Google Scholar] [CrossRef] [Green Version]
  47. 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]
  48. Xu, T.; Zhu, Y.; Xiong, Y.; Ge, Y.-Y.; Yun, J.-P.; Zhuang, S.-M. MicroRNA-195 suppresses tumorigenicity and regulates G1/S transition of human hepatocellular carcinoma cells. Hepatology 2009, 50, 113–121. [Google Scholar] [CrossRef] [PubMed]
  49. 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]
  50. 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] [PubMed] [Green Version]
  51. 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]
  52. Yang, Y.-L.; Tsai, M.-C.; Chang, Y.-H.; Wang, C.-C.; Chu, P.-Y.; Lin, H.-Y.; Huang, Y.-H. MIR29A Impedes Metastatic Behaviors in Hepatocellular Carcinoma via Targeting LOX, LOXL2, and VEGFA. Int. J. Mol. Sci. 2021, 22, 6001. [Google Scholar] [CrossRef] [PubMed]
  53. Yang, Y.-L.; Chang, Y.-H.; Li, C.-J.; Huang, Y.-H.; Tsai, M.-C.; Chu, P.-Y.; Lin, H.-Y. New Insights into the Role of miR-29a in Hepatocellular Carcinoma: Implications in Mechanisms and Theragnostics. J. Pers. Med. 2021, 11, 219. [Google Scholar] [CrossRef]
  54. You, Y.; Que, K.; Zhou, Y.; Zhang, Z.; Zhao, X.; Gong, J.; Liu, Z. MicroRNA-766-3p Inhibits Tumour Progression by Targeting Wnt3a in Hepatocellular Carcinoma. Mol. Cells 2018, 41, 830–841. [Google Scholar] [CrossRef]
  55. Nie, X.; Liu, Y.; Chen, W.-D.; Wang, Y.-D. Interplay of miRNAs and Canonical Wnt Signaling Pathway in Hepatocellular Carcinoma. Front. Pharmacol. 2018, 9, 657. [Google Scholar] [CrossRef]
  56. 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.e26. [Google Scholar] [CrossRef]
  57. 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] [Green Version]
  58. Gramantieri, L.; Pollutri, D.; Gagliardi, M.; Giovannini, C.; Quarta, S.; Ferracin, M.; Casadei-Gardini, A.; Callegari, E.; De Carolis, S.; Marinelli, S.; et al. MiR-30e-3p Influences Tumor Phenotype through MDM2/TP53 Axis and Predicts Sorafenib Resistance in Hepatocellular Carcinoma. Cancer Res. 2020, 80, 1720–1734. [Google Scholar] [CrossRef]
  59. Gramantieri, L.; Giovannini, C.; Piscaglia, F.; Fornari, F. MicroRNAs as Modulators of Tumor Metabolism, Microenvironment, and Immune Response in Hepatocellular Carcinoma. J. Hepatocell. Carcinoma 2021, 8, 369–385. [Google Scholar] [CrossRef]
  60. Ye, Y.; Wei, Y.; Xu, Y.; Li, Y.; Wang, R.; Chen, J.; Zhou, Y.; Fu, Z.; Chen, Y.; Wang, X.; et al. Induced MiR-1249 expression by aberrant activation of Hedegehog signaling pathway in hepatocellular carcinoma. Exp. Cell Res. 2017, 355, 9–17. [Google Scholar] [CrossRef] [PubMed]
  61. Gatfield, D.; Le Martelot, G.; Vejnar, C.E.; Gerlach, D.; Schaad, O.; Fleury-Olela, F.; Ruskeepää, A.-L.; Oresic, M.; Esau, C.C.; Zdobnov, E.M.; et al. Integration of microRNA miR-122 in hepatic circadian gene expression. Genes Dev. 2009, 23, 1313–1326. [Google Scholar] [CrossRef] [Green Version]
  62. Castro-Muñoz, L.J.; Ulloa, E.V.; Sahlgren, C.; Lizano, M.; De La Cruz-Hernández, E.; Contreras-Paredes, A. Modulating epigenetic modifications for cancer therapy (Review). Oncol. Rep. 2023, 49, 59. [Google Scholar] [CrossRef] [PubMed]
  63. Fornari, F.; Gramantieri, L.; Giovannini, C.; Veronese, A.; Ferracin, M.; Sabbioni, S.; Calin, G.A.; Grazi, G.L.; Croce, C.M.; Tavolari, S.; et al. MiR-122/Cyclin G1 Interaction Modulates p53 Activity and Affects Doxorubicin Sensitivity of Human Hepatocarcinoma Cells. Cancer Res. 2009, 69, 5761–5767. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Stella, L.; Santopaolo, F.; Gasbarrini, A.; Pompili, M.; Ponziani, F.R. Viral hepatitis and hepatocellular carcinoma: From molecular pathways to the role of clinical surveillance and antiviral treatment. World J. Gastroenterol. 2022, 28, 2251–2281. [Google Scholar] [CrossRef]
  65. Loureiro, D.; Tout, I.; Narguet, S.; Benazzouz, S.M.; Mansouri, A.; Asselah, T. miRNAs as Potential Biomarkers for Viral Hepatitis B and C. Viruses 2020, 12, 1440. [Google Scholar] [CrossRef] [PubMed]
  66. Fan, H.-X.; Tang, H. Complex interactions between microRNAs and hepatitis B/C viruses. World J. Gastroenterol. 2014, 20, 13477–13492. [Google Scholar] [CrossRef] [PubMed]
  67. Kunden, R.D.; Khan, J.Q.; Ghezelbash, S.; Wilson, J.A. The Role of the Liver-Specific microRNA, miRNA-122 in the HCV Replication Cycle. Int. J. Mol. Sci. 2020, 21, 5677. [Google Scholar] [CrossRef] [PubMed]
  68. Ottosen, S.; Parsley, T.B.; Yang, L.; Zeh, K.; van Doorn, L.-J.; van der Veer, E.; Raney, A.K.; Hodges, M.R.; Patick, A.K. In Vitro Antiviral Activity and Preclinical and Clinical Resistance Profile of Miravirsen, a Novel Anti-Hepatitis C Virus Therapeutic Targeting the Human Factor miR-122. Antimicrob. Agents Chemother. 2015, 59, 599–608. [Google Scholar] [CrossRef] [Green Version]
  69. Li, H.; Jiang, J.-D.; Peng, Z.-G. MicroRNA-mediated interactions between host and hepatitis C virus. World J. Gastroenterol. 2016, 22, 1487–1496. [Google Scholar] [CrossRef]
  70. Ahmad, I.; Valverde, A.; Siddiqui, H.; Schaller, S.; Naqvi, A.R. Viral MicroRNAs: Interfering the Interferon Signaling. Curr. Pharm. Des. 2020, 26, 446–454. [Google Scholar] [CrossRef]
  71. Sartorius, K.; Makarova, J.; Sartorius, B.; An, P.; Winkler, C.; Chuturgoon, A.; Kramvis, A. The Regulatory Role of MicroRNA in Hepatitis-B Virus-Associated Hepatocellular Carcinoma (HBV-HCC) Pathogenesis. Cells 2019, 8, 1504. [Google Scholar] [CrossRef] [Green Version]
  72. Mahmoudian-Sani, M.R.; Asgharzade, S.; Alghasi, A.; Saeedi-Boroujeni, A.; Sadati, S.J.A.; Moradi, M.T. MicroRNA-122 in patients with hepatitis B and hepatitis B virus-associated hepatocellular carcinoma. J. Gastrointest. Oncol. 2019, 10, 789–796. [Google Scholar] [CrossRef]
  73. Sartorius, K.; Swadling, L.; An, P.; Makarova, J.; Winkler, C.; Chuturgoon, A.; Kramvis, A. The Multiple Roles of Hepatitis B Virus X Protein (HBx) Dysregulated MicroRNA in Hepatitis B Virus-Associated Hepatocellular Carcinoma (HBV-HCC) and Immune Pathways. Viruses 2020, 12, 746. [Google Scholar] [CrossRef]
  74. Xie, K.-L.; Zhang, Y.-G.; Liu, J.; Zeng, Y.; Wu, H. MicroRNAs Associated with HBV Infection and HBV-related HCC. Theranostics 2014, 4, 1176–1192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Yang, G.; Wan, P.; Zhang, Y.; Tan, Q.; Qudus, M.S.; Yue, Z.; Luo, W.; Zhang, W.; Ouyang, J.; Li, Y.; et al. Innate Immunity, Inflammation, and Intervention in HBV Infection. Viruses 2022, 14, 2275. [Google Scholar] [CrossRef] [PubMed]
  76. Dong, M.-L.; Wen, X.; He, X.; Ren, J.-H.; Yu, H.-B.; Qin, Y.-P.; Yang, Z.; Yang, M.-L.; Zhou, C.-Y.; Zhang, H.; et al. HBx Mediated Increase of DDX17 Contributes to HBV-Related Hepatocellular Carcinoma Tumorigenesis. Front. Immunol. 2022, 13, 871558. [Google Scholar] [CrossRef]
  77. Fang, Z.; Dou, G.; Wang, L. MicroRNAs in the Pathogenesis of Nonalcoholic Fatty Liver Disease. Int. J. Biol. Sci. 2021, 17, 1851–1863. [Google Scholar] [CrossRef]
  78. Torres, J.-L.; Novo-Veleiro, I.; Manzanedo, L.; Alvela-Suárez, L.; Macías, R.; Laso, F.-J.; Marcos, M. Role of microRNAs in alcohol-induced liver disorders and non-alcoholic fatty liver disease. World J. Gastroenterol. 2018, 24, 4104–4118. [Google Scholar] [CrossRef]
  79. Chai, C.; Rivkin, M.; Berkovits, L.; Simerzin, A.; Zorde-Khvalevsky, E.; Rosenberg, N.; Klein, S.; Yaish, D.; Durst, R.; Shpitzen, S.; et al. Metabolic Circuit Involving Free Fatty Acids, microRNA 122, and Triglyceride Synthesis in Liver and Muscle Tissues. Gastroenterology 2017, 153, 1404–1415. [Google Scholar] [CrossRef]
  80. Li, M.; Tang, Y.; Wu, L.; Mo, F.; Wang, X.; Li, H.; Qi, R.; Zhang, H.; Srivastava, A.; Ling, C. The hepatocyte-specific HNF4α/miR-122 pathway contributes to iron overload–mediated hepatic inflammation. Blood 2017, 130, 1041–1051. [Google Scholar] [CrossRef] [Green Version]
  81. Chai, C.; Cox, B.; Yaish, D.; Gross, D.; Rosenberg, N.; Amblard, F.; Shemuelian, Z.; Gefen, M.; Korach, A.; Tirosh, O.; et al. Agonist of RORA Attenuates Nonalcoholic Fatty Liver Progression in Mice via Up-regulation of MicroRNA 122. Gastroenterology 2020, 159, 999–1014.e9. [Google Scholar] [CrossRef]
  82. De Sousa, M.C.; Calo, N.; Sobolewski, C.; Gjorgjieva, M.; Clément, S.; Maeder, C.; Dolicka, D.; Fournier, M.; Vinet, L.; Montet, X.; et al. Mir-21 Suppression Promotes Mouse Hepatocarcinogenesis. Cancers 2021, 13, 4983. [Google Scholar] [CrossRef] [PubMed]
  83. Fodor, A.; Lazar, A.; Buchman, C.; Tiperciuc, B.; Orasan, O.; Cozma, A. MicroRNAs: The Link between the Metabolic Syndrome and Oncogenesis. Int. J. Mol. Sci. 2021, 22, 6337. [Google Scholar] [CrossRef]
  84. McDaniel, K.; Huang, L.; Sato, K.; Wu, N.; Annable, T.; Zhou, T.; Ramos-Lorenzo, S.; Wan, Y.; Huang, Q.; Francis, H.; et al. The let-7/Lin28 axis regulates activation of hepatic stellate cells in alcoholic liver injury. J. Biol. Chem. 2017, 292, 11336–11347. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Mohr, R.; Özdirik, B.; Lambrecht, J.; Demir, M.; Eschrich, J.; Geisler, L.; Hellberg, T.; Loosen, S.H.; Luedde, T.; Tacke, F.; et al. From Liver Cirrhosis to Cancer: The Role of Micro-RNAs in Hepatocarcinogenesis. Int. J. Mol. Sci. 2021, 22, 1492. [Google Scholar] [CrossRef] [PubMed]
  86. Liu, X.; Ma, H.; Wu, R.; Wang, H.; Xu, H.; Li, S.; Wang, G.; Lv, G.; Niu, J. Identification of Liver Fibrosis-Related MicroRNAs in Human Primary Hepatic Stellate Cells Using High-Throughput Sequencing. Genes 2022, 13, 2201. [Google Scholar] [CrossRef] [PubMed]
  87. Ma, T.; Cai, X.; Wang, Z.; Huang, L.; Wang, C.; Jiang, S.; Hua, Y.; Liu, Q. miR-200c Accelerates Hepatic Stellate Cell-Induced Liver Fibrosis via Targeting the FOG2/PI3K Pathway. BioMed Res. Int. 2017, 2017, 2670658. [Google Scholar] [CrossRef] [Green Version]
  88. Trifylli, E.-M.; Kriebardis, A.G.; Koustas, E.; Papadopoulos, N.; Deutsch, M.; Aloizos, G.; Fortis, S.P.; Papageorgiou, E.G.; Tsagarakis, A.; Manolakopoulos, S. The Emerging Role of Extracellular Vesicles and Autophagy Machinery in NASH—Future Horizons in NASH Management. Int. J. Mol. Sci. 2022, 23, 12185. [Google Scholar] [CrossRef]
  89. Kogure, A.; Kosaka, N.; Ochiya, T.; Kogure, A.; Kosaka, N.; Ochiya, T. Cross-talk between cancer cells and their neighbors via miRNA in extracellular vesicles: An emerging player in cancer metastasis. J. Biomed. Sci. 2019, 26, 7. [Google Scholar] [CrossRef]
  90. Liu, L.; Lu, L.; Zheng, A.; Xie, J.; Xue, Q.; Wang, F.; Wang, X.; Zhou, H.; Tong, X.; Li, Y.; et al. MiR-199a-5p and let-7c cooperatively inhibit migration and invasion by targeting MAP4K3 in hepatocellular carcinoma. Oncotarget 2017, 8, 13666–13677. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. 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, 67169–67180. [Google Scholar] [CrossRef] [Green Version]
  92. Song, T.; Li, L.; Wu, S.; Liu, Y.; Guo, C.; Wang, W.; Dai, L.; Zhang, T.; Wu, H.; Su, B. Peripheral Blood Genetic Biomarkers for the Early Diagnosis of Hepatocellular Carcinoma. Front. Oncol. 2021, 11, 583714. [Google Scholar] [CrossRef]
  93. Nagy, Á.; Lánczky, A.; Menyhárt, O.; Győrffy, B. Validation of miRNA prognostic power in hepatocellular carcinoma using expression data of independent datasets. Sci. Rep. 2018, 8, 9227. [Google Scholar] [CrossRef] [Green Version]
  94. Morishita, A.; Oura, K.; Tadokoro, T.; Fujita, K.; Tani, J.; Masaki, T. MicroRNAs in the Pathogenesis of Hepatocellular Carcinoma: A Review. Cancers 2021, 13, 514. [Google Scholar] [CrossRef] [PubMed]
  95. Fu, X.; Liu, M.; Qu, S.; Ma, J.; Zhang, Y.; Shi, T.; Wen, H.; Yang, Y.; Wang, S.; Wang, J.; et al. 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] [Green Version]
  96. Sasaki, R.; Kanda, T.; Yokosuka, O.; Kato, N.; Matsuoka, S.; Moriyama, M. Exosomes and Hepatocellular Carcinoma: From Bench to Bedside. Int. J. Mol. Sci. 2019, 20, 1406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Fu, Y.; Liu, M.; Li, F.; Qian, L.; Zhang, P.; Lv, F.; Cheng, W.; Hou, R. MiR-221 Promotes Hepatocellular Carcinoma Cells Migration via Targeting PHF2. BioMed Res. Int. 2019, 2019, 4371405. [Google Scholar] [CrossRef]
  98. Yuan, B.; Liang, Y.; Wang, D.; Luo, F. MiR-940 inhibits hepatocellular carcinoma growth and correlates with prognosis of hepatocellular carcinoma patients. Cancer Sci. 2015, 106, 819–824. [Google Scholar] [CrossRef] [Green Version]
  99. Wei, Q.; Zhao, L.; Jiang, L.; Bi, J.; Yu, Z.; Zhao, L.; Song, X.; Sun, M.; Chen, Y.; Wei, M. Prognostic relevance of miR-137 and its liver microenvironment regulatory target gene AFM in hepatocellular carcinoma. J. Cell. Physiol. 2019, 234, 11888–11899. [Google Scholar] [CrossRef] [PubMed]
  100. Wang, L.; Bo, X.; Zheng, Q.; Xiao, X.; Wu, L.; Li, B. miR-296 inhibits proliferation and induces apoptosis by targeting FGFR1 in human hepatocellular carcinoma. FEBS Lett. 2016, 590, 4252–4262. [Google Scholar] [CrossRef] [PubMed]
  101. Jun, L.; Yang, G.; Zhisu, L. The utility of serum exosomal microRNAs in hepatocellular carcinoma. Biomed. Pharmacother. 2019, 111, 1221–1227. [Google Scholar] [CrossRef] [PubMed]
  102. Yokota, Y.; Noda, T.; Okumura, Y.; Kobayashi, S.; Iwagami, Y.; Yamada, D.; Tomimaru, Y.; Akita, H.; Gotoh, K.; Takeda, Y.; et al. Serum exosomal miR-638 is a prognostic marker of HCC via downregulation of VE-cadherin and ZO-1 of endothelial cells. Cancer Sci. 2021, 112, 1275–1288. [Google Scholar] [CrossRef] [PubMed]
  103. Zhang, Y.; Zhang, D.; Jiang, J.; Dong, L. Loss of miR-638 promotes invasion and epithelial-mesenchymal transition by targeting SOX2 in hepatocellular carcinoma. Oncol. Rep. 2017, 37, 323–332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Yang, J.; Li, B.; Zhao, S.; Du, H.; Du, Y. Exosomal miR-638 Inhibits Hepatocellular Carcinoma Progression by Targeting SP1. OncoTargets Ther. 2020, 13, 6709–6720. [Google Scholar] [CrossRef]
  105. Chen, S.; Fu, Z.; Wen, S.; Yang, X.; Yu, C.; Zhou, W.; Lin, Y.; Lv, Y. Expression and Diagnostic Value of miR-497 and miR-1246 in Hepatocellular Carcinoma. Front. Genet. 2021, 12, 666306. [Google Scholar] [CrossRef]
  106. Peng, Y.; Huang, D.; Qing, X.; Tang, L.; Shao, Z. Investigation of MiR-92a as a Prognostic Indicator in Cancer Patients: A Meta-Analysis. J. Cancer 2019, 10, 4430–4441. [Google Scholar] [CrossRef] [Green Version]
  107. Cao, X.; Yang, Q.; Yu, Q. Increased Expression of miR-487b Is Associated With Poor Prognosis and Tumor Progression of HBV-Related Hepatocellular Carcinoma. Open Forum Infect. Dis. 2020, 7, ofaa498. [Google Scholar] [CrossRef] [PubMed]
  108. Han, J.; Li, J.; Qian, Y.; Liu, W.; Liang, J.; Huang, Z.; Wang, S.; Zhao, C. Identification of plasma miR-148a as a noninvasive biomarker for hepatocellular carcinoma. Clin. Res. Hepatol. Gastroenterol. 2019, 43, 585–593. [Google Scholar] [CrossRef] [PubMed]
  109. Pan, L.; Huang, S.; He, R.; Rong, M.; Dang, Y.; Chen, G. Decreased expression and clinical significance of miR-148a in hepatocellular carcinoma tissues. Eur. J. Med. Res. 2014, 19, 68. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Li, Y.; Deng, X.; Zeng, X.; Peng, X. The Role of Mir-148a in Cancer. J. Cancer 2016, 7, 1233–1241. [Google Scholar] [CrossRef] [Green Version]
  111. Fu, X.; Calin, G.A. miR-122 and hepatocellular carcinoma: From molecular biology to therapeutics. Ebiomedicine 2018, 37, 17–18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Mariam, A.; Miller-Atkins, G.; Moro, A.; Rodarte, A.I.; Siddiqi, S.; Acevedo-Moreno, L.-A.; Brown, J.M.; Allende, D.S.; Aucejo, F.; Rotroff, D.M. Salivary miRNAs as non-invasive biomarkers of hepatocellular carcinoma: A pilot study. PeerJ 2022, 10, e12715. [Google Scholar] [CrossRef] [PubMed]
  113. Jin, Y.; Wong, Y.S.; Goh, B.K.P.; Chan, C.Y.; Cheow, P.C.; Chow, P.K.H.; Lim, T.K.H.; Goh, G.B.B.; Krishnamoorthy, T.L.; Kumar, R.; et al. Circulating microRNAs as Potential Diagnostic and Prognostic Biomarkers in Hepatocellular Carcinoma. Sci. Rep. 2019, 9, 10464. [Google Scholar] [CrossRef] [Green Version]
  114. Dat, V.H.X.; Nhung, B.T.H.; Chau, N.N.B.; Cuong, P.H.; Hieu, V.D.; Linh, N.T.M.; Quoc, N.B. Identification of potential microRNA groups for the diagnosis of hepatocellular carcinoma (HCC) using microarray datasets and bioinformatics tools. Heliyon 2022, 8, e08987. [Google Scholar] [CrossRef]
  115. Zhang, L.; Lou, J.; Lu, M.; Gao, C.; Zhao, S.; Li, B.; Liang, S.; Li, Y.; Li, D.; Liu, M. Suppression of miR-199a maturation by HuR is crucial for hypoxia-induced glycolytic switch in hepatocellular carcinoma. EMBO J. 2015, 34, 2671–2685. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Dabbish, A.M.; Abdelzaher, H.M.; Abohawya, M.; Shamma, S.; Mahmoud, Y.H.; Maged, A.; Manaa, M.; Hassany, M.; Kobeissy, F.; Bazgir, O.; et al. Prognostic MicroRNA Panel for HCV-Associated HCC: Integrating Computational Biology and Clinical Validation. Cancers 2022, 14, 3036. [Google Scholar] [CrossRef]
  117. Di Bella, M.A. Overview and Update on Extracellular Vesicles: Considerations on Exosomes and Their Application in Modern Medicine. Biology 2022, 11, 804. [Google Scholar] [CrossRef] [PubMed]
  118. Bartel, S.; Deshane, J.; Wilkinson, T.; Gabrielsson, S. Extracellular Vesicles as Mediators of Cellular Cross Talk in the Lung Microenvironment. Front. Med. 2020, 7, 326. [Google Scholar] [CrossRef] [PubMed]
  119. Tsuchiya, A.; Natsui, K.; Ishii, Y.; Koseki, Y.; Takeda, N.; Tomiyoshi, K.; Yamazaki, F.; Yoshida, Y.; Terai, S. Small extracellular vesicles and liver diseases: From diagnosis to therapy. World J. Hepatol. 2022, 14, 1307–1318. [Google Scholar] [CrossRef]
  120. Liu, C.; Wu, H.; Mao, Y.; Chen, W.; Chen, S. Exosomal microRNAs in hepatocellular carcinoma. Cancer Cell Int. 2021, 21, 254. [Google Scholar] [CrossRef]
  121. 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]
  122. Sorop, A.; Constantinescu, D.; Cojocaru, F.; Dinischiotu, A.; Cucu, D.; Dima, S. Exosomal microRNAs as Biomarkers and Therapeutic Targets for Hepatocellular Carcinoma. Int. J. Mol. Sci. 2021, 22, 4997. [Google Scholar] [CrossRef]
  123. Wang, H.; Yu, L.; Huang, P.; Zhou, Y.; Zheng, W.; Meng, N.; He, R.; Xu, Y.; Keong, T.S.; Cui, Y. Tumor-associated Exosomes Are Involved in Hepatocellular Carcinoma Tumorigenesis, Diagnosis, and Treatment. J. Clin. Transl. Hepatol. 2022, 10, 496–508. [Google Scholar] [CrossRef]
  124. Ghosh, S.; Bhowmik, S.; Majumdar, S.; Goswami, A.; Chakraborty, J.; Gupta, S.; Aggarwal, S.; Ray, S.; Chatterjee, R.; Bhattacharyya, S.; et al. The exosome encapsulated microRNAs as circulating diagnostic marker for hepatocellular carcinoma with low alpha-fetoprotein. Int. J. Cancer 2020, 147, 2934–2947. [Google Scholar] [CrossRef]
  125. Wang, H.; Hou, L.; Li, A.; Duan, Y.; Gao, H.; Song, X. Expression of Serum Exosomal MicroRNA-21 in Human Hepatocellular Carcinoma. BioMed Res. Int. 2014, 2014, 864894. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Sorop, A.; Iacob, R.; Iacob, S.; Constantinescu, D.; Chitoiu, L.; Fertig, T.E.; Dinischiotu, A.; Chivu-Economescu, M.; Bacalbasa, N.; Savu, L.; et al. Plasma Small Extracellular Vesicles Derived miR-21-5p and miR-92a-3p as Potential Biomarkers for Hepatocellular Carcinoma Screening. Front. Genet. 2020, 11, 712. [Google Scholar] [CrossRef]
  127. Nakano, T.; Chen, I.-H.; Wang, C.-C.; Chen, P.-J.; Tseng, H.-P.; Huang, K.-T.; Hu, T.-H.; Li, L.-C.; Goto, S.; Cheng, Y.-F.; et al. Circulating exosomal miR-92b: Its role for cancer immunoediting and clinical value for prediction of posttransplant hepatocellular carcinoma recurrence. Am. J. Transplant. 2019, 19, 3250–3262. [Google Scholar] [CrossRef] [PubMed]
  128. Sun, L.; Xu, M.; Zhang, G.; Dong, L.; Wu, J.; Wei, C.; Xu, K.; Zhang, L. Identification of Circulating Exosomal miR-101 and miR-125b Panel Act as a Potential Biomarker for Hepatocellular Carcinoma. Int. J. Genom. 2021, 2021, 1326463. [Google Scholar] [CrossRef] [PubMed]
  129. Liu, W.; Hu, J.; Zhou, K.; Chen, F.; Wang, Z.; Liao, B.; Dai, Z.; Cao, Y.; Fan, J.; Zhou, J. Serum exosomal miR-125b is a novel prognostic marker for hepatocellular carcinoma. OncoTargets Ther. 2017, 10, 3843–3851. [Google Scholar] [CrossRef] [Green Version]
  130. Abdelaziz, H.E.; El-Guendy, N.; Radwan, E.M.; Sharawi, S.K.; E El-Houseini, M.; Abdel-Wahab, A.-H.A. Possible Use of miR-223-3p as a Prognostic Marker in Transarterial Chemoembolization Treatment of Hepatocellular Carcinoma Patients. Asian Pac. J. Cancer Prev. 2022, 23, 4125–4135. [Google Scholar] [CrossRef]
  131. Sohn, W.; Kim, J.; Kang, S.H.; Yang, S.R.; Cho, J.-Y.; Cho, H.C.; Shim, S.G.; Paik, Y.-H. Serum exosomal microRNAs as novel biomarkers for hepatocellular carcinoma. Exp. Mol. Med. 2015, 47, e184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Fründt, T.; Krause, L.; Hussey, E.; Steinbach, B.; Köhler, D.; von Felden, J.; Schulze, K.; Lohse, A.; Wege, H.; Schwarzenbach, H. Diagnostic and Prognostic Value of miR-16, miR-146a, miR-192 and miR-221 in Exosomes of Hepatocellular Carcinoma and Liver Cirrhosis Patients. Cancers 2021, 13, 2484. [Google Scholar] [CrossRef] [PubMed]
  133. Deng, P.; Li, M.; Wu, Y. The Predictive Efficacy of Serum Exosomal microRNA-122 and microRNA-148a for Hepatocellular Carcinoma Based on Smart Healthcare. J. Healthc. Eng. 2022, 2022, 5914541. [Google Scholar] [CrossRef]
  134. Suehiro, T.; Miyaaki, H.; Kanda, Y.; Shibata, H.; Honda, T.; Ozawa, E.; Miuma, S.; Taura, N.; Nakao, K. Serum exosomal microRNA-122 and microRNA-21 as predictive biomarkers in transarterial chemoembolization-treated hepatocellular carcinoma patients. Oncol. Lett. 2018, 16, 3267–3273. [Google Scholar] [CrossRef] [Green Version]
  135. Ma, Y.-S.; Liu, J.-B.; Lin, L.; Zhang, H.; Wu, J.-J.; Shi, Y.; Jia, C.-Y.; Zhang, D.-D.; Yu, F.; Wang, H.-M.; et al. Exosomal microRNA-15a from mesenchymal stem cells impedes hepatocellular carcinoma progression via downregulation of SALL4. Cell Death Discov. 2021, 7, 224. [Google Scholar] [CrossRef]
  136. 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] [Green Version]
  137. Liu, J.; Fan, L.; Yu, H.; Zhang, J.; He, Y.; Feng, D.; Wang, F.; Li, X.; Liu, Q.; Li, Y.; et al. Endoplasmic Reticulum Stress Causes Liver Cancer Cells to Release Exosomal miR-23a-3p and Up-regulate Programmed Death Ligand 1 Expression in Macrophages. Hepatology 2019, 70, 241–258. [Google Scholar] [CrossRef] [PubMed]
  138. Feng, X.; Zou, B.; Nan, T.; Zheng, X.; Zheng, L.; Lan, J.; Chen, W.; Yu, J. MiR-25 enhances autophagy and promotes sorafenib resistance of hepatocellular carcinoma via targeting FBXW7. Int. J. Med. Sci. 2022, 19, 257–266. [Google Scholar] [CrossRef]
  139. Koustas, E.; Sarantis, P.; Papavassiliou, A.G.; Karamouzis, M.V. Upgraded role of autophagy in colorectal carcinomas. World J. Gastrointest. Oncol. 2018, 10, 367–369. [Google Scholar] [CrossRef]
  140. Fang, J.H.; Zhang, Z.J.; Shang, L.R.; Luo, Y.W.; Lin, Y.F.; Yuan, Y.; Zhuang, S.M. Hepatoma cell-secreted exosomal microRNA-103 increases vascular permeability and promotes metastasis by targeting junction proteins. Hepatology 2018, 68, 1459–1475. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Lou, G.; Song, X.; Yang, F.; Wu, S.; Wang, J.; Chen, Z.; Liu, Y. Exosomes derived from miR-122-modified adipose tissue-derived MSCs increase chemosensitivity of hepatocellular carcinoma. J. Hematol. Oncol. 2015, 8, 122. [Google Scholar] [CrossRef] [Green Version]
  142. Li, S.; Chen, L. Exosomes in Pathogenesis, Diagnosis, and Treatment of Hepatocellular Carcinoma. Front. Oncol. 2022, 12, 793432. [Google Scholar] [CrossRef] [PubMed]
  143. Kim, J.; Jiang, J.; Badawi, M.; Schmittgen, T.D. miR-221 regulates CD44 in hepatocellular carcinoma through the PI3K-AKT-mTOR pathway. Biochem. Biophys. Res. Commun. 2017, 487, 709–715. [Google Scholar] [CrossRef]
  144. Zhang, Z.; Li, X.; Sun, W.; Yue, S.; Yang, J.; Li, J.; Ma, B.; Wang, J.; Yang, X.; Pu, M.; et al. Loss of exosomal miR-320a from cancer-associated fibroblasts contributes to HCC proliferation and metastasis. Cancer Lett. 2017, 397, 33–42. [Google Scholar] [CrossRef]
  145. Sun, W.; Fu, S.; Wu, S.; Tu, R. Growing Evidence of Exosomal MicroRNA-Related Metastasis of Hepatocellular Carcinoma. BioMed Res. Int. 2020, 2020, 4501454. [Google Scholar] [CrossRef] [PubMed]
  146. Fang, T.; Lv, H.; Lv, G.; Li, T.; Wang, C.; Han, Q.; Yu, L.; Su, B.; Guo, L.; Huang, S.; et al. 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] [Green Version]
  147. Dong, S.-S.; Dong, D.-D.; Yang, Z.-F.; Zhu, G.-Q.; Gao, D.-M.; Chen, J.; Zhao, Y.; Liu, B.-B. Exosomal miR-3682-3p Suppresses Angiogenesis by Targeting ANGPT1 via the RAS-MEK1/2-ERK1/2 Pathway in Hepatocellular Carcinoma. Front. Cell Dev. Biol. 2021, 9, 633358. [Google Scholar] [CrossRef]
  148. Majeed, S.T.; Majeed, R.; Andrabi, K.I. Expanding the view of the molecular mechanisms of autophagy pathway. J. Cell. Physiol. 2022, 237, 3257–3277. [Google Scholar] [CrossRef]
  149. Koustas, E.; Sarantis, P.; Karamouzis, M.; Vielh, P.; Theocharis, S. The Controversial Role of Autophagy in Ewing Sarcoma Pathogenesis—Current Treatment Options. Biomolecules 2021, 11, 355. [Google Scholar] [CrossRef] [PubMed]
  150. Koustas, E.; Trifylli, E.-M.; Sarantis, P.; Papadopoulos, N.; Papanikolopoulos, K.; Aloizos, G.; Damaskos, C.; Garmpis, N.; Garmpi, A.; Matthaios, D.; et al. Exploiting Autophagy-Dependent Neoantigen Presentation in Tumor Microenvironment. Genes 2023, 14, 474. [Google Scholar] [CrossRef]
  151. Shan, C.; Chen, X.; Cai, H.; Hao, X.; Li, J.; Zhang, Y.; Gao, J.; Zhou, Z.; Li, X.; Liu, C.; et al. The Emerging Roles of Autophagy-Related MicroRNAs in Cancer. Int. J. Biol. Sci. 2021, 17, 134–150. [Google Scholar] [CrossRef] [PubMed]
  152. Akkoc, Y.; Gozuacik, D. MicroRNAs as major regulators of the autophagy pathway. Biochim. Biophys. Acta (BBA) Mol. Cell Res. 2020, 1867, 118662. [Google Scholar] [CrossRef]
  153. Wang, P.; Liu, X.-M.; Ding, L.; Zhang, X.-J.; Ma, Z.-L. mTOR signaling-related MicroRNAs and Cancer involvement. J. Cancer 2018, 9, 667–673. [Google Scholar] [CrossRef] [Green Version]
  154. Lu, G.; Wang, Y.; Shi, Y.; Zhang, Z.; Huang, C.; He, W.; Wang, C.; Shen, H. Autophagy in health and disease: From molecular mechanisms to therapeutic target. Medcomm 2022, 3, e150. [Google Scholar] [CrossRef]
  155. Yang, Y.; Liang, C.; Isidoro, C. MicroRNAs: An Emerging Player In Autophagy. Sci. Res. 2015, 10, 14293. [Google Scholar] [CrossRef]
  156. Chen, H.; Zhang, Z.; Lu, Y.; Song, K.; Liu, X.; Xia, F.; Sun, W. Downregulation of ULK 1 by micro RNA -372 inhibits the survival of human pancreatic adenocarcinoma cells. Cancer Sci. 2017, 108, 1811–1819. [Google Scholar] [CrossRef] [Green Version]
  157. Roy, S.G. Regulation of autophagy by miRNAs in human diseases. Nucleus 2021, 64, 317–329. [Google Scholar] [CrossRef]
  158. Zhu, H.; Wu, H.; Liu, X.; Li, B.; Chen, Y.; Ren, X.; Liu, C.-G.; Yang, J.-M. Regulation of autophagy by a beclin 1-targeted microRNA, miR-30a, in cancer cells. Autophagy 2009, 5, 816–823. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Gozuacik, D.; Akkoc, Y.; Ozturk, D.G.; Kocak, M. Autophagy-Regulating microRNAs and Cancer. Front. Oncol. 2017, 7, 65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Hyttinen, J.M.; Blasiak, J.; Felszeghy, S.; Kaarniranta, K. MicroRNAs in the regulation of autophagy and their possible use in age-related macular degeneration therapy. Ageing Res. Rev. 2021, 67, 101260. [Google Scholar] [CrossRef]
  161. Zhao, Y.; Wang, Z.; Zhang, W.; Zhang, L. MicroRNAs play an essential role in autophagy regulation in various disease phenotypes. Biofactors 2019, 45, 844–856. [Google Scholar] [CrossRef]
  162. Xu, X.; Tao, Y.; Shan, L.; Chen, R.; Jiang, H.; Qian, Z.; Cai, F.; Ma, L.; Yu, Y. The Role of MicroRNAs in Hepatocellular Carcinoma. J. Cancer 2018, 9, 3557–3569. [Google Scholar] [CrossRef]
  163. 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] [Green Version]
  164. 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]
  165. Xu, Y.; An, Y.; Wang, Y.; Zhang, C.; Zhang, H.; Huang, C.; Jiang, H.; Wang, X.; Li, X. miR-101 inhibits autophagy and enhances cisplatin-induced apoptosis in hepatocellular carcinoma cells. Oncol. Rep. 2013, 29, 2019–2024. [Google Scholar] [CrossRef] [Green Version]
  166. 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] [PubMed] [Green Version]
  167. Yang, J.; He, Y.; Zhai, N.; Ding, S.; Li, J.; Peng, Z.Z. MicroRNA-181a inhibits autophagy by targeting Atg5 in hepatocellular carcinoma. Front. Biosci. 2018, 23, 388–396. [Google Scholar]
  168. 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] [PubMed] [Green Version]
  169. Yang, S.; Yang, L.; Li, X.; Li, B.; Li, Y.; Zhang, X.; Ma, Y.; Peng, X.; Jin, H.; Li, H. New Insights into Autophagy in Hepatocellular Carcinoma: Mechanisms and Therapeutic Strategies. Am. J. Cancer Res. 2019, 9, 1329–1353. [Google Scholar]
  170. 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] [PubMed]
  171. Kaur, R.; Kanthaje, S.; Taneja, S.; Dhiman, R.K.; Chakraborti, A. miR-23b-3p Modulating Cytoprotective Autophagy and Glutamine Addiction in Sorafenib Resistant HepG2, a Hepatocellular Carcinoma Cell Line. Genes 2022, 13, 1375. [Google Scholar] [CrossRef]
  172. Xie, C.-M.; Sun, Y. The MTORC1-mediated autophagy is regulated by the FBXW7-SHOC2-RPTOR axis. Autophagy 2019, 15, 1470–1472. [Google Scholar] [CrossRef]
  173. Stiuso, P.; Potenza, N.; Lombardi, A.; Ferrandino, I.; Monaco, A.; Zappavigna, S.; Vanacore, D.; Mosca, N.; Castiello, F.; Porto, S.; et al. MicroRNA-423-5p Promotes Autophagy in Cancer Cells and Is Increased in Serum from Hepatocarcinoma Patients Treated with Sorafenib. Mol. Ther. Nucleic Acids 2015, 4, e233. [Google Scholar] [CrossRef]
  174. Jin, W.; Liang, Y.; Li, S.; Lin, G.; Liang, H.; Zhang, Z.; Zhang, W.; Nie, R. MiR-513b-5p represses autophagy during the malignant progression of hepatocellular carcinoma by targeting PIK3R3. Aging 2021, 13, 16072–16087. [Google Scholar] [CrossRef] [PubMed]
  175. Chang, Y.; Yan, W.; He, X.; Zhang, L.; Li, C.; Huang, H.; Nace, G.; Geller, D.A.; Lin, J.; Tsung, A. miR-375 Inhibits Autophagy and Reduces Viability of Hepatocellular Carcinoma Cells under Hypoxic Conditions. Gastroenterology 2012, 143, 177–187.e8. [Google Scholar] [CrossRef] [PubMed]
  176. Lan, S.-H.; Wu, S.-Y.; Zuchini, R.; Lin, X.-Z.; Su, I.-J.; Tsai, T.-F.; Lin, Y.-J.; Wu, C.-T.; Liu, H.-S. Autophagy-preferential degradation of MIR224 participates in hepatocellular carcinoma tumorigenesis. Autophagy 2014, 10, 1687–1689. [Google Scholar] [CrossRef] [Green Version]
  177. Tripathi, A.; Debelius, J.; Brenner, D.A.; Karin, M.; Loomba, R.; Schnabl, B.; Knight, R. The gut–liver axis and the intersection with the microbiome. Nat. Rev. Gastroenterol. Hepatol. 2018, 15, 397–411. [Google Scholar] [CrossRef] [PubMed]
  178. Cani, P.D. Human gut microbiome: Hopes, threats and promises. Gut 2018, 67, 1716–1725. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Kiriyama, Y.; Nochi, H. Physiological Role of Bile Acids Modified by the Gut Microbiome. Microorganisms 2021, 10, 68. [Google Scholar] [CrossRef]
  180. Trifylli, E.-M.; Koustas, E.; Papadopoulos, N.; Sarantis, P.; Aloizos, G.; Damaskos, C.; Garmpis, N.; Garmpi, A.; Karamouzis, M.V. An Insight into the Novel Immunotherapy and Targeted Therapeutic Strategies for Hepatocellular Carcinoma and Cholangiocarcinoma. Life 2022, 12, 665. [Google Scholar] [CrossRef]
  181. Rinninella, E.; Raoul, P.; Cintoni, M.; Franceschi, F.; Miggiano, G.A.D.; Gasbarrini, A.; Mele, M.C. What Is the Healthy Gut Microbiota Composition? A Changing Ecosystem across Age, Environment, Diet, and Diseases. Microorganisms 2019, 7, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Rattan, P.; Minacapelli, C.D.; Rustgi, V. The Microbiome and Hepatocellular Carcinoma. Liver Transplant. 2020, 26, 1316–1327. [Google Scholar] [CrossRef] [PubMed]
  183. Allegra, A.; Musolino, C.; Tonacci, A.; Pioggia, G.; Gangemi, S. Interactions between the MicroRNAs and Microbiota in Cancer Development: Roles and Therapeutic Opportunities. Cancers 2020, 12, 805. [Google Scholar] [CrossRef] [Green Version]
  184. Tian, Y.; Zhang, M.; Fan, M.; Xu, H.; Wu, S.; Zou, S.; Wang, Y.; Tang, D.; Zhang, C.; Han, W.; et al. A miRNA-mediated attenuation of hepatocarcinogenesis in both hepatocytes and Kupffer cells. Mol. Ther. Nucleic Acids 2022, 30, 1–12. [Google Scholar] [CrossRef] [PubMed]
  185. Zhang, H.; Zhang, W.; Jiang, L.; Chen, Y. Recent advances in systemic therapy for hepatocellular carcinoma. Biomark. Res. 2022, 10, 3. [Google Scholar] [CrossRef]
  186. Pratama, M.Y.; Pascut, D.; Massi, M.N.; Tiribelli, C. The role of microRNA in the resistance to treatment of hepatocellular carcinoma. Ann. Transl. Med. 2019, 7, 577. [Google Scholar] [CrossRef]
  187. Guo, W.; Qiu, Z.; Wang, Z.; Wang, Q.; Tan, N.; Chen, T.; Chen, Z.; Huang, S.; Gu, J.; Li, J.; et al. MiR-199a-5p is negatively associated with malignancies and regulates glycolysis and lactate production by targeting hexokinase 2 in liver cancer. Hepatology 2015, 62, 1132–1144. [Google Scholar] [CrossRef]
  188. Bozzato, A.M.; Martingano, P.; Mucelli, R.A.P.; Cavallaro, M.F.M.; Cesarotto, M.; Marcello, C.; Tiribelli, C.; Pascut, D.; Pizzolato, R.; Mucelli, F.P.; et al. MicroRNAs Related to TACE Treatment Response: A Review of the Literature from a Radiological Point of View. Diagnostics 2022, 12, 374. [Google Scholar] [CrossRef]
  189. Hassan, M.; Elzallat, M.; Aboushousha, T.; Elhusseny, Y.; El-Ahwany, E. MicroRNA-122 mimic/microRNA-221 inhibitor combination as a novel therapeutic tool against hepatocellular carcinoma. Noncoding RNA Res. 2023, 8, 126–134. [Google Scholar] [CrossRef] [PubMed]
  190. 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.; et al. 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]
  191. Zhang, L.; Liao, Y.; Tang, L. MicroRNA-34 family: A potential tumor suppressor and therapeutic candidate in cancer. J. Exp. Clin. Cancer Res. 2019, 38, 53. [Google Scholar] [CrossRef] [Green Version]
  192. Sousa, D.P.; Conde, J. Gold Nanoconjugates for MiRNA Modulation in Cancer Therapy: From MiRNA Silencing to MiRNA Mimics. ACS Mater. Au 2022, 2, 626–640. [Google Scholar] [CrossRef] [PubMed]
  193. Mo, Y.; He, L.; Lai, Z.; Wan, Z.; Chen, Q.; Pan, S.; Li, L.; Li, D.; Huang, J.; Xue, F.; et al. Gold Nano-Particles (AuNPs) Carrying MiR-326 Targets PDK1/AKT/c-Myc Axis in Hepatocellular Carcinoma. Artif. Cells Nanomed. Biotechnol. 2019, 47, 2830–2837. [Google Scholar] [CrossRef]
  194. Huang, S.-T.; Hsu, W.-F.; Huang, H.-S.; Yen, J.-H.; Lin, M.-C.; Peng, C.-Y.; Yen, H.-R. Improved Survival in Hepatocellular Carcinoma Patients with Cardiac Arrhythmia by Amiodarone Treatment through Autophagy. Int. J. Mol. Sci. 2019, 20, 3978. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. A schematic presentation of miRNA biogenesis. The mechanism of miRNAs biogenesis starts with DNA transcription of specific miRNAs coding genes, under the enzymatic action of RNA polymerase II inside the nucleus. Pri-miRNA arises and is cleaved under the action of DGCR8–Drosha (ribonuclease III), giving rise to pre-miRNA, which is relocated to the cytoplasm via Exportin 5 and Ran-GTP. Pre-miRNA then is cleaved by the Dicer-TRBP, forming the duplex miRNA, which is further loaded on the RISC complex. The active miRNA strand (leading strand/mature miRNA) is loaded on RISC (miRISC) and the other is the so-called passenger, which becomes degraded. The miRISC (seed sequence) binds a target mRNA, which leads to the repression of the mRNA translation, as well as its silencing or degradation. The figure was created with “BioRender.com” (agreement number OS252V998Q).
Figure 1. A schematic presentation of miRNA biogenesis. The mechanism of miRNAs biogenesis starts with DNA transcription of specific miRNAs coding genes, under the enzymatic action of RNA polymerase II inside the nucleus. Pri-miRNA arises and is cleaved under the action of DGCR8–Drosha (ribonuclease III), giving rise to pre-miRNA, which is relocated to the cytoplasm via Exportin 5 and Ran-GTP. Pre-miRNA then is cleaved by the Dicer-TRBP, forming the duplex miRNA, which is further loaded on the RISC complex. The active miRNA strand (leading strand/mature miRNA) is loaded on RISC (miRISC) and the other is the so-called passenger, which becomes degraded. The miRISC (seed sequence) binds a target mRNA, which leads to the repression of the mRNA translation, as well as its silencing or degradation. The figure was created with “BioRender.com” (agreement number OS252V998Q).
Ijms 24 07168 g001
Table 1. A list of oncomiRs, their expression levels, as well as their targets and functional effects in HCC.
Table 1. A list of oncomiRs, their expression levels, as well as their targets and functional effects in HCC.
OncomiRsExpression Levels in HCCEffectsTarget
miR-18aEnhances HCC cell proliferation and tumor migratory behaviorBcl2l10 [34]
miR-21Tumor proliferation, migration, and EMT PI3K/AKT, STAT3, TGF-β/SMADs HEPN1 [32,33]
Involvement in cell cycle
miR-92aTumor proliferation/invasionFOXA2 [29]
miR-155-5pHCC progressionPTEN [35]
miR-203a-3p.1Metastatic disseminationIL24 [30]
miR-302dTumor proliferationTGFBR2 [38]
miR-331-3pHCC progressionE2F1 [36]
miR-346Tumor growth and progressionBRMS1 [41]
miR-454Recurrent HCC, metastatic dissemination resistance to
therapeutic modalities
TICs [40]
miR-487aTumor migration, metastatic disseminationSPRED2 PIK3R1 [39]
miR-765Tumor proliferationINPP4B [37]
miR-873HCC growth and metastatic disseminationAKT/mTOR [43]
miR-3910Tumor progression and
migration
MST1 [42]
Hippo-YAP pathway
miR-4417Inhibits apoptosis and tumor proliferationPKM2, TRIM35 [31]
Up-regulated (↑): increased expression levels.
Table 2. A list of tumor-suppressive miRNAs in HCC and their effects and targets.
Table 2. A list of tumor-suppressive miRNAs in HCC and their effects and targets.
Tumor-Suppressing miRNAExpression Levels
in HCC
EffectTarget
miR-29aSuppresses invasionVEGFA, LOXL2 LOX [52,53]
and metastasis
miR-30e-3pTumor suppressionp53 [58,59]
Activates p53
miR-122↓ or Regulate lipid metabolism mdm2, p53 [61,62,63]
knockdownTumor suppression
miR-148bTumor growth suppressionWNT1 [55]
miR-193a-5pInhibits HCC formation and progressionNucleolar- and
spindle-associated protein [56,57]
miR-195Suppresses HCC development via interacting the malignant hepatocyte cell cycle.CDC25A, CDK6, CDK4, CCNE1 [48]
miR-199aImpairment of tumor growth,
proliferation
Aerobic
Glycolysis [45]
miR-199a-5pGlycometabolism in hepatocytes3′ UTR HK2 mRNA [44]
Reprograms the glycolysis of the malignant cells
miR-199a/b-3pTumor inhibitionMAPK/ERK PAK4 [46]
miR-206Inhibits HCC development in miceCDK6, CCND1, cMET [51]
miR-296-5pSuppresses the HCC stem cell lines/EMT phenomenonNeuregulin-ERBB pathway [49,50]
miR-766-3pTumor suppressionWnt/β-catenin
signaling pathway [54]
miR-1249Tumor suppressionHedgehog pathway
PTCH1 [60]
↑ increased/↓ decreased.
Table 3. Summary of miRNAs expression levels and their utilization in HCC management.
Table 3. Summary of miRNAs expression levels and their utilization in HCC management.
UtilizationMiRNAsExpression LevelsIsolationReferences
Diagnostic markersmir-92bDecreased in saliva/
increased in tumor tissue
Saliva and
Tumor tissue
[112]
mir-548i-2Decreased in saliva/
decreased in tumor tissue
mir-548lDecreased in saliva/
decreased in tumor tissue
Diagnostic markers miR-1972Increased only in HCC Plasma[113]
miR-193a-5pIncreased only in HCC
miR-214-3p Increased only in HCC
miR-365a-3pIncreased only in HCC
Poor
prognostic markers
miR-137Decreased Tumor tissue[99]
miR-296 DecreasedTumor tissue[100]
miR-638 DecreasedSerum[101]
miR-940 DecreasedTumor tissue[98]
miR-32-5p Increased Tumor tissue[95]
miR-92a Increased Tumor tissue[106]
miR-221 Increased Tumor tissue[97]
miR-1246 Increased Serum[105]
Good
prognostic markers
miR-296-5pNotably decreased Tumor tissue[49,50]
Monitoring markers miR-199a-5p [18F]-FDG PET-CT performance[115]
Follow up for reoccurrence miR-148 Serum[108]
Screening markersmiR-148a in ↓ or absent AFP Increased Serum[108]
StagingPre-metastaticIncreasedSerum[102]
miR-638
miR-663
miR-3648
miR-4258
Circulatory miRNAs HCV and HCCmiR-142
miR-150
Decreased
Decreased
Blood[116]
miR-183Increased
miR-199bDecreased
miR-215Decreased
miR-217Increased
miR-224Increased
miR-424 Decreased
miR-3607Decreased
Bioptic specimensmiR-142 Tumor tissue[116]
miR-183
miR-217
miR-3607
Table 4. EV-contained miRNA cargoes and their utilization as diagnostic, prognostic, and predictive biomarkers in HCC management.
Table 4. EV-contained miRNA cargoes and their utilization as diagnostic, prognostic, and predictive biomarkers in HCC management.
Exosomal miRNAIsolationExpression LevelsUtilization
miR-21Serum↑ in HCC patientsDiagnostic tool [125]
miR-21-5pPlasma↑ in HCC patientsDiagnostic tool and monitoring
together with serum AFP [126]
miR-92bSerum↑levels—HCC recurrence
after liver transplantation
Predictive biomarker [127]
miR-92-3pPlasma↑ in HCC patients/AFP low levelsDiagnostic tool and monitoring [126]
miR-101Serum↓ in HCC patientsDiagnostic tool [128]
miR-125bSerum↓ in HCC patients
↓levels—↓ recurrence time
↓overall survival


Diagnostic [128] and prognostic tool [129]
miR-223-3pSerum↓↓ levels—non-responders to TACE treatmentPredictive biomarker [130]
miR-665Serum↑ levels—HCC patients
↑↑levels—poor prognosis
(stage, tumor size, survival)
Diagnostic tool and prognostic biomarker [121]
miR-718Serum↓levels—HCC recurrence after liver transplantationPredictive biomarker [121]
miR-18a,
miR-221,
miR-222,
miR-224



Serum



↑↑levels—HCC patients



Diagnostic tool [131]
miR-16,
miR-146,
miR-192
Plasma↑levels—HCC and cirrhosisDiagnostic tool and prognostic biomarker [132]
miR-122,
miR-148a
Serum Predictive biomarkers [133]
miR-122Serum↓↓ levels post-TACE in
Cirrhotic-HCC patients
Predictive biomarkers [134]
↑↑significantly increased/↑ increased/↓↓ significantly decreased/↓ decreased.
Table 5. EV-contained miRNA cargoes and their functional role in HCC.
Table 5. EV-contained miRNA cargoes and their functional role in HCC.
EV-Contained miRNA CargoFunctional Role in HCC
EVs-miR let-7b [126]↓tumor inflammation by targeting
interleukin 6
EVs-miR 15a [135]Inhibits tumor proliferation/migration
EVs-miR-21 [136]PDK1/AKT pathway activation
Conversion of hepatic stellate cells to CAFs
Tumor migration
Neoangiogenesis
EVs-miR-23a-3p [137]Uptake by macrophages
Promoted tumor escape
phenomenon
T-cell function suppression
EVs-miR-25 [121,138]Modification of BAX, BCL2 expression
(apoptotic markers for sorafenib)
Drug resistance in sorafenib
EVs-miR-32-5p [139]Multi-drug resistance
EVs-miR-103 [140]Uptake by endothelial cells
Modified integrity of vessels
Tumor invasion
EVs-miR-122 [141]Suppressed tumor proliferation
↑chemosensitivity
EVs-miR-125a [142]Targets HCC stem cells CD90
Suppressed tumor proliferation and
Migration
EVs-miR-221 [143]CD44 regulation via interaction with
PI3K-AKT-mTOR
Tumor proliferation/migration
EVs-miR-320a [123,144]Inhibition of MAPK pathway
Suppressed tumor migration and invasion
EVs-miR-429 [145]↑E2F1 transcription activity
Tumor proliferation
EVs-miR-1247-3p [146]Conversion of fibroblasts to CAFs
Promoted EMT phenomenon
Metastatic dissemination and proliferation
EVs-miR 3682 3p [147]Inhibited neoangiogenesis
↑: increased/↓: decreased.
Table 6. The role of autophagomiRs in HCC.
Table 6. The role of autophagomiRs in HCC.
AutophagomiRsTargetRole in HCC
miR-7 [163]ATG5Inhibits invasion and metastasis
miR-25 [138]FBXW7Autophagy enhancement
Sorafenib resistance
miR-23b-3p [171]ATG12↓survival and ↑apoptosis of sorafenib-resistant HepG2 cells
miR-101 [165]ATG4D RAB5A, STMN1Inhibits autophagy
Enhances cisplatin-induced apoptosis
miR-142-3p [166]ATG5/ATG16L1Autophagy inhibition
Sensitization of HCC cells to sorafenib
miR-181a [167]ATG5autophagy inhibition
↑apoptosis in HepG2 cells
↓tumor growth (HepG2 cell xenograft tumor models)
miR-375 [170]ATG7Inhibits autophagy
↓HCC survival under hypoxia
miR-423-5p [173]Autophagy
cell cycle
↑autophagy in HCC cells-treated with
sorafenib
Partially remitted/stable disease 6 months after the treatment initiation (predictive biomarker)
miR-513b-5p [174]PIK3R3Inhibits proliferation, migration/invasion,
Induces HCC cell apoptosis
miR-519d [168]Rab10Autophagy induction
↑HCC cell Apoptosis
AMPK signaling pathway activation via Rab10
miR-559 [164]PARD3Inhibits proliferation, neoangiogenesis,
inhibits autophagy
↑: increased, ↓: decreased.
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

Koustas, E.; Trifylli, E.-M.; Sarantis, P.; Papadopoulos, N.; Papanikolopoulos, K.; Aloizos, G.; Damaskos, C.; Garmpis, N.; Garmpi, A.; Matthaios, D.; et al. An Insight into the Arising Role of MicroRNAs in Hepatocellular Carcinoma: Future Diagnostic and Therapeutic Approaches. Int. J. Mol. Sci. 2023, 24, 7168. https://doi.org/10.3390/ijms24087168

AMA Style

Koustas E, Trifylli E-M, Sarantis P, Papadopoulos N, Papanikolopoulos K, Aloizos G, Damaskos C, Garmpis N, Garmpi A, Matthaios D, et al. An Insight into the Arising Role of MicroRNAs in Hepatocellular Carcinoma: Future Diagnostic and Therapeutic Approaches. International Journal of Molecular Sciences. 2023; 24(8):7168. https://doi.org/10.3390/ijms24087168

Chicago/Turabian Style

Koustas, Evangelos, Eleni-Myrto Trifylli, Panagiotis Sarantis, Nikolaos Papadopoulos, Konstantinos Papanikolopoulos, Georgios Aloizos, Christos Damaskos, Nikolaos Garmpis, Anna Garmpi, Dimitris Matthaios, and et al. 2023. "An Insight into the Arising Role of MicroRNAs in Hepatocellular Carcinoma: Future Diagnostic and Therapeutic Approaches" International Journal of Molecular Sciences 24, no. 8: 7168. https://doi.org/10.3390/ijms24087168

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