Next Article in Journal
Identification of Novel Cathepsin B Inhibitors with Implications in Alzheimer’s Disease: Computational Refining and Biochemical Evaluation
Next Article in Special Issue
Feedback Signaling between Cholangiopathies, Ductular Reaction, and Non-Alcoholic Fatty Liver Disease
Previous Article in Journal
The Arbuscular Mycorrhizal Fungus Glomus viscosum Improves the Tolerance to Verticillium Wilt in Artichoke by Modulating the Antioxidant Defense Systems
Previous Article in Special Issue
Cyclic AMP Signaling in Biliary Proliferation: A Possible Target for Cholangiocarcinoma Treatment?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Circular RNA as An Epigenetic Regulator in Chronic Liver Diseases

1
Lab of Gastroenterology and Hepatology, West China Hospital, Sichuan University, NO. 1, 4th Keyuan Road, Chengdu 610093, China
2
Department of Gastroenterology, West China Hospital, Sichuan University, Chengdu 610041, China
*
Author to whom correspondence should be addressed.
Cells 2021, 10(8), 1945; https://doi.org/10.3390/cells10081945
Submission received: 12 May 2021 / Revised: 19 July 2021 / Accepted: 28 July 2021 / Published: 30 July 2021
(This article belongs to the Special Issue New Aspects and Mechanisms in Liver Diseases)

Abstract

:
Circular RNA (circRNA) is a type of non-coding RNA characterized by a covalently closed continuous loop. CircRNA is generated by pre-mRNA through back-splicing and is probably cleared up by extracellular vesicles. CircRNAs play a pivotal role in the epigenetic regulation of gene expression at transcriptional and post-transcriptional levels. Recently, circRNAs have been demonstrated to be involved in the regulation of liver homeostasis and diseases. However, the epigenetic role and underlying mechanisms of circRNAs in chronic liver diseases remain unclear. This review discussed the role of circRNAs in non-neoplastic chronic liver diseases, including alcoholic liver disease (ALD), metabolic-associated fatty liver disease (MAFLD), viral hepatitis, liver injury and regeneration, liver cirrhosis, and autoimmune liver disease. The review also highlighted that further efforts are urgently needed to develop circRNAs as novel diagnostics and therapeutics for chronic liver diseases.

1. Introduction

The liver is a critical organ for humans. It is composed of hepatocytes, biliary epithelial cells (cholangiocytes), hepatic stellate cells (HSCs), Kupffer cells, mast cells, and liver sinusoidal endothelial cells (LSECs). These cells cooperatively regulate hepatic function in numerous physiological processes, including glucose metabolism, lipid and cholesterol homeostasis, maintenance of energy balance, blood volume regulation, immune responses, and detoxification [1,2]; intercellular communication also contributes to the pathogenesis of liver diseases [3,4,5]. The liver has a powerful and unique regenerative capacity to maintain the constant size required for body homeostasis, regardless of injury [6,7]. However, various pathogenic agents can damage the liver, resulting in acute or chronic liver disease, such as alcoholic liver disease (ALD), metabolic-associated fatty liver disease (MAFLD), viral hepatitis, hepatocellular carcinoma (HCC), and their complications [8,9]. Liver cirrhosis accounts for approximately one million deaths worldwide annually [10]. The leading causes of cirrhosis in Western countries are alcohol and MAFLD, whereas viral hepatitis B is the main cause in China and other Asian countries [8]. However, the pathogenesis of liver diseases is still not fully understood.
Epigenetic research has accelerated rapidly in the past two decades, and achievements in epigenetics have advanced our knowledge of the pathogenesis of liver diseases [11,12,13]. In addition, epigenetic modifications have a significant impact on regulating gene expression and occurrence, and the development of diseases [14]. The major carriers of epigenetic information include non-coding RNAs, DNA methylation, histone acetylation, heterochromatin components, and polycomb proteins [15].
Among non-coding RNAs, microRNA (miRNA), long non-coding RNA (lncRNA) and circular RNA (circRNA) are the most well-studied non-coding RNAs. miRNAs are non-coding RNAs of only 20–22 nucleotides in length; they repress gene expression by the RNA-induced silencing complex (RISC) [16]. A single miRNA can correspond to many different mRNAs, whereas a single mRNA is regulated by multiple miRNAs [17]. lncRNAs are considered to be longer than 200 nucleotides, which is distinctly different from mRNAs, miRNAs, small nuclear RNAs (snoRNAs) and tRNAs. They are able to regulate gene expression at the level of chromatin modification, transcription and post-transcriptional processing through a diversity of mechanisms [16,18]. The ability of lncRNAs to regulate associated protein-coding genes might contribute to many diseases. The competitive endogenous RNA (ceRNA) hypothesis is a theory about lncRNA function that is gaining significant attention; it states that competition for miRNA binding sites regulates the expression of a target gene. The role of lncRNA in chronic liver diseases has previously been described and is not the focus of this review [19,20,21,22].
circRNAs are characterized by a covalently closed continuous loop [23]. They have the advantages of stability, conservation, and tissue specificity [24], generated by pre-mRNA through back-splicing (Figure 1).
Degradation of circRNAs cannot be achieved through many classical RNA degradation pathways. They might be eliminated from cells by exocytosis [23,24,25]. Enclosing circRNAs or circRNA complexes in vesicles and releasing them into the extracellular space is a way to move circRNAs out of the cytoplasm. However, with the protection of vesicles, circRNAs or circRNA complexes can reach other cells or tissues to act as messenger molecules or to fulfill other unknown functions [23]. The degradation of circRNA theoretically involves endonuclease. A recent study found that the N6-methylation of adenosine (m6A) modification facilitates the recruitment of endonucleases, affecting the degradation of circRNAs via a YTHDF2-HRSP12-RNase P/MRP axis [26].
circRNAs act through several mechanisms (Figure 2), including: (1) miRNA sponges or ceRNA mechanisms [16], which alleviate the inhibitory effects of miRNAs on target genes, thereby increasing the levels of the target genes [27]; (2) protein sponge, such as sponging RNA-binding protein components (RBPs); (3) regulating the expression of a parental gene; and (4) translating into peptides [24,28,29].
Through novel bioinformatics approaches integrated with biochemical enrichment strategies and deep sequencing, circRNAs can be studied in a comprehensive and in-depth perspective [30,31]. Many online resources have been developed, such as identification tools, databases, detection software systems, websites, and recommended nomenclature of circRNAs [28,32,33,34]. Gene Ontology (GO) and the Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway have generally been used to reveal the potential role of the differentially expressed circRNAs. As a rising star in the past few years, circRNAs have been proven to be involved in regulating liver homeostasis and diseases [35]. A certain number of circRNAs have been discovered as potential biomarkers or therapeutic targets for HCC [36,37,38,39,40,41,42,43]. As well as for HCC, circRNAs play an important role in non-neoplastic chronic liver diseases, such as ALD, MAFLD, liver fibrosis, and autoimmune diseases through epigenetic mechanisms [11,12,44,45,46].
Here, we conduct a literature review to summarize the role of circRNAs in ALD, MAFLD, viral hepatitis, liver injury and regeneration, liver cirrhosis, and autoimmune liver disease.

2. Circular RNAs in Chronic Liver Diseases

2.1. Alcoholic Liver Disease (ALD)

ALD, ranging from asymptomatic liver steatosis to alcoholic hepatitis and cirrhosis, is the most prevalent type of chronic liver disease worldwide [47]. ALD is characterized by fat accumulation, hepatocyte apoptosis, and inflammatory infiltration [48]. The pathogenesis of ALD involves several factors, including lipid deposition, excessive generation of reactive oxygen species (ROS), hepatocyte apoptosis, increased gut permeability, translocation of bacteria-derived lipopolysaccharides (LPS) from the gut into the liver, and excessive inflammatory responses that share some relationships with miRNAs and circRNAs [47,49,50].
miR-122 is one of the most well-studied miRNAs. The target genes of miR-122 include genes that are related to lipid metabolism (HMGCG, ApoE, MTTP, and PGC1a), liver fibrosis (P4HA1), hepatocellular cancer-relevant genes (Igf1R, ADAM10, cyclin G1, and KLF6) [51], hypoxia-related genes (HIF1α) [52], and inflammatory responses-associated genes (TNFRSF13C) [49]. miR-122 is a suppressive factor in ALD, and is sponged by circRNA_1639. circRNA_1639 is a pro-inflammatory factor identified in primary Kupffer cells in CCl4-induced liver fibrosis. circRNA_1639 induces ALD via activating the NF-κB signaling pathway through circRNA_1639/miR-122/TNFRSF13C axis [49]. In other studies of alcohol-induced animal models, mm9_circ_018725 [53] and mou_circ_1657 [54] have contributed to ALD by releasing pro-inflammatory cytokines and activating HSCs via interacting with miR-96-5p, respectively. miR-21 also plays a protective role against ALD through the anti-extrinsic apoptotic signaling pathway [55], but no circRNA corresponding to it under the ceRNA mechanism has been reported yet.
Another pivotal miRNA, miR-155, is a crucial regulator for inflammation and is increased in the liver and circulation in ALD [47]. Kupffer cell activation and sensitization to LPS are regulated by miRNA-155 in ALD [51]. Bioinformatics analysis predicted that circRNA-0067835 acts as an miR-155 sponge to regulate FOXO3a expression in the progress of liver fibrosis [56]. However, there is no report in which circRNAs specifically interact with miR-155 in ALD.
In summary, ALD is caused by various factors, including genetics, epigenetics, and non-genetic factors. The ceRNA mechanism might also play an essential role in the development of ALD, an area which needs further study. Additionally, alcohol intake can lead to changes in the composition of the gut microbiota. Functional alterations in the microbiota have been proven to be associated with ALD progression [50]. Similarly, gut microbiota regulates tumor lung metastasis via a circRNA/miRNA network [57]. Accordingly, gut microbiota might also interact with the circRNA/miRNA network in ALD, making it a promising research direction.

2.2. Metabolic-Associated Fatty Liver Disease (MAFLD)

Nonalcoholic fatty liver disease (NAFLD) has become the most common cause of chronic liver diseases [58]. Nonalcoholic steatohepatitis (NASH), currently recognized as part of NAFLD, progresses into liver cirrhosis and further decompensation liver cirrhosis in some individuals [59]. Recently, international experts reached a consensus that the disease abbreviation should be changed from NAFLD to metabolic dysfunction-associated fatty liver disease (MAFLD) [60]. Insulin resistance and abnormal lipid metabolism [61] are the hallmarks of MAFLD, and many circRNAs have effects on oxidative stress, including mitochondrial dysfunction and endoplasmic reticulum (ER) stress [62], and on autophagy or lipophagy in fatty livers [63]. For example, LPIN1 is a crucial gene in regulating triglyceride and phospholipid biosynthesis [64]. The circRNA_021412/miR-1972/LPIN1 signaling pathway reflected the regulatory mechanisms underlying the steatosis-related circRNA–miRNA–mRNA network [65]. CircScd1 promotes fatty liver disease by increasing hepatocellular lipidosis in NAFLD via the JAK2/STAT5 pathway [66]. Low-density lipoprotein (LDL), peroxisome proliferator-activated receptor α (PPARα), and their corresponding genes are critical in lipometabolisim. circRNA_0049392 is differentially expressed in an NAFLD mouse model [67], whereas target miRNAs of circRNA_0049392 are miR-6919-5p and miR-7037-5p, which are validated to regulate LDL [67]. miR-34a, an inhibitor of PPARα, is significantly up-regulated in different models of rodent hepatic steatosis. circRNA_0046366 and circRNA_0046367, as the sponges of miR-34a, abolish the inhibitory effect of miR-34a on PPARα, thus ameliorating the lipoxidative stress [68,69]. The circRNA_002581/miR-122/CPEB1 axis promoting NASH is associated with the dysregulation of autophagy, which can be inhibited through the PTEN/AMPK/mTOR signaling pathway [70].
Liver fibrosis is also recognized as the hallmark of disease progression in NASH [58], and fibroblasts, along with HSCs, contribute to the occurrence and development of liver fibrosis. circRNA_29981, a potential regulator of HSC activation, was identified in the murine NASH model [71]. Moreover, a steatohepatitis-associated mitochondrial circRNA-SCAR can shut down the mitochondrial permeability transition pore (mPTP) by binding to ATP5B, thereby inhibiting mitochondrial ROS (mtROS) output and fibroblast activation [62]. It seems that circRNAs might also be involved in the progress of NASH to liver fibrosis. circRNAs are also involved in the efficacy and mechanism of NAFLD treatment. Metformin is now recommended for NAFLD treatment to improve insulin resistance in NAFLD [58]. Metformin systematically alleviates the transcriptome alterations induced by a high-fat diet, while the circRNA–miRNA–mRNA network may attribute to the beneficial effect of metformin on NAFLD [72]. The traditional Chinese medicine prescription, Qianggan formula, has also been reported to improve NASH by modulating the lncRNA/circRNA immune ceRNA network [73].
Altogether, there is increasing evidence suggesting that circRNAs, as well as ceRNA mechanisms, play a crucial role in the development and progression of MAFLD. It is a network where advances in the study of one molecule may provide ideas for the study of other disease mechanisms. For example, miR-34a also functions as a profibrotic factor that promotes alcohol-induced liver fibrosis by reducing HSC senescence and increasing the senescence of hepatocytes [74]. It is possible that in ALD, miR-34a could interact with certain circRNAs to promote or inhibit disease progression. Further studies are needed to deepen the mechanism.

2.3. Viral Hepatitis (Hepatitis B and Hepatitis C)

Chronic hepatitis B virus (HBV) infection is one of the leading causes of liver fibrosis, cirrhosis, and hepatocellular carcinoma [75,76]. Interferons (IFNs) are used in anti-HBV therapy. Recently, Zhang et al. found that the circRNA hsa_circ_0004812 could regulate the levels of IFNs and thus their anti-HBV efficiency [77]. Knockdown of hsa_circ_0004812 in HBV-infected cells resulted in a significant increase in mRNA and protein levels of IFN-α and IFN-β [77]. TGFβ2 has suppressive effects on IL-2-dependent T-cell growth and is significantly down-regulated in chronic hepatitis B patients [78]. Zhou et al. found a strong positive correlation between hsa_circ_0000650 and TGFβ2 in the liver biopsies from chronic hepatitis B patients compared with the control group, whereas a robust negative correlation between hsa_circ_0000650 and miR-6873-3p was also established [79].
Hepatitis C virus (HCV) is also an important pathogen of hepatitis, cirrhosis and HCC. HCV is hepatocyte-specific, partly due to its dependence on the host miR-122, which is exceptionally abundant in hepatocytes [80]. miR-122 is thought to enhance the viral RNA abundance and has already been utilized as a drug target for anti-HCV therapy [81]. Jost et al. designed artificial circRNAs containing an array of miR-122 binding sites, thus efficiently sequestering miR-122 [82]. They also found that these circRNAs might work as protein sponges containing binding sites derived from SELEX- or CLIP-data, and were available for multiple RNA-binding proteins [82]. Besides, circPSD3, inhibiting viral RNA abundance at a post-translational step, was shown to promote HCV infection [83].
As circRNAs are more stable than linear RNAs, they might play crucial roles in viral infection pathogenesis. Interestingly, there are also many covalently closed circular DNAs (cccDNA) in the genome of the hepatitis virus itself. For example, HBV cccDNA produced during HBV replication is predominant in the liver of CHB patients [84], and their levels can be regulated by the RNA-binding protein, DHX9 [85]. However, whether the circRNA originated from cccDNA remains unclear.

2.4. Liver Injury and Liver Regeneration

The liver has a remarkable capacity for regeneration when stimulated with physical (loss of volume or ischemia) or chemical (medicine or infection) damage [86]. In the 2/3 partial hepatectomy (PH) murine model, a ceRNA network containing 5 circRNAs, 28 target genes, and 533 miRNAs was identified during the DNA synthesis phase (36 h) of liver regeneration [87]. Knockdown of circ_0002498, circ_0000281, and circ_0003822 significantly restrained the proliferation of hepatocytes. In contrast, the knockdown of circ_0003745 and circ_0004245 accelerated the proliferation of hepatocytes [87]. Similarly, a circRNA–miRNA interactions network was also established in the rat PH model. Four significantly changed circRNAs (circ432, circ2077, circ1366, and circ15) were screened as key circRNAs during the priming phase of liver regeneration. These circRNAs might regulate the rat liver regeneration by controlling the expression of MAPK14, FN1, TNFRSF21, and GOT1, respectively [88].
Hepatic ischemia and reperfusion injury (IRI) are major complications during surgical procedures such as liver transplantation, liver resection, and trauma surgery [89]. The comprehensive analysis of circRNAs during hepatic IRI was performed using next-generation RNA sequencing [90,91], and several circRNAs were selected as potential targets against hepatic IRI. In the hepatic IRI model established by hepatic artery occlusion, circRNA_017753–miR-218-5p–Jade1, circRNA_017753–miR-7002-3p–Jade1, and circRNA_017753–miR-7008-3p–Jade1 ceRNA signaling pathways may play important roles in the mechanisms of ischemic preconditioning protection [92].
For the radiation-induced liver injury, circRSF1 was upregulated in irradiated LX2 cells, and promoted inflammatory and fibrotic phenotypes of HSC by sponging miR-146a-5p [93]. Similarly, circTUBD1 acted as an miR-146a-5p sponge to affect the viability and pro-inflammatory cytokine production of LX-2 cells through the Toll-like receptor 4 (TLR4)/NF-κB signaling pathway [94]. These two studies suggest that circRSF1 and circTUBD1 are potential targets for radiation-induced liver disease therapy.
Drug-induced liver injury is one of the major causes of liver disease. For acetaminophen (APAP)-induced liver injury, p66Shc, a master regulator of mtROS, is upregulated in liver tissues in response to APAP. circ-CBFB acts as the sponge of miR-185-5p, triggering mitochondrial dynamics perturbation via p66Shc [95]. For the hydrogen peroxide (H2O2)-induced injury, circRNA-4099 augments H2O2-induced damage by inhibiting miR-706 and triggering Keap1/Nrf2 and p38MAPK in the L02 cells [96]. For anti-tuberculosis drug-induced liver injury, circMARS may participate in the compensation and regeneration of the liver through the circMARS–miR-6808-5p/miR-6874-3p/miR-3157-5p–KMT2C–EGFR functional axis [97].
For infection-induced liver injury, the mmu_circRNA_005186/miR-124-3p/Epha2 axis is associated with LPS-induced inflammation, and silencing mmu_circRNA_005186 attenuates this kind of inflammation [98]; circ_0003420 also mediates LPS-induced cellular injury and inflammation, mainly by targeting the neuronal PAS domain protein 4 (NPAS4) mRNA [99].
The current knowledge mentioned above is somewhat about the critical role of circRNAs in liver injury and liver regeneration. However, further studies that deepen the underlying mechanisms in these liver injuries are still urgently needed.

2.5. Liver Fibrosis/Cirrhosis

During chronic liver diseases, persistent inflammatory responses and parenchymal injury eventually lead to liver fibrosis and liver cirrhosis [100,101]. Chronic loss of hepatocytes occurs in chronic liver diseases of any etiology and is associated with the activation of HSCs and an abnormal liver microenvironment [6]. Thus, inhibiting HSC activation has been considered as one of the most potent approaches to alleviate the progression of liver cirrhosis.
Many circRNAs have been found to suppress the activation of HSCs. Telomerase-associated protein 1 (TEP1), a component of telomerase and target of hsa-miR-660-3p [102], plays a key role in the development of liver fibrosis and cirrhosis [103]. Overexpression of hsa_circ_0004018, a sponge of hsa-miR-660-3p, increases the expression of TEP1 and significantly suppresses the proliferation and activation of HSCs [104]. Consistently, hsa_circ_0007874 (cMTO1) could inhibit HSC activation through sponging miR-181b-5p and miR-17-5p [105,106]. Besides, hsa_circ_0070963 inhibits liver fibrosis via the regulation of miR-223-3p, which is related to HSC activation [107]. By targeting miR-18b-3p, circFBXW4 attenuates murine liver fibrogenesis with an anti-inflammation effect via inhibiting the activation and proliferation of HSCs, and inducing HSC apoptosis [108]. Moreover, mmu_circ_34116 can also inhibit the HSC activation via mmu_circ_34116/miR-22-3P/BMP7 signal axis [109,110]. Recently, circPSD3 was found to inhibit the activation and proliferation of HSCs by targeting the miR-92b-3p/Smad7 axis [111].
As mentioned above, these circRNAs keep HSCs in quiescent status. However, there are still some circRNAs that promote the activation of HSCs through different mechanisms. Follistatin-like 1 (Fstl1) contributes to TGF-β-induced HSC proliferation, the expression of α-SMA, and type I collagen, while TLR4 contributes to LPS-induced HSC activation [112]. circ-PWWP2A sponges miR-203 and miR-223, and subsequently increases the expression of Fstl1 and TLR4, respectively, eventually promoting the activation and proliferation of HSCs [112,113]. circRNA-0067835 acts as a sponge of miR-155 to promote FOXO3a expression, subsequently promoting fibrosis progression [56]. hsa_circ_0071410 decreases the expression of miR-9-5p, resulting in the progression of irradiation-induced HSC activation [114,115].
lncRNAs have also been shown to regulate the process of liver fibrosis by inhibiting miRNAs. For example, lncRNA-p21 can inhibit miR-181b and miR-17-5p, which in turn inhibit the activation of HSCs through two different signaling pathways [20]. Interestingly, both cMTO1 and lncRNA-p21 sponge miR-181b and miR-17-5p, which indicates the potential circRNA–miRNA–lncRNA network in the development of liver fibrosis.
Liver cirrhosis is a common ultimate consequence of chronic liver diseases. As HSCs are supposed to be a major source of collagen deposit during liver cirrhosis, the role of circRNAs in HSC activation attracts the most attention during chronic liver diseases. The hepatic cells are mainly composed of hepatocytes, LSECs, HSCs, cholangiocytes, and Kupffer cells, which precisely “cross-talk” to maintain normal hepatic functions and to aid cell survival [116]. Thus, the functions of circRNAs from other hepatic cells should also be investigated during chronic liver diseases and liver cirrhosis.

2.6. Autoimmune Liver Disease

Currently, the most common autoimmune liver diseases in clinical practice are primary biliary cholangitis (PBC), primary sclerosing cholangitis (PSC), and autoimmune hepatitis (AIH) [117]. They have a high heterogeneity of clinical presentation and result from the interactions of multiple genetic and environmental factors. circRNAs can contribute to the development of autoimmune diseases, mainly by serving as miRNA sponges to regulate DNA methylation, immune response and inflammatory response [118]. miR-522-3p might be involved in chronic inflammatory disorders, and miR-943 is a participant in the repair of DNA double-strand breaks [119]. Zheng et al. found that hsa_circ_402458, the sponge of hsa-miR-522-3p and hsa-miR-943, was significantly higher in PBC patients not treated with ursodeoxycholic acid (UDCA) than in those treated with it [119]. Liu et al. recently reported that the pathogenesis of AIH may involve the mmu_circ_0001520/mmu-miR-193b-3p/MAPK10 axis [120]. However, the role of circRNAs in PSC has not yet been reported. Overall, there are relatively few studies focusing on circRNAs and autoimmune liver disease. The etiology of autoimmune diseases is complex [121]. In different diseases, there is a significant overlap of susceptibility factors. Thus, the role of circRNAs in autoimmune liver disease may be complicated and would benefit from further study.

3. Conclusions and Prospects

In this literature review, we discussed the epigenetic role of circRNAs in non-neoplastic chronic liver diseases (Table 1). Almost all current studies suggest that circRNAs act in chronic liver diseases through the ceRNA mechanism, i.e., through interactions with miRNAs, supporting a certain association between circRNAs and chronic liver diseases (Figure 3). Some miRNAs can interact with both circRNAs and lncRNAs in hepatic fibrosis, but in other chronic liver diseases, such intermediate miRNAs have not been reported.
However, the ceRNA network is more than just the interaction of circRNAs and miRNAs. Expressed 3′ untranslated regions (3′ UTRs), pseudogenes and lncRNAs may also act as ceRNAs to sequester miRNA activity by competing for miRNA binding sites. Investigating further connections among circRNA, miRNA and lncRNA in liver diseases could be a promising research direction in the future. Besides, circRNAs can be obtained from tissues, serum, and exosomes. The expression and function of circRNAs in the sites mentioned above may vary due to the different microenvironments. The clearance of circRNAs can be seen under a possible mechanism that is eliminating circRNAs from cells into the extracellular space by extracellular vesicles [24,25]. However, there are few studies on the role and function of circRNAs in serum and exosomes. Research on large samples from different sources is warranted to understand these circRNAs further.
Another important part of epigenetics is m6A modification. m6A modification plays a unique role in critical physiological hepatic functions and various liver diseases. It participates in glucose and lipid homeostasis involving the development of MAFLD and liver cancer pathogenesis and metastasis [122,123]. m6A is the most abundant internal modification known on DNA, mRNAs and lncRNAs [124,125]. Almost all studies have focused on the effects of m6A and mRNA, miRNA or lncRNA in MAFLD, liver fibrosis, PBC, and HCC [45,123,126,127,128,129]. m6A is widespread in circRNAs [130], regulating circRNA translation and facilitating circRNA degradation [131]. However, the relationship between m6A and circRNAs in chronic liver diseases remains unclear. Recently, Zhou et al. defined thousands of m6A circRNAs with cell-type-specific expression [130]. Although much of the same cellular machinery is shared between m6A-modified circRNAs and mRNAs, m6A circRNAs exhibit distinct patterns of m6A modifications compared with mRNAs. The discovery of m6A circRNAs raises many questions that will need to be addressed in future studies.
Liver diseases are a heavy burden on healthcare and account for significant costs worldwide. Clinicians and public health professionals may be more interested in how to use the characteristics of circRNAs and their interconnections with other molecules to diagnose and treat liver diseases. The articles published so far only provide new theories or ideas from the perspective of basic research, and there is still a long way to go before any clinical application. Therefore, it is worthwhile to reveal the full biological relevance of circRNAs and their potential therapeutics.
In conclusion, circRNAs play an increasingly crucial role in the epigenetic regulation of the onset and progression of chronic liver diseases. Further efforts are urgently needed in order to develop circRNAs as novel diagnostics and therapeutics.

Author Contributions

J.G. conceived and supervised the study; X.Z., X.Y., Q.C. and C.T. analyzed articles and finalized the figures; X.Z., C.T. and J.G. wrote the manuscript with input from all the authors. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Fund of China (81670551, U1702281, 81873584, 82000613 and 82000574); Sichuan Science and Technology Program (2020YJ0084 and 2021YFS0147); National Key R&D Program of China (2017YFA0205404).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kubes, P.; Jenne, C. Immune Responses in the Liver. Annu. Rev. Immunol. 2018, 36, 247–277. [Google Scholar] [CrossRef]
  2. Trefts, E.; Gannon, M.; Wasserman, D.H. The liver. Curr. Biol. 2017, 27, R1147–R1151. [Google Scholar] [CrossRef]
  3. Sato, K.; Kennedy, L.; Liangpunsakul, S.; Kusumanchi, P.; Yang, Z.; Meng, F.; Glaser, S.; Francis, H.; Alpini, G. Intercellular Communication between Hepatic Cells in Liver Diseases. Int. J. Mol. Sci. 2019, 20, 2180. [Google Scholar] [CrossRef] [Green Version]
  4. Jarido, V.; Kennedy, L.; Hargrove, L.; Demieville, J.; Thomson, J.; Stephenson, K.; Francis, H. The emerging role of mast cells in liver disease. Am. J. Physiol. Gastrointest. Liver Physiol. 2017, 313, G89–G101. [Google Scholar] [CrossRef] [PubMed]
  5. Francis, H.; Meininger, C.J. A review of mast cells and liver disease: What have we learned? Dig. Liver Dis. 2010, 42, 529–536. [Google Scholar] [CrossRef] [PubMed]
  6. Michalopoulos, G.K.; Bhushan, B. Liver regeneration: Biological and pathological mechanisms and implications. Nat. Rev. Gastroenterol. Hepatol. 2021, 18, 40–55. [Google Scholar] [CrossRef] [PubMed]
  7. Duncan, A.W.; Dorrell, C.; Grompe, M. Stem cells and liver regeneration. Gastroenterology 2009, 137, 466–481. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Asrani, S.K.; Devarbhavi, H.; Eaton, J.; Kamath, P.S. Burden of liver diseases in the world. J. Hepatol. 2019, 70, 151–171. [Google Scholar] [CrossRef] [PubMed]
  9. Wang, F.S.; Fan, J.G.; Zhang, Z.; Gao, B.; Wang, H.Y. The global burden of liver disease: The major impact of China. Hepatology 2014, 60, 2099–2108. [Google Scholar] [CrossRef]
  10. Mokdad, A.A.; Lopez, A.D.; Shahraz, S.; Lozano, R.; Mokdad, A.H.; Stanaway, J.; Murray, C.J.; Naghavi, M. Liver cirrhosis mortality in 187 countries between 1980 and 2010: A systematic analysis. BMC Med. 2014, 12, 145. [Google Scholar] [CrossRef] [Green Version]
  11. Hardy, T.; Mann, D.A. Epigenetics in liver disease: From biology to therapeutics. Gut 2016, 65, 1895–1905. [Google Scholar] [CrossRef] [PubMed]
  12. Mann, D.A. Epigenetics in liver disease. Hepatology 2014, 60, 1418–1425. [Google Scholar] [CrossRef] [Green Version]
  13. Blum, H.E. Gastrointestinal and liver diseases: Genetic and epigenetic markers. Gut 2011, 60, 1630–1634. [Google Scholar] [CrossRef]
  14. Zhang, L.; Lu, Q.; Chang, C. Epigenetics in Health and Disease. Adv. Exp. Med. Biol. 2020, 1253, 3–55. [Google Scholar] [CrossRef]
  15. Cavalli, G.; Heard, E. Advances in epigenetics link genetics to the environment and disease. Nature 2019, 571, 489–499. [Google Scholar] [CrossRef] [Green Version]
  16. Thomson, D.W.; Dinger, M.E. Endogenous microRNA sponges: Evidence and controversy. Nat. Rev. Genet. 2016, 17, 272–283. [Google Scholar] [CrossRef] [PubMed]
  17. Lai, S.; Iwakiri, Y. Is miR-21 a potent target for liver fibrosis? Hepatology 2018, 67, 2082–2084. [Google Scholar] [CrossRef]
  18. Mercer, T.R.; Dinger, M.E.; Mattick, J.S. Long non-coding RNAs: Insights into functions. Nat. Rev. Genet. 2009, 10, 155–159. [Google Scholar] [CrossRef] [PubMed]
  19. Shabgah, A.G.; Norouzi, F.; Hedayati-Moghadam, M.; Soleimani, D.; Pahlavani, N.; Navashenaq, J.G. A comprehensive review of long non-coding RNAs in the pathogenesis and development of non-alcoholic fatty liver disease. Nutr. Metab. 2021, 18, 22. [Google Scholar] [CrossRef] [PubMed]
  20. Peng, H.; Wan, L.Y.; Liang, J.J.; Zhang, Y.Q.; Ai, W.B.; Wu, J.F. The roles of lncRNA in hepatic fibrosis. Cell Biosci. 2018, 8, 63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Takahashi, K.; Yan, I.; Haga, H.; Patel, T. Long noncoding RNA in liver diseases. Hepatology 2014, 60, 744–753. [Google Scholar] [CrossRef]
  22. Wang, Y.; Hylemon, P.B.; Zhou, H. Long non-coding RNA H19: A key player in liver diseases. Hepatology 2021. [Google Scholar] [CrossRef] [PubMed]
  23. Patop, I.L.; Wüst, S.; Kadener, S. Past, present, and future of circRNAs. Embo. J. 2019, 38, e100836. [Google Scholar] [CrossRef] [PubMed]
  24. Yao, T.; Chen, Q.; Fu, L.; Guo, J. Circular RNAs: Biogenesis, properties, roles, and their relationships with liver diseases. Hepatol. Res. 2017, 47, 497–504. [Google Scholar] [CrossRef]
  25. Lasda, E.; Parker, R. Circular RNAs Co-Precipitate with Extracellular Vesicles: A Possible Mechanism for circRNA Clearance. PLoS ONE 2016, 11, e0148407. [Google Scholar] [CrossRef] [Green Version]
  26. Park, O.H.; Ha, H.; Lee, Y.; Boo, S.H.; Kwon, D.H.; Song, H.K.; Kim, Y.K. Endoribonucleolytic Cleavage of m(6)A-Containing RNAs by RNase P/MRP Complex. Mol. Cell 2019, 74, 494–507.e498. [Google Scholar] [CrossRef] [PubMed]
  27. Hansen, T.B.; Jensen, T.I.; Clausen, B.H.; Bramsen, J.B.; Finsen, B.; Damgaard, C.K.; Kjems, J. Natural RNA circles function as efficient microRNA sponges. Nature 2013, 495, 384–388. [Google Scholar] [CrossRef] [PubMed]
  28. Haque, S.; Harries, L.W. Circular RNAs (circRNAs) in Health and Disease. Genes 2017, 8, 353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Fu, L.Y.; Wang, S.W.; Hu, M.Y.; Jiang, Z.L.; Shen, L.L.; Zhou, Y.P.; Guo, J.M.; Hu, Y.R. Circular RNAs in liver diseases: Mechanisms and therapeutic targets. Life Sci. 2021, 264, 118707. [Google Scholar] [CrossRef] [PubMed]
  30. Hansen, T.B.; Venø, M.T.; Damgaard, C.K.; Kjems, J. Comparison of circular RNA prediction tools. Nucleic Acids Res. 2016, 44, e58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Jeck, W.R.; Sharpless, N.E. Detecting and characterizing circular RNAs. Nat. Biotechnol. 2014, 32, 453–461. [Google Scholar] [CrossRef] [PubMed]
  32. Liu, M.; Wang, Q.; Shen, J.; Yang, B.B.; Ding, X. Circbank: A comprehensive database for circRNA with standard nomenclature. RNA Biol. 2019, 16, 899–905. [Google Scholar] [CrossRef]
  33. Glažar, P.; Papavasileiou, P.; Rajewsky, N. circBase: A database for circular RNAs. RNA 2014, 20, 1666–1670. [Google Scholar] [CrossRef] [Green Version]
  34. Wang, J.; Samuels, D.C.; Zhao, S.; Xiang, Y.; Zhao, Y.Y.; Guo, Y. Current Research on Non-Coding Ribonucleic Acid (RNA). Genes 2017, 8, 366. [Google Scholar] [CrossRef] [Green Version]
  35. Song, M.; Xia, L.; Sun, M.; Yang, C.; Wang, F. Circular RNA in Liver: Health and Diseases. Adv. Exp. Med. Biol. 2018, 1087, 245–257. [Google Scholar] [CrossRef] [PubMed]
  36. Wen, J.; Liao, J.; Liang, J.; Chen, X.P.; Zhang, B.; Chu, L. Circular RNA HIPK3: A Key Circular RNA in a Variety of Human Cancers. Front. Oncol. 2020, 10, 773. [Google Scholar] [CrossRef]
  37. Li, D.; Zhang, J.; Li, J. Role of miRNA sponges in hepatocellular carcinoma. Clin. Chim. Acta 2020, 500, 10–19. [Google Scholar] [CrossRef]
  38. Shang, W.; Adzika, G.K.; Li, Y.; Huang, Q.; Ding, N.; Chinembiri, B.; Rashid, M.S.; Machuki, J.O. Molecular mechanisms of circular RNAs, transforming growth factor-β, and long noncoding RNAs in hepatocellular carcinoma. Cancer Med. 2019, 8, 6684–6699. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Qiu, L.; Xu, H.; Ji, M.; Shang, D.; Lu, Z.; Wu, Y.; Tu, Z.; Liu, H. Circular RNAs in hepatocellular carcinoma: Biomarkers, functions and mechanisms. Life Sci. 2019, 231, 116660. [Google Scholar] [CrossRef]
  40. Wang, M.; Yu, F.; Li, P. Circular RNAs: Characteristics, Function and Clinical Significance in Hepatocellular Carcinoma. Cancers 2018, 10, 258. [Google Scholar] [CrossRef] [Green Version]
  41. Qiu, L.P.; Wu, Y.H.; Yu, X.F.; Tang, Q.; Chen, L.; Chen, K.P. The Emerging Role of Circular RNAs in Hepatocellular Carcinoma. J. Cancer 2018, 9, 1548–1559. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Hu, J.; Li, P.; Song, Y.; Ge, Y.X.; Meng, X.M.; Huang, C.; Li, J.; Xu, T. Progress and prospects of circular RNAs in Hepatocellular carcinoma: Novel insights into their function. J. Cell Physiol. 2018, 233, 4408–4422. [Google Scholar] [CrossRef] [PubMed]
  43. Fu, L.; Jiang, Z.; Li, T.; Hu, Y.; Guo, J. Circular RNAs in hepatocellular carcinoma: Functions and implications. Cancer Med. 2018, 7, 3101–3109. [Google Scholar] [CrossRef]
  44. Tanaka, A.; Leung, P.S.C.; Gershwin, M.E. The genetics of primary biliary cholangitis. Curr. Opin Gastroenterol. 2019, 35, 93–98. [Google Scholar] [CrossRef]
  45. Eslam, M.; Valenti, L.; Romeo, S. Genetics and epigenetics of NAFLD and NASH: Clinical impact. J. Hepatol. 2018, 68, 268–279. [Google Scholar] [CrossRef]
  46. Chien, Y.; Tsai, P.H.; Lai, Y.H.; Lu, K.H.; Liu, C.Y.; Lin, H.F.; Huang, C.S.; Wu, W.W.; Wang, C.Y. CircularRNA as novel biomarkers in liver diseases. J. Chin. Med. Assoc. 2020, 83, 15–17. [Google Scholar] [CrossRef] [PubMed]
  47. Seitz, H.K.; Bataller, R.; Cortez-Pinto, H.; Gao, B.; Gual, A.; Lackner, C.; Mathurin, P.; Mueller, S.; Szabo, G.; Tsukamoto, H. Alcoholic liver disease. Nat. Rev. Dis. Primers 2018, 4, 16. [Google Scholar] [CrossRef] [PubMed]
  48. Sehrawat, T.S.; Liu, M.; Shah, V.H. The knowns and unknowns of treatment for alcoholic hepatitis. Lancet Gastroenterol. Hepatol. 2020, 5, 494–506. [Google Scholar] [CrossRef]
  49. Lu, X.; Liu, Y.; Xuan, W.; Ye, J.; Yao, H.; Huang, C.; Li, J. Circ_1639 induces cells inflammation responses by sponging miR-122 and regulating TNFRSF13C expression in alcoholic liver disease. Toxicol. Lett. 2019, 314, 89–97. [Google Scholar] [CrossRef]
  50. Bajaj, J.S. Alcohol, liver disease and the gut microbiota. Nat. Rev. Gastroenterol. Hepatol. 2019, 16, 235–246. [Google Scholar] [CrossRef]
  51. Szabo, G.; Satishchandran, A. MicroRNAs in alcoholic liver disease. Semin. Liver Dis. 2015, 35, 36–42. [Google Scholar] [CrossRef] [Green Version]
  52. Satishchandran, A.; Ambade, A.; Rao, S.; Hsueh, Y.C.; Iracheta-Vellve, A.; Tornai, D.; Lowe, P.; Gyongyosi, B.; Li, J.; Catalano, D.; et al. MicroRNA 122, Regulated by GRLH2, Protects Livers of Mice and Patients From Ethanol-Induced Liver Disease. Gastroenterology 2018, 154, 238–252.e237. [Google Scholar] [CrossRef] [Green Version]
  53. Meng, H.; Wang, L.; You, H.; Huang, C.; Li, J. Circular RNA expression profile of liver tissues in an EtOH-induced mouse model of alcoholic hepatitis. Eur. J. Pharmacol. 2019, 862, 172642. [Google Scholar] [CrossRef] [PubMed]
  54. Dou, X.; Feng, L.; Ying, N.; Ding, Q.; Song, Q.; Jiang, F.; Wang, C.; Li, S. RNA Sequencing Reveals a Comprehensive Circular RNA Expression Profile in a Mouse Model of Alcoholic Liver Disease. Alcohol. Clin. Exp. Res. 2020, 44, 415–422. [Google Scholar] [CrossRef]
  55. Francis, H.; McDaniel, K.; Han, Y.; Liu, X.; Kennedy, L.; Yang, F.; McCarra, J.; Zhou, T.; Glaser, S.; Venter, J.; et al. Regulation of the extrinsic apoptotic pathway by microRNA-21 in alcoholic liver injury. J. Biol. Chem. 2014, 289, 27526–27539. [Google Scholar] [CrossRef] [Green Version]
  56. Zhu, L.; Ren, T.; Zhu, Z.; Cheng, M.; Mou, Q.; Mu, M.; Liu, Y.; Yao, Y.; Cheng, Y.; Zhang, B.; et al. Thymosin-β4 Mediates Hepatic Stellate Cell Activation by Interfering with CircRNA-0067835/miR-155/FoxO3 Signaling Pathway. Cell Physiol. Biochem. 2018, 51, 1389–1398. [Google Scholar] [CrossRef] [PubMed]
  57. Zhu, Z.; Huang, J.; Li, X.; Xing, J.; Chen, Q.; Liu, R.; Hua, F.; Qiu, Z.; Song, Y.; Bai, C.; et al. Gut microbiota regulate tumor metastasis via circRNA/miRNA networks. Gut Microbes 2020, 12, 1788891. [Google Scholar] [CrossRef] [PubMed]
  58. Sanyal, A.J. Past, present and future perspectives in nonalcoholic fatty liver disease. Nat. Rev. Gastroenterol. Hepatol. 2019, 16, 377–386. [Google Scholar] [CrossRef]
  59. Neuschwander-Tetri, B.A. Non-alcoholic fatty liver disease. BMC Med. 2017, 15, 45. [Google Scholar] [CrossRef] [Green Version]
  60. Eslam, M.; Newsome, P.N.; Sarin, S.K.; Anstee, Q.M.; Targher, G.; Romero-Gomez, M.; Zelber-Sagi, S.; Wai-Sun Wong, V.; Dufour, J.F.; Schattenberg, J.M.; et al. A new definition for metabolic dysfunction-associated fatty liver disease: An international expert consensus statement. J. Hepatol. 2020, 73, 202–209. [Google Scholar] [CrossRef]
  61. Yu, G.; Yang, Z.; Peng, T.; Lv, Y. Circular RNAs: Rising stars in lipid metabolism and lipid disorders. J. Cell Physiol. 2021, 236, 4797–4806. [Google Scholar] [CrossRef]
  62. Zhao, Q.; Liu, J.; Deng, H.; Ma, R.; Liao, J.Y.; Liang, H.; Hu, J.; Li, J.; Guo, Z.; Cai, J.; et al. Targeting Mitochondria-Located circRNA SCAR Alleviates NASH via Reducing mROS Output. Cell 2020, 183, 76–93. [Google Scholar] [CrossRef]
  63. Chen, C.; Zhang, X.; Deng, Y.; Cui, Q.; Zhu, J.; Ren, H.; Liu, Y.; Hu, X.; Zuo, J.; Peng, Y. Regulatory roles of circRNAs in adipogenesis and lipid metabolism: Emerging insights into lipid-related diseases. FEBS J. 2020. [Google Scholar] [CrossRef] [PubMed]
  64. Csaki, L.S.; Reue, K. Lipins: Multifunctional lipid metabolism proteins. Annu. Rev. Nutr. 2010, 30, 257–272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Guo, X.Y.; He, C.X.; Wang, Y.Q.; Sun, C.; Li, G.M.; Su, Q.; Pan, Q.; Fan, J.G. Circular RNA Profiling and Bioinformatic Modeling Identify Its Regulatory Role in Hepatic Steatosis. Biomed. Res. Int. 2017, 2017, 5936171. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Li, P.; Shan, K.; Liu, Y.; Zhang, Y.; Xu, L.; Xu, L. CircScd1 Promotes Fatty Liver Disease via the Janus Kinase 2/Signal Transducer and Activator of Transcription 5 Pathway. Dig. Dis. Sci. 2019, 64, 113–122. [Google Scholar] [CrossRef]
  67. Yuan, X.; Diao, J.; Du, A.; Wen, S.; Zhou, L.; Pan, Y. Circular RNA expression profiles and features in NAFLD mice: A study using RNA-seq data. J. Transl. Med. 2020, 18, 476. [Google Scholar] [CrossRef]
  68. Guo, X.Y.; Sun, F.; Chen, J.N.; Wang, Y.Q.; Pan, Q.; Fan, J.G. circRNA_0046366 inhibits hepatocellular steatosis by normalization of PPAR signaling. World J. Gastroenterol. 2018, 24, 323–337. [Google Scholar] [CrossRef]
  69. Guo, X.Y.; Chen, J.N.; Sun, F.; Wang, Y.Q.; Pan, Q.; Fan, J.G. circRNA_0046367 Prevents Hepatoxicity of Lipid Peroxidation: An Inhibitory Role against Hepatic Steatosis. Oxid Med. Cell Longev. 2017, 2017, 3960197. [Google Scholar] [CrossRef]
  70. Jin, X.; Gao, J.; Zheng, R.; Yu, M.; Ren, Y.; Yan, T.; Huang, Y.; Li, Y. Antagonizing circRNA_002581-miR-122-CPEB1 axis alleviates NASH through restoring PTEN-AMPK-mTOR pathway regulated autophagy. Cell Death Dis. 2020, 11, 123. [Google Scholar] [CrossRef]
  71. Ou, Q.; Zhao, Y.; Zhou, J.; Wu, X. Comprehensive circular RNA expression profiles in a mouse model of nonalcoholic steatohepatitis. Mol. Med. Rep. 2019, 19, 2636–2648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Guo, J.; Zhou, Y.; Cheng, Y.; Fang, W.; Hu, G.; Wei, J.; Lin, Y.; Man, Y.; Guo, L.; Sun, M.; et al. Metformin-Induced Changes of the Coding Transcriptome and Non-Coding RNAs in the Livers of Non-Alcoholic Fatty Liver Disease Mice. Cell Physiol. Biochem. 2018, 45, 1487–1505. [Google Scholar] [CrossRef]
  73. Zhu, M.; Li, M.; Zhou, W.; Yang, Y.; Li, F.; Zhang, L.; Ji, G. Qianggan extract improved nonalcoholic steatohepatitis by modulating lncRNA/circRNA immune ceRNA networks. BMC Complement. Altern Med. 2019, 19, 156. [Google Scholar] [CrossRef] [Green Version]
  74. Wan, Y.; McDaniel, K.; Wu, N.; Ramos-Lorenzo, S.; Glaser, T.; Venter, J.; Francis, H.; Kennedy, L.; Sato, K.; Zhou, T.; et al. Regulation of Cellular Senescence by miR-34a in Alcoholic Liver Injury. Am. J. Pathol. 2017, 187, 2788–2798. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Yuen, M.-F.; Chen, D.-S.; Dusheiko, G.M.; Janssen, H.L.A.; Lau, D.T.Y.; Locarnini, S.A.; Peters, M.G.; Lai, C.-L. Hepatitis B virus infection. Nat. Rev. Dis. Primers 2018, 4, 18035. [Google Scholar] [CrossRef]
  76. Seto, W.K.; Lo, Y.R.; Pawlotsky, J.M.; Yuen, M.F. Chronic hepatitis B virus infection. Lancet 2018, 392, 2313–2324. [Google Scholar] [CrossRef]
  77. Zhang, L.; Wang, Z. Circular RNA hsa_circ_0004812 impairs IFN-induced immune response by sponging miR-1287-5p to regulate FSTL1 in chronic hepatitis B. Virol. J. 2020, 17, 40. [Google Scholar] [CrossRef] [Green Version]
  78. Honda, M.; Kaneko, S.; Kawai, H.; Shirota, Y.; Kobayashi, K. Differential gene expression between chronic hepatitis B and C hepatic lesion. Gastroenterology 2001, 120, 955–966. [Google Scholar] [CrossRef]
  79. Zhou, T.C.; Li, X.; Chen, L.J.; Fan, J.H.; Lai, X.; Tang, Y.; Zhang, L.; Wei, J. Differential expression profile of hepatic circular RNAs in chronic hepatitis B. J. Viral Hepat. 2018, 25, 1341–1351. [Google Scholar] [CrossRef]
  80. Sarnow, P.; Sagan, S.M. Unraveling the Mysterious Interactions Between Hepatitis C Virus RNA and Liver-Specific MicroRNA-122. Annu. Rev. Virol. 2016, 3, 309–332. [Google Scholar] [CrossRef] [PubMed]
  81. 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]
  82. Jost, I.; Shalamova, L.A.; Gerresheim, G.K.; Niepmann, M.; Bindereif, A.; Rossbach, O. Functional sequestration of microRNA-122 from Hepatitis C Virus by circular RNA sponges. RNA Biol. 2018, 15, 1032–1039. [Google Scholar] [CrossRef]
  83. Chen, T.C.; Tallo-Parra, M.; Cao, Q.M.; Kadener, S.; Böttcher, R.; Pérez-Vilaró, G.; Boonchuen, P.; Somboonwiwat, K.; Díez, J.; Sarnow, P. Host-derived circular RNAs display proviral activities in Hepatitis C virus-infected cells. PLoS Pathog. 2020, 16, e1008346. [Google Scholar] [CrossRef] [PubMed]
  84. Wong, D.K.; Yuen, M.F.; Yuan, H.; Sum, S.S.; Hui, C.K.; Hall, J.; Lai, C.L. Quantitation of covalently closed circular hepatitis B virus DNA in chronic hepatitis B patients. Hepatology 2004, 40, 727–737. [Google Scholar] [CrossRef] [PubMed]
  85. Sekiba, K.; Otsuka, M.; Ohno, M.; Kishikawa, T.; Yamagami, M.; Suzuki, T.; Ishibashi, R.; Seimiya, T.; Tanaka, E.; Koike, K. DHX9 regulates production of hepatitis B virus-derived circular RNA and viral protein levels. Oncotarget 2018, 9, 20953–20964. [Google Scholar] [CrossRef] [Green Version]
  86. Itoh, T. Stem/progenitor cells in liver regeneration. Hepatology 2016, 64, 663–668. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Zhang, J.; Xu, L.; Wang, P.; Zheng, X.; Hu, Y.; Luo, J.; Zhang, M.; Xu, M. RNA-seq Used to Explore circRNA Expression and Identify Key circRNAs During the DNA Synthesis Phase of Mice Liver Regeneration. DNA Cell Biol. 2020, 39, 2059–2076. [Google Scholar] [CrossRef]
  88. Li, L.; Guo, J.; Chen, Y.; Chang, C.; Xu, C. Comprehensive CircRNA expression profile and selection of key CircRNAs during priming phase of rat liver regeneration. BMC Genom. 2017, 18, 80. [Google Scholar] [CrossRef] [Green Version]
  89. Peralta, C.; Jiménez-Castro, M.B.; Gracia-Sancho, J. Hepatic ischemia and reperfusion injury: Effects on the liver sinusoidal milieu. J. Hepatol. 2013, 59, 1094–1106. [Google Scholar] [CrossRef] [Green Version]
  90. Ye, Z.; Kong, Q.; Han, J.; Deng, J.; Wu, M.; Deng, H. Circular RNAs are differentially expressed in liver ischemia/reperfusion injury model. J. Cell Biochem. 2018, 119, 7397–7405. [Google Scholar] [CrossRef]
  91. Qu, X.; Zheng, C.; Wang, B.; Wang, F.; Sun, X.; Gao, Y.; Xia, Q.; Kong, X. Comprehensive analysis of circular RNAs from steatotic livers after ischemia and reperfusion injury by next-generation RNA sequencing. FEBS Lett. 2020, 595, 99–109. [Google Scholar] [CrossRef]
  92. Tian, X.; Hu, Y.; Liu, Y.; Yang, Z.; Xie, H.; Zhou, L.; Zheng, S. Circular RNA Microarray Analyses in Hepatic Ischemia-Reperfusion Injury With Ischemic Preconditioning Prevention. Front. Med. 2021, 8, 626948. [Google Scholar] [CrossRef] [PubMed]
  93. Chen, Y.; Yuan, B.; Chen, G.; Zhang, L.; Zhuang, Y.; Niu, H.; Zeng, Z. Circular RNA RSF1 promotes inflammatory and fibrotic phenotypes of irradiated hepatic stellate cell by modulating miR-146a-5p. J. Cell Physiol. 2020, 235, 8270–8282. [Google Scholar] [CrossRef]
  94. Niu, H.; Zhang, L.; Chen, Y.H.; Yuan, B.Y.; Wu, Z.F.; Cheng, J.C.; Lin, Q.; Zeng, Z.C. Circular RNA TUBD1 Acts as the miR-146a-5p Sponge to Affect the Viability and Pro-Inflammatory Cytokine Production of LX-2 Cells through the TLR4 Pathway. Radiat Res. 2020, 193, 383–393. [Google Scholar] [CrossRef] [PubMed]
  95. Wang, Z.; Zhao, Y.; Sun, R.; Sun, Y.; Liu, D.; Lin, M.; Chen, Z.; Zhou, J.; Lv, L.; Tian, X.; et al. circ-CBFB upregulates p66Shc to perturb mitochondrial dynamics in APAP-induced liver injury. Cell Death Dis. 2020, 11, 953. [Google Scholar] [CrossRef]
  96. Li, Y.; Gao, X.; Wang, Z.; Liu, W.; Xu, F.; Hu, Y.; Li, Y.; Shi, L. Circular RNA 4099 aggravates hydrogen peroxide-induced injury by down-regulating microRNA-706 in L02 cells. Life Sci. 2020, 241, 116826. [Google Scholar] [CrossRef]
  97. Li, B. The Diagnostic Value and Mechanism of Circular RNA circMARS and circITPR1 in Anti-Tuberculosis Drug- Induced Liver Injury. Master’s Thesis, North China University of Science and Technology, Tangshan, China, 2019. (In Chinese). [Google Scholar]
  98. Zhang, P.; Ming, Y.; Ye, Q.; Niu, Y. Comprehensive circRNA expression profile during ischemic postconditioning attenuating hepatic ischemia/reperfusion injury. Sci. Rep. 2019, 9, 264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Xiong, H.; Wang, H.; Yu, Q. Circular RNA circ_0003420 mediates inflammation in sepsis-induced liver damage by downregulating neuronal PAS domain protein 4. Immunopharmacol. Immunotoxicol. 2021, 43, 271–282. [Google Scholar] [CrossRef]
  100. Parola, M.; Pinzani, M. Liver fibrosis: Pathophysiology, pathogenetic targets and clinical issues. Mol. Aspects Med. 2019, 65, 37–55. [Google Scholar] [CrossRef] [PubMed]
  101. Tsochatzis, E.A.; Bosch, J.; Burroughs, A.K. Liver cirrhosis. Lancet 2014, 383, 1749–1761. [Google Scholar] [CrossRef]
  102. Harrington, L.; McPhail, T.; Mar, V.; Zhou, W.; Oulton, R.; Bass, M.B.; Arruda, I.; Robinson, M.O. A mammalian telomerase-associated protein. Science 1997, 275, 973–977. [Google Scholar] [CrossRef] [PubMed]
  103. Barnard, A.; Moch, A.; Saab, S. Relationship between Telomere Maintenance and Liver Disease. Gut Liver 2019, 13, 11–15. [Google Scholar] [CrossRef] [Green Version]
  104. Li, S.; Song, F.; Lei, X.; Li, J.; Li, F.; Tan, H. hsa_circ_0004018 suppresses the progression of liver fibrosis through regulating the hsa-miR-660-3p/TEP1 axis. Aging 2020, 12, 11517–11529. [Google Scholar] [CrossRef]
  105. Jin, H.; Li, C.; Dong, P.; Huang, J.; Yu, J.; Zheng, J. Circular RNA cMTO1 Promotes PTEN Expression Through Sponging miR-181b-5p in Liver Fibrosis. Front. Cell Dev. Biol. 2020, 8, 714. [Google Scholar] [CrossRef]
  106. Wang, W.; Dong, R.; Guo, Y.; He, J.; Shao, C.; Yi, P.; Yu, F.; Gu, D.; Zheng, J. CircMTO1 inhibits liver fibrosis via regulation of miR-17-5p and Smad7. J. Cell Mol. Med. 2019, 23, 5486–5496. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Ji, D.; Chen, G.F.; Wang, J.C.; Ji, S.H.; Wu, X.W.; Lu, X.J.; Chen, J.L.; Li, J.T. Hsa_circ_0070963 inhibits liver fibrosis via regulation of miR-223-3p and LEMD3. Aging 2020, 12, 1643–1655. [Google Scholar] [CrossRef]
  108. Chen, X.; Li, H.D.; Bu, F.T.; Li, X.F.; Chen, Y.; Zhu, S.; Wang, J.N.; Chen, S.Y.; Sun, Y.Y.; Pan, X.Y.; et al. Circular RNA circFBXW4 suppresses hepatic fibrosis via targeting the miR-18b-3p/FBXW7 axis. Theranostics 2020, 10, 4851–4870. [Google Scholar] [CrossRef] [PubMed]
  109. Zhou, Y.; Lv, X.; Qu, H.; Zhao, K.; Fu, L.; Zhu, L.; Ye, G.; Guo, J. Differential expression of circular RNAs in hepatic tissue in a model of liver fibrosis and functional analysis of their target genes. Hepatol. Res. 2019, 49, 324–334. [Google Scholar] [CrossRef] [PubMed]
  110. Zhou, Y.; Lv, X.; Qu, H.; Zhao, K.; Fu, L.; Zhu, L.; Ye, G.; Guo, J. Preliminary screening and functional analysis of circular RNAs associated with hepatic stellate cell activation. Gene 2018, 677, 317–323. [Google Scholar] [CrossRef]
  111. Bu, F.T.; Zhu, Y.; Chen, X.; Wang, A.; Zhang, Y.F.; You, H.M.; Yang, Y.; Yang, Y.R.; Huang, C.; Li, J. Circular RNA circPSD3 alleviates hepatic fibrogenesis by regulating the miR-92b-3p/Smad7 axis. Mol. Ther. Nucleic Acids 2021, 23, 847–862. [Google Scholar] [CrossRef]
  112. Liu, W.; Feng, R.; Li, X.; Li, D.; Zhai, W. TGF-β- and lipopolysaccharide-induced upregulation of circular RNA PWWP2A promotes hepatic fibrosis via sponging miR-203 and miR-223. Aging 2019, 11, 9569–9580. [Google Scholar] [CrossRef] [PubMed]
  113. Riaz, F.; Li, D. Non-coding RNA Associated Competitive Endogenous RNA Regulatory Network: Novel Therapeutic Approach in Liver Fibrosis. Curr. Gene Ther. 2019, 19, 305–317. [Google Scholar] [CrossRef] [PubMed]
  114. Yao, J.; Dai, Q.; Liu, Z.; Zhou, L.; Xu, J. Circular RNAs in Organ Fibrosis. Adv. Exp. Med. Biol. 2018, 1087, 259–273. [Google Scholar] [CrossRef] [PubMed]
  115. Chen, Y.; Yuan, B.; Wu, Z.; Dong, Y.; Zhang, L.; Zeng, Z. Microarray profiling of circular RNAs and the potential regulatory role of hsa_circ_0071410 in the activated human hepatic stellate cell induced by irradiation. Gene 2017, 629, 35–42. [Google Scholar] [CrossRef]
  116. Marrone, G.; Shah, V.H.; Gracia-Sancho, J. Sinusoidal communication in liver fibrosis and regeneration. J. Hepatol. 2016, 65, 608–617. [Google Scholar] [CrossRef] [Green Version]
  117. Carbone, M.; Neuberger, J.M. Autoimmune liver disease, autoimmunity and liver transplantation. J. Hepatol. 2014, 60, 210–223. [Google Scholar] [CrossRef] [Green Version]
  118. Zhou, Z.; Sun, B.; Huang, S.; Zhao, L. Roles of circular RNAs in immune regulation and autoimmune diseases. Cell Death Dis. 2019, 10, 503. [Google Scholar] [CrossRef] [Green Version]
  119. Zheng, J.; Li, Z.; Wang, T.; Zhao, Y.; Wang, Y. Microarray Expression Profile of Circular RNAs in Plasma from Primary Biliary Cholangitis Patients. Cell Physiol. Biochem. 2017, 44, 1271–1281. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Liu, Y.; Li, Z.; Hao, J.; Chen, H.; Hou, T.; Hao, H. Circular RNAs associated with a mouse model of concanavalin A-induced autoimmune hepatitis: Preliminary screening and comprehensive functional analysis. FEBS Open Bio 2020, 10, 2350–2362. [Google Scholar] [CrossRef]
  121. Virani, S.; Akers, A.; Stephenson, K.; Smith, S.; Kennedy, L.; Alpini, G.; Francis, H. Comprehensive Review of Molecular Mechanisms during Cholestatic Liver Injury and Cholangiocarcinoma. J. Liver 2018, 7, 231. [Google Scholar] [CrossRef]
  122. Zhao, Z.; Meng, J.; Su, R.; Zhang, J.; Chen, J.; Ma, X.; Xia, Q. Epitranscriptomics in liver disease: Basic concepts and therapeutic potential. J. Hepatol. 2020, 73, 664–679. [Google Scholar] [CrossRef] [PubMed]
  123. Yi, Y.C.; Chen, X.Y.; Zhang, J.; Zhu, J.S. Novel insights into the interplay between m(6)A modification and noncoding RNAs in cancer. Mol. Cancer 2020, 19, 121. [Google Scholar] [CrossRef] [PubMed]
  124. Gilbert, W.V.; Bell, T.A.; Schaening, C. Messenger RNA modifications: Form, distribution, and function. Science 2016, 352, 1408–1412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Meyer, K.D.; Jaffrey, S.R. The dynamic epitranscriptome: N6-methyladenosine and gene expression control. Nat. Rev. Mol. Cell Biol. 2014, 15, 313–326. [Google Scholar] [CrossRef] [Green Version]
  126. Chi, F.; Cao, Y.; Chen, Y. Analysis and Validation of circRNA-miRNA Network in Regulating m(6)A RNA Methylation Modulators Reveals CircMAP2K4/miR-139-5p/YTHDF1 Axis Involving the Proliferation of Hepatocellular Carcinoma. Front. Oncol. 2021, 11, 560506. [Google Scholar] [CrossRef]
  127. Pirola, C.J.; Sookoian, S. Epigenetics factors in nonalcoholic fatty liver disease. Expert Rev. Gastroenterol. Hepatol. 2020, 1–16. [Google Scholar] [CrossRef] [PubMed]
  128. Xu, K.; Sun, Y.; Sheng, B.; Zheng, Y.; Wu, X.; Xu, K. Role of identified RNA N6-methyladenosine methylation in liver. Anal. Biochem. 2019, 578, 45–50. [Google Scholar] [CrossRef]
  129. Joshita, S.; Umemura, T.; Tanaka, E.; Ota, M. Genetics and epigenetics in the pathogenesis of primary biliary cholangitis. Clin. J. Gastroenterol. 2018, 11, 11–18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Zhou, C.; Molinie, B.; Daneshvar, K.; Pondick, J.V.; Wang, J.; Van Wittenberghe, N.; Xing, Y.; Giallourakis, C.C.; Mullen, A.C. Genome-Wide Maps of m6A circRNAs Identify Widespread and Cell-Type-Specific Methylation Patterns that Are Distinct from mRNAs. Cell Rep. 2017, 20, 2262–2276. [Google Scholar] [CrossRef] [Green Version]
  131. Zhang, L.; Hou, C.; Chen, C.; Guo, Y.; Yuan, W.; Yin, D.; Liu, J.; Sun, Z. The role of N(6)-methyladenosine (m(6)A) modification in the regulation of circRNAs. Mol. Cancer 2020, 19, 105. [Google Scholar] [CrossRef]
Figure 1. Formation of circRNAs. (a) Three types of circRNAs according to their sources of pre-mRNA. (b) RNA-binding protein (RBP) induces circulation. Circular exons are generated by pre-mRNA through back-splicing between the splice donor site of a downstream exon (the 5′ end of exon 2) and the splice acceptor site of an upstream exon (the 3′ end of exon 3). This event can be mediated by RBPs or specific sequence elements, resulting in the production of a circular RNA and a linear RNA. Blue lines represent introns.
Figure 1. Formation of circRNAs. (a) Three types of circRNAs according to their sources of pre-mRNA. (b) RNA-binding protein (RBP) induces circulation. Circular exons are generated by pre-mRNA through back-splicing between the splice donor site of a downstream exon (the 5′ end of exon 2) and the splice acceptor site of an upstream exon (the 3′ end of exon 3). This event can be mediated by RBPs or specific sequence elements, resulting in the production of a circular RNA and a linear RNA. Blue lines represent introns.
Cells 10 01945 g001
Figure 2. The main functions of circRNAs. (a) miRNA sponge; (b) protein sponge; (c) translating into peptides; (d) regulating the expression of the parental gene via interacting with RNA polymerase II and regulating transcription; (e) miRNA sponges alleviate the inhibitory effects of miRNAs on target mRNA.
Figure 2. The main functions of circRNAs. (a) miRNA sponge; (b) protein sponge; (c) translating into peptides; (d) regulating the expression of the parental gene via interacting with RNA polymerase II and regulating transcription; (e) miRNA sponges alleviate the inhibitory effects of miRNAs on target mRNA.
Cells 10 01945 g002
Figure 3. Examples of circRNAs in chronic liver diseases. ALD: alcoholic liver disease; MAFLD: metabolic-associated fatty liver disease; AIH: autoimmune hepatitis; HSC: hepatic stellate cell; I/R: ischemia/reperfusion; APAP: acetaminophen; IFN: interferon.
Figure 3. Examples of circRNAs in chronic liver diseases. ALD: alcoholic liver disease; MAFLD: metabolic-associated fatty liver disease; AIH: autoimmune hepatitis; HSC: hepatic stellate cell; I/R: ischemia/reperfusion; APAP: acetaminophen; IFN: interferon.
Cells 10 01945 g003
Table 1. Summary of circRNAs involved in chronic liver diseases.
Table 1. Summary of circRNAs involved in chronic liver diseases.
DiseasesCircRNA or LncRNAPossible MechanismsExpression ChangeRefs.
ALDcircRNA_1639circ_1639/miR-122/TNFRSF13C axis and NF-κB pathwayup-regulation[49]
mou_circ_1657sponge miR-96-5pup-regulation[54]
MAFLDcircScd1JAK2/STAT5 pathway down-regulation[66]
circRNA_0049392target miR-7037-5p and miR-6919-5pup-regulation[67]
circRNA_0046366, circRNA_0046367circRNA_0046366,7/miR-34a/PPAR α axisdown-regulation[68,69]
circRNA_021412circRNA_021412/miR-1972/LPIN1 axisdown-regulation[65]
circRNA SCARshut down mPTP by binding to ATP5Bdown-regulation[62]
Viral hepatitishsa_circ_0004812circ_0004812/miR-1287-5p/FSTL1 axisup-regulation[77]
hsa_circ_0000650hsa_circ_0000650/miR-6873-3p and TGF β 2 axisdown-regulation[79]
circPSD3bind factor eIF4A3; nonsense-mediated decay (NMD) pathwayup-regulation[83]
Liver injury and regenerationcirc432, circ2077, circ1366 and circ15controlling the expression level of MAPK14, FN1, TNFRSF21 and GOT1up-regulation[88]
circRNA_017753circRNA_017753/miR-218-5p, miR-7002-3p, miR-7008-3p/Jade1up-regulation[92]
circRSF1promoting inflammatory and fibrotic phenotypes of HSC by sponging miR-146a-5pup-regulation[93]
circTUBD1circTUBD1/miR-146a-5p/Toll-like receptor 4/NF-κB axisup-regulation[94]
circ-CBFBcirc-CBFB/miR-185-5p/p66Shc axisup-regulation[95]
circRNA-4099circRNA-4099/miR-706/keap1/Nrf2 and p38MAPK axisdown-regulation[96]
circMARScircMARS—miR-6808-5p/-6874-3p/-3157-5p—KMT2C—EGFR functional axisup-regulation[97]
mmu_circRNA_005186mmu_circRNA_005186/miR-124-3p/Epha2 axisup-regulation[98]
Liver fibrosis/Cirrhosishsa_circ_0004018 hsa_circ_0004018/hsa-miR-660-3p/TEP1 axis; inhibits liver fibrosisdown-regulation[104]
hsa_circ_0007874 (cMTO1)inhibit HSC activation through sponging miR-181b-5p and miR-17-5pdown-regulation[105,106]
hsa_circ_0070963 inhibits liver fibrosis via regulation of miR-223-3p that related to HSC activationdown-regulation[107]
circFBXW4circFBXW4/miR-18b-3p/FBXW7 axis; inhibit fibrosisdown-regulation[108]
mmu_circ_34116mmu_circ_34116/miR-22-3P/BMP7 signal axis; inhibit HSC activationdown-regulation[109,110]
circPSD3circPSD3/miR-92b-3p/Smad7down-regulation[111]
circ-PWWP2Acirc-PWWP2A/miR-203 and miR-223/Fstl 1 and TLR4 axis; promote HSC activationup-regulation[112,113]
circRNA-0067835circRNA-0067835/miR-155/FoxO3 axis; promote HSC activationup-regulation[56]
hsa_circ_0071410sponge miR-9-5p; promote irradiation-induced HSC activationup-regulation[114,115]
Autoimmune liver diseasehsa_circ_402458 sponge hsa-miR-522-3p and hsa-miR-943up-regulation[119]
mmu_circ_0001520mmu_circ_0001520/mmu-miR-193b-3p/MAPK10up-regulation[120]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zeng, X.; Yuan, X.; Cai, Q.; Tang, C.; Gao, J. Circular RNA as An Epigenetic Regulator in Chronic Liver Diseases. Cells 2021, 10, 1945. https://doi.org/10.3390/cells10081945

AMA Style

Zeng X, Yuan X, Cai Q, Tang C, Gao J. Circular RNA as An Epigenetic Regulator in Chronic Liver Diseases. Cells. 2021; 10(8):1945. https://doi.org/10.3390/cells10081945

Chicago/Turabian Style

Zeng, Xianhui, Xianglei Yuan, Qiuyu Cai, Chengwei Tang, and Jinhang Gao. 2021. "Circular RNA as An Epigenetic Regulator in Chronic Liver Diseases" Cells 10, no. 8: 1945. https://doi.org/10.3390/cells10081945

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