Next Article in Journal
Unraveling Presenilin 2 Functions in a Knockout Zebrafish Line to Shed Light into Alzheimer’s Disease Pathogenesis
Next Article in Special Issue
Genetic Characterization of Rat Hepatic Stellate Cell Line PAV-1
Previous Article in Journal
Rapid Identification of Infectious Pathogens at the Single-Cell Level via Combining Hyperspectral Microscopic Images and Deep Learning
Previous Article in Special Issue
12-O-tetradecanoylphorbol-13-acetate Reduces Activation of Hepatic Stellate Cells by Inhibiting the Hippo Pathway Transcriptional Coactivator YAP
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Regulatory Functions and Mechanisms of Circular RNAs in Hepatic Stellate Cell Activation and Liver Fibrosis

by
Archittapon Nokkeaw
1,2,3,†,
Pannathon Thamjamrassri
1,2,3,†,
Pisit Tangkijvanich
1,2,* and
Chaiyaboot Ariyachet
1,2,*
1
Department of Biochemistry, Faculty of Medicine, Chulalongkorn University, Bangkok 10330, Thailand
2
Center of Excellence in Hepatitis and Liver Cancer, Faculty of Medicine, Chulalongkorn University, Bangkok 10330, Thailand
3
Medical Biochemistry Program, Department of Biochemistry, Faculty of Medicine, Chulalongkorn University, Bangkok 10330, Thailand
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Cells 2023, 12(3), 378; https://doi.org/10.3390/cells12030378
Submission received: 9 December 2022 / Revised: 6 January 2023 / Accepted: 9 January 2023 / Published: 19 January 2023

Abstract

:
Chronic liver injury induces the activation of hepatic stellate cells (HSCs) into myofibroblasts, which produce excessive amounts of extracellular matrix (ECM), resulting in tissue fibrosis. If the injury persists, these fibrous scars could be permanent and disrupt liver architecture and function. Currently, effective anti-fibrotic therapies are lacking; hence, understanding molecular mechanisms that control HSC activation could hold a key to the development of new treatments. Recently, emerging studies have revealed roles of circular RNAs (circRNAs), a class of non-coding RNAs that was initially assumed to be the result of splicing errors, as new regulators in HSC activation. These circRNAs can modulate the activity of microRNAs (miRNAs) and their interacting protein partners involved in regulating fibrogenic signaling cascades. In this review, we will summarize the current knowledge of this class of non-coding RNAs for their molecular function in HSC activation and liver fibrosis progression.

1. Introduction

Chronic liver disease (CLD) represents a gradual deterioration of liver functions, which results from a spectrum of aetiologies including toxins, viral hepatitis, alcohol abuse, and non-alcoholic fatty liver disease (NAFLD) [1]. CLD involves a replacement of healthy liver tissues with fibrotic scars, and hepatic stellate cells (HSCs) play a major role in this process at a cellular level. HSCs are a liver mesenchymal cell type residing in Disse space. In normal livers, HSCs are quiescent and store retinol in cytoplasmic lipid droplets, but upon liver injury, HSCs become activated, lose retinol storage, and trans-differentiate into myofibroblasts that produce extracellular matrix (ECM) mainly type I collagen [2,3]. Excessive ECM accumulation is one of the critical hallmarks of liver fibrosis, disrupting liver architecture and forming fibrous scars [4]. For a short-term injury, fibrosis could be reversed by HSC inactivation and/or apoptosis [5]. However, persistence of injury promotes the continuous production of ECM and eventually leads to irreversible liver fibrosis and cirrhosis [6]. Cirrhosis is the end-stage CLD where the liver structures and functions are severely compromised [7,8].
Several signals play essential roles in HSC activation, such as damage-associated molecular patterns (DAMPs) [9], transforming growth factor beta (TGF-β) [10], platelet-derived growth factor (PDGF) [11], NLRP3 inflammasome [9], and WNT/β-catenin [12]. Recent studies reveal the roles of non-coding RNAs (ncRNAs) in controlling these signals and that modulation of these ncRNAs can either accelerate or inhibit fibrosis progression [13]. For example, microRNAs (miRNAs) including miR-155, miR-150-5p, and miR-130b-5p are reported as liver fibrosis promoters [14,15,16]. On the other hand, miR-130a-3p, miR-29a, and miR-101 are found to be anti-fibrotic factors [17,18,19]. In addition, long non-coding RNAs (lncRNAs) such as HOXA distal transcript antisense RNA (HOTTIP), HOX antisense intergenic RNA (HOTAIR), and hepatocellular carcinoma upregulated lncRNA (HULC) have been identified as regulators of HSC activation [20,21,22,23]. Recently, a novel type of ncRNAs, circular RNAs (circRNAs), has been found to modulate liver fibrosis by regulating the activity of miRNA targets or interacting with RNA-binding proteins (RBPs) [24,25]. In this review, we will focus on the roles of circRNAs and their molecular mechanisms in HSC activation, which causes the development and progression of liver fibrosis.

2. Non-Coding RNA (ncRNA)

Only approximately 1.5% of our genome encodes 21,000 proteins via the central dogma [26,27]. The remaining genomic DNA are non-coding, but a large portion can be transcribed into RNA strands, known as non-coding RNAs (ncRNAs) [26]. Unlike mRNAs, ncRNAs do not translate into proteins but can function as structural, catalytic, and regulatory RNAs [28,29]. Several types of ncRNAs assist gene expression such as small nuclear RNAs (snRNAs), transfer RNAs (tRNAs), and ribosomal RNAs (rRNAs) [27]. Other ncRNAs including microRNAs (miRNAs), long non-coding RNAs (lncRNAs), and circular RNA (circRNAs), play a crucial role in epigenetic mechanisms [25,30,31].

3. Circular RNAs (circRNAs)

CircRNAs, a newly discovered type of ncRNAs, are single-stranded, covalently closed RNAs without 3′ poly(A) tails and 5′ end caps and were first identified in RNA viruses as viroids [32]. CircRNAs are previously believed to be the consequence of splicing errors [33]. However, accumulating evidence show that many circRNAs are highly conserved across species with differential expression patterns in developmental stages and physiological conditions, indicating their potential roles in biological processes [34,35]. Due to their circular structure, which lacks free ends, circRNAs are more stable than linear mRNAs. Most mRNAs exhibit an average half-life of approximately 9 h [36], while the half-life of circRNAs can reach 48 h [37]. This prolonged lifespan of circRNAs may be due to their resistance to degradation by exonucleases [38].
CircRNAs could consist of a single exon or, more often, several exons of protein-coding sequences; this class of circRNAs is known as exonic circRNAs (ecircRNAs) [39]. Some circRNAs might contain only introns instead of exons; these types of circRNAs are called circular intronic RNAs (ciRNAs) [40]. Furthermore, some circRNAs may include both introns and exons which are known as exon-intron circRNAs (EIciRNAs) [41]. Interestingly, circRNAs can also derive from the intron region of a tRNA gene. These circRNAs are called tRNA intronic circular RNAs (tricRNAs) [42].

4. Biogenesis and Transportation of circRNAs

CircRNAs are often formed via back-splicing of precursor mRNAs (pre-mRNAs) in which a downstream 3′ splice-donor is covalently connected to an upstream 5′ splice-acceptor. CircRNAs may also be formed through splicing intermediates called lariat precursors of pre-mRNAs resulting from exon-skipping or intronic lariat precursors that escape the debranching [43,44]. Back-splicing is naturally unfavorable [45]; consequently, most circRNAs are less abundant than their linear mRNA isoforms [46]. Nevertheless, cells possess cis- and trans-acting elements to facilitate back-splicing by bringing the downstream splice-donor site and the upstream splice-acceptor site in proximity [45,47,48]. Particularly, reverse complementary sequences (such as Alu elements) play crucial roles in the formation of circRNAs in a process called intron-pairing-driven circularization [34,37,49]. Moreover, RNA-binding proteins (RBPs) can act as trans-acting elements to promote circularization by bringing the splice sites within proximity of each other or stabilizing intronic RNA pairs [50,51]. These RBPs include muscleblind (MBL) [50], quaking (QKI) [52], heterogeneous nuclear ribonucleoprotein L (HNRNPL) [53], nuclear factor 90/110 (NF90/NF110) [51] and fused in sarcoma (FUS) [54]. Some trans-acting elements, for instance, DEAH (Asp-Glu-Ala-His) box helicase 9 (DHX9) and adenosine deaminase acting on RNA 1 (ADAR1), have been reported to decrease circRNA biogenesis by disrupting base pairing between flanking inverted repeat elements [34,55,56]. Intron-pairing-driven and/or RBP-mediated circularization results in the formation of EIciRNAs or ecircRNAs if introns are further removed [57]. Some circRNAs can be formed via lariat-driven circularization [37,44], enabling the formation of EIciRNAs, which can further undergo internal splicing to remove the intronic sequence and generate ecircRNAs [41,44]. Moreover, lariat-driven circularization can also promote ciRNAs generation [58].
Except for those with introns (EIciRNAs and ciRNAs), most circRNAs are transported to the cytoplasm after biogenesis [41,59,60]. ATP-dependent RNA helicase DDX39A (URH49) and spliceosome RNA helicase DDX39B (UAP56) play key roles in this process to export circRNAs with a small and large size, respectively, from the nucleus to the cytoplasm [61].

5. Degradation of circRNAs

Unlike mechanisms of circRNA biogenesis, details of how circRNAs are degraded in cells remain largely unknown. CircRNAs are resistant to exonuclease catalysis because they lack free 5′ or 3′ ends, so endonuclease is required for the degradation process [62]. RNase L, an endoribonuclease, is identified as a potential enzyme that can globally degrade circRNAs [63]. Moreover, other mechanisms in the degradation of circRNAs have been reported, including (i) Argonaute 2 (AGO2)-mediated cleavage [64], (ii) degradation by the RNase P/mitochondrial RNA processing (MRP) Complex [59], (iii) the up-frameshift suppressor 1 homolog (UPF1)/endo-ribonuclease Ras GTPase-activating protein-binding protein 1 (G3BP1)-mediated RNA decay [65], (iv) RNase H1-mediated degradation [66], (v) TMAO-mediated degradation [67], and (vi) exosome-mediated degradation [68]. These mechanisms are considered to be able to decrease the levels of circRNAs.

6. Role of circRNAs

CircRNAs serve an essential non-coding function in various cellular processes. Other than competing with pre-mRNA splicing and lowering the levels of linear mRNA isoforms, leading to altered gene expression [43,47,50], some speculate that they act as miRNA sponges if they contain multiple miRNA binding sites [69,70,71,72]. Moreover, since circRNAs possess protein binding sites [39], circRNAs can affect cellular function by controlling activity of their protein partners [50,73,74]. Despite the lack of a 5′ cap structure and a 3′ Poly(A) tail, circRNAs can be translated into proteins via the internal ribosome entry site (IRES) and N6-methyladenosines (m6A)-mediated cap-independent translation [75,76].

6.1. Acting as miRNA Sponges

MiRNAs are a class of small and highly conserved non-coding RNA molecules that can regulate gene expression [77]. Most miRNAs are transcribed from DNA into primary miRNAs (pri-miRNAs) by RNA polymerases II and III [78]. Then, a class 2 ribonuclease enzyme process these pri-miRNAs into precursor miRNAs (pre-miRNAs), following by a cascade of cleavage steps to generate mature miRNAs [78,79]. Then, a mature miRNA duplex is incorporated into the AGO protein family to form the miRNA-induced silencing complex (miRISC) that can target mRNAs and cause mRNA degradation and/or translational repression [77,78,79,80,81], resulting in altered gene expression. CircRNAs appear to play an important role in the miRNA-mediated regulation of gene expression, possibly by sponging specific miRNAs and thus increasing protein levels of miRNA target genes (Figure 1a) [82,83,84,85]. Nevertheless, a caution should be taken that several miRNA binding sites within a circRNA are required for efficient sequestration of a particular miRNA [86].

6.2. Interacting with Proteins

CircRNAs may bind with RBPs via RBP-binding sites [50]. RBPs play essential roles in the control of gene expression by regulating splicing, mRNA maturation, transport, localization, translation, and decay [87,88,89]. Various circRNAs have been discovered to interact with RBPs by acting as protein sponges or decoys, thus activating, or disrupting their biological functions (Figure 1b) [90,91,92]. Some circRNAs serve as a protein scaffold to promote the colocalization of two or more proteins, such as enzymes and their substrates, enhancing their activity (Figure 1c) [74,93]. Some circRNAs can recruit certain proteins to specific cellular locations and spatially modulate their function (Figure 1e) [94,95,96]. EIciRNAs and ciRNAs, which are located in the nucleus, are involved in transcriptional regulation by promoting RNA Pol II activity, which enhances target gene transcription (Figure 1f) [40,41]. Moreover, it has been revealed that some nuclear retained ecircRNAs influence splicing by exon-skipping variants, consequently affecting the balance between circRNAs and their linear isoforms (Figure 1g) [97].

6.3. Translated into Proteins

Linear mRNA translation is usually a cap-dependent process that requires a 7-methylguanosine (m7G) cap at the 5′ end and a poly(A) tail at the 3′ end [98]. Recent findings suggest that circRNA translation can occur in a cap-independent manner through IRES or m6A RNA modification in the 5′ untranslated region (UTR) (Figure 1d) [99,100]. CircMbl-encoded proteins are expressed in small amounts under normal conditions, but following starvation, their expression notably increases [101]. These results suggest that circRNA-derived peptides may play a role in maintaining homeostasis.

7. CircRNAs and Hepatic Stellate Cell Activation

HSC activation is associated with several cytokines and regulatory networks and promotes expression of HSC activation markers, including α-smooth muscle actin (α-SMA) and type I collagen [102]. Many signaling pathways have been linked to increasing expression levels of α-SMA and/or type I collagen, including TGF-β [103], JAK/STAT [104], PDGF [105], PI3K/Akt [106], Wnt/β-catenin [107], Notch [108], Hedgehog [109,110], Hippo [111], and inflammasome signaling pathways [112]. Thus, circRNAs targeting these signaling pathways can regulate HSC activation and ultimately affect liver fibrosis (Table 1, Figure 2 and Figure 3).

7.1. Anti-Fibrotic circRNAs

TGF-β signaling is regarded as the primary fibrogenic pathway that stimulates HSC activation and ECM synthesis [103]. TGF-β is found in trace amounts in healthy livers [133]. Following liver damage, macrophages start producing TGF-β and PDGF, which can activate excessive ECM production from HSCs and lead to the development of liver fibrosis [134]. Most reported circRNAs involved in HSC activation target miRNAs and proteins in the TGF-β pathway. One of these circRNAs is circPSD3 (mmu_circ_0001682), which is downregulated in primary HSCs and liver tissues from mice with carbon tetrachloride (CCl4)-induced liver fibrosis [113]. CircPSD3 can function as a miR-92b-3p sponge, consequently promoting Smad7 expression [113]. Smad7 can block the activation of receptor-regulated Smads (R-Smads) and thus inhibits the TGF-β signaling pathway [135,136]. Furthermore, circPSD3 can preclude HSC proliferation and forestall fibrosis progression in vivo [113].
CircCREBBP acts as a miR-1291 sponge, consequently increasing the expression of left-right determination factor 2 (LEFTY2) [114], which can inhibit the phosphorylation of Smad2/3 [137,138], a key signaling molecule in the TGF-β pathway [139]. Moreover, circCREBBP can also inhibit cell proliferation and arrest the cell cycle in HSCs [114]. With a similar mode of action, hsa_circ_0070963, also known as circSCLT1, has the ability to sponge miR-223-3p, which can target LEM domain containing 3 (LEMD3) [115], an inhibitory molecule that can antagonize Smad2/3 signaling and perturb the TGF-β signaling pathway [140]. In addition to suppressing HSC activation, the overexpression of circSCLT1 can induce cell cycle arrest and suppress cell proliferation in HSCs [115]. However, the circSCLT1/miR-223 axis is subject to further investigation as most studies suggest the anti-fibrotic roles of miR-223 in various fibrosis models [141,142,143,144]. Another circRNA is mmu_circ_34116, which is shown to suppress HSC activation upon overexpression [145]. Using bioinformatic techniques, mmu_circ_34116 is predicted to target the miR-22-3p/BMP7 axis [145]. Bone morphogenetic protein 7 (BMP7) can activate Smad1/5/8 but inhibits Smad3 activation, thus antagonizing TGF-β signaling [146,147].
CircMTO1 has been extensively studied for their anti-fibrotic roles. Its expression levels are observed to downregulate in liver tissues of fibrosis patients [116]. Patients with higher liver fibrosis stages have significantly less circMTO1 expression [116]. By interacting with miR-17-5p, circMTO1 can positively regulate the expression of Smad7 and thereby negatively regulate the TGF-β signaling pathway [116]. Moreover, circMTO1 is found to act as a miR-181b-5p sponge [117]. miR-181b-5p can target phosphatase and tensin homolog (PTEN), a negative regulator in the phosphatidylinositol 3-kinase/protein kinase B (PI3K/Akt) signaling pathway [117], which promotes HSC activation [106,148]. Therefore, circMTO1 plays an anti-fibrotic role by downregulating miR-181b-5p and enhancing PTEN activity [117]. Additionally, Akt signaling can also lead to the activation of NF-κB [149], protecting activated HSCs against apoptosis and sustaining cell survival [150]. As a consequence, by suppressing PI3K/Akt signaling via PTEN, circMTO1 overexpression potentially enhances HSC apoptosis and diminishes ECM production.
Some circRNAs play a significant role in other liver cell types but can indirectly have an impact on HSC activation. For example, in normal hepatocytes, circBNC2 expression is high but drastically decreased upon liver injury [118]. Interestingly, expression of activation markers (α-SMA and type I collagen) is significantly increased in HSCs incubated with conditional medium from circBNC2 knockout hepatocytes, which appear to contain high levels of the pro-fibrotic cytokines including TGF-β [118]. In contrast, the overexpression of circBNC2 in hepatocytes can reduce expression levels of these cytokines upon liver injury, and conditional medium from these hepatocytes can suppress expression of α-SMA and type I collagen in cultured HSCs, indicating the anti-fibrotic roles of circBNC2 [118]. Mechanistically, circBNC2 contains an open reading frame (ORF) and an IRES, suggesting their function as a protein template [118]. The circBNC2-translated protein (ctBNC2) is a protein product derived from circBNC2 translation [118]. This 681-amino acid protein can bind to CDK1 and cyclin B1 and promote CDK/cyclin complex translocation into the nucleus, a critical step to initiate mitosis and prevent apoptosis [118]. As ctBNC2 levels decrease in a damaged liver, the translocation of the CDK/cyclin complex could be impaired and induce apoptosis, which triggers the secretion of TGF-β and DAMPs from hepatocytes [151,152]. As previously mentioned, these molecules can activate ECM production from HSCs and promote fibrosis progression. Therefore, by protecting hepatocytes from apoptosis upon injury, circBNC2 could play an anti-fibrotic role by suppressing production of inflammatory cytokines [118].
Another key signaling pathway that involves HSC activation is the network of PTEN/PI3K/Akt [106]. PI3K/Akt signaling can activate the serine/threonine kinase p70 ribosomal protein S6 kinase (p70S6K) via the mammalian target of rapamycin complex 1 (mTORC1), whereas PTEN inhibits activation of PI3K/Akt signaling [106,153]. Evidently, p70S6K facilitates HSC proliferation and collagen expression [153], and thus, the PI3K/Akt signaling pathway is considered pro-fibrotic [106]. A recent report shows that circDIDO1 inhibits HSC activation through the miR-141-3p/PTEN axis [124]. The overexpression of circDIDO1 sponges miR-141-3p and increases PTEN expression, which impairs PI3K/Akt signaling and subsequently suppresses HSC activation [124]. Similarly, circCDK13 forestalls liver fibrosis by repressing the activation and proliferation of HSCs via the miR-17-5p/KAT2B/MFGE8/PTEN axis [125]. By sponging miR-17-5p, circCDK13 can promote the expression of lysine acetyltransferase 2B (KAT2B), a protein that can regulate histone protein acetylation [125]. The acetylation of histones can enhance the accessibility of transcriptional factors to DNA templates, thereby increasing transcriptional activity [154]. Specifically, KAT2B facilitates milk fat globulin-epidermal growth factor 8 (MFGE8) transcription by promoting histone H3 acetylation [125]. Because MFGE8 can inhibit PTEN ubiquitination and degradation [155], KAT2B promotes PTEN stability, which in turn suppresses the PI3K/Akt signaling pathway [125]. Inactivation of Akt can also lead to the reduction of NF-κB activity, thus impairing the survival of fibrogenic HSCs [125,156]. Collectively, circDIDO1 and circCDK13 can exert the anti-fibrotic mechanism via suppression of PI3K/Akt signaling.
In addition to TGF-β and PI3K/Akt signaling pathways, Wnt/β-catenin and hippo signaling could be regulated by circRNAs in HSCs. For example, circ608 increases PTEN-induced kinase 1 (PINK1) expression through sponging miR-222 [127]. PINK1 can then phosphorylate Parkin, an E3 ubiquitin-protein ligase, which in turn triggers ubiquitination and degradation of β-catenin, thereby inhibiting Wnt/β-catenin transduction [157,158]. PINK1 can also promote expression of Mps one binder kinase activator 1B (MOB1B) [159], a regulator of large tumor suppressor kinases 1/2 (LATS1/2) [160]. The MOB1/LATS complex can inactivate YAP and TAZ [160], the effectors in the Hippo signaling pathway [161], therefore inhibiting the Hippo signaling cascade and subsequently HSC activation. PINK1/Parkin can promote mitophagy [159], which is suppressed in fibrogenic HSCs [162]. Taken together, circ608 is an anti-fibrotic mediator via PINK1-mediated suppression of Wnt/β-catenin and hippo signaling.
For the Notch signaling pathway, circFBXW4 positively regulates F-box/WD repeat-containing protein 7 (FBXW7) [128], which induces Notch intracellular domain (NICD) degradation [163]. By interacting with miR-18b-3p, circFBXW4 can upregulate FBXW7 expression and inhibit Notch signaling, thereby inhibiting HSC activation and proliferation as well as promoting apoptosis [128].
A mitochondrial circRNA called circSCAR is found to be downregulated in NASH patients [130]. Lipid accumulation can induce endoplasmic reticulum (ER) stress, thus increasing the expression of the ER stress mediator CCAAT-enhancer-binding protein homologous protein (CHOP) [130]. CHOP can inhibit the expression of PGC-1α, which positively regulates circSCAR transcription and thus decreases circSCAR levels [130]. CircSCAR can interact with ATP5B, a regulator of the mitochondrial permeability transition pore (mPTP) [130]. Binding between circSCAR and ATP5B hinders the opening of mPTP and consequently the efflux of reactive oxygen species (ROS) from mitochondria into the cytosol [130]. Previous studies suggest that ROS promotes NF-κB signaling pathway and HSC activation [164,165]. Therefore, circSCAR may play an anti-fibrotic role by preventing leakage of ROS from mitochondria [130].
Some circRNAs are found to suppress HSC activation, but their underlying mechanisms are not fully understood. One example is hsa_circ_0004018 or circSMYD4, which is proposed to suppress HSC activation and proliferation via miR-660-3p [132]. By using bioinformatics tools, telomerase-associated protein 1 (TEP1) is predicted to be a target of miR-660-3p and experimentally confirmed by a luciferase assay [132]. While functions of TEP1 in HSCs have not been reported, higher levels of TEP1 in hepatocytes can indirectly suppress the proliferation and activation of HSCs, potentially because TEP1 prevents formation of critically short telomeres and reduces DNA-damage response, which normally triggers hepatocyte senescence and sends apoptotic signals to activate HSCs [132,166]. Intriguingly, although TEP1 play a role in telomere length regulation in many cell types [167,168], a study shows that TEP1 is not essential for telomere length regulation in murine liver [169]. This finding suggests the roles of TEP1 in other cellular functions that could modulate HSC activation [169]. In addition to being a component of the telomerase complex, TEP1 has been found in vault ribonucleoprotein complexes (VRCs) [170]. Although the function of VRCs in HSC activation is not yet known, major vault protein knock-out mice can intensify hepatic steatosis and induce fibrotic responses [171]. Further investigations are needed to expand the knowledge of how the circSMYD4/miR-660-3p/TEP1 axis inhibits HSC activation. Other examples are hsa_circ_0089761 and hsa_circ_0089763, which are two mitochondrial-encoded circRNAs downregulated in HSCs during LPS stimulation [172]. Their downregulation indicates their potential roles as anti-fibrotic circRNAs, but their mechanisms and targets remain to be further studied. Lastly, in a hepatitis B virus (HBV)-induced activation model of LX-2 cells, circMTM1 is upregulated along with interleukin 7 receptor (IL7R), potentially via absorbing miR-122-5p [173]. This study shows that circMTM1 knockdown expression and miR-122-5p overexpression reduce the expression levels of activated HSC markers, while the upregulation of IL7R attenuates the anti-fibrotic function of miR-122-5p [173]. Normally, IL7R is expressed in T cells and plays a pivotal role in T cell homeostasis [174]. Despite interesting results, the exact roles of circMTM1 and IL7R in HSC functions need a further investigation.

7.2. Pro-Fibrotic circRNAs

While most reported circRNAs impedes the activation of TGF-β signaling pathway in HSCs, some circRNAs play the opposite roles. These circRNAs include circPWWP2A [119], circUBE2k [120], and circTUBD1 [121,122]. CircPWWP2A can enhance TGF-β signaling by decreasing miR-203 and miR-223 expression and thus increasing expression levels of follistatin-like 1 (Fstl1) and Toll-like receptor 4 (TLR4), respectively [119]. Fstl1 is a glycoprotein ligand and can promote TGF-β signaling by binding to TGF-β1 type II receptors (TβRII) [175,176]. TLR4 can recognize lipopolysaccharide (LPS) and sequentially trigger signaling cascades that produce the NF-κBp50 homodimer to interact with histone deacetylase 1 (HDAC1). This protein complex can transcriptionally suppress BMP and activin membrane-bound inhibitor (BAMBI), a pseudoreceptor of TGF-β [119,177,178]. Since BAMBI can inhibit TGF-β/Smad signaling [177], TLR4-mediated downregulation of BAMBI can induce fibrotic HSCs via amplification of TGF-β signaling [119]. In addition, TLR4 can activate NF-κB to support the survival of the activated HSCs [179,180]. Together, circPWWP2A can promote HSC activation via sponging miRNAs that suppress the expression of Fstl1 and TLR4.
The overexpression of circUBE2K induces liver fibrosis by sponging miR-149-5p, which negatively regulates expression of TGF-β2 [120]. Conversely, inhibiting circUBE2K expression can impair TGF-β signaling and induce cell cycle arrest in HSCs, confirming the pro-fibrotic roles of circUBE2K in liver fibrosis progression [120]. Finally, circTUBD1 can interact with both miR-203a-3p and miR-146a-5p [121,122]. By sponging miR-203a-3p, circTUBD1 can positively regulate TGF-β signaling by increasing Smad3 expression [121]. Moreover, Smad3 is found to positively regulate circTUBD1 via a feedback loop, thus further enhancing Smad3 expression [121]. TLR4 is targeted by miR-146a-5p, and by downregulating miR-146a-5p via the circTUBD1 sponge, expression of TLR4 increases, thus promoting TGF-β and NF-κB signaling pathways, which can help activate and support cell survival of HSCs, respectively [122,177,179,180,181]. Furthermore, circTUBD1 can also promote HSC proliferation and inhibit the apoptosis of HSCs by increasing the expression of the anti-apoptotic protein B-cell lymphoma-2 (BCL-2), thereby amplifying the population of fibrogenic cells [122].
Interleukin-6 (IL-6) can activate HSC through the MAPK and JAK/STAT signaling pathways [104]. CircCHD2 can enhance hepatic leukemia factor (HLF) expression by interacting with miR-200b-3p [123]. HLF can promote the expression of IL-6 that can bind to the IL-6 receptor and activate the JAK/STAT3 pathway [182]. The JAK/STAT pathway is also a part of a non-SMAD pathway of TGF-β signal transduction that is essential for HSC activation [183]. The JAK/STAT3 pathway also promotes HLF expression, thereby regulating signal transduction in a feed-forward circuit manner [182].
Additionally, hsa_circ_0071410 (circPALLD) and hsa-circ-0067835 (circIFT80) are pro-fibrotic circRNAs that promote the PI3K/Akt signaling pathway [126,184,185]. CircPALLD can induce HSC activation and increase cell viability by interacting with miR-9-5p [184], which is reported to target annexin A2 (ANXA2) [185]. ANXA2 has been considered to play a role in HSC activation, proliferation, and apoptosis [185], possibly by activating the FAK/PI3K/Akt signaling pathway [186]. CircIFT80 can activate HSC by regulating the miR-155/Forkhead box O3 (FOXO3)/Akt axis [126,187]. As CircIFT80 alleviates FOXO3 expression by sponging miR-155, this transcription factor can transactivate PI3K/Akt through a positive feedback loop and promote PI3K/Akt signal transduction in fibrogenic HSCs [188,189].
The Hedgehog signaling pathway plays a significant role in HSC activation [109,110]. A recent study shows that circRSF1 sponges miR-146a-5p. Subsequently, this sequestration promotes Ras-related C3 botulinum toxin substrate 1 (RAC1) expression and Hedgehog signal transduction, resulting in HSC activation and proliferation [129]. Rac1 has been shown to be involved in the Hedgehog signaling pathway [190]. Activated Rac1 induces glioma-associated oncogene (Gli) nuclear translocation, which is necessary for the Hedgehog signaling pathway [191]. Rac1 can also influence NF-κB and JNK signaling by promoting their transduction [192,193]. JNK signal transduction can phosphorylate Smad2 at the C-terminal and linker regions [194]. Initial findings reveal that phosphorylation of Smad2 in the linker region hinders its nuclear translocation and cellular signaling, but recent evidence shows that phosphorylation of the Smad linker can stimulate expression of fibrotic genes [194,195]. One study discovers that the phosphorylated Smad2 linker region is associated with increased expression of glycosaminoglycans (GAG) [196], and hyaluronan (HA), a class of GAG, is found to be able to activate HSCs via Notch1 [197], so the phosphorylation of the Smad2 linker region may be associated with HSC activation by increasing the expression of HA and promoting the Notch1 fibrogenic signaling pathway. However, this proposed mechanism of circRSF1/miR-146a-5p/Rac1 with subsequent signal transductions remains to be further explored.
Moreover, the direct regulation of circRNAs on fibrotic genes have been reported. For instance, the circARID1A/miR-185-3p axis post-transcriptionally regulates expression of COL1A1, a key marker of HSC activation [131]. The pro-fibrotic circARID1A can also promote the proliferation and migration of HSCs as well as inhibit their apoptosis [131]. Mechanistically, increased type I collagen expression can lead to a positive feedback loop of HSC activation in which accumulation of collagen further activates HSCs by increasing ECM stiffness [198]. This mechanical tension in turn leads to YAP activation in the Hippo signaling pathway and Akt activation in the PI3K/Akt signaling pathway [198,199].
The above discussion mainly focuses on the intrinsic expression of circRNAs in HSCs as most in vitro studies rely on the single cell type for their analysis. However, apart from HSCs, liver is a complex organ comprising of several cell types including hepatocytes, Kupffer cells, liver sinusoidal endothelial cells, and cholangiocytes, which are known to communicate with one another to maintain liver functions [200]. Some of these cells were reported to modulate liver fibrosis by transferring circRNAs to HSCs such as hepatocyte-derived circBNC2 [118]. Alternatively, circRNAs that regulate the production of inflammatory cytokines in Kupffer cells can have an indirect impact on the state of HSC activation. These circRNAs include circMcph1 [201,202] and circ1639 [203,204,205,206]. Nevertheless, a study of cell-cell interaction in mediating liver fibrosis through circRNAs is still lacking and need to be further investigated.

8. Conclusions and Future Perspectives

HSC activation is the major contribution to liver fibrosis, which can trigger the development of cirrhosis and ultimately hepatocellular carcinoma (HCC). Molecular mechanisms underlying the activation process are currently being explored in the hope of identifying new therapeutic targets for liver fibrosis. CircRNAs have been reported to regulate HSC activation through modulating signal transduction in fibrogenic signaling pathways. Depending on which miRNAs or proteins they target, circRNAs can be either anti- or pro-fibrotic. Although most discovered circRNAs so far are mechanistically explained as miRNA sponges for their functions, some circRNAs linked to HSC activation can act as protein templates or protein sponges, but only a few studies have reported this mode of action. Given that most circRNAs do not carry multiple binding sites for a single miRNA, the sequestration of miRNAs may not be the primary function of circRNAs [207]. Future studies should explore new regulatory mechanisms of circRNAs other than miRNA sponges in HSC activation. Recent technology based on proximity labeling enzymes in couple with CRISPR/Cas allows the isolation of proteins interacting with specific RNA transcripts [208]. The identification of these proteins could yield novel functions of circRNAs in fibrogenic HSCs.
The number of newly discovered circRNAs has increased over the past years thanks to advances in nucleotide sequencing technologies. Oxford nanopore sequencing combined with circRNA identifiers using long-read sequencing data (CIRI-long) could detect additional splicing events and alternative circularization of circRNAs at a higher efficiency compared with that of Illumina RNA-sequencing methods [209]. This advanced approach might allow the identification of new circRNAs that regulate liver fibrosis. To study roles of the newly discovered circRNAs, loss-of-function and gain-of-function experiments are essential. However, current methods for such genetic manipulation have significant drawback when studying circRNAs in terms of efficiency and specificity. A recent study describes a new technique known as twister-optimized RNA for durable the overexpression (Tornado) expression system, which enables a higher rate of RNA circularization than that of traditional methods [210]. Conversely, utilizing base editors to alter single nucleotide sequences in back-splice sites could impair circRNA biogenesis and reduce its expression while avoiding off-target effects on the cognate linear mRNA [211]. These new techniques could be applied for functional analyses of circRNAs in HSC activation. Lastly, recent studies illustrate cell-to-cell communications via circRNA-containing exosomes from other liver cell types to HSCs, adding another layer of complexity in the modulation of HSC fates [124,125]. More mechanistic understanding of circRNAs in HSC activation could provide new knowledge in biology of non-coding RNA and potentially shed light on development of a new therapeutic strategy for liver fibrosis.

Author Contributions

Conceptualization, C.A. and P.T. (Pisit Tangkijvanich); writing—original draft preparation, A.N. and P.T. (Pannathon Thamjamrassri); writing—review and editing, C.A. and P.T. (Pisit Tangkijvanich); visualization, A.N. and P.T. (Pannathon Thamjamrassri); supervision, C.A. and P.T. (Pisit Tangkijvanich); project administration, C.A. and P.T. (Pisit Tangkijvanich); funding acquisition, C.A. and P.T. (Pisit Tangkijvanich). All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by Thailand Science Research and Innovation Fund Chulalongkorn University (CU_FRB65_hea (46)_053_30_34), Ratchadapiseksompotch Fund, Faculty of Medicine, Chulalongkorn University (Grant No. RA 66/017), Thailand Research Fund (TRF) Senior Research Scholar (Grant No. RTA6280004), and Center of Excellence in Hepatitis and Liver Cancer, Faculty of Medicine, Chulalongkorn University.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We would like to thank all members in Center of Excellence in Hepatitis and Liver Cancer, Faculty of Medicine, Chulalongkorn University for their assistance in writing this review.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bataller, R.; Brenner, D.A. Liver fibrosis. J. Clin. Investig. 2005, 115, 209–218. [Google Scholar] [CrossRef]
  2. Friedman, S.L. Hepatic stellate cells: Protean, multifunctional, and enigmatic cells of the liver. Physiol. Rev. 2008, 88, 125–172. [Google Scholar] [CrossRef]
  3. Hernandez-Gea, V.; Friedman, S.L. Pathogenesis of liver fibrosis. Annu. Rev. Pathol. 2011, 6, 425–456. [Google Scholar] [CrossRef]
  4. Friedman, S.L. Liver fibrosis—From bench to bedside. J. Hepatol. 2003, 38 Suppl. 1, S38–S53. [Google Scholar] [CrossRef]
  5. Elpek, G. Cellular and molecular mechanisms in the pathogenesis of liver fibrosis: An update. World J. Gastroenterol. 2014, 20, 7260–7276. [Google Scholar] [CrossRef]
  6. Roehlen, N.; Crouchet, E.; Baumert, T.F. Liver Fibrosis: Mechanistic Concepts and Therapeutic Perspectives. Cells 2020, 9, 875. [Google Scholar] [CrossRef] [Green Version]
  7. Pinzani, M.; Rosselli, M.; Zuckermann, M. Liver cirrhosis. Best Pract. Res. Clin. Gastroenterol. 2011, 25, 281–290. [Google Scholar] [CrossRef]
  8. Anthony, P.P.; Ishak, K.G.; Nayak, N.C.; Poulsen, H.E.; Scheuer, P.J.; Sobin, L.H. The morphology of cirrhosis. Recommendations on definition, nomenclature, and classification by a working group sponsored by the World Health Organization. J. Clin. Pathol. 1978, 31, 395–414. [Google Scholar] [CrossRef] [Green Version]
  9. Krenkel, O.; Tacke, F. Liver macrophages in tissue homeostasis and disease. Nat. Rev. Immunol. 2017, 17, 306–321. [Google Scholar] [CrossRef]
  10. Xu, F.; Liu, C.; Zhou, D.; Zhang, L. TGF-β/SMAD Pathway and Its Regulation in Hepatic Fibrosis. J. Histochem. Cytochem. 2016, 64, 157–167. [Google Scholar] [CrossRef]
  11. Ying, H.Z.; Chen, Q.; Zhang, W.Y.; Zhang, H.H.; Ma, Y.; Zhang, S.Z.; Fang, J.; Yu, C.H. PDGF signaling pathway in hepatic fibrosis pathogenesis and therapeutics (Review). Mol. Med. Rep. 2017, 16, 7879–7889. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Nishikawa, K.; Osawa, Y.; Kimura, K. Wnt/β-Catenin Signaling as a Potential Target for the Treatment of Liver Cirrhosis Using Antifibrotic Drugs. Int. J. Mol. Sci. 2018, 19, 3103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Zhou, C.; York, S.R.; Chen, J.Y.; Pondick, J.V.; Motola, D.L.; Chung, R.T.; Mullen, A.C. Long noncoding RNAs expressed in human hepatic stellate cells form networks with extracellular matrix proteins. Genome Med. 2016, 8, 31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Bala, S.; Csak, T.; Saha, B.; Zatsiorsky, J.; Kodys, K.; Catalano, D.; Satishchandran, A.; Szabo, G. The pro-inflammatory effects of miR-155 promote liver fibrosis and alcohol-induced steatohepatitis. J. Hepatol. 2016, 64, 1378–1387. [Google Scholar] [CrossRef] [Green Version]
  15. Chen, W.; Yan, X.; Yang, A.; Xu, A.; Huang, T.; You, H. miRNA-150-5p promotes hepatic stellate cell proliferation and sensitizes hepatocyte apoptosis during liver fibrosis. Epigenomics 2020, 12, 53–67. [Google Scholar] [CrossRef]
  16. Wang, H.; Wang, Z.; Wang, Y.; Li, X.; Yang, W.; Wei, S.; Shi, C.; Qiu, J.; Ni, M.; Rao, J.; et al. miRNA-130b-5p promotes hepatic stellate cell activation and the development of liver fibrosis by suppressing SIRT4 expression. J. Cell Mol. Med. 2021, 25, 7381–7394. [Google Scholar] [CrossRef]
  17. Liu, L.; Wang, P.; Wang, Y.S.; Zhang, Y.N.; Li, C.; Yang, Z.Y.; Liu, Z.H.; Zhan, T.Z.; Xu, J.; Xia, C.M. MiR-130a-3p Alleviates Liver Fibrosis by Suppressing HSCs Activation and Skewing Macrophage to Ly6C(lo) Phenotype. Front. Immunol. 2021, 12, 696069. [Google Scholar] [CrossRef]
  18. Lin, H.Y.; Wang, F.S.; Yang, Y.L.; Huang, Y.H. MicroRNA-29a Suppresses CD36 to Ameliorate High Fat Diet-Induced Steatohepatitis and Liver Fibrosis in Mice. Cells 2019, 8, 1298. [Google Scholar] [CrossRef] [Green Version]
  19. Lei, Y.; Wang, Q.L.; Shen, L.; Tao, Y.Y.; Liu, C.H. MicroRNA-101 suppresses liver fibrosis by downregulating PI3K/Akt/mTOR signaling pathway. Clin. Res. Hepatol. Gastroenterol. 2019, 43, 575–584. [Google Scholar] [CrossRef]
  20. Zheng, J.; Mao, Y.; Dong, P.; Huang, Z.; Yu, F. Long noncoding RNA HOTTIP mediates SRF expression through sponging miR-150 in hepatic stellate cells. J. Cell Mol. Med. 2019, 23, 1572–1580. [Google Scholar] [CrossRef]
  21. Li, Z.; Wang, J.; Zeng, Q.; Hu, C.; Zhang, J.; Wang, H.; Yan, J.; Li, H.; Yu, Z. Long Noncoding RNA HOTTIP Promotes Mouse Hepatic Stellate Cell Activation via Downregulating miR-148a. Cell Physiol. Biochem. 2018, 51, 2814–2828. [Google Scholar] [CrossRef] [PubMed]
  22. Yu, F.; Chen, B.; Dong, P.; Zheng, J. HOTAIR Epigenetically Modulates PTEN Expression via MicroRNA-29b: A Novel Mechanism in Regulation of Liver Fibrosis. Mol. Ther. 2017, 25, 205–217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Shen, X.; Guo, H.; Xu, J.; Wang, J. Inhibition of lncRNA HULC improves hepatic fibrosis and hepatocyte apoptosis by inhibiting the MAPK signaling pathway in rats with nonalcoholic fatty liver disease. J. Cell Physiol. 2019, 234, 18169–18179. [Google Scholar] [CrossRef]
  24. Kristensen, L.S.; Andersen, M.S.; Stagsted, L.V.W.; Ebbesen, K.K.; Hansen, T.B.; Kjems, J. The biogenesis, biology and characterization of circular RNAs. Nat. Rev. Genet. 2019, 20, 675–691. [Google Scholar] [CrossRef] [PubMed]
  25. Zeng, X.; Yuan, X.; Cai, Q.; Tang, C.; Gao, J. Circular RNA as An Epigenetic Regulator in Chronic Liver Diseases. Cells 2021, 10, 1945. [Google Scholar] [CrossRef]
  26. Richard Boland, C. Non-coding RNA: It's Not Junk. Dig. Dis. Sci. 2017, 62, 1107–1109. [Google Scholar] [CrossRef]
  27. Hombach, S.; Kretz, M. Non-coding RNAs: Classification, Biology and Functioning. Adv. Exp. Med. Biol. 2016, 937, 3–17. [Google Scholar] [CrossRef]
  28. Eddy, S.R. Noncoding RNA genes. Curr. Opin. Genet. Dev. 1999, 9, 695–699. [Google Scholar] [CrossRef]
  29. Erdmann, V.A.; Barciszewska, M.Z.; Hochberg, A.; De Groot, N.; Barciszewski, J. Regulatory RNAs. Cell. Mol. Life Sci. 2001, 58, 960–977. [Google Scholar] [CrossRef]
  30. Rotondo, J.C.; Mazziotta, C.; Lanzillotti, C.; Tognon, M.; Martini, F. Epigenetic Dysregulations in Merkel Cell Polyomavirus-Driven Merkel Cell Carcinoma. Int. J. Mol. Sci. 2021, 22, 11464. [Google Scholar] [CrossRef]
  31. Wang, C.; Wang, L.; Ding, Y.; Lu, X.; Zhang, G.; Yang, J.; Zheng, H.; Wang, H.; Jiang, Y.; Xu, L. LncRNA Structural Characteristics in Epigenetic Regulation. Int. J. Mol. Sci. 2017, 18, 2659. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Huang, A.; Zheng, H.; Wu, Z.; Chen, M.; Huang, Y. Circular RNA-protein interactions: Functions, mechanisms, and identification. Theranostics 2020, 10, 3503–3517. [Google Scholar] [CrossRef] [PubMed]
  33. Nigro, J.M.; Cho, K.R.; Fearon, E.R.; Kern, S.E.; Ruppert, J.M.; Oliner, J.D.; Kinzler, K.W.; Vogelstein, B. Scrambled exons. Cell 1991, 64, 607–613. [Google Scholar] [CrossRef] [PubMed]
  34. Ivanov, A.; Memczak, S.; Wyler, E.; Torti, F.; Porath, H.T.; Orejuela, M.R.; Piechotta, M.; Levanon, E.Y.; Landthaler, M.; Dieterich, C.; et al. Analysis of Intron Sequences Reveals Hallmarks of Circular RNA Biogenesis in Animals. Cell Rep. 2015, 10, 170–177. [Google Scholar] [CrossRef] [Green Version]
  35. Rybak-Wolf, A.; Stottmeister, C.; Glažar, P.; Jens, M.; Pino, N.; Giusti, S.; Hanan, M.; Behm, M.; Bartok, O.; Ashwal-Fluss, R.; et al. Circular RNAs in the Mammalian Brain Are Highly Abundant, Conserved, and Dynamically Expressed. Mol. Cell 2015, 58, 870–885. [Google Scholar] [CrossRef] [Green Version]
  36. Schwanhäusser, B.; Busse, D.; Li, N.; Dittmar, G.; Schuchhardt, J.; Wolf, J.; Chen, W.; Selbach, M. Global quantification of mammalian gene expression control. Nature 2011, 473, 337–342. [Google Scholar] [CrossRef] [Green Version]
  37. Jeck, W.R.; Sorrentino, J.A.; Wang, K.; Slevin, M.K.; Burd, C.E.; Liu, J.; Marzluff, W.F.; Sharpless, N.E. Circular RNAs are abundant, conserved, and associated with ALU repeats. RNA 2013, 19, 141–157. [Google Scholar] [CrossRef] [Green Version]
  38. Suzuki, H.; Tsukahara, T. A View of Pre-mRNA Splicing from RNase R Resistant RNAs. Int. J. Mol. Sci. 2014, 15, 9331–9342. [Google Scholar] [CrossRef] [Green Version]
  39. Guo, J.U.; Agarwal, V.; Guo, H.; Bartel, D.P. Expanded identification and characterization of mammalian circular RNAs. Genome Biol. 2014, 15, 409. [Google Scholar] [CrossRef]
  40. Zhang, Y.; Zhang, X.-O.; Chen, T.; Xiang, J.-F.; Yin, Q.-F.; Xing, Y.-H.; Zhu, S.; Yang, L.; Chen, L.-L. Circular Intronic Long Noncoding RNAs. Mol. Cell 2013, 51, 792–806. [Google Scholar] [CrossRef]
  41. Li, Z.; Huang, C.; Bao, C.; Chen, L.; Lin, M.; Wang, X.; Zhong, G.; Yu, B.; Hu, W.; Dai, L.; et al. Exon-intron circular RNAs regulate transcription in the nucleus. Nat. Struct. Mol. Biol. 2015, 22, 256–264. [Google Scholar] [CrossRef]
  42. Noto, J.J.; Schmidt, C.A.; Matera, A.G. Engineering and expressing circular RNAs via tRNA splicing. RNA Biol. 2017, 14, 978–984. [Google Scholar] [CrossRef] [Green Version]
  43. Kelly, S.; Greenman, C.; Cook, P.R.; Papantonis, A. Exon Skipping Is Correlated with Exon Circularization. J. Mol. Biol. 2015, 427, 2414–2417. [Google Scholar] [CrossRef]
  44. Barrett, S.P.; Wang, P.L.; Salzman, J. Circular RNA biogenesis can proceed through an exon-containing lariat precursor. Elife 2015, 4, e07540. [Google Scholar] [CrossRef] [PubMed]
  45. Chen, L.L.; Yang, L. Regulation of circRNA biogenesis. RNA Biol. 2015, 12, 381–388. [Google Scholar] [CrossRef] [PubMed]
  46. Enuka, Y.; Lauriola, M.; Feldman, M.E.; Sas-Chen, A.; Ulitsky, I.; Yarden, Y. Circular RNAs are long-lived and display only minimal early alterations in response to a growth factor. Nucleic Acids Res. 2016, 44, 1370–1383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Zhang, X.-O.; Wang, H.-B.; Zhang, Y.; Lu, X.; Chen, L.-L.; Yang, L. Complementary Sequence-Mediated Exon Circularization. Cell 2014, 159, 134–147. [Google Scholar] [CrossRef] [Green Version]
  48. Geng, X.; Jia, Y.; Zhang, Y.; Shi, L.; Li, Q.; Zang, A.; Wang, H. Circular RNA: Biogenesis, degradation, functions and potential roles in mediating resistance to anticarcinogens. Epigenomics 2020, 12, 267–283. [Google Scholar] [CrossRef]
  49. Chen, L.-L.; Yang, L. ALU ternative Regulation for Gene Expression. Trends Cell Biol. 2017, 27, 480–490. [Google Scholar] [CrossRef] [Green Version]
  50. Ashwal-Fluss, R.; Meyer, M.; Nagarjuna, I.A.; Bartok, O.; Hanan, M..; Evantal, N.; Memczak, S.; Rajewsky, N.; Kadener, S. circRNA Biogenesis Competes with Pre-mRNA Splicing. Mol. Cell 2014, 56, 55–66. [Google Scholar] [CrossRef]
  51. Li, X.; Liu, C.-X.; Xue, W.; Zhang, Y.; Jiang, S.; Yin, Q.-F.; Wei, J.; Yao, R.-W.; Yang, L.; Chen, L.-L. Coordinated circRNA Biogenesis and Function with NF90/NF110 in Viral Infection. Mol. Cell 2017, 67, 214–227. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Simon, S.K.T.J.; Simon, S.K.T.J.; Vanessa, S.M.; Caroline, R.S.; Andreas, W.S.; Philip, A.G.; Gregory, J.G. The RNA Binding Protein Quaking Regulates Formation of circRNAs. Cell 2015, 160, 1125–1134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Fei, T.; Chen, Y.; Xiao, T.; Li, W.; Cato, L.; Zhang, P.; Cotter, M.B.; Bowden, M.; Lis, R.T.; Zhao, S.G.; et al. Genome-wide CRISPR screen identifies HNRNPL as a prostate cancer dependency regulating RNA splicing. Proc. Natl. Acad. Sci. USA 2017, 114, E5207–E5215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Errichelli, L.; Dini Modigliani, S.; Laneve, P.; Colantoni, A.; Legnini, I.; Capauto, D.; Rosa, A.; De Santis, R.; Scarfò, R.; Peruzzi, G.; et al. FUS affects circular RNA expression in murine embryonic stem cell-derived motor neurons. Nat. Commun. 2017, 8, 14741. [Google Scholar] [CrossRef] [Green Version]
  55. Aktaş, T.; Avşar Ilık, İ.; Maticzka, D.; Bhardwaj, V.; Pessoa Rodrigues, C.; Mittler, G.; Manke, T.; Backofen, R.; Akhtar, A. DHX9 suppresses RNA processing defects originating from the Alu invasion of the human genome. Nature 2017, 544, 115–119. [Google Scholar] [CrossRef]
  56. Eisenberg, E.; Levanon, E.Y. A-to-I RNA editing—Immune protector and transcriptome diversifier. Nat. Rev. Genet. 2018, 19, 473–490. [Google Scholar] [CrossRef]
  57. Wu, J.; Qi, X.; Liu, L.; Hu, X.; Liu, J.; Yang, J.; Yang, J.; Lu, L.; Zhang, Z.; Ma, S.; et al. Emerging Epigenetic Regulation of Circular RNAs in Human Cancer. Mol. Ther. Nucleic Acids 2019, 16, 589–596. [Google Scholar] [CrossRef] [Green Version]
  58. Qu, S.; Yang, X.; Li, X.; Wang, J.; Gao, Y.; Shang, R.; Sun, W.; Dou, K.; Li, H. Circular RNA: A new star of noncoding RNAs. Cancer Lett. 2015, 365, 141–148. [Google Scholar] [CrossRef]
  59. Park, O.H.; Ha, H.; Lee, Y.; Boo, S.H.; Kwon, D.H.; Song, H.K.; Kim, Y.K. Endoribonucleolytic Cleavage of m6A-Containing RNAs by RNase P/MRP Complex. Mol. Cell 2019, 74, 494–507. [Google Scholar] [CrossRef]
  60. Lu, Z.; Filonov, G.; Noto, J.; Schmidt, C.; Hatkevich, T.; Wen, Y.; Jaffrey, S.; Matera, A.G. Metazoan tRNA introns generate stable circular RNAs in vivo. RNA 2015, 21, 1554–1565. [Google Scholar] [CrossRef]
  61. Huang, C.; Liang, D.; Tatomer, D.C.; Wilusz, J.E. A length-dependent evolutionarily conserved pathway controls nuclear export of circular RNAs. Genes Dev. 2018, 32, 639–644. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Guria, A.; Sharma, P.; Natesan, S.; Pandi, G. Circular RNAs—The Road Less Traveled. Front. Mol. Biosci. 2020, 6, 146. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Liu, C.-X.; Li, X.; Nan, F.; Jiang, S.; Gao, X.; Guo, S.-K.; Xue, W.; Cui, Y.; Dong, K.; Ding, H.; et al. Structure and Degradation of Circular RNAs Regulate PKR Activation in Innate Immunity. Cell 2019, 177, 865–880. [Google Scholar] [CrossRef] [PubMed]
  64. Hansen, T.B.; Wiklund, E.D.; Bramsen, J.B.; Villadsen, S.B.; Statham, A.L.; Clark, S.J.; Kjems, J. miRNA-dependent gene silencing involving Ago2-mediated cleavage of a circular antisense RNA. EMBO J. 2011, 30, 4414–4422. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Fischer, J.W.; Busa, V.F.; Shao, Y.; Leung, A.K.L. Structure-Mediated RNA Decay by UPF1 and G3BP1. Mol. Cell 2020, 78, 70–84. [Google Scholar] [CrossRef]
  66. Li, X.; Zhang, J.L.; Lei, Y.N.; Liu, X.Q.; Xue, W.; Zhang, Y.; Nan, F.; Gao, X.; Zhang, J.; Wei, J.; et al. Linking circular intronic RNA degradation and function in transcription by RNase H1. Sci. China Life Sci. 2021, 64, 1795–1809. [Google Scholar] [CrossRef]
  67. Guo, Y.; Zhu, X.; Zeng, M.; Qi, L.; Tang, X.; Wang, D.; Zhang, M.; Xie, Y.; Li, H.; Yang, X.; et al. A diet high in sugar and fat influences neurotransmitter metabolism and then affects brain function by altering the gut microbiota. Transl. Psychiatry 2021, 11, 328. [Google Scholar] [CrossRef]
  68. 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] [PubMed] [Green Version]
  69. Zhang, X.; Wang, S.; Wang, H.; Cao, J.; Huang, X.; Chen, Z.; Xu, P.; Sun, G.; Xu, J.; Lv, J.; et al. Circular RNA circNRIP1 acts as a microRNA-149-5p sponge to promote gastric cancer progression via the AKT1/mTOR pathway. Mol. Cancer 2019, 18, 20. [Google Scholar] [CrossRef] [Green Version]
  70. Lin, J.; Zhang, Y.; Zeng, X.; Xue, C.; Lin, X. CircRNA CircRIMS Acts as a MicroRNA Sponge to Promote Gastric Cancer Metastasis. ACS Omega 2020, 5, 23237–23246. [Google Scholar] [CrossRef]
  71. Zhang, L.; Liu, Y.; Tao, H.; Zhu, H.; Pan, Y.; Li, P.; Liang, H.; Zhang, B.; Song, J. Circular RNA circUBE2J2 acts as the sponge of microRNA-370-5P to suppress hepatocellular carcinoma progression. Cell Death Dis. 2021, 12, 985. [Google Scholar] [CrossRef]
  72. Liang, H.F.; Zhang, X.Z.; Liu, B.G.; Jia, G.T.; Li, W.L. Circular RNA circ-ABCB10 promotes breast cancer proliferation and progression through sponging miR-1271. Am. J. Cancer Res. 2017, 7, 1566–1576. [Google Scholar] [PubMed]
  73. Du, W.W.; Fang, L.; Yang, W.; Wu, N.; Awan, F.M.; Yang, Z.; Yang, B.B. Induction of tumor apoptosis through a circular RNA enhancing Foxo3 activity. Cell Death Differ. 2017, 24, 357–370. [Google Scholar] [CrossRef] [PubMed]
  74. Du, W.W.; Yang, W.; Liu, E.; Yang, Z.; Dhaliwal, P.; Yang, B.B. Foxo3 circular RNA retards cell cycle progression via forming ternary complexes with p21 and CDK2. Nucleic Acids Res. 2016, 44, 2846–2858. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Lei, M.; Zheng, G.; Ning, Q.; Zheng, J.; Dong, D. Translation and functional roles of circular RNAs in human cancer. Mol. Cancer 2020, 19, 30. [Google Scholar] [CrossRef] [Green Version]
  76. Yang, Y.; Gao, X.; Zhang, M.; Yan, S.; Sun, C.; Xiao, F.; Huang, N.; Yang, X.; Zhao, K.; Zhou, H.; et al. Novel Role of FBXW7 Circular RNA in Repressing Glioma Tumorigenesis. J. Natl. Cancer Inst. 2018, 110, 304–315. [Google Scholar] [CrossRef] [Green Version]
  77. Hammond, S.M. An overview of microRNAs. Adv. Drug Deliv. Rev. 2015, 87, 3–14. [Google Scholar] [CrossRef] [Green Version]
  78. Macfarlane, L.A.; Murphy, P.R. MicroRNA: Biogenesis, Function and Role in Cancer. Curr. Genom. 2010, 11, 537–561. [Google Scholar] [CrossRef] [Green Version]
  79. 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]
  80. Fabian, M.R.; Sonenberg, N. The mechanics of miRNA-mediated gene silencing: A look under the hood of miRISC. Nat. Struct. Mol. Biol. 2012, 19, 586–593. [Google Scholar] [CrossRef]
  81. Ramchandran, R.; Chaluvally-Raghavan, P. miRNA-Mediated RNA Activation in Mammalian Cells. Adv. Exp. Med. Biol. 2017, 983, 81–89. [Google Scholar] [CrossRef] [PubMed]
  82. Xie, H.; Ren, X.; Xin, S.; Lan, X.; Lu, G.; Lin, Y.; Yang, S.; Zeng, Z.; Liao, W.; Ding, Y.Q.; et al. Emerging roles of circRNA_001569 targeting miR-145 in the proliferation and invasion of colorectal cancer. Oncotarget 2016, 7, 26680–26691. [Google Scholar] [CrossRef] [PubMed]
  83. Xu, X.W.; Zheng, B.A.; Hu, Z.M.; Qian, Z.Y.; Huang, C.J.; Liu, X.Q.; Wu, W.D. Circular RNA hsa_circ_000984 promotes colon cancer growth and metastasis by sponging miR-106b. Oncotarget 2017, 8, 91674–91683. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. 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]
  85. Fu, L.; Chen, Q.; Yao, T.; Li, T.; Ying, S.; Hu, Y.; Guo, J. Hsa_circ_0005986 inhibits carcinogenesis by acting as a miR-129-5p sponge and is used as a novel biomarker for hepatocellular carcinoma. Oncotarget 2017, 8, 43878–43888. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Jarlstad Olesen, M.T.; L, S.K. Circular RNAs as microRNA sponges: Evidence and controversies. Essays Biochem. 2021, 65, 685–696. [Google Scholar] [CrossRef]
  87. Oliveira, C.; Faoro, H.; Alves, L.R.; Goldenberg, S. RNA-binding proteins and their role in the regulation of gene expression in Trypanosoma cruzi and Saccharomyces cerevisiae. Genet. Mol. Biol. 2017, 40, 22–30. [Google Scholar] [CrossRef] [Green Version]
  88. Conlon, E.G.; Manley, J.L. RNA-binding proteins in neurodegeneration: Mechanisms in aggregate. Genes Dev. 2017, 31, 1509–1528. [Google Scholar] [CrossRef]
  89. Newman, R.; McHugh, J.; Turner, M. RNA binding proteins as regulators of immune cell biology. Clin. Exp. Immunol. 2016, 183, 37–49. [Google Scholar] [CrossRef] [Green Version]
  90. Abdelmohsen, K.; Panda, A.C.; Munk, R.; Grammatikakis, I.; Dudekula, D.B.; De, S.; Kim, J.; Noh, J.H.; Kim, K.M.; Martindale, J.L.; et al. Identification of HuR target circular RNAs uncovers suppression of PABPN1 translation by CircPABPN1. RNA Biol. 2017, 14, 361–369. [Google Scholar] [CrossRef]
  91. Xia, P.; Wang, S.; Ye, B.; Du, Y.; Li, C.; Xiong, Z.; Qu, Y.; Fan, Z. A Circular RNA Protects Dormant Hematopoietic Stem Cells from DNA Sensor cGAS-Mediated Exhaustion. Immunity 2018, 48, 688–701. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Du, W.W.; Yang, W.; Chen, Y.; Wu, Z.K.; Foster, F.S.; Yang, Z.; Li, X.; Yang, B.B. Foxo3 circular RNA promotes cardiac senescence by modulating multiple factors associated with stress and senescence responses. Eur. Heart J. 2017, 38, 1402–1412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Zeng, Y.; Du, W.W.; Wu, Y.; Yang, Z.; Awan, F.M.; Li, X.; Yang, W.; Zhang, C.; Yang, Q.; Yee, A.; et al. A Circular RNA Binds To and Activates AKT Phosphorylation and Nuclear Localization Reducing Apoptosis and Enhancing Cardiac Repair. Theranostics 2017, 7, 3842–3855. [Google Scholar] [CrossRef] [PubMed]
  94. Chen, N.; Zhao, G.; Yan, X.; Lv, Z.; Yin, H.; Zhang, S.; Song, W.; Li, X.; Li, L.; Du, Z.; et al. A novel FLI1 exonic circular RNA promotes metastasis in breast cancer by coordinately regulating TET1 and DNMT1. Genome Biol. 2018, 19, 218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Yang, Z.G.; Awan, F.M.; Du, W.W.; Zeng, Y.; Lyu, J.; Wu, D.; Gupta, S.; Yang, W.; Yang, B.B. The Circular RNA Interacts with STAT3, Increasing Its Nuclear Translocation and Wound Repair by Modulating Dnmt3a and miR-17 Function. Mol. Ther. 2017, 25, 2062–2074. [Google Scholar] [CrossRef] [Green Version]
  96. Sun, Y.M.; Wang, W.T.; Zeng, Z.C.; Chen, T.Q.; Han, C.; Pan, Q.; Huang, W.; Fang, K.; Sun, L.Y.; Zhou, Y.F.; et al. circMYBL2, a circRNA from MYBL2, regulates FLT3 translation by recruiting PTBP1 to promote FLT3-ITD AML progression. Blood 2019, 134, 1533–1546. [Google Scholar] [CrossRef] [Green Version]
  97. Conn, V.M.; Hugouvieux, V.; Nayak, A.; Conos, S.A.; Capovilla, G.; Cildir, G.; Jourdain, A.; Tergaonkar, V.; Schmid, M.; Zubieta, C.; et al. A circRNA from SEPALLATA3 regulates splicing of its cognate mRNA through R-loop formation. Nat. Plants 2017, 3, 17053. [Google Scholar] [CrossRef]
  98. Richter, J.D.; Sonenberg, N. Regulation of cap-dependent translation by eIF4E inhibitory proteins. Nature 2005, 433, 477–480. [Google Scholar] [CrossRef]
  99. Abe, N.; Matsumoto, K.; Nishihara, M.; Nakano, Y.; Shibata, A.; Maruyama, H.; Shuto, S.; Matsuda, A.; Yoshida, M.; Ito, Y.; et al. Rolling Circle Translation of Circular RNA in Living Human Cells. Sci. Rep. 2015, 5, 16435. [Google Scholar] [CrossRef] [Green Version]
  100. Yang, Y.; Fan, X.; Mao, M.; Song, X.; Wu, P.; Zhang, Y.; Jin, Y.; Yang, Y.; Chen, L.-L.; Wang, Y.; et al. Extensive translation of circular RNAs driven by N6-methyladenosine. Cell Res. 2017, 27, 626–641. [Google Scholar] [CrossRef]
  101. Pamudurti, N.R.; Bartok, O.; Jens, M.; Ashwal-Fluss, R.; Stottmeister, C.; Ruhe, L.; Hanan, M.; Wyler, E.; Perez-Hernandez, D.; Ramberger, E.; et al. Translation of CircRNAs. Mol. Cell 2017, 66, 9–21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Bennett, R.G.; Kharbanda, K.K.; Tuma, D.J. Inhibition of markers of hepatic stellate cell activation by the hormone relaxin. Biochem. Pharmacol. 2003, 66, 867–874. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Dewidar, B.; Meyer, C.; Dooley, S.; Meindl-Beinker, A.N. TGF-β in Hepatic Stellate Cell Activation and Liver Fibrogenesis-Updated 2019. Cells 2019, 8, 1419. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Kagan, P.; Sultan, M.; Tachlytski, I.; Safran, M.; Ben-Ari, Z. Both MAPK and STAT3 signal transduction pathways are necessary for IL-6-dependent hepatic stellate cells activation. PLoS One 2017, 12, e0176173. [Google Scholar] [CrossRef] [Green Version]
  105. Park, H.-J.; Kim, H.-G.; Wang, J.-H.; Choi, M.-K.; Han, J.-M.; Lee, J.-S.; Son, C.-G. Comparison of TGF-β, PDGF, and CTGF in hepatic fibrosis models using DMN, CCl4, and TAA. Drug Chem. Toxicol. 2016, 39, 111–118. [Google Scholar] [CrossRef] [PubMed]
  106. Cai, C.X.; Buddha, H.; Castelino-Prabhu, S.; Zhang, Z.; Britton, R.S.; Bacon, B.R.; Neuschwander-Tetri, B.A. Activation of Insulin-PI3K/Akt-p70S6K Pathway in Hepatic Stellate Cells Contributes to Fibrosis in Nonalcoholic Steatohepatitis. Dig. Dis. Sci. 2017, 62, 968–978. [Google Scholar] [CrossRef] [PubMed]
  107. Ge, W.-S.; Wang, Y.-J.; Wu, J.-X.; Fan, J.-G.; Chen, Y.-W.; Zhu, L. β-catenin is overexpressed in hepatic fibrosis and blockage of Wnt/β-catenin signaling inhibits hepatic stellate cell activation. Mol. Med. Rep. 2014, 9, 2145–2151. [Google Scholar] [CrossRef] [Green Version]
  108. Bansal, R.; van Baarlen, J.; Storm, G.; Prakash, J. The interplay of the Notch signaling in hepatic stellate cells and macrophages determines the fate of liver fibrogenesis. Sci. Rep. 2015, 5, 18272. [Google Scholar] [CrossRef] [Green Version]
  109. Sicklick, J.K.; Li, Y.-X.; Choi, S.S.; Qi, Y.; Chen, W.; Bustamante, M.; Huang, J.; Zdanowicz, M.; Camp, T.; Torbenson, M.S.; et al. Role for Hedgehog signaling in hepatic stellate cell activation and viability. Lab. Investig. 2005, 85, 1368–1380. [Google Scholar] [CrossRef] [Green Version]
  110. Xie, G.; Karaca, G.; Swiderska-Syn, M.; Michelotti, G.A.; Krüger, L.; Chen, Y.; Premont, R.T.; Choi, S.S.; Diehl, A.M. Cross-talk between Notch and Hedgehog regulates hepatic stellate cell fate in mice. Hepatology 2013, 58, 1801–1813. [Google Scholar] [CrossRef]
  111. Mannaerts, I.; Leite, S.B.; Verhulst, S.; Claerhout, S.; Eysackers, N.; Thoen, L.F.; Hoorens, A.; Reynaert, H.; Halder, G.; van Grunsven, L.A. The Hippo pathway effector YAP controls mouse hepatic stellate cell activation. J. Hepatol. 2015, 63, 679–688. [Google Scholar] [CrossRef] [PubMed]
  112. Tao, Y.; Wang, N.; Qiu, T.; Sun, X. The Role of Autophagy and NLRP3 Inflammasome in Liver Fibrosis. Biomed. Res. Int. 2020, 2020, 7269150. [Google Scholar] [CrossRef] [PubMed]
  113. 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]
  114. Yang, Y.R.; Hu, S.; Bu, F.T.; Li, H.; Huang, C.; Meng, X.M.; Zhang, L.; Lv, X.W.; Li, J. Circular RNA CREBBP Suppresses Hepatic Fibrosis Via Targeting the hsa-miR-1291/LEFTY2 Axis. Front. Pharmacol. 2021, 12, 741151. [Google Scholar] [CrossRef]
  115. 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] [PubMed]
  116. 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] [Green Version]
  117. 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]
  118. Wang, P.; Huang, Z.; Peng, Y.; Li, H.; Lin, T.; Zhao, Y.; Hu, Z.; Zhou, Z.; Zhou, W.; Liu, Y.; et al. Circular RNA circBNC2 inhibits epithelial cell G2-M arrest to prevent fibrotic maladaptive repair. Nat. Commun. 2022, 13, 1–19. [Google Scholar] [CrossRef]
  119. 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]
  120. Zhu, S.; Chen, X.; Wang, J.N.; Xu, J.J.; Wang, A.; Li, J.J.; Wu, S.; Wu, Y.Y.; Li, X.F.; Huang, C.; et al. Circular RNA circUbe2k promotes hepatic fibrosis via sponging miR-149-5p/TGF-2 axis. Faseb J. 2021, 35, e21622. [Google Scholar] [CrossRef]
  121. Niu, H.; Zhang, L.; Wang, B.; Zhang, G.C.; Liu, J.; Wu, Z.F.; Du, S.S.; Zeng, Z.C. CircTUBD1 Regulates Radiation-induced Liver Fibrosis Response via a circTUBD1/micro-203a-3p/Smad3 Positive Feedback Loop. J. Clin. Transl. Hepatol. 2022, 10, 680–691. [Google Scholar] [CrossRef] [PubMed]
  122. Niu, H.; Zhang, L.; Chen, Y.-H.; Yuan, B.-Y.; Wu, Z.-F.; Cheng, J.C.-H.; 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]
  123. Hu, P.; Guo, J.; Zhao, B.; Zhang, Z.; Zhu, J.; Liu, F. CircCHD2/miR-200b-3p/HLF Axis Promotes Liver Cirrhosis. J. Environ. Pathol. Toxicol. Oncol. 2022, 41, 1–10. [Google Scholar] [CrossRef] [PubMed]
  124. Ma, L.; Wei, J.; Zeng, Y.; Liu, J.; Xiao, E.; Kang, Y.; Kang, Y. Mesenchymal stem cell-originated exosomal circDIDO1 suppresses hepatic stellate cell activation by miR-141-3p/PTEN/AKT pathway in human liver fibrosis. Drug Deliv. 2022, 29, 440–453. [Google Scholar] [CrossRef]
  125. Ma, J.; Li, Y.; Chen, M.; Wang, W.; Zhao, Q.; He, B.; Zhang, M.; Jiang, Y. hMSCs-derived exosome circCDK13 inhibits liver fibrosis by regulating the expression of MFGE8 through miR-17-5p/KAT2B. Cell Biol. Toxicol. 2022, 1–22. [Google Scholar] [CrossRef] [PubMed]
  126. Zhu, L.; Ren, T.; Zhu, Z.; Cheng, M.; Mou, Q.; Mu, M.; Liu, Y.; Yao, Y.; Cheng, Y.; Zhang, B.; et al. Thymosin-4 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]
  127. Xu, Z.-X.; Li, J.-Z.; Li, Q.; Xu, M.-Y.; Li, H.-Y. CircRNA608-microRNA222-PINK1 axis regulates the mitophagy of hepatic stellate cells in NASH related fibrosis. Biochem. Biophys. Res. Commun. 2022, 610, 35–42. [Google Scholar] [CrossRef] [PubMed]
  128. 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]
  129. 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]
  130. 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.e22. [Google Scholar] [CrossRef]
  131. Li, B.; Zhou, J.; Luo, Y.; Tao, K.; Zhang, L.; Zhao, Y.; Lin, Y.; Zeng, X.; Yu, H. Suppressing circ_0008494 inhibits HSCs activation by regulating the miR-185-3p/Col1a1 axis. Front. Pharmacol. 2022, 13, 1050093. [Google Scholar] [CrossRef] [PubMed]
  132. 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] [PubMed]
  133. Milani, S.; Herbst, H.; Schuppan, D.; Stein, H.; Surrenti, C. Transforming growth factors beta 1 and beta 2 are differentially expressed in fibrotic liver disease. Am. J. Pathol. 1991, 139, 1221–1229. [Google Scholar]
  134. Sun, Y.-Y.; Li, X.-F.; Meng, X.-M.; Huang, C.; Zhang, L.; Li, J. Macrophage Phenotype in Liver Injury and Repair. Scand. J. Immunol. 2017, 85, 166–174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Yan, X.; Liu, Z.; Chen, Y. Regulation of TGF-beta signaling by Smad7. Acta Biochim. Biophys. Sin. 2009, 41, 263–272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Dooley, S.; Hamzavi, J.; Breitkopf, K.; Wiercinska, E.; Said, H.M.; Lorenzen, J.; Dijke, P.T.; Gressner, A.M. Smad7 prevents activation of hepatic stellate cells and liver fibrosis in rats. Gastroenterology 2003, 125, 178–191. [Google Scholar] [CrossRef]
  137. Yang, Y.-r.; Bu, F.-t.; Yang, Y.; Li, H.; Huang, C.; Meng, X.-m.; Zhang, L.; Lv, X.-w.; Li, J. LEFTY2 alleviates hepatic stellate cell activation and liver fibrosis by regulating the TGF-1/Smad3 pathway. Mol. Immunol. 2020, 126, 31–39. [Google Scholar] [CrossRef]
  138. Ulloa, L.; Tabibzadeh, S. Lefty inhibits receptor-regulated Smad phosphorylation induced by the activated transforming growth factor-beta receptor. J. Biol. Chem. 2001, 276, 21397–21404. [Google Scholar] [CrossRef] [Green Version]
  139. Lin, A.H.; Luo, J.; Mondshein, L.H.; ten Dijke, P.; Vivien, D.; Contag, C.H.; Wyss-Coray, T. Global analysis of Smad2/3-dependent TGF-beta signaling in living mice reveals prominent tissue-specific responses to injury. J. Immunol. 2005, 175, 547–554. [Google Scholar] [CrossRef] [Green Version]
  140. Chambers, D.M.; Moretti, L.; Zhang, J.J.; Cooper, S.W.; Chambers, D.M.; Santangelo, P.J.; Barker, T.H. LEM domain-containing protein 3 antagonizes TGF-SMAD2/3 signaling in a stiffness-dependent manner in both the nucleus and cytosol. J. Biol. Chem. 2018, 293, 15867–15886. [Google Scholar] [CrossRef] [Green Version]
  141. Ariyachet, C.; Chuaypen, N.; Kaewsapsak, P.; Chantaravisoot, N.; Jindatip, D.; Potikanond, S.; Tangkijvanich, P. MicroRNA-223 Suppresses Human Hepatic Stellate Cell Activation Partly via Regulating the Actin Cytoskeleton and Alleviates Fibrosis in Organoid Models of Liver Injury. Int. J. Mol. Sci. 2022, 23, 9380. [Google Scholar] [CrossRef]
  142. Wang, X.; Seo, W.; Park, S.H.; Fu, Y.; Hwang, S.; Rodrigues, R.M.; Feng, D.; Gao, B.; He, Y. MicroRNA-223 restricts liver fibrosis by inhibiting the TAZ-IHH-GLI2 and PDGF signaling pathways via the crosstalk of multiple liver cell types. Int. J. Biol. Sci. 2021, 17, 1153–1167. [Google Scholar] [CrossRef] [PubMed]
  143. Wang, L.; Wang, Y.; Quan, J. Exosomal miR-223 derived from natural killer cells inhibits hepatic stellate cell activation by suppressing autophagy. Mol. Med. 2020, 26, 1–9. [Google Scholar] [CrossRef]
  144. Calvente, C.J.; Del Pilar, H.; Tameda, M.; Johnson, C.D.; Feldstein, A.E. MicroRNA 223 3p Negatively Regulates the NLRP3 Inflammasome in Acute and Chronic Liver Injury. Mol. Ther. 2019, 28, 653–663. [Google Scholar] [CrossRef] [PubMed]
  145. 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] [PubMed]
  146. Zou, G.L.; Zuo, S.; Lu, S.; Hu, R.H.; Lu, Y.Y.; Yang, J.; Deng, K.S.; Wu, Y.T.; Mu, M.; Zhu, J.J.; et al. Bone morphogenetic protein-7 represses hepatic stellate cell activation and liver fibrosis via regulation of TGF-/Smad signaling pathway. World J. Gastroenterol. 2019, 25, 4222–4234. [Google Scholar] [CrossRef]
  147. Meng, X.M.; Chung, A.C.; Lan, H.Y. Role of the TGF-/BMP-7/Smad pathways in renal diseases. Clin. Sci. 2013, 124, 243–254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Takashima, M.; Parsons, C.J.; Ikejima, K.; Watanabe, S.; White, E.S.; Rippe, R.A. The tumor suppressor protein PTEN inhibits rat hepatic stellate cell activation. J. Gastroenterol. 2009, 44, 847–855. [Google Scholar] [CrossRef] [Green Version]
  149. Dan, H.C.; Cooper, M.J.; Cogswell, P.C.; Duncan, J.A.; Ting, J.P.; Baldwin, A.S. Akt-dependent regulation of NF-{kappa}B is controlled by mTOR and Raptor in association with IKK. Genes Dev. 2008, 22, 1490–1500. [Google Scholar] [CrossRef] [Green Version]
  150. Gao, R.; Brigstock, D.R. Activation of nuclear factor kappa B (NF-kappaB) by connective tissue growth factor (CCN2) is involved in sustaining the survival of primary rat hepatic stellate cells. Cell Commun. Signal. 2005, 3, 14. [Google Scholar] [CrossRef] [Green Version]
  151. Paik, Y.H.; Kim, J.; Aoyama, T.; De Minicis, S.; Bataller, R.; Brenner, D.A. Role of NADPH Oxidases in Liver Fibrosis. Antioxid. Redox Signal. 2014, 20, 2854–2872. [Google Scholar] [CrossRef]
  152. Ignat, S.-R.; Dinescu, S.; Hermenean, A.; Costache, M. Cellular Interplay as a Consequence of Inflammatory Signals Leading to Liver Fibrosis Development. Cells 2020, 9, 461. [Google Scholar] [CrossRef] [Green Version]
  153. Reiter, F.P.; Ye, L.; Ofner, A.; Schiergens, T.S.; Ziesch, A.; Brandl, L.; Ben Khaled, N.; Hohenester, S.; Wimmer, R.; Artmann, R.; et al. p70 Ribosomal Protein S6 Kinase Is a Checkpoint of Human Hepatic Stellate Cell Activation and Liver Fibrosis in Mice. Cell. Mol. Gastroenterol. Hepatol. 2021, 13, 95–112. [Google Scholar] [CrossRef] [PubMed]
  154. Struhl, K. Histone acetylation and transcriptional regulatory mechanisms. Genes Dev. 1998, 12, 599–606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Khalifeh-Soltani, A.; Gupta, D.; Ha, A.; Podolsky, M.J.; Datta, R.; Atabai, K. The Mfge8-α8β1;1-PTEN pathway regulates airway smooth muscle contraction in allergic inflammation. FASEB J. 2018, 32, 5927–5936. [Google Scholar] [CrossRef] [PubMed]
  156. Kane, L.P.; Shapiro, V.S.; Stokoe, D.; Weiss, A. Induction of NF-κB by the Akt/PKB kinase. Curr. Biol. 1999, 9, 601–604. [Google Scholar] [CrossRef] [Green Version]
  157. Koyano, F.; Okatsu, K.; Kosako, H.; Tamura, Y.; Go, E.; Kimura, M.; Kimura, Y.; Tsuchiya, H.; Yoshihara, H.; Hirokawa, T.; et al. Ubiquitin is phosphorylated by PINK1 to activate parkin. Nature 2014, 510, 162–166. [Google Scholar] [CrossRef]
  158. Berwick, D.; Harvey, K. The regulation and deregulation of Wnt signaling by PARK genes in health and disease. J. Mol. Cell Biol. 2013, 6, 3–12. [Google Scholar] [CrossRef] [Green Version]
  159. Fan, S.; Price, T.; Huang, W.; Plue, M.; Warren, J.; Sundaramoorthy, P.; Paul, B.; Feinberg, D.; Maciver, N.; Chao, N.; et al. PINK1-Dependent Mitophagy Regulates the Migration and Homing of Multiple Myeloma Cells via the MOB1B-Mediated Hippo-YAP/TAZ Pathway. Adv. Sci. 2020, 7, 1900860. [Google Scholar] [CrossRef] [Green Version]
  160. Kulaberoglu, Y.; Lin, K.; Holder, M.; Gai, Z.; Gomez, M.; Shifa, B.A.; Mavis, M.; Hoa, L.; Sharif, A.A.D.; Lujan, C.; et al. Stable MOB1 interaction with Hippo/MST is not essential for development and tissue growth control. Nat. Commun. 2017, 8, 1–16. [Google Scholar] [CrossRef] [Green Version]
  161. Piccolo, S.; Dupont, S.; Cordenonsi, M. The Biology of YAP/TAZ: Hippo Signaling and Beyond. Physiol. Rev. 2014, 94, 1287–1312. [Google Scholar] [CrossRef]
  162. Dou, S.D.; Zhang, J.N.; Xie, X.L.; Liu, T.; Hu, J.L.; Jiang, X.Y.; Wang, M.M.; Jiang, H.D. MitoQ inhibits hepatic stellate cell activation and liver fibrosis by enhancing PINK1/parkin-mediated mitophagy. Open Med. 2021, 16, 1718–1727. [Google Scholar] [CrossRef]
  163. Kar, R.; Jha, S.K.; Ojha, S.; Sharma, A.; Dholpuria, S.; Raju, V.S.R.; Prasher, P.; Chellappan, D.K.; Gupta, G.; Singh, S.K.; et al. The FBXW7-NOTCH interactome: A ubiquitin proteasomal system-induced crosstalk modulating oncogenic transformation in human tissues. Cancer Rep. 2021, 4, e1369. [Google Scholar] [CrossRef] [PubMed]
  164. Xie, H.; Xie, D.; Zhang, J.; Jin, W.; Li, Y.; Yao, J.; Pan, Z.; Xie, D. ROS/NF-kB Signaling Pathway-Mediated Transcriptional Activation of TRIM37 Promotes HBV-Associated Hepatic Fibrosis. Mol. Ther.-Nucleic Acids 2020, 22, 114–123. [Google Scholar] [CrossRef] [PubMed]
  165. Ramos-Tovar, E.; Muriel, P. Molecular Mechanisms that Link Oxidative Stress, Inflammation, and Fibrosis in the Liver. Antioxidants 2020, 9, 1279. [Google Scholar] [CrossRef]
  166. Chaiteerakij, R.; Roberts, L.R. Telomerase mutation: A genetic risk factor for cirrhosis. Hepatology 2011, 53, 1430–1432. [Google Scholar] [CrossRef] [PubMed]
  167. Duan, X.; Wang, H.; Yang, Y.; Wang, P.; Zhang, H.; Liu, B.; Wei, W.; Yao, W.; Zhou, X.; Zhao, J.; et al. Genetic variants in telomerase-associated protein 1 are associated with telomere damage in PAH-exposed workers. Ecotoxicol. Environ. Saf. 2021, 223, 112558. [Google Scholar] [CrossRef]
  168. Chang, J.T.-C.; Chen, Y.-L.; Yang, H.-T.; Chen, C.-Y.; Cheng, A.-J. Differential regulation of telomerase activity by six telomerase subunits. JBIC J. Biol. Inorg. Chem. 2002, 269, 3442–3450. [Google Scholar] [CrossRef]
  169. Liu, Y.; Snow, B.; Hande, P.; Baerlocher, G.; Kickhoefer, V.; Yeung, D.; Wakeham, A.; Itie, A.; Siderovski, D.; Lansdorp, P.; et al. Telomerase-Associated Protein TEP1 Is Not Essential for Telomerase Activity or Telomere Length Maintenance In Vivo. Mol. Cell. Biol. 2000, 20, 8178–8184. [Google Scholar] [CrossRef]
  170. Kickhoefer, V.A.; Stephen, A.G.; Harrington, L.; Robinson, M.; Rome, L. Vaults and Telomerase Share a Common Subunit, TEP1. J. Biol. Chem. 1999, 274, 32712–32717. [Google Scholar] [CrossRef] [Green Version]
  171. Hahne, J.C.; Lampis, A.; Valeri, N. Vault RNAs: Hidden gems in RNA and protein regulation. Cell. Mol. Life Sci. 2020, 78, 1487–1499. [Google Scholar] [CrossRef] [PubMed]
  172. Liu, B.; Tian, Y.; He, J.; Gu, Q.; Jin, B.; Shen, H.; Li, W.; Shi, L.; Yu, H.; Shan, G.; et al. The potential of mecciRNA in hepatic stellate cell to regulate progression of nonalcoholic hepatitis. J. Transl. Med. 2022, 20, 1–17. [Google Scholar] [CrossRef]
  173. Li, B.; Li, Y.; Li, S.; Li, H.; Liu, L.; Yu, H. Circ_MTM1 knockdown inhibits the progression of HBV-related liver fibrosis via regulating IL7R expression through targeting miR-122-5p. Am. J. Transl. Res. 2022, 14, 2199–2211. [Google Scholar] [PubMed]
  174. Jacobs, S.R.; Michalek, R.D.; Rathmell, J.C. IL-7 Is Essential for Homeostatic Control of T Cell Metabolism In Vivo. J. Immunol. 2010, 184, 3461–3469. [Google Scholar] [CrossRef] [Green Version]
  175. Li, X.; Fang, Y.; Jiang, D.; Dong, Y.; Liu, Y.; Zhang, S.; Guo, J.; Qi, C.; Zhao, C.; Jiang, F.; et al. Targeting FSTL1 for Multiple Fibrotic and Systemic Autoimmune Diseases. Mol. Ther. 2020, 29, 347–364. [Google Scholar] [CrossRef]
  176. Shang, H.; Liu, X.; Guo, H. Knockdown of Fstl1 attenuates hepatic stellate cell activation through the TGF-1/Smad3 signaling pathway. Mol. Med. Rep. 2017, 16, 7119–7123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Liu, C.; Chen, X.; Yang, L.; Kisseleva, T.; Brenner, D.A.; Seki, E. Transcriptional Repression of the Transforming Growth Factor (TGF-) Pseudoreceptor BMP and Activin Membrane-bound Inhibitor (BAMBI) by Nuclear Factor Signaling in Hepatic Stellate Cells*. J. Biol. Chem. 2014, 289, 7082–7091. [Google Scholar] [CrossRef] [Green Version]
  178. Seki, E.; De Minicis, S.; Österreicher, C.H.; Kluwe, J.; Osawa, Y.; Brenner, D.; Schwabe, R.F. TLR4 enhances TGF-β signaling and hepatic fibrosis. Nat. Med. 2007, 13, 1324–1332. [Google Scholar] [CrossRef]
  179. Lang, A.; Schoonhoven, R.; Tuvia, S.; A Brenner, D.; A Rippe, R. Nuclear factor κB in proliferation, activation, and apoptosis in rat hepatic stellate cells. J. Hepatol. 2000, 33, 49–58. [Google Scholar] [CrossRef]
  180. Luedde, T.; Schwabe, R.F. NF-B in the liver–linking injury, fibrosis and hepatocellular carcinoma. Nat. Rev. Gastroenterol. Hepatol. 2011, 8, 108–118. [Google Scholar] [CrossRef] [Green Version]
  181. Liu, W.; Wu, Y.-H.; Zhang, L.; Xue, B.; Wang, Y.; Liu, B.; Liu, X.-Y.; Zuo, F.; Yang, X.-Y.; Chen, F.-Y.; et al. MicroRNA-146a suppresses rheumatoid arthritis fibroblastlike synoviocytes proliferation and inflammatory responses by inhibiting the TLR4/NFkB signaling. Oncotarget 2018, 9, 24050. [Google Scholar] [CrossRef] [PubMed]
  182. Xiang, D.-M.; Sun, W.; Ning, B.-F.; Zhou, T.-F.; Li, X.-F.; Zhong, W.; Cheng, Z.; Xia, M.-Y.; Wang, X.; Deng, X.; et al. The HLF/IL-6/STAT3 feedforward circuit drives hepatic stellate cell activation to promote liver fibrosis. Gut 2018, 67, 1704–1715. [Google Scholar] [CrossRef] [PubMed]
  183. Tang, L.-Y.; Heller, M.; Meng, Z.; Yu, L.-R.; Tang, Y.; Zhou, M.; Zhang, Y.E. Transforming Growth Factor-β (TGF-β) Directly Activates the JAK1-STAT3 Axis to Induce Hepatic Fibrosis in Coordination with the SMAD Pathway. J. Biol. Chem. 2017, 292, 4302–4312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. 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] [PubMed]
  185. Liao, J.; Zhang, Z.; Yuan, Q.; Luo, L.; Hu, X. The mouse Anxa6/miR-9-5p/Anxa2 axis modulates TGF-1-induced mouse hepatic stellate cell (mHSC) activation and CCl4-caused liver fibrosis. Toxicol. Lett. 2022, 362, 38–49. [Google Scholar] [CrossRef]
  186. Wang, J.; He, Z.; Liu, X.; Xu, J.; Jiang, X.; Quan, G.; Jiang, J. LINC00941 promotes pancreatic cancer malignancy by interacting with ANXA2 and suppressing NEDD4L-mediated degradation of ANXA2. Cell Death Dis. 2022, 13, 1–15. [Google Scholar] [CrossRef]
  187. Liu, Y.; Ao, X.; Ding, W.; Ponnusamy, M.; Wu, W.; Hao, X.; Yu, W.; Wang, Y.; Li, P.; Wang, J. Critical role of FOXO3a in carcinogenesis. Mol. Cancer 2018, 17, 1–12. [Google Scholar] [CrossRef] [Green Version]
  188. Hui, R.C.-Y.; Gomes, A.R.; Constantinidou, D.; Costa, J.R.; Karadedou, C.T.; de Mattos, S.F.; Wymann, M.P.; Brosens, J.J.; Schulze, A.; Lam, E.W.-F. The Forkhead Transcription Factor FOXO3a Increases Phosphoinositide-3 Kinase/Akt Activity in Drug-Resistant Leukemic Cells through Induction of PIK3CA Expression. Mol. Cell. Biol. 2008, 28, 5886–5898. [Google Scholar] [CrossRef] [Green Version]
  189. Lu, M.; Hartmann, D.; Braren, R.; Gupta, A.; Wang, B.; Wang, Y.; Mogler, C.; Cheng, Z.; Wirth, T.; Friess, H.; et al. Oncogenic Akt-FOXO3 loop favors tumor-promoting modes and enhances oxidative damage-associated hepatocellular carcinogenesis. BMC Cancer 2019, 19, 1–13. [Google Scholar] [CrossRef] [Green Version]
  190. Choi, S.S.; Witek, R.P.; Yang, L.; Omenetti, A.; Syn, W.-K.; Moylan, C.A.; Jung, Y.; Karaca, G.F.; Teaberry, V.S.; Pereira, T.A.; et al. Activation of Rac1 promotes hedgehog-mediated acquisition of the myofibroblastic phenotype in rat and human hepatic stellate cells. Hepatology 2010, 52, 278–290. [Google Scholar] [CrossRef] [Green Version]
  191. Tang, C.; Wu, X.; Ren, Q.; Yao, M.; Xu, S.; Yan, Z. Hedgehog signaling is controlled by Rac1 activity. Theranostics 2022, 12, 1303–1320. [Google Scholar] [CrossRef] [PubMed]
  192. Arbibe, L.; Mira, J.P.; Teusch, N.; Kline, L.; Guha, M.; Mackman, N.; Godowski, P.J.; Ulevitch, R.J.; Knaus, U.G. Toll-like receptor 2-mediated NF-kappa B activation requires a Rac1-dependent pathway. Nat. Immunol. 2000, 1, 533–540. [Google Scholar] [CrossRef]
  193. Coso, O.A.; Chiariello, M.; Yu, J.-C.; Teramoto, H.; Crespo, P.; Xu, N.; Miki, T.; Gutkind, J.S. The small GTP-binding proteins Rac1 and Cdc42regulate the activity of the JNK/SAPK signaling pathway. Cell 1995, 81, 1137–1146. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Wang, J.; Liu, G.; Li, Q.; Wang, F.; Xie, F.; Zhai, R.; Guo, Y.; Chen, T.; Zhang, N.; Ni, W.; et al. Mucin1 promotes the migration and invasion of hepatocellular carcinoma cells via JNK-mediated phosphorylation of Smad2 at the C-terminal and linker regions. Oncotarget 2015, 6, 19264–19278. [Google Scholar] [CrossRef] [Green Version]
  195. Kamato, D.; Little, P.J. Smad2 linker region phosphorylation is an autonomous cell signalling pathway: Implications for multiple disease pathologies. Biomed. Pharmacother. 2020, 124, 109854. [Google Scholar] [CrossRef]
  196. Afroz, R.; Zhou, Y.; Little, P.J.; Xu, S.; Mohamed, R.; Stow, J.; Kamato, D. Toll-like Receptor 4 Stimulates Gene Expression via Smad2 Linker Region Phosphorylation in Vascular Smooth Muscle Cells. ACS Pharmacol. Transl. Sci. 2020, 3, 524–534. [Google Scholar] [CrossRef] [Green Version]
  197. Yang, Y.M.; Noureddin, M.; Liu, C.; Ohashi, K.; Kim, S.Y.; Ramnath, D.; Powell, E.E.; Sweet, M.J.; Roh, Y.S.; Hsin, I.-F.; et al. Hyaluronan synthase 2–mediated hyaluronan production mediates Notch1 activation and liver fibrosis. Sci. Transl. Med. 2019, 11, aat9824. [Google Scholar] [CrossRef]
  198. Zhubanchaliyev, A.; Temirbekuly, A.; Kongrtay, K.; Wanshura, L.C.; Kunz, J. Targeting Mechanotransduction at the Transcriptional Level: YAP and BRD4 Are Novel Therapeutic Targets for the Reversal of Liver Fibrosis. Front. Pharmacol. 2016, 7, 462. [Google Scholar] [CrossRef] [Green Version]
  199. Liu, Z.; Mo, H.; Liu, R.; Niu, Y.; Chen, T.; Xu, Q.; Tu, K.; Yang, N. Matrix stiffness modulates hepatic stellate cell activation into tumor-promoting myofibroblasts via E2F3-dependent signaling and regulates malignant progression. Cell Death Dis. 2021, 12, 1134. [Google Scholar] [CrossRef]
  200. 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]
  201. Wesche, H.; Gao, X.; Li, X.; Kirschning, C.J.; Stark, G.R.; Cao, Z. IRAK-M is a novel member of the Pelle/interleukin-1 receptor-associated kinase (IRAK) family. J. Biol. Chem. 1999, 274, 19403–19410. [Google Scholar] [CrossRef] [PubMed]
  202. Xu, J.J.; Chen, X.; Zhu, S.; Jiang, L.F.; Ma, W.X.; Chen, S.Y.; Meng, X.M.; Huang, C.; Li, J. Myc-mediated circular RNA circMcph1/miR-370-3p/Irak2 axis is a progressive regulator in hepatic fibrosis. Life Sci. 2023, 312, 121182. [Google Scholar] [CrossRef] [PubMed]
  203. Csak, T.; Bala, S.; Lippai, D.; Satishchandran, A.; Catalano, D.; Kodys, K.; Szabo, G. microRNA-122 regulates hypoxia-inducible factor-1 and vimentin in hepatocytes and correlates with fibrosis in diet-induced steatohepatitis. Liver Int. 2015, 35, 532–541. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Lagos-Quintana, M.; Rauhut, R.; Yalcin, A.; Meyer, J.; Lendeckel, W.; Tuschl, T. Identification of tissue-specific microRNAs from mouse. Curr. Biol. 2002, 12, 735–739. [Google Scholar] [CrossRef] [Green Version]
  205. Halász, T.; Horváth, G.; Pár, G.; Werling, K.; Kiss, A.; Schaff, Z.; Lendvai, G. miR-122 negatively correlates with liver fibrosis as detected by histology and FibroScan. World J. Gastroenterol. 2015, 21, 7814–7823. [Google Scholar] [CrossRef] [Green Version]
  206. 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]
  207. Li, H.-M.; Ma, X.-L.; Li, H.-G. Intriguing circles: Conflicts and controversies in circular RNA research. WIREs RNA 2019, 10, e1538. [Google Scholar] [CrossRef]
  208. Weissinger, R.; Heinold, L.; Akram, S.; Jansen, R.P.; Hermesh, O. RNA Proximity Labeling: A New Detection Tool for RNA-Protein Interactions. Molecules 2021, 26, 2270. [Google Scholar] [CrossRef] [PubMed]
  209. Zhang, J.; Hou, L.; Zuo, Z.; Ji, P.; Zhang, X.; Xue, Y.; Zhao, F. Comprehensive profiling of circular RNAs with nanopore sequencing and CIRI-long. Nat. Biotechnol. 2021, 39, 836–845. [Google Scholar] [CrossRef]
  210. Litke, J.L.; Jaffrey, S.R. Highly efficient expression of circular RNA aptamers in cells using autocatalytic transcripts. Nat. Biotechnol. 2019, 37, 667–675. [Google Scholar] [CrossRef]
  211. Gao, X.; Ma, X.K.; Li, X.; Li, G.W.; Liu, C.X.; Zhang, J.; Wang, Y.; Wei, J.; Chen, J.; Chen, L.L.; et al. Knockout of circRNAs by base editing back-splice sites of circularized exons. Genome Biol. 2022, 23, 16. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Biological roles of circRNAs. (a) Circular RNAs (circRNAs) can function as microRNA (miRNA) sponges, promoting their degradation and increasing the expression of target mRNAs. (b) CircRNAs with RNA-binding protein (RBP) binding sites have the potential to decoy these proteins and disrupt their functions. (c) Some circRNAs have been demonstrated to serve as protein scaffolds, allowing the colocalization of proteins (e.g., enzymes and their substrates). (d) CircRNAs containing an internal ribosome entry site (IRES) or N6-methyladenosine (m6A) and a start codon can be translated to proteins under certain conditions. (e) Additionally, circRNAs may attract certain proteins to a particular location in the cell. (f) CircRNAs can alter gene expression by interacting with RNA polymerase II (RNA Pol II). (g) Some circRNAs can affect the splicing of the host gene and influence the ratio between mRNAs and circRNAs. Abbreviation: Tet1: tet methylcytosine dioxygenase 1; U1 snRNP: U1 small nuclear ribonucleoprotein particle; EIcircRNA: exon-intron circRNAs; ecircRNA: exonic circRNA; ciRNA: circular intronic RNAs; UAP56: spliceosome RNA helicase DDX39B; URH49: ATP-dependent RNA helicase DDX39A.
Figure 1. Biological roles of circRNAs. (a) Circular RNAs (circRNAs) can function as microRNA (miRNA) sponges, promoting their degradation and increasing the expression of target mRNAs. (b) CircRNAs with RNA-binding protein (RBP) binding sites have the potential to decoy these proteins and disrupt their functions. (c) Some circRNAs have been demonstrated to serve as protein scaffolds, allowing the colocalization of proteins (e.g., enzymes and their substrates). (d) CircRNAs containing an internal ribosome entry site (IRES) or N6-methyladenosine (m6A) and a start codon can be translated to proteins under certain conditions. (e) Additionally, circRNAs may attract certain proteins to a particular location in the cell. (f) CircRNAs can alter gene expression by interacting with RNA polymerase II (RNA Pol II). (g) Some circRNAs can affect the splicing of the host gene and influence the ratio between mRNAs and circRNAs. Abbreviation: Tet1: tet methylcytosine dioxygenase 1; U1 snRNP: U1 small nuclear ribonucleoprotein particle; EIcircRNA: exon-intron circRNAs; ecircRNA: exonic circRNA; ciRNA: circular intronic RNAs; UAP56: spliceosome RNA helicase DDX39B; URH49: ATP-dependent RNA helicase DDX39A.
Cells 12 00378 g001
Figure 2. A diagram illustrating the regulatory networks of HSC activation modulated by anti-fibrotic circRNAs. Arrows represent activation, whereas bars symbolize inhibition. Abbreviation: COL1A1: collagen type I alpha 1 chain; CypD: Cyclophilin D; FBXW7: F-box/WD repeat-containing protein 7; HLF: hepatic leukemia factor; KAT2B: lysine acetyltransferase 2B; LEMD3: LEM domain containing 3; Lefty2: left-right determination factor 2; mPTP: mitochondrial permeability transition pore; PINK1: PTEN-induced kinase 1; PTEN: phosphatase and tensin homolog; TGF-β: transforming growth factor beta; TLR4: toll-like receptor 4.
Figure 2. A diagram illustrating the regulatory networks of HSC activation modulated by anti-fibrotic circRNAs. Arrows represent activation, whereas bars symbolize inhibition. Abbreviation: COL1A1: collagen type I alpha 1 chain; CypD: Cyclophilin D; FBXW7: F-box/WD repeat-containing protein 7; HLF: hepatic leukemia factor; KAT2B: lysine acetyltransferase 2B; LEMD3: LEM domain containing 3; Lefty2: left-right determination factor 2; mPTP: mitochondrial permeability transition pore; PINK1: PTEN-induced kinase 1; PTEN: phosphatase and tensin homolog; TGF-β: transforming growth factor beta; TLR4: toll-like receptor 4.
Cells 12 00378 g002
Figure 3. A diagram illustrating the regulatory networks of HSC activation modulated by pro-fibrotic circRNAs. Arrows represent activation, whereas bars symbolize inhibition. Abbreviation: COL1A1: collagen type I alpha 1 chain; HLF: hepatic leukemia factor; HDAC1: histone deacetylase 1; IL-6: interleukin 6; PTEN: phosphatase and tensin homolog; RAC1: Ras-related C3 botulinum toxin substrate 1; TGF-β: transforming growth factor beta; TLR4: toll-like receptor 4.
Figure 3. A diagram illustrating the regulatory networks of HSC activation modulated by pro-fibrotic circRNAs. Arrows represent activation, whereas bars symbolize inhibition. Abbreviation: COL1A1: collagen type I alpha 1 chain; HLF: hepatic leukemia factor; HDAC1: histone deacetylase 1; IL-6: interleukin 6; PTEN: phosphatase and tensin homolog; RAC1: Ras-related C3 botulinum toxin substrate 1; TGF-β: transforming growth factor beta; TLR4: toll-like receptor 4.
Cells 12 00378 g003
Table 1. Overview of the circular RNA (circRNA) associated with the activation of hepatic stellate cells (HSC).
Table 1. Overview of the circular RNA (circRNA) associated with the activation of hepatic stellate cells (HSC).
CircRNAsFunctionPathwayTarget miRNATarget ProteinMolecular MechanismRoleRef.
TGF-β signaling pathway
CircPSD3miRNA spongeTGF-β signaling pathwaymiR-92b-3pSmad7Inhibit TGF-β signal transduction by increasing the expression of Smad7, a negative regulator in TGF-β/Smad pathwayAnti-fibrotic[113]
CircCREBBPmiRNA spongeTGF-β signaling pathwaymiR-1291LEFTY2Inhibit TGF-β signal transduction by increasing the expression of LEFTY2, a negative regulator in the TGF-β/Smad pathwayAnti-fibrotic[114]
CircSCLT1miRNA spongeTGF-β signaling pathwaymiR-223-3pLEMD3Inhibit TGF-β signal transduction by increasing the expression of LEMD3, a negative regulator in TGF-β/Smad pathwayAnti-fibrotic[115]
CircMTO1miRNA sponge

miRNA sponge
TGF-β signaling pathway

PI3K/Akt signaling pathway
miR-17-5p

miR-181b-5p
Smad7


PTEN
Inhibit TGF-β signal transduction by increasing the expression of Smad7, a negative regulator in TGF-β/Smad pathway

Inhibit PI3K/Akt signal transduction by increasing the expression of PTEN, an inhibitory signaling molecule in the PI3K/Akt pathway
Anti-fibrotic

Anti-fibrotic
[116]


[117]
CircBNC2Protein templateTGF-β signaling pathway-ctBNC2Reduces G2/M cell arrest of hepatocytes, which inhibits the release of pro-fibrotic factors that can activate HSCs such as TGF-β and DAMPsAnti-fibrotic[118]
CircPWWP2AmiRNA sponge

miRNA sponge
TGF-β signaling pathway

TGF-β signaling pathway
miR-203


miR-223
Fstl1


TLR4
Promote TGF-β signal transduction by increasing the expression of Fstl1, which can bind to TGF-β receptors

Promote TGF-β signal transduction by suppressing the expression of BAMBI, a negative regulator in the TGF-β pathway
Pro-fibrotic

Pro-fibrotic
[119]

[119]
CircUbe2kmiRNA spongeTGF-β signaling pathwaymiR-149-5pTGF-β2Promote TGF-β signal transduction by increasing the expression of TGF-β2, a fibrogenesis mediatorPro-fibrotic[120]
CircTUBD1miRNA sponge

miRNA sponge
TGF-β signaling pathway

TGF-β and NF-κB signaling pathway
micro-203a-3p


miR-146a-5p
Smad3


TLR4
Promote TGF-β transduction by increasing the expression of Smad3 and regulating circTUBD1 biogenesis via a positive feedback loop

Promote TGF-β signal transduction by suppressing the expression of BAMBI, a negative regulator in the TGF-β pathway, and promoting NF-κB signaling pathway
Pro-fibrotic

Pro-fibrotic
[121]


[122]
JAK/STAT signaling pathway
CircCHD2miRNA spongeJAK/STAT signaling pathwaymiR-200b-3pHLFPromote JAK/STAT signal transduction by increasing the expression of HLF, which can promote IL-6 expression, thus increasing IL-6/JAK/STAT3 signaling pathwayPro-fibrotic[123]
PI3K/Akt signaling pathway
CircDIDO1miRNA spongePI3K/Akt signaling pathwaymiR-141-3pPTENInhibit PI3K/Akt signal transduction by increasing the expression of PTEN, an inhibitory signaling molecule in the PI3K/Akt pathwayAnti-fibrotic[124]
CircCDK13miRNA spongePI3K/Akt and NF-κB signaling pathwaymiR-17-5pKAT2BInhibit PI3K/Akt signal transduction by increasing the expression of PTEN, an inhibitory signaling molecule in the PI3K/Akt pathway, via inhibiting PTEN degradationAnti-fibrotic[125]
CircIFT80miRNA spongePI3K/Akt signaling pathwaymiR-155FOXO3Promote PI3K/Akt signal transduction by increasing FOXO3 expressionPro-fibrotic[126]
Wnt/β-catenin signaling pathway
Circ608miRNA spongeWnt/β-catenin and Hippo signaling pathwaymiR-222PINK1Inhibit Wnt/β-catenin signal transduction by promoting the degradation of β-catenin, and inhibit Hippo signal transduction by promoting the expression of a negative regulator in the Hippo signaling pathwayAnti-fibrotic[127]
Notch signaling pathway
CircFBXW4miRNA spongeNotch signaling pathwaymiR-18b-3pFBXW7Inhibit Notch signal transduction by promoting the degradation of NICDAnti-fibrotic[128]
Hedgehog signaling pathway
CircRSF1miRNA spongeHedgehog and NF-κB signaling pathwaymiR-146a-5pRAC1Promote Hedgehog signal transduction as well as NF-κB and JNK signaling by increasing the expression of RAC1Pro-fibrotic[129]
Other pathways
CircSCARProtein spongeROS ATP5BShutting down the mPTP, which interferes with ROS release and thus inhibits HSC activationAnti-fibrotic[130]
CircARID1AmiRNA sponge miR-185-3pCol1a1Directly promote the expression of Col1a1, a key indicator of HSC activationPro-fibrotic[131]
CircSMYD4miRNA sponge miR-660-3pTEP1Possibly by maintaining telomere length, which is found to be shorten in liver diseasesAnti-fibrotic[132]
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

Nokkeaw, A.; Thamjamrassri, P.; Tangkijvanich, P.; Ariyachet, C. Regulatory Functions and Mechanisms of Circular RNAs in Hepatic Stellate Cell Activation and Liver Fibrosis. Cells 2023, 12, 378. https://doi.org/10.3390/cells12030378

AMA Style

Nokkeaw A, Thamjamrassri P, Tangkijvanich P, Ariyachet C. Regulatory Functions and Mechanisms of Circular RNAs in Hepatic Stellate Cell Activation and Liver Fibrosis. Cells. 2023; 12(3):378. https://doi.org/10.3390/cells12030378

Chicago/Turabian Style

Nokkeaw, Archittapon, Pannathon Thamjamrassri, Pisit Tangkijvanich, and Chaiyaboot Ariyachet. 2023. "Regulatory Functions and Mechanisms of Circular RNAs in Hepatic Stellate Cell Activation and Liver Fibrosis" Cells 12, no. 3: 378. https://doi.org/10.3390/cells12030378

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