Next Article in Journal
Molecular and Clinical Epidemiology of SARS-CoV-2 Infection among Vaccinated and Unvaccinated Individuals in a Large Healthcare Organization from New Jersey
Previous Article in Journal
Interchangeability of the Assays Used to Assess the Activity of Anti-SARS-CoV-2 Monoclonal Antibodies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Functional Involvement of circRNAs in the Innate Immune Responses to Viral Infection

1
Key Laboratory of Animal Pathogen Infection and Immunology of Fujian Province, College of Animal Sciences, Fujian Agriculture and Forestry University, Fuzhou 350002, China
2
CAS Key Laboratory of Pathogenic Microbiology and Immunology, Institute of Microbiology, Chinese Academy of Sciences (CAS), Beijing 100101, China
3
Department of Virology, Faculty of Veterinary Medicine, Suez Canal University, Ismailia 41522, Egypt
4
Fujian Province Joint Laboratory of Animal Pathogen Prevention and Control of the “Belt and Road”, College of Animal Sciences, Fujian Agriculture and Forestry University, Fuzhou 350002, China
5
Department of Microbiology, ShiGan International College of Science and Technology/ShiGan Health Foundation, Narayangopal Chowk, Kathmandu 44600, Nepal
*
Author to whom correspondence should be addressed.
Submission received: 13 July 2023 / Revised: 2 August 2023 / Accepted: 3 August 2023 / Published: 5 August 2023
(This article belongs to the Section Viral Immunology, Vaccines, and Antivirals)

Abstract

:
Effective viral clearance requires fine-tuned immune responses to minimize undesirable inflammatory responses. Circular RNAs (circRNAs) are a class of non-coding RNAs that are abundant and highly stable, formed by backsplicing pre-mRNAs, and expressed ubiquitously in eukaryotic cells, emerging as critical regulators of a plethora of signaling pathways. Recent progress in high-throughput sequencing has enabled a better understanding of the physiological and pathophysiological functions of circRNAs, overcoming the obstacle of the sequence overlap between circRNAs and their linear cognate mRNAs. Some viruses also encode circRNAs implicated in viral replication or disease progression. There is increasing evidence that viral infections dysregulate circRNA expression and that the altered expression of circRNAs is critical in regulating viral infection and replication. circRNAs were shown to regulate gene expression via microRNA and protein sponging or via encoding small polypeptides. Recent studies have also highlighted the potential role of circRNAs as promising diagnostic and prognostic biomarkers, RNA vaccines and antiviral therapy candidates due to their higher stability and lower immunogenicity. This review presents an up-to-date summary of the mechanistic involvement of circRNAs in innate immunity against viral infections, the current understanding of their regulatory roles, and the suggested applications.

1. Introduction

The human genome project showed that less than 2% of the human genome comprises protein-coding genes [1]. Nevertheless, most genomic DNA is a transcription template, indicating that the human transcriptome predominantly contains non-coding RNAs (ncRNAs) [2,3]. ncRNAs have been demonstrated to play crucial roles in the regulation of gene expression by impacting target genes’ transcription or post-transcriptional modifications. ncRNAs can be divided into two major categories: small ncRNAs (sncRNAs) and long ncRNAs (lncRNAs) [4]. sncRNAs are shorter than 200 nucleotides (nt) in length, and these include piwi-interacting RNAs, microRNAs (miRNAs), transcription initiation RNAs, and endogenous small interfering RNAs [5]. LncRNAs are longer than 200 nt in length and constitute most of the non-coding transcriptome in mammals [6,7].
Circular RNAs (circRNAs) were initially discovered in RNA viruses such as the Sendai virus, hepatitis D virus (HDV), and plant viroids in the 1970s [8,9,10,11]. Nevertheless, circRNAs were thought to be viral genomes or the results of pre-mRNA alternative splicing. Hence, they received little attention in the past [12]. Current advances in high-throughput sequencing technologies have enabled scientists to undertake comprehensive investigations of the structure, expression profile, and functions of circRNAs, as well as the mechanisms underlying their roles [13]. CircRNAs have been discovered in a wide range of plant and animal species [14]. Some circRNAs are evolutionarily conserved across related species [15]. The extensive presence of circRNAs in eukaryotic cells indicates that circRNAs are not only unintended products of RNA splicing events but rather an essential component of the ncRNA family [16].
CircRNAs are now established ncRNA family members formed via backsplicing introns, exons, or both [17]. CircRNAs are distinguished by their closed circular structure, and lack of 5′ N7-methylguanosine (m7G) cap ends or 3′ Poly (A) tails [18]. CircRNAs are a highly stable ncRNA form because they are naturally resistant to RNA nucleases [19]. Recently, the physiological and pathological roles of circRNAs have drawn researchers’ attention, as an increasing number of data demonstrate that circRNAs can serve various functions through multiple mechanisms [20]. It has been shown that circRNAs play numerous roles in cellular processes, such as the modulation of gene expression and alternative splicing, sponging to miRNAs or proteins, providing translation templates, rRNA and tRNA synthesis modulators, and so on [21]. Remarkably, many circRNAs exhibit altered expression levels in response to specific disease conditions or infections with pathogens, implying a link between circRNAs and the emergence and progression of human and animal disorders [22].
Interestingly, despite the significant focus on the study of the relationship between circRNAs and cancer, multiple reports have also suggested the involvement of circRNAs in innate immunity against viral infection. Typically, viral infection dysregulates the expression of circRNAs, which, in turn, could regulate viral replication by modulating innate immunity, providing novel insights into the diagnosis and treatment of viral infectious diseases. Here, in this review, we will discuss the taxonomy, biosynthesis, functions, and mechanism underlying the action of circRNAs and highlight their relevance to antiviral immunity and potential applications, such as antiviral therapeutics, vaccine candidates, and diagnostic and prognostic biomarkers.

2. Physical and Chemical Characteristics of Circular RNAs

2.1. RNA Circularization

Unlike conventional linear splicing, which forms a linear 5′ to 3′ mRNA of joined exons, circRNAs feature a covalently closed structure lacking the 5′-cap and 3′-poly(A) tail [23]. The conventional splicing patterns of circRNAs, known as exon-skipping and backsplicing [17], have been observed both in vivo and in vitro. Nevertheless, evidence suggests that backsplicing is more significant, as this pattern is widely reported [16].
Based on origin and composition, circRNAs may be categorized into three categories: circular intronic RNAs (ciRNAs), exonic circRNAs (ecircRNAs), or exon-intron circRNAs (eiciRNAs) [19]. EcircRNAs, composed of one or multiple exons, constitute more than 80% of the detectable circRNAs and are preferentially generated by intron pairing-driven circularization or lariat-driven circularization [24,25]. The flanking introns adjacent to the backspliced exons are crucial for circRNA biogenesis, as they contain complementary sequences that base-pair and form hairpin-like structures, allowing the 5′ and 3′ splicing sites to be closer for circularization to occur [26]. These sequences may consist of Alu repeats of ~300 nt in length or non-repetitive elements [26]. Alu repeats-driven circularization is often complex as the inverted Alu repeats may pair across introns and induce different exon circularization events; consequently, one single gene locus can produce various circRNAs [18,26].
EiciRNAs utilize an exon-skipping strategy for circularization, and contain flanking intron sequences on the exonic core sequence’s off-side [27]. CiRNAs are generated via a lariat-derived process that requires a consensus GU-rich domain close to the 5′-end splicing site and a C-rich domain close to the breakpoint, and then the uncircularized intron sequences are then sequestered [23]. CircRNAs are predominantly cytoplasmic, but ciRNAs and eiciRNAs are exclusively nuclear, suggesting roles in nuclear processes, such as transcriptional regulation [28].
Since circRNAs and their cognate mRNA share the same pre-mRNA and most splicing sites, circRNA backsplicing and mRNA splicing often compete for transcription against each other [8,29]. Nevertheless, due to unfavorable spliceosome assembly at backsplicing sites, the efficiency of backsplicing is significantly lower than that of canonical splicing [30].

2.2. circRNAs versus mRNAs

circRNAs exhibit the same nucleotide sequence as their corresponding linear RNAs, except for the back-splicing junction (BSJ) site. Thus, distinguishing the expression of a particular circRNA from its linear cognate has been challenging on multiple scales, including identification, validation, and loss- and gain-of-function investigations [19]. Nevertheless, the BSJ site enables the utilization of divergent primers to detect specific circRNAs while avoiding the undesired signal of their cognate linear mRNAs [31].
In addition, most circRNAs are stable, with half-lives ranging between 18.8 and 23.7 h [30,32], which is much longer than the range of 4.0–7.4 h that their linear RNA cognates have [32]. This stability is presumably a result of their resistance to linear RNA degradation machinery. Consequently, in slowly dividing or nondividing cells and tissues, certain circRNAs could accumulate to high levels [19,33]. However, it is suggested that following viral infection or poly(I:C) stimulation, the RNase L could degrade the transcribed circRNAs, a mechanism necessary for early innate immune responses [34]. Notably, after being transcribed, circRNAs could undergo N6-methyladenosine (m6A) modifications [35]. m6A-circRNAs can then be recognized by YTHDF2 and HRSP12, resulting in circRNA degradation by the RNAse P/MRP ribonuclease complex [36]. Some circRNAs are degraded after perfect sponging to miRNA via the AGO2-mediated cleavage of circRNA [37]. Additionally, another degradation mechanism is that of the G3BP1 endonuclease, as it complexes with the RNA-binding protein UPF1 and decays highly structured circRNAs [38].
Another functionally important distinction between circRNAs and their cognate mRNAs is immunogenicity [39,40,41]. Whereas self-circRNAs are not immunogenic, the RNA pattern recognition receptors, Toll-like receptor 7/8 and retinoic-acid-inducible gene-I (RIG-I), may be activated by exogenous circRNAs, making the circRNA itself immunostimulatory [41]. The immunogenicity of circRNA could be unfavorable to the circRNAs’ translation efficiency, their half-life stability, and biomedical applications. It has been demonstrated that impure circRNA formulas elicit potent immunological responses in cells [42,43,44].

2.3. Mechanisms Underlying Action of circRNAs

Studies have established that circRNAs play crucial roles in the modulation of physiological processes and the progression of numerous diseases [45]. In the past few years, the mechanisms underlying the action of circRNAs, such as the regulation of gene expression, microRNA reservoirs, and microRNA sponges, the ability to encode proteins, etc., have been progressively reported [19].
MicroRNA response elements (MREs) are miRNA-complementary sequences in the 3′-UTRs of target mRNAs [46]. Similarly, circRNAs, primarily cytoplasmic, include several MREs that may bind to target miRNAs, thereby functioning as competitive endogenous RNAs (ceRNAs) [47]. Additionally, lacking the 5′-cap or 3′-poly(A) tail, circRNAs are resistant to RNase degradation, enabling them to serve as potent sponges for miRNA [48]. On the other hand, some circRNAs possess the capacity to stabilize or stimulate the actions of miRNAs, a property known as miRNA reservoirs [49].
Moreover, circRNAs can bind to specific proteins called RNA-binding proteins (RBPs), which serve as protein sponges [19]. RBPs exhibit complementary sequences to bind with their target RNA, and circRNAs can modulate protein activity by docking to the active sites of RBPs [50].
Of note is that in 2015, circRNAs were reported in fruit flies to encode translatable proteins and peptides [51]. This discovery established the theoretical foundation that paved the way for circRNA vaccine development [52]. Multiple reports have demonstrated that, unlike conventional mRNAs, circRNAs initiate translation not at the 5′ cap but rather at the m6A-induced ribosome engagement site or the internal ribosomal entry site (IRES) [53,54,55].

3. Involvement of circRNAs in Innate Immunity against Viral Infection

Innate immunity provides the first line of defense against infections with pathogens [56,57], and it possesses a crucial function in virus recognition and the subsequent activation of the adaptive immune response. Pathogen-associated molecular patterns (PAMPs) are viral components that can be sensed by the innate immune molecules’ pathogen recognition receptors (PRRs). PAMPs include, for instance, viral double-stranded RNA, viral single-stranded RNA, and viral DNA. PPRs, such as RIG-I, Toll-like receptors, and nucleotide oligomerization domain (NOD)-like receptors, are indispensable for the activation of innate immune signaling that ultimately induces the production of various cytokines and antiviral molecules, including interferons (IFNs) [56,58] (Figure 1).
Interestingly, recent studies have demonstrated the mechanistic involvement of circRNAs in viral infections, the regulation of innate immune responses, and other biological processes of antiviral immunity [31,59]. Numerous virus- and host-derived circRNAs are shown to regulate antiviral immune responses [21,60,61,62].
Figure 1. Mechanistic involvement of some circRNAs in innate immunity against some representative viral infections. As reviewed by Rai et al. [57], virus detection by PRRs triggers innate immune signaling via the activation of specific adaptor proteins (MYD88, MAVS, TRIF, STING, etc.) to subsequently activate other transcriptional factors, including NF-κB, IRF3/5/7, and others. The translocation of activated transcriptional factors into the nucleus induces the expression of IFNs. The mechanistic involvement of several circRNAs in the modulation of immune responses has been revealed. For instance, circ_0000479 sponged miR-149-5p and regulated RIG-I expression, thus impacting HTNV and SARS-CoV2 viral replication [63,64]. CircRNA AIVR inhibited IAV replication by predominantly absorbing miR-330-3p and enhancing CREBBP expression, thus facilitating the production of IFN-β [59]. CircSIAE suppressed CVB3 replication by targeting miR-331-3p and TAOK2, and impacting the levels of p-NF-κB [65]. CircEZH2 promoted the activation of NF-κB via sponging miR-22 after TGEV infection [66,67]. HTNV: Hantaan virus; HBV: Hepatitis B virus; TGEV: Transmissible Gastroenteritis Virus; IAV: Influenza A virus; CVB3: Coxsackievirus B3. Created with BioRender.com (accessed on 4 August 2023).
Figure 1. Mechanistic involvement of some circRNAs in innate immunity against some representative viral infections. As reviewed by Rai et al. [57], virus detection by PRRs triggers innate immune signaling via the activation of specific adaptor proteins (MYD88, MAVS, TRIF, STING, etc.) to subsequently activate other transcriptional factors, including NF-κB, IRF3/5/7, and others. The translocation of activated transcriptional factors into the nucleus induces the expression of IFNs. The mechanistic involvement of several circRNAs in the modulation of immune responses has been revealed. For instance, circ_0000479 sponged miR-149-5p and regulated RIG-I expression, thus impacting HTNV and SARS-CoV2 viral replication [63,64]. CircRNA AIVR inhibited IAV replication by predominantly absorbing miR-330-3p and enhancing CREBBP expression, thus facilitating the production of IFN-β [59]. CircSIAE suppressed CVB3 replication by targeting miR-331-3p and TAOK2, and impacting the levels of p-NF-κB [65]. CircEZH2 promoted the activation of NF-κB via sponging miR-22 after TGEV infection [66,67]. HTNV: Hantaan virus; HBV: Hepatitis B virus; TGEV: Transmissible Gastroenteritis Virus; IAV: Influenza A virus; CVB3: Coxsackievirus B3. Created with BioRender.com (accessed on 4 August 2023).
Viruses 15 01697 g001

3.1. Host-Coded circRNAs Involved in Immune Responses

The dysregulation of cellular gene expression occurs in responses to multiple stimuli, including infections [68]. Among them, viral infections have been reported to induce a differential expression of host circRNAs, potentially enhancing or inhibiting the innate immune response or impacting viral pathogenesis. Although the precise mechanisms underlying the involvement of circRNAs in antiviral innate and adaptive immune responses and viral pathogenesis are elusive, several studies have revealed their significance. For example, multiple reports have revealed various differentially expressed host circRNAs and established a circRNA–miRNA–mRNA regulatory network in immune responses to infections with multiple viruses, including Hantaan Virus, Human Immunodeficiency Virus, Coxsackievirus B5, Coxsackievirus A16, Rabies Virus, Peste-Des-Petits-Ruminants Virus, Japanese Encephalitis Virus, Porcine Endemic Diarrhea Virus, Pseudorabies Virus Type II, Foot-And-Mouth Disease Virus, Human Papillomavirus E7, Avian Leukemia, Influenza Virus, Infectious Bursal Disease Virus, Enterovirus A71, Swine Hepatitis E Virus, Or Middle East Respiratory Syndrome Coronavirus, SARS-CoV-2 virus, and so on [63,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87]. In addition, plant viruses are also reported to alter circRNAs expression. For instance, Maize Iranian mosaic virus (MIMV) induced the dysregulation of circRNAs expression in MIMV-infected maize [88]. Nevertheless, the precise functions of the altered circRNAs during viral infections are still the subject of ongoing research.
Various circRNAs have been implicated in host–virus interactions. Here, we present a brief overview of the most up-to-date comprehension of the roles of circRNAs in modulating immune responses and impacting the pathogenesis of viruses with medical or veterinary importance (Table 1 and Figure 1).

3.1.1. Hepatitis B Virus (HBV)

HBV is an exclusively hepatotropic virus that can cause persistent infections (chronic hepatitis B (CHB)) and, in extreme cases, cirrhosis and hepatocellular cancer (HCC). It was shown that hsa_circ_0004812 was highly upregulated in HCC tissues from CHB patients. The robust expression of hsa_circ_0004812 enhanced HBV-induced immunosuppression by sponging to miR-1287-5p, whereas knocking down hsa_circ_0004812 promoted interferon (IFN)-α/β production and suppressed viral propagation and multiplication, suggesting a putative HBV therapeutic target [90].
It was also found that circ-ATP5H was upregulated in HCC-HBV-infected tissues. Circ-ATP5H knockdown impacted HBV replication through sponging miR-138-5p. Circ-ATP5H modulates TNFAIP3 by binding to miR-138-5p. Interestingly, circ-ATP5H boosted HBV multiplication by altering the miR-138-5p/TNFAIP3 axis, revealing a potential novel biomarker for HBV-positive HCC therapy [92].
In addition, HCC tissues, HepG2, and Huh7 cell lines showed elevated levels of circBACH1 and MAP3K2 and diminished levels of miR-200a-3p. CircBACH1 depletion or miR-200a-3p overexpression significantly impaired HBV replication. Studies revealed that circBACH1 governs HBV replication via the miR-200a-3p/MAP3K2 pathway [91].

3.1.2. Influenza Virus

Influenza is a contagious respiratory disease caused by the influenza A (IAV) and B viruses. Yu et al. [95] reported that following IAV H1N1 infection, A549 cells showed a dramatic increase in circ-GATAD2A expression. It was observed that circ-GATAD2A overexpression facilitated H1N1 replication via the inhibition of VPS34-dependent autophagy, whereas circ-GATAD2A silencing reduced H1N1 titers [95]. It was also reported that circRNA_0050463 sponged to miR-33b-5p, therefore promoting the expression of eukaryotic translation elongation factor 1 alpha 1 and boosting the replication of IAV [94]. In addition, circRNA AIVR was demonstrated to inhibit IAV multiplication principally by sponging to miR-330-3p, inhibiting its binding to the mRNA of the CREB-binding protein, thereby accelerating IFN production (Figure 1) [59].
Recently, circMertk was reported by Qiu et al. [31] as a novel circRNA derived from pre-MerTK. Interestingly, the overexpression or silencing of circMerTK enhanced or inhibited the replication of the IAV and Sendai viruses, respectively. CircMerTK silencing stimulated the secretion of type I IFNs and the expression of interferon-stimulating genes, while the robust expression of circMerTK impaired their expression at both mRNA and protein levels [31]. Additionally, in another report, RNA-seq of A549 cells in response to avian influenza (AIV) or IAV infections led to the identification of multiple sets of altered circRNAs’ expression, and the authors selected six circRNAs (hsa_circ_0005870, hsa_circ_0006104, hsa_circ_0009609, hsa_circ_0060300, hsa_circ_0009365, and hsa_circ_0003428) for further analysis [96]. The authors suggested that the selected circRNAs may influence the cell cycle process and the endocytosis pathway via an in silico-established ceRNA network [96].

3.1.3. Middle East Respiratory Syndrome Coronavirus (MERS-CoV)

MERS-CoV is an extremely pathogenic zoonotic virus that was first reported in humans in Saudi Arabia and Jordan in 2012, with a high mortality rate and unpredictable incidence [100]. Interestingly, Calu-3 cells infected with MERS-CoV showed elevated circFNDC3B and circCNOT expression levels. Further studies showed that silencing these particular circRNAs decreased the cellular viral load and indirectly influenced MAPK signaling pathways [69]. Additionally, another study observed that the inhibition of hsa_circ_0004445 significantly reduced MERS-CoV replication by 50–75% via binding to hnRNP C [93].

3.1.4. Hepatitis C Virus (HCV)

HCV, a member of the Flaviviridae family, is the causative agent of hepatitis C, a worldwide health concern with an estimated 7.1 million people chronically infected with HCV [101]. HCV viral infectivity was significantly diminished after the depletion of circTIAL and circEXOSC in HCV-infected cells [97]. Additionally, the upregulation of circPSD3 in response to hepatitis C virus infection was reported to impact HCV viral abundances significantly, thus acting as a proviral factor in the post-translational regulation of HCV RNA amplification [97].

3.1.5. Transmissible Gastroenteritis Coronavirus (TGEV)

Transmissible gastroenteritis is an infectious disease in swine, particularly piglets, characterized by severe vomiting and diarrhea caused by TGEV [66]. CircEZH2 was found to be downregulated after TGEV infection. CircEZH2 promoted NF-κB activation by targeting miR-22 in intestinal porcine enterocyte (IPEC-J2) cells (Figure 1). The mitochondrial permeability transmission pore (mPTP) opening of IPEC-J2 was induced by TGEV infection and repressed by miR-22 sponging to circEZH2. Interleukin 6 (IL-6) and hexokinase 2 (HK2) were identified as miR-22 targets. TGEV-induced mPTP opening might be controlled by two pathways: the circEZH2/miR-22/IL-6/NF-κB axis and the circEZH2/miR-22/HK2 axis [66,67].

3.1.6. Ebola Virus (EBOV)

EBOV causes severe and typically fatal Ebola virus disease (EVD) [102]. Wang et al. [103] demonstrated that circ-chr19 boosted the claudin-18 (CLDN18) expression by sponging miR-30b-3p, thereby serving as a ceRNA during EBOV infection [103]. Since CLDN18 regulates cellular permeability, the basis of EBOV pathogenesis, circ-chr19 could be a promising target for EVD therapy.

3.1.7. Human Immunodeficiency Virus (HIV)

Over 75 million individuals have contracted HIV worldwide. Untreated HIV infections are associated with progressive CD4+ T cell depletion and numerous immunological abnormalities [104]. Interestingly, the expression pattern and function of circRNAs in the pathophysiology of HIV were analyzed in peripheral blood mononuclear cells obtained from early HIV-infected patients [87]. The authors identified 15,145 unique circRNA transcripts, and the circRNA–miRNA–mRNA network uncovered that dysregulated circRNAs contributed to HIV-1 multiplication by regulating the expression of CCNK, CDKN1A, and IL-15 genes. Previous studies proved that circRNAs played a role in HIV replication and indicated their potential therapeutic application [87].

3.1.8. Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2)

SARS-CoV-2 is a highly pathogenic and contagious coronavirus that evolved in 2019 and resulted in a global pandemic of acute respiratory illness [52]. A recent study by Firoozi et al. [64] has suggested that hsa_circ_0000479, induced by SARS-CoV-2, may regulate the immune response to SARS-CoV-2 by binding to hsa-miR-149-5p and influencing the expression of IL-6 and RIG-I (Figure 1) [64].

3.1.9. Kaposi Sarcoma-Associated Herpesvirus (KSHV)

KSHV, or Human Herpesvirus-8, is the causative agent of Kaposi sarcoma [105]. According to several reports, KSHV, Epstein–Barr virus (EBV), and human cytomegalovirus induced hsa_circ_0001400 expression. The upregulation of hsa_circ_0001400 promoted the expression of tumor necrosis factor-alpha and diminished the production of two viral genes (replication and transcription activator and latency-associated nuclear antigen), which govern latent and lytic infection, respectively [89]. hsa_circ_0001400 promoted the cell cycle, suppressed apoptosis, and activated multiple immune genes’ expression (TAP2, ICAM1, CD40, etc.). hsa_circ_0001400’s capacity to impact viral genes implies that it may exhibit potential antiviral properties, given that numerous viruses trigger circ_0001400 expression, substantiating its importance [106].

3.1.10. Herpes Simplex Virus Type 1 (HSV-1)

HSV-1 is a herpesvirus causing vesicular eruptions, most commonly in the orolabial and genital mucosa [107]. The analysis of top dysregulated circRNAs and their circRNA-miRNA–mRNA regulatory axis after HSV-1 infection (circRNA14189, circRNA14556, circRNA15053, and circRNA15655) revealed that a substantial number of genes related to immunity in the NOD-like receptor/JAK-STAT signaling pathways could be governed by HSV-1-induced circRNAs [107].

3.1.11. Non-Mammalian Viruses

The majority of the studies on circRNAs focus on mammalian circRNAs, but numerous reports have addressed that non-mammalian species can also encode circRNAs that are implicated in immunity against viral infections. For instance, in teleost fish infected with Siniperca chuatsi rhabdovirus, circBCL2L1 was found to be significantly upregulated [108]. CircBCL2L1 could act as a ceRNA, enhancing the innate immune response by sponging to miR-30c-3-3p and influencing TRAF6, thereby inducing NF-κB/IRF3-mediated innate immunity and inflammatory pathways [108].

3.2. Virus-Coded circRNAs Affecting Innate Immunity

Circular RNAs were first discovered in a small number of viruses in the 1970s [11]. Since then, multiple DNA and RNA viruses have been shown to encode viral circRNAs [109,110,111,112], and several reports have investigated the crucial role of virus-coded circRNAs in the intricate virus–host interaction. Nonetheless, the precise biological roles of virus-coded circRNAs are still elusive [113,114].
For instance, circBART2.2, an EBV-encoded circRNA, was reported to be significantly upregulated in nasopharyngeal carcinoma (NPC), where it upregulates programmed death-ligand 1 (PD-L1) expression levels and inhibits T-cell functions in vivo and in vitro [115]. circBART2.2 assisted in immune escape by binding to the RIG-I helicase domain and activating transcription factors NF-κB and IRF3, resulting in increased PD-L1 transcription and the inhibition of the activation of the effector T lymphocytes [115].
In addition, it has been recognized that circRNAs generated by SARS-CoV-1, SARS-CoV-2, and MERS-CoV induce the expression of genes linked to mRNA processing and splicing during the initial stages of viral infection. In contrast, late-stage circRNAs modulate genes implicated in metabolism, autophagy, cancer, and viral infection [112].
It has also been observed that Merkel cell polyomavirus (MCV) expresses multiple circRNAs [116]. Of note, is that circMCV-T, the most highly expressed MCV-circRNA, modulated MCV replication by sponging to the miR-M1-loaded RISC complex, thus stabilizing the transcripts of linear T-Ag, and enhancing the expression of T-Ag. The depletion of circMCV-T was accompanied by the degradation of T-Ag linear transcripts via the miR-M1-induced RISC complex, inhibiting T-Ag production and, consequently, affecting MCV replication [116].
Additionally, some non-mammalian viruses have been shown to encode circRNAs. For example, Bombyx mori cypovirus (BmCPV) is an RNA virus affecting silkworms that can cause developmental retardation and severe economic losses [117]. BmCPV was reported to encode circRNA-vSP27, which can translate a small viral peptide, vSP27, that activates NF-κB signaling, suppressing BmCPV infection [108]. Bombyx mori Nucleopolyhedrovirus (BmNPV) is a critically important virus in silkworms, causing a significant economic impact on the silk production industry [118]. BmNPV was found to express multiple circRNAs; of them, circRNA-000010 could encode a small viral peptide termed VSP39 that acts as a proviral agent, boosting BmNPV virus replication [119]. Additionally, gibel carp (Carassius gibelio) is a uniquely important globally cultured freshwater fish species [120]. Cyprinid herpesvirus 2 (CyHV-2), which causes gill hemorrhagic disease and severe mortalities to gibel carps, encodes circ-udg, which can promote CyHV-2 proliferation and propagation [121].

4. Practical Applications of Circular RNAs

circRNAs differ from linear RNAs in conformation, immunogenicity, and stability. In particular, circRNAs are arguably more stable and less immunogenic than are other types of linear RNAs. These characteristics render circRNAs superior to linear RNA for practical applications. Therefore, several efforts have been made to create circRNA-based formulas, including non-coding aptamers, antisense RNAs, templates for sustained translation, modulators of innate immune responses and miRNAs, diagnostic and prognostic biomarkers for viral infections, and vaccine candidates [19].
The immunogenicity of synthetic circRNAs must be assessed and, if necessary, tailored for optimal biomedical uses. Recently, self-circRNAs were demonstrated to be commonly programmed by introns and paired with RBPs, denoting their origins. Nonself-circRNAs are distinguishable, and RIG-I-mediated signaling triggers antiviral immune responses after sensing them [122]. Breuer et al. [123] demonstrated that synthetic circRNAs might be used as miRNA sponges if they are generated in a cell-free system via in vitro transcription and ligation and purified via gel extraction [123].

4.1. Vaccine Candidates

Being chemically stable and less immunogenic than linear mRNAs [42,124], it has been suggested that synthetic and adeno-associated viral-based translatable circular RNAs can be incorporated into circRNA-based therapies [42,125,126,127]. Additionally, lipid nanoparticles (LNPs) are employed to deliver circRNA vaccines and therapeutics [42,52]. In mouse adipose cells and tissues, the nano-formulated administration of unmodified IRES-containing circRNAs improved the length of translation duration relative to that of linear mRNAs [42,124]. Of note is that circRNA-LNPs demonstrated significantly greater thermostability than did linear mRNA-LNPs [52]. Unlike current mRNA vaccines, necessitating strict transportation and storage conditions [128], circRNA vaccines encapsulated in LNP can be effectively kept for four weeks at 4 °C and up to two weeks at room temperature [52].
Intriguingly, it was hypothesized that the circRNA vaccine could exploit the observation that non-self circRNAs can activate RIG-I and PKR signaling pathways to function as a self-adjuvant and further boost the immune response induced by vaccination [42,43,115]. Nevertheless, its immunogenicity must be optimized to maintain the desired adjuvant effect without inducing adverse reactions that could significantly reduce circRNA vaccine efficacy [129].
circRNA vaccines expressing the receptor-binding domain (RBD) of the SARS-CoV-2 spike protein demonstrated sustained antigen production, cellular and humoral immune responses, neutralizing antibody formation in mice and monkeys, and distinct Th1-biased immune responses (Figure 2A) [52,55,130]. These recent findings substantiate the therapeutic benefits of circRNAs for sustained expression. Nevertheless, a number of crucial concerns must first be resolved, including IRES optimization and other strategies facilitating cap-independent translation [127,131], as well as reducing the cellular immune responses induced by the administration of the circRNA vaccine [43,55].

4.2. Therapeutic Agents

circRNAs can absorb miRNAs, reducing their bioaccessibility and activity against their intended mRNA targets. This led to the introduction of ectopically produced or in vitro-generated circular RNAs expressing partial MREs sites, enabling a reduction in disease- or virus-related miRNA activity in vitro and in vivo [48,136], giving a potential therapeutics alternative to the existing gold standard, antagomirs [137].
The HCV functional sequestration of miRNA-122 in cells is an excellent example of this approach. miRNA-122 is critical for HCV replication and propagation as it binds to the HCV RNA 5′-end, stabilizing and protecting the HCV genome from nucleolytic degradation and boosting HCV viral replication. Miravirsen, a locked nucleic acid (LNA)-modified DNA phosphorothioate antisense oligonucleotide complementary to miRNA-122, currently undergoing clinical trials, disrupts miRNA-122’s protective role on HCV RNA [138]. Briefly, 5′-3′-end-ligation using T4 RNA ligase 1 was used to generate a circular RNA sponge containing four miRNA-122 binding sites, thus sequestering miRNA-122 and suppressing HCV viral protein synthesis more effectively than did Miravirsen (Figure 2B) [48].
Additionally, SARS-CoV-2 viral replication was significantly suppressed in a cell culture by a panel of antisense circular RNAs that were designed to target the structurally conserved 5′-UTR of the virus’s genomic RNA [132]. Antisense-circRNAs targeting distinct regions of the 5′-UTR of SARS-CoV-2 efficiently inhibited virus replication by up to 90% compared to the control, and the durability was at least 48 h (Figure 2B) [132].
Additionally, two interesting Chinese herbs (Oldenlandia diffusa (Willd.) and Scutellaria barbata D.Don (SB)) were shown to dramatically reduce HBV activity, and HCC growth, migration, and invasion both in vitro and in vivo [139]. This activity may have been due to the modulation of the circRNA–miRNA–gene expression network [139].

4.3. Viral Infection Biomarker

circRNAs are detected as enriched detectable components in exosomes [140,141] and as a component of physiological body fluids (such as saliva, blood, and urine); they are also prevalent in peripheral tissues [142,143,144,145,146]. These properties make circRNAs stable and resistant to environmental fluctuations, rendering them potential biomarkers for detecting various diseases and infections [147].
For instance, diagnosing NPC, a disease typically caused by EBV, can be challenging; consequently, early detection can benefit therapeutic management. In this context, host hsa_circRNA_001387 and EBV-encoded circRPMS1 are significantly expressed in EBV-positive NPC samples. Hence, EBV-circRPMS1 and cellular hsa_circRNA_001387 could be valuable biomarkers for the diagnosis and prognosis of NPC [133,134]. Additionally, it was reported that the overexpression of circEAF2 in EBV-positive B lymphoma cells induces cell apoptosis and sensitizes lymphoma cells to epirubicin. circEAF2’s preferential target is the EBV-encoded miR-BART19-3p, which upregulates the tumor suppressor adenomatous polyposis coli (APC) and inhibits downstream β-catenin production, leading to the inactivation of the Wnt signaling pathway and suppression of EBV and DLBCL cell proliferation. CircEAF2 impacted the miR-BART19-3p/APC/β-catenin axis, and consequently inhibited EBV and large B-cell lymphoma progression, indicating that it is a potential prognostic biomarker (Figure 2C) [99].
Recently, four circRNAs (hsa_circ_0026579, hsa_circ_0018429, hsa_circ_0099188, and hsa_circ_0125357) were shown to be very sensitive and specific biomarkers for diagnosing community-acquired pneumonia (CAP). Interestingly, hsa_circ_0026579 was proposed as a circRNA biomarker that can distinguish the causative agent of CAP to be either viral/bacterial or mixed infection (Figure 2C) [135]. In addition, He et al. [148] reported that patients with dengue fever showed considerable upregulation of hsa_circ_0015962 and significant downregulation of hsa_circ_0006459. The upregulation of hsa_circ_0015962 and downregulation of hsa_circ_0006459 influence the therapeutic response to dengue fever and are promising biomarkers in dengue fever patients [148].

5. Summary

The biological functions of circRNAs have drawn the scientific community’s attention in the past years. Importantly, recent research has revealed that circRNAs could play a role in inducing or dampening antiviral immunity. It has been shown that circRNAs potentially regulate the expression of genes implicated in innate immunity, serving as either antiviral or proviral host factors. These findings contribute to the ever-growing comprehension of physiological and pathophysiological functions of such ncRNAs. However, our current knowledge and advancements in circRNA research are limited. For instance, the mammalian transcriptome contains a vast number of circRNAs with unknown functions that remain to be determined. Moreover, during circRNA synthesis, the backsplicing of pre-mRNA is frequently accompanied by alternative splicing. How the splicing machinery decides and chooses between RNA splicing, alternative splicing, and backsplicing to generate circRNAs is poorly understood. Consequently, it is essential to have a thorough understanding of the processes underlying circRNA synthesis and degradation.
Although progress has been made in the understanding of circRNAs’ involvement in the innate immune response to viral infection, the exact mechanism of how they regulate innate immunity is still unclear. Extensive studies of circRNA expression and function in response to viral infections may provide a solid basis for a better understanding of regulatory networks that protein-centric research might have underestimated. CircRNAs could potentially serve as proper biomarkers for a number of infections and diseases, therapeutic agents, and vaccine candidates, but these need to be further investigated. One of the challenges confronting circRNAs research is the lack of a well-established in vivo system for depleting circRNAs of interest. Since circRNAs share the same pre-mRNA with their cognate mRNA, it is challenging to alter the expression of only circRNAs without influencing the expression of the linear mRNA. Limited success has been achieved in developing some specific circRNA-depleted animal models. However, establishing a methodology to knock out merely circRNA expression effectively is an ongoing task. The development of such an in vivo system would be a breakthrough in investigating the biogenesis and functions of circRNAs in vivo.

Author Contributions

Conceptualization, J.-L.C. and M.M.; writing—original draft preparation, M.M.; review and editing, M.M., K.R.R., L.W., Y.W., Y.C., M.F. and J.-L.C.; supervision, critical comments, and suggestions, and manuscript revision, J.-L.C.; funding acquisition, J.-L.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by National Natural Science Foundation of China (32030110), National Key Research and Development Program of China (2021YFD1800205), National Natural Science Foundation of China (32102688), and the Chinese Academy of Sciences President’s International Fellowship for Postdoctoral Researchers, grant no. 2021PB0039.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

APCAdenomatous polyposis coli
BmCPVBombyx mori cypovirus
BmNPVBombyx mori Nucleopolyhedrovirus
BSJBacksplice junction
Calu-3Human lung adenocarcinoma cells
CAPCommunity-acquired pneumonia
CCNKCyclin-K
ceRNAsCompeting endogenous RNAs
CHBChronic Hepatitis B
circRNAsCircular RNAs
ciRNAsCircular intronic RNAs
CLDN18Claudin 18
COVID-19Coronavirus Disease 2019
CREBBP CREB-binding protein
CVB3Coxsackievirus B3
CVB3Coxsackievirus B3
CyHV-2Cyprinid herpesvirus 2
DENVDengue virus
DLBCLDiffuse Large B-Cell Lymphoma
EBVEpstein–Barr virus.
ecircRNAsCircular exonic RNAs
eiciRNAsExon–intron circRNAs
EVDEbola Virus Disease
G3BP1 GTPase-activating protein SH3 domain-binding protein 1
HBVHepatitis B virus
HCCHepatocellular carcinoma
HCVHepatitis C virus
HDVHepatitis D virus
HepG2Human hepatoma cell line
HIVHuman immunodeficiency virus
HK2Hexokinase 2
hnRNP CHeterogeneous nuclear ribonucleoprotein C
HRSP12Heat-responsive protein 12
HTNVHantaan virus
Huh7 Human hepatocarcinoma cell line
IAVInfluenza A virus
ICAM1Intercellular adhesion molecule 1
IFNInterferon
IL-15Interleukin-15
IL-6Interleukin-6
IPEC-J2Intestinal porcine enterocytes cell line
IRF3Interferon Regulatory Factor 3
JAK-STAT Janus Kinase/Signal Transducer and Activator of Transcription
KSHVKaposi sarcoma-associated herpesvirus
LNALocked nucleic acid
lncRNAsLong non-coding RNAs
LNPLipid-based nanoparticle
MAP3K2 Mitogen-Activated Protein Kinase Kinase Kinase 2
MAPK Mitogen-activated protein kinase
MERS-CoVMiddle East respiratory syndrome coronavirus
MIMVmaize Iranian mosaic virus
miRNAsMicroRNAs
mPTPMitochondrial permeability transition pore
MREs MiRNA response elements
mRNAsMessenger RNAs
ncRNAsNon-coding RNAs
NF-κB Nuclear factor kappa light chain enhancer of activated B cells
NODNucleotide-binding oligomerization domain
NPCNasopharyngeal carcinoma
p53 Tumor protein P53
PD-L1Programmed death ligand 1
PKR Protein kinase R
RBDReceptor binding domain
sncRNAsSmall non-coding RNAs
SRSF1 Serine And Arginine-Rich Splicing Factor 1
TAOK2Thousand-And-One Kinase 2
TAP2Transporter 2
TGEVTransmissible Gastroenteritis Virus
TNFAIP3 TNF alpha-induced protein 3
TRAF6TNF receptor associated factor 6
TRIM59Tripartite Motif Containing 59
VPS34Vacuolar protein sorting 34
vSP27Viral small peptide 27
VSP39Viral small peptide 39
YTHDF2 YTH N6-Methyladenosine RNA-Binding Protein F2

References

  1. Lander, E.S.; Linton, L.M.; Birren, B.; Nusbaum, C.; Zody, M.C.; Baldwin, J.; Devon, K.; Dewar, K.; Doyle, M.; FitzHugh, W.; et al. Initial sequencing and analysis of the human genome. Nature 2001, 409, 860–921. [Google Scholar] [CrossRef] [PubMed]
  2. Carninci, P.; Kasukawa, T.; Katayama, S.; Gough, J.; Frith, M.C.; Maeda, N.; Oyama, R.; Ravasi, T.; Lenhard, B.; Wells, C.; et al. The transcriptional landscape of the mammalian genome. Science 2005, 309, 1559–1563. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Consortium, E.P. An integrated encyclopedia of DNA elements in the human genome. Nature 2012, 489, 57–74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Awan, H.M.; Shah, A.; Rashid, F.; Shan, G. Primate-specific Long Non-coding RNAs and MicroRNAs. Genom. Proteom. Bioinform. 2017, 15, 187–195. [Google Scholar] [CrossRef]
  5. Kota, S.K.; Kota, S.B. Noncoding RNA and epigenetic gene regulation in renal diseases. Drug Discov. Today 2017, 22, 1112–1122. [Google Scholar] [CrossRef]
  6. Ponting, C.P.; Oliver, P.L.; Reik, W. Evolution and functions of long noncoding RNAs. Cell 2009, 136, 629–641. [Google Scholar] [CrossRef] [Green Version]
  7. Rai, K.R.; Liao, Y.; Cai, M.; Qiu, H.; Wen, F.; Peng, M.; Wang, S.; Liu, S.; Guo, G.; Chi, X.; et al. MIR155HG Plays a Bivalent Role in Regulating Innate Antiviral Immunity by Encoding Long Noncoding RNA-155 and microRNA-155-5p. Mbio 2022, 13, e0251022. [Google Scholar] [CrossRef]
  8. 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]
  9. Sanger, H.L.; Klotz, G.; Riesner, D.; Gross, H.J.; Kleinschmidt, A.K. Viroids are single-stranded covalently closed circular RNA molecules existing as highly base-paired rod-like structures. Proc. Natl. Acad. Sci. USA 1976, 73, 3852–3856. [Google Scholar] [CrossRef]
  10. Kolakofsky, D. Isolation and characterization of Sendai virus DI-RNAs. Cell 1976, 8, 547–555. [Google Scholar] [CrossRef]
  11. Kos, A.; Dijkema, R.; Arnberg, A.C.; van der Meide, P.H.; Schellekens, H. The hepatitis delta (delta) virus possesses a circular RNA. Nature 1986, 323, 558–560. [Google Scholar] [CrossRef] [PubMed]
  12. Lyu, D.; Huang, S. The emerging role and clinical implication of human exonic circular RNA. RNA Biol. 2017, 14, 1000–1006. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Hsiao, K.Y.; Sun, H.S.; Tsai, S.J. Circular RNA-New member of noncoding RNA with novel functions. Exp. Biol. Med. 2017, 242, 1136–1141. [Google Scholar] [CrossRef] [Green Version]
  14. Salzman, J.; Chen, R.E.; Olsen, M.N.; Wang, P.L.; Brown, P.O. Cell-type specific features of circular RNA expression. PLoS Genet. 2013, 9, e1003777. [Google Scholar] [CrossRef]
  15. Barrett, S.P.; Salzman, J. Circular RNAs: Analysis, expression and potential functions. Development 2016, 143, 1838–1847. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Salzman, J.; Gawad, C.; Wang, P.L.; Lacayo, N.; Brown, P.O. Circular RNAs are the predominant transcript isoform from hundreds of human genes in diverse cell types. PLoS ONE 2012, 7, e30733. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Jeck, W.R.; Sharpless, N.E. Detecting and characterizing circular RNAs. Nat. Biotechnol. 2014, 32, 453–461. [Google Scholar] [CrossRef]
  18. Liang, D.; Wilusz, J.E. Short intronic repeat sequences facilitate circular RNA production. Genes. Dev. 2014, 28, 2233–2247. [Google Scholar] [CrossRef] [Green Version]
  19. Liu, C.X.; Chen, L.L. Circular RNAs: Characterization, cellular roles, and applications. Cell 2022, 185, 2390. [Google Scholar] [CrossRef]
  20. Zhao, Z.J.; Shen, J. Circular RNA participates in the carcinogenesis and the malignant behavior of cancer. RNA Biol. 2017, 14, 514–521. [Google Scholar] [CrossRef] [Green Version]
  21. Li, I.; Chen, Y.G. Emerging roles of circular RNAs in innate immunity. Curr. Opin. Immunol. 2021, 68, 107–115. [Google Scholar] [CrossRef]
  22. Greene, J.; Baird, A.M.; Brady, L.; Lim, M.; Gray, S.G.; McDermott, R.; Finn, S.P. Circular RNAs: Biogenesis, Function and Role in Human Diseases. Front. Mol. Biosci. 2017, 4, 38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Li, X.; Yang, L.; Chen, L.L. The Biogenesis, Functions, and Challenges of Circular RNAs. Mol. Cell 2018, 71, 428–442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. 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]
  25. Matera, A.G.; Wang, Z. A day in the life of the spliceosome. Nat. Rev. Mol. Cell Biol. 2014, 15, 108–121. [Google Scholar] [CrossRef] [Green Version]
  26. 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]
  27. 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]
  28. 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] [PubMed] [Green Version]
  29. Ashwal-Fluss, R.; Meyer, M.; Pamudurti, N.R.; Ivanov, 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] [Green Version]
  30. Zhang, Y.; Xue, W.; Li, X.; Zhang, J.; Chen, S.; Zhang, J.L.; Yang, L.; Chen, L.L. The Biogenesis of Nascent Circular RNAs. Cell Rep. 2016, 15, 611–624. [Google Scholar] [CrossRef] [Green Version]
  31. Qiu, H.; Yang, B.; Chen, Y.; Zhu, Q.; Wen, F.; Peng, M.; Wang, G.; Guo, G.; Chen, B.; Maarouf, M.; et al. Influenza A Virus-Induced circRNA circMerTK Negatively Regulates Innate Antiviral Responses. Microbiol. Spectr. 2023, 11, e0363722. [Google Scholar] [CrossRef] [PubMed]
  32. 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]
  33. Szabo, L.; Morey, R.; Palpant, N.J.; Wang, P.L.; Afari, N.; Jiang, C.; Parast, M.M.; Murry, C.E.; Laurent, L.C.; Salzman, J. Statistically based splicing detection reveals neural enrichment and tissue-specific induction of circular RNA during human fetal development. Genome Biol. 2015, 16, 126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. 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.e21. [Google Scholar] [CrossRef]
  35. Zhou, C.; Molinie, B.; Daneshvar, K.; Pondick, J.V.; Wang, J.; Van Wittenberghe, N.; Xing, Y.; Giallourakis, C.C.; Mullen, A.C. Genome-Wide Maps of m6A circRNAs Identify Widespread and Cell-Type-Specific Methylation Patterns that Are Distinct from mRNAs. Cell Rep. 2017, 20, 2262–2276. [Google Scholar] [CrossRef] [Green Version]
  36. Park, O.H.; Ha, H.; Lee, Y.; Boo, S.H.; Kwon, D.H.; Song, H.K.; Kim, Y.K. Endoribonucleolytic Cleavage of m(6)A-Containing RNAs by RNase P/MRP Complex. Mol. Cell 2019, 74, 494–507.e498. [Google Scholar] [CrossRef]
  37. 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] [Green Version]
  38. 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.e76. [Google Scholar] [CrossRef]
  39. Kariko, K.; Buckstein, M.; Ni, H.; Weissman, D. Suppression of RNA recognition by Toll-like receptors: The impact of nucleoside modification and the evolutionary origin of RNA. Immunity 2005, 23, 165–175. [Google Scholar] [CrossRef] [Green Version]
  40. Verbeke, R.; Hogan, M.J.; Lore, K.; Pardi, N. Innate immune mechanisms of mRNA vaccines. Immunity 2022, 55, 1993–2005. [Google Scholar] [CrossRef]
  41. Chen, Y.G.; Kim, M.V.; Chen, X.; Batista, P.J.; Aoyama, S.; Wilusz, J.E.; Iwasaki, A.; Chang, H.Y. Sensing Self and Foreign Circular RNAs by Intron Identity. Mol. Cell 2017, 67, 228–238. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Wesselhoeft, R.A.; Kowalski, P.S.; Parker-Hale, F.C.; Huang, Y.X.; Bisaria, N.; Anderson, D.G. RNA Circularization Diminishes Immunogenicity and Can Extend Translation Duration In Vivo. Mol. Cell 2019, 74, 508–520.e4. [Google Scholar] [CrossRef] [PubMed]
  43. Liu, C.X.; Guo, S.K.; Nan, F.; Xu, Y.F.; Yang, L.; Chen, L.L. RNA circles with minimized immunogenicity as potent PKR inhibitors. Mol. Cell 2022, 82, 420–434.e6. [Google Scholar] [CrossRef] [PubMed]
  44. Tai, J.; Chen, Y.G. Differences in the immunogenicity of engineered circular RNAs. J. Mol. Cell Biol. 2023, 15, mjad002. [Google Scholar] [CrossRef]
  45. 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]
  46. Fabian, M.R.; Sonenberg, N.; Filipowicz, W. Regulation of mRNA translation and stability by microRNAs. Annu. Rev. Biochem. 2010, 79, 351–379. [Google Scholar] [CrossRef] [Green Version]
  47. Chen, S.; Zhao, Y. Circular RNAs: Characteristics, function, and role in human cancer. Histol. Histopathol. 2018, 33, 887–893. [Google Scholar] [CrossRef]
  48. Jost, I.; Shalamova, L.A.; Gerresheim, G.K.; Niepmann, M.; Bindereif, A.; Rossbach, O. Functional sequestration of microRNA-122 from Hepatitis C Virus by circular RNA sponges. RNA Biol. 2018, 15, 1032–1039. [Google Scholar] [CrossRef]
  49. 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.e217. [Google Scholar] [CrossRef] [Green Version]
  50. Du, W.W.; Zhang, C.; Yang, W.; Yong, T.; Awan, F.M.; Yang, B.B. Identifying and Characterizing circRNA-Protein Interaction. Theranostics 2017, 7, 4183–4191. [Google Scholar] [CrossRef]
  51. Wang, Y.; Wang, Z. Efficient backsplicing produces translatable circular mRNAs. RNA 2015, 21, 172–179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Qu, L.; Yi, Z.Y.; Shen, Y.; Lin, L.R.; Chen, F.; Xu, Y.Y.; Wu, Z.G.; Tang, H.X.; Zhang, X.X.; Tian, F.; et al. Circular RNA vaccines against SARS-CoV-2 and emerging variants. Cell 2022, 185, 1728–1744.e16. [Google Scholar] [CrossRef]
  53. Chen, R.; Wang, S.K.; Belk, J.A.; Amaya, L.; Li, Z.; Cardenas, A.; Abe, B.T.; Chen, C.K.; Wender, P.A.; Chang, H.Y. Engineering circular RNA for enhanced protein production. Nat. Biotechnol. 2023, 41, 262–272. [Google Scholar] [CrossRef]
  54. Wang, X.; Ma, R.; Zhang, X.; Cui, L.; Ding, Y.; Shi, W.; Guo, C.; Shi, Y. Crosstalk between N6-methyladenosine modification and circular RNAs: Current understanding and future directions. Mol. Cancer 2021, 20, 121. [Google Scholar] [CrossRef] [PubMed]
  55. Chen, C.; Wei, H.; Zhang, K.; Li, Z.; Wei, T.; Tang, C.; Yang, Y.; Wang, Z. A flexible, efficient, and scalable platform to produce circular RNAs as new therapeutics. bioRxiv 2022, 05, 494115. [Google Scholar] [CrossRef]
  56. Chen, X.; Liu, S.; Goraya, M.U.; Maarouf, M.; Huang, S.; Chen, J.L. Host Immune Response to Influenza A Virus Infection. Front. Immunol. 2018, 9, 320. [Google Scholar] [CrossRef] [Green Version]
  57. Rai, K.R.; Shrestha, P.; Yang, B.; Chen, Y.; Liu, S.; Maarouf, M.; Chen, J.L. Acute Infection of Viral Pathogens and Their Innate Immune Escape. Front. Microbiol. 2021, 12, 672026. [Google Scholar] [CrossRef]
  58. Maarouf, M.; Rai, K.R.; Goraya, M.U.; Chen, J.L. Immune Ecosystem of Virus-Infected Host Tissues. Int. J. Mol. Sci. 2018, 19, 1379. [Google Scholar] [CrossRef] [Green Version]
  59. Qu, Z.Y.; Meng, F.; Shi, J.Z.; Deng, G.H.; Zeng, X.Y.; Ge, J.Y.; Li, Y.B.; Liu, L.L.; Chen, P.C.; Jiang, Y.P.; et al. A Novel Intronic Circular RNA Antagonizes Influenza Virus by Absorbing a microRNA That Degrades CREBBP and Accelerating IFN-beta Production. Mbio 2021, 12, e0101721. [Google Scholar] [CrossRef]
  60. Cadena, C.; Hur, S. Antiviral Immunity and Circular RNA: No End in Sight. Mol. Cell 2017, 67, 163–164. [Google Scholar] [CrossRef] [Green Version]
  61. Awan, F.M.; Yang, B.B.; Naz, A.; Hanif, A.; Ikram, A.; Obaid, A.; Malik, A.; Janjua, H.A.; Ali, A.; Sharif, S. The emerging role and significance of circular RNAs in viral infections and antiviral immune responses: Possible implication as theranostic agents. RNA Biol. 2021, 18, 1–15. [Google Scholar] [CrossRef] [PubMed]
  62. Choudhary, A.; Madbhagat, P.; Sreepadmanabh, M.; Bhardwaj, V.; Chande, A. Circular RNA as an Additional Player in the Conflicts Between the Host and the Virus. Front. Immunol. 2021, 12, 602006. [Google Scholar] [CrossRef]
  63. Lu, S.; Zhu, N.; Guo, W.; Wang, X.; Li, K.; Yan, J.; Jiang, C.; Han, S.; Xiang, H.; Wu, X.; et al. RNA-Seq Revealed a Circular RNA-microRNA-mRNA Regulatory Network in Hantaan Virus Infection. Front. Cell Infect. Microbiol. 2020, 10, 97. [Google Scholar] [CrossRef] [Green Version]
  64. Firoozi, Z.; Mohammadisoleimani, E.; Shahi, A.; Naghizadeh, M.M.; Mirzaei, E.; Asad, A.G.; Salmanpour, Z.; Javad Nouri, S.M.; Mansoori, Y. Hsa_circ_0000479/Hsa-miR-149-5p/RIG-I, IL-6 Axis: A Potential Novel Pathway to Regulate Immune Response against COVID-19. Can. J. Infect. Dis. Med. Microbiol. 2022, 2022, 2762582. [Google Scholar] [CrossRef]
  65. Yang, Q.; Li, Y.; Wang, Y.; Qiao, X.; Liu, T.; Wang, H.; Shen, H. The circRNA circSIAE Inhibits Replication of Coxsackie Virus B3 by Targeting miR-331-3p and Thousand and One Amino-Acid Kinase 2. Front. Cell Infect. Microbiol. 2021, 11, 779919. [Google Scholar] [CrossRef]
  66. Zhao, X.; Ma, X.; Guo, J.; Mi, M.; Wang, K.; Zhang, C.; Tang, X.; Chang, L.; Huang, Y.; Tong, D. Circular RNA CircEZH2 Suppresses Transmissible Gastroenteritis Coronavirus-induced Opening of Mitochondrial Permeability Transition Pore via Targeting MiR-22 in IPEC-J2. Int. J. Biol. Sci. 2019, 15, 2051–2064. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Ma, X.; Zhao, X.; Zhang, Z.; Guo, J.; Guan, L.; Li, J.; Mi, M.; Huang, Y.; Tong, D. Differentially expressed non-coding RNAs induced by transmissible gastroenteritis virus potentially regulate inflammation and NF-kappaB pathway in porcine intestinal epithelial cell line. BMC Genom. 2018, 19, 747. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Jenner, R.G.; Young, R.A. Insights into host responses against pathogens from transcriptional profiling. Nat. Rev. Microbiol. 2005, 3, 281–294. [Google Scholar] [CrossRef] [PubMed]
  69. Zhang, X.; Chu, H.; Wen, L.; Shuai, H.; Yang, D.; Wang, Y.; Hou, Y.; Zhu, Z.; Yuan, S.; Yin, F.; et al. Competing endogenous RNA network profiling reveals novel host dependency factors required for MERS-CoV propagation. Emerg. Microbes Infect. 2020, 9, 733–746. [Google Scholar] [CrossRef] [Green Version]
  70. Li, J.; Yang, H.; Shi, H.; Zhang, J.; Chen, W. Expression Profiles of Differentially Expressed Circular RNAs and circRNA-miRNA-mRNA Regulatory Networks in SH-SY5Y Cells Infected with Coxsackievirus B5. Int. J. Genom. 2022, 2022, 9298149. [Google Scholar] [CrossRef]
  71. Zhao, W.; Su, J.; Wang, N.; Zhao, N.; Su, S. Expression Profiling and Bioinformatics Analysis of CircRNA in Mice Brain Infected with Rabies Virus. Int. J. Mol. Sci. 2021, 22, 6537. [Google Scholar] [CrossRef] [PubMed]
  72. Tanuj, G.N.; Khan, O.; Malla, W.A.; Rajak, K.K.; Chandrashekar, S.; Kumar, A.; Dhara, S.K.; Gupta, P.K.; Mishra, B.P.; Dutt, T.; et al. Integrated analysis of long-noncoding RNA and circular RNA expression in Peste-des-Petits-Ruminants Virus (PPRV) infected marmoset B lymphocyte (B95a) cells. Microb. Pathog. 2022, 170, 105702. [Google Scholar] [CrossRef] [PubMed]
  73. Chen, J.; Wang, H.; Jin, L.; Wang, L.; Huang, X.; Chen, W.; Yan, M.; Liu, G. Profile analysis of circRNAs induced by porcine endemic diarrhea virus infection in porcine intestinal epithelial cells. Virology 2019, 527, 169–179. [Google Scholar] [CrossRef]
  74. Li, H.; Tang, W.; Jin, Y.; Dong, W.; Yan, Y.; Zhou, J. Differential CircRNA Expression Profiles in PK-15 Cells Infected with Pseudorabies Virus Type II. Virol. Sin. 2021, 36, 75–84. [Google Scholar] [CrossRef]
  75. Yang, M.; Qi, M.; Xu, L.; Huang, P.; Wang, X.; Sun, J.; Shi, J.; Hu, Y. Differential host circRNA expression profiles in human lung epithelial cells infected with SARS-CoV-2. Infect. Genet. Evol. 2021, 93, 104923. [Google Scholar] [CrossRef] [PubMed]
  76. Yang, J.; Yang, B.; Wang, Y.; Zhang, T.; Hao, Y.; Cui, H.; Zhao, D.; Yuan, X.; Chen, X.; Shen, C.; et al. Profiling and functional analysis of differentially expressed circular RNAs identified in foot-and-mouth disease virus infected PK-15 cells. Vet. Res. 2022, 53, 24. [Google Scholar] [CrossRef] [PubMed]
  77. Huang, X.; Zhang, J.; Liu, Z.; Wang, M.; Fan, X.; Wang, L.; Zhou, H.; Jiang, Y.; Cui, W.; Qiao, X.; et al. Genome-wide analysis of differentially expressed mRNAs, lncRNAs, and circRNAs in chicken bursae of Fabricius during infection with very virulent infectious bursal disease virus. BMC Genom. 2020, 21, 724. [Google Scholar] [CrossRef]
  78. Liu, Z.; Guo, Y.; Zhao, L.; Liu, Q.; Tian, M.; Huang, N.; Fan, M.; Yu, M.; Xia, H.; Ping, J. Analysis of the circRNAs expression profile in mouse lung with H7N9 influenza A virus infection. Genomics 2021, 113, 716–727. [Google Scholar] [CrossRef]
  79. Jiao, H.; Zhao, Y.; Zhou, Z.; Li, W.; Li, B.; Gu, G.; Luo, Y.; Shuai, X.; Fan, C.; Wu, L.; et al. Identifying Circular RNAs in HepG2 Expressing Genotype IV Swine Hepatitis E Virus ORF3 Via Whole Genome Sequencing. Cell Transpl. 2021, 30, 9636897211055042. [Google Scholar] [CrossRef]
  80. Hu, Y.; Xu, Y.; Deng, X.; Wang, R.; Li, R.; You, L.; Song, J.; Zhang, Y. Comprehensive analysis of the circRNA expression profile and circRNA-miRNA-mRNA network in the pathogenesis of EV-A71 infection. Virus Res. 2021, 303, 198502. [Google Scholar] [CrossRef]
  81. Wang, J.; Zhang, Y.; Zhu, F.; Chen, L.; Wei, Y.; Zhu, Q.; Jiang, J.; Huang, J.; Guo, Q.; Yang, X. CircRNA expression profiling and bioinformatics analysis indicate the potential biological role and clinical significance of circRNA in influenza A virus-induced lung injury. J. Biosci. 2021, 46, 38. [Google Scholar] [CrossRef]
  82. Zheng, S.R.; Zhang, H.R.; Zhang, Z.F.; Lai, S.Y.; Huang, L.J.; Liu, J.; Bai, X.; Ding, K.; Zhou, J.Y. Human papillomavirus 16 E7 oncoprotein alters the expression profiles of circular RNAs in Caski cells. J. Cancer 2018, 9, 3755–3764. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Li, Y.; Ashraf, U.; Chen, Z.; Zhou, D.; Imran, M.; Ye, J.; Chen, H.; Cao, S. Genome-wide profiling of host-encoded circular RNAs highlights their potential role during the Japanese encephalitis virus-induced neuroinflammatory response. BMC Genom. 2020, 21, 409. [Google Scholar] [CrossRef] [PubMed]
  84. Chen, L.; Li, G.; Tian, Y.; Zeng, T.; Xu, W.; Gu, T.; Lu, L. RNA Sequencing Reveals circRNA Expression Profiles in Chicken DF1 Cells Infected with H5N1 Influenza Virus. Animals 2022, 12, 158. [Google Scholar] [CrossRef] [PubMed]
  85. Yang, T.; Qiu, L.; Bai, M.; Wang, L.; Hu, X.; Huang, L.; Chen, G.; Chang, G. Identification, biogenesis and function prediction of novel circRNA during the chicken ALV-J infection. Anim. Biotechnol. 2022, 33, 981–991. [Google Scholar] [CrossRef]
  86. Hu, Y.; Yang, R.; Zhao, W.; Liu, C.; Tan, Y.; Pu, D.; Song, J.; Zhang, Y. circRNA expression patterns and circRNA-miRNA-mRNA networks during CV-A16 infection of SH-SY5Y cells. Arch. Virol. 2021, 166, 3023–3035. [Google Scholar] [CrossRef]
  87. Zhang, Y.; Zhang, H.; An, M.H.; Zhao, B.; Ding, H.B.; Zhang, Z.N.; He, Y.W.; Shang, H.; Han, X.X. Crosstalk in competing endogenous RNA networks reveals new circular RNAs involved in the pathogenesis of early HIV infection. J. Transl. Med. 2018, 16, 3023–3035. [Google Scholar] [CrossRef]
  88. Ghorbani, A.; Izadpanah, K.; Peters, J.R.; Dietzgen, R.G.; Mitter, N. Detection and profiling of circular RNAs in uninfected and maize Iranian mosaic virus-infected maize. Plant Sci. 2018, 274, 402–409. [Google Scholar] [CrossRef] [Green Version]
  89. Tagawa, T.; Gao, S.; Koparde, V.N.; Gonzalez, M.; Spouge, J.L.; Serquiña, A.P.; Lurain, K.; Ramaswami, R.; Uldrick, T.S.; Yarchoan, R.; et al. Discovery of Kaposi’s sarcoma herpesvirus-encoded circular RNAs and a human antiviral circular RNA. Proc. Natl. Acad. Sci. USA 2018, 115, 12805–12810. [Google Scholar] [CrossRef] [Green Version]
  90. Zhang, L.; Wang, Z. Circular RNA hsa_circ_0004812 impairs IFN-induced immune response by sponging miR-1287-5p to regulate FSTL1 in chronic hepatitis B. Virol. J. 2020, 17, 40. [Google Scholar] [CrossRef] [Green Version]
  91. Du, N.; Li, K.; Wang, Y.; Song, B.; Zhou, X.; Duan, S. CircRNA circBACH1 facilitates hepatitis B virus replication and hepatoma development by regulating the miR-200a-3p/MAP3K2 axis. Histol. Histopathol. 2022, 37, 863–877. [Google Scholar] [CrossRef]
  92. Jiang, W.; Wang, L.; Zhang, Y.; Li, H. Circ-ATP5H Induces Hepatitis B Virus Replication and Expression by Regulating miR-138-5p/TNFAIP3 Axis. Cancer Manag. Res. 2020, 12, 11031–11040. [Google Scholar] [CrossRef] [PubMed]
  93. Zhang, X.; Chu, H.; Chik, K.K.; Wen, L.; Shuai, H.; Yang, D.; Wang, Y.; Hou, Y.; Yuen, T.T.; Cai, J.P.; et al. hnRNP C modulates MERS-CoV and SARS-CoV-2 replication by governing the expression of a subset of circRNAs and cognitive mRNAs. Emerg. Microbes Infect. 2022, 11, 519–531. [Google Scholar] [CrossRef] [PubMed]
  94. Shi, N.; Zhang, S.; Guo, Y.; Yu, X.; Zhao, W.; Zhang, M.; Guan, Z.; Duan, M. CircRNA_0050463 promotes influenza A virus replication by sponging miR-33b-5p to regulate EEF1A1. Vet. Microbiol. 2021, 254, 108995. [Google Scholar] [CrossRef]
  95. Yu, T.; Ding, Y.; Zhang, Y.; Liu, Y.; Li, Y.; Lei, J.; Zhou, J.; Song, S.; Hu, B. Circular RNA GATAD2A promotes H1N1 replication through inhibiting autophagy. Vet. Microbiol. 2019, 231, 238–245. [Google Scholar] [CrossRef]
  96. Min, J.; Cao, Y.; Liu, H.; Liu, D.; Liu, W.; Li, J. RNA Sequencing Demonstrates That Circular RNA Regulates Avian Influenza Virus Replication in Human Cells. Int. J. Mol. Sci. 2022, 23, 9901. [Google Scholar] [CrossRef]
  97. Chen, T.C.; Tallo-Parra, M.; Cao, Q.M.; Kadener, S.; Bottcher, R.; Perez-Vilaro, G.; Boonchuen, P.; Somboonwiwat, K.; Diez, J.; Sarnow, P. Host-derived circular RNAs display proviral activities in Hepatitis C virus-infected cells. PLoS Pathog. 2020, 16, e1008346. [Google Scholar] [CrossRef] [PubMed]
  98. Bhardwaj, V.; Singh, A.; Dalavi, R.; Ralte, L.; Chawngthu, R.L.; Kumar, N.S.; Vijay, N.; Chande, A. HIV-1 Vpr induces ciTRAN to prevent transcriptional silencing of the provirus. bioRxiv 2022, 11, 515166. [Google Scholar] [CrossRef]
  99. Zhao, C.X.; Yan, Z.X.; Wen, J.J.; Fu, D.; Xu, P.P.; Wang, L.; Cheng, S.; Hu, J.D.; Zhao, W.L. CircEAF2 counteracts Epstein-Barr virus-positive diffuse large B-cell lymphoma progression via miR-BART19-3p/APC/beta-catenin axis. Mol. Cancer 2021, 20, 153. [Google Scholar] [CrossRef]
  100. Memish, Z.A.; Perlman, S.; Van Kerkhove, M.D.; Zumla, A. Middle East respiratory syndrome. Lancet 2020, 395, 1063–1077. [Google Scholar] [CrossRef]
  101. Spearman, C.W.; Dusheiko, G.M.; Hellard, M.; Sonderup, M. Hepatitis C. Lancet 2019, 394, 1451–1466. [Google Scholar] [CrossRef]
  102. Jacob, S.T.; Crozier, I.; Fischer, W.A.; Hewlett, A.; Kraft, C.S.; Vega, M.-A.d.L.; Soka, M.J.; Wahl, V.; Griffiths, A.; Bollinger, L.; et al. Ebola virus disease. Nat. Rev. Dis. Primers 2020, 6, 13. [Google Scholar] [CrossRef] [Green Version]
  103. Wang, Z.Y.; Guo, Z.D.; Li, J.M.; Zhao, Z.Z.; Fu, Y.Y.; Zhang, C.M.; Zhang, Y.; Liu, L.N.; Qian, J.; Liu, L.N. Genome-Wide Search for Competing Endogenous RNAs Responsible for the Effects Induced by Ebola Virus Replication and Transcription Using a trVLP System. Front. Cell Infect. Microbiol. 2017, 7, 479. [Google Scholar] [CrossRef] [PubMed]
  104. Deeks, S.G.; Overbaugh, J.; Phillips, A.; Buchbinder, S. HIV infection. Nat. Rev. Dis. Primers 2015, 1, 15035. [Google Scholar] [CrossRef] [PubMed]
  105. Holmes, K.K.; Bertozzi, S.; Bloom, B.R.; Jha, P.; Piot, P. Major Infectious Diseases, 3rd ed.; World Bank Group: Washington, DC, USA, 2017; p. xviii. 486p. [Google Scholar]
  106. Tagawa, T.; Oh, D.; Dremel, S.; Mahesh, G.; Koparde, V.N.; Duncan, G.; Andresson, T.; Ziegelbauer, J.M. A virus-induced circular RNA maintains latent infection of Kaposi’s sarcoma herpesvirus. Proc. Natl. Acad. Sci. USA 2023, 120, e2212864120. [Google Scholar] [CrossRef] [PubMed]
  107. Shi, J.D.; Hu, N.Z.; Mo, L.; Zeng, Z.P.; Sun, J.; Hu, Y.Z. Deep RNA Sequencing Reveals a Repertoire of Human Fibroblast Circular RNAs Associated with Cellular Responses to Herpes Simplex Virus 1 Infection. Cell Physiol. Biochem. 2018, 47, 2031–2045. [Google Scholar] [CrossRef] [PubMed]
  108. Zheng, W.; Sun, L.; Yang, L.; Xu, T. The circular RNA circBCL2L1 regulates innate immune responses via microRNA-mediated downregulation of TRAF6 in teleost fish. J. Biol. Chem. 2021, 297, 101199. [Google Scholar] [CrossRef] [PubMed]
  109. Zhao, J.; Lee, E.E.; Kim, J.; Yang, R.; Chamseddin, B.; Ni, C.; Gusho, E.; Xie, Y.; Chiang, C.M.; Buszczak, M.; et al. Transforming activity of an oncoprotein-encoding circular RNA from human papillomavirus. Nat. Commun. 2019, 10, 2300. [Google Scholar] [CrossRef] [Green Version]
  110. Abere, B.; Li, J.; Zhou, H.; Toptan, T.; Moore, P.S.; Chang, Y. Kaposi’s Sarcoma-Associated Herpesvirus-Encoded circRNAs Are Expressed in Infected Tumor Tissues and Are Incorporated into Virions. Mbio 2020, 11, e03027-19. [Google Scholar] [CrossRef] [Green Version]
  111. Yang, S.; Zhou, H.; Liu, M.; Jaijyan, D.; Cruz-Cosme, R.; Ramasamy, S.; Subbian, S.; Liu, D.; Xu, J.; Niu, X.; et al. SARS-CoV-2, SARS-CoV, and MERS-CoV encode circular RNAs of spliceosome-independent origin. J. Med. Virol. 2022, 94, 3203–3222. [Google Scholar] [CrossRef]
  112. Cai, Z.; Lu, C.; He, J.; Liu, L.; Zou, Y.; Zhang, Z.; Zhu, Z.; Ge, X.; Wu, A.; Jiang, T.; et al. Identification and characterization of circRNAs encoded by MERS-CoV, SARS-CoV-1 and SARS-CoV-2. Brief. Bioinform. 2021, 22, 1297–1308. [Google Scholar] [CrossRef] [PubMed]
  113. Zhang, X.; Liang, Z.; Wang, C.; Shen, Z.; Sun, S.; Gong, C.; Hu, X. Viral Circular RNAs and Their Possible Roles in Virus-Host Interaction. Front. Immunol. 2022, 13, 939768. [Google Scholar] [CrossRef] [PubMed]
  114. Tan, K.E.; Lim, Y.Y. Viruses join the circular RNA world. FEBS J. 2021, 288, 4488–4502. [Google Scholar] [CrossRef] [PubMed]
  115. Ge, J.; Wang, J.; Xiong, F.; Jiang, X.; Zhu, K.; Wang, Y.; Mo, Y.; Gong, Z.; Zhang, S.; He, Y.; et al. Epstein-Barr Virus-Encoded Circular RNA CircBART2.2 Promotes Immune Escape of Nasopharyngeal Carcinoma by Regulating PD-L1. Cancer Res. 2021, 81, 5074–5088. [Google Scholar] [CrossRef]
  116. Abere, B.; Zhou, H.; Li, J.; Cao, S.; Toptan, T.; Grundhoff, A.; Fischer, N.; Moore, P.S.; Chang, Y. Merkel Cell Polyomavirus Encodes Circular RNAs (circRNAs) Enabling a Dynamic circRNA/microRNA/mRNA Regulatory Network. Mbio 2020, 11, e03059-20. [Google Scholar] [CrossRef]
  117. Zhang, Z.; Zhao, Z.; Lin, S.; Wu, W.; Tang, W.; Dong, Y.; Shen, M.; Wu, P.; Guo, X. Identification of long noncoding RNAs in silkworm larvae infected with Bombyx mori cypovirus. Arch. Insect Biochem. Physiol. 2021, 106, 1–12. [Google Scholar] [CrossRef]
  118. Dong, W.T.; Ling, X.D.; Xiao, L.F.; Hu, J.J.; Zhao, X.X.; Liu, J.X.; Zhang, Y. Effects of Bombyx mori nuclear polyhedrosis virus on serpin and antibacterial peptide expression in B. mori. Microb. Pathog. 2019, 130, 137–145. [Google Scholar] [CrossRef]
  119. Zhang, Y.; Zhang, X.; Shen, Z.; Qiu, Q.; Tong, X.; Pan, J.; Zhu, M.; Hu, X.; Gong, C. BmNPV circular RNA-encoded peptide VSP39 promotes viral replication. Int. J. Biol. Macromol. 2023, 228, 299–310. [Google Scholar] [CrossRef]
  120. Gui, J.; Zhou, L. Genetic basis and breeding application of clonal diversity and dual reproduction modes in polyploid Carassius auratus gibelio. Sci. China Life Sci. 2010, 53, 409–415. [Google Scholar] [CrossRef]
  121. Zhu, M.; Dai, Y.; Tong, X.; Zhang, Y.; Zhou, Y.; Cheng, J.; Jiang, Y.; Yang, R.; Wang, X.; Cao, G.; et al. Circ-Udg Derived from Cyprinid Herpesvirus 2 Promotes Viral Replication. Microbiol. Spectr. 2022, 10, e0094322. [Google Scholar] [CrossRef]
  122. Wang, M.; Yu, F.; Wu, W.; Zhang, Y.; Chang, W.; Ponnusamy, M.; Wang, K.; Li, P. Circular RNAs: A novel type of non-coding RNA and their potential implications in antiviral immunity. Int. J. Biol. Sci. 2017, 13, 1497–1506. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Breuer, J.; Barth, P.; Noe, Y.; Shalamova, L.; Goesmann, A.; Weber, F.; Rossbach, O. What goes around comes around: Artificial circular RNAs bypass cellular antiviral responses. Mol. Ther.-Nucleic Acids 2022, 28, 623–635. [Google Scholar] [CrossRef] [PubMed]
  124. Wesselhoeft, R.A.; Kowalski, P.S.; Anderson, D.G. Engineering circular RNA for potent and stable translation in eukaryotic cells. Nat. Commun. 2018, 9, 2629. [Google Scholar] [CrossRef] [Green Version]
  125. Banerji, A.; Wickner, P.G.; Saff, R.; Stone, C.A.; Robinson, L.B.; Long, A.A.; Wolfson, A.R.; Williams, P.; Khan, D.A.; Phillips, E.; et al. mRNA Vaccines to Prevent COVID-19 Disease and Reported Allergic Reactions: Current Evidence and Suggested Approach. J. Allergy Clin. Immunol. Pract. 2021, 9, 1423–1437. [Google Scholar] [CrossRef] [PubMed]
  126. Polack, F.P.; Thomas, S.J.; Kitchin, N.; Absalon, J.; Gurtman, A.; Lockhart, S.; Perez, J.L.; Marc, G.P.; Moreira, E.D.; Zerbini, C.; et al. Safety and Efficacy of the BNT162b2 mRNA Covid-19 Vaccine. N. Engl. J. Med. 2020, 383, 2603–2615. [Google Scholar] [CrossRef]
  127. Meganck, R.M.; Borchardt, E.K.; Rivera, R.M.C.; Scalabrino, M.L.; Wilusz, J.E.; Marzluff, W.F.; Asokan, A. Tissue-Dependent Expression and Translation of Circular RNAs with Recombinant AAV Vectors In Vivo. Mol. Ther. Nucleic Acids 2018, 13, 89–98. [Google Scholar] [CrossRef] [Green Version]
  128. Pardi, N.; Hogan, M.J.; Porter, F.W.; Weissman, D. mRNA vaccines-a new era in vaccinology. Nat. Rev. Drug Discov. 2018, 17, 261–279. [Google Scholar] [CrossRef] [Green Version]
  129. Bai, Y.; Liu, D.; He, Q.; Liu, J.; Mao, Q.; Liang, Z. Research progress on circular RNA vaccines. Front. Immunol. 2022, 13, 1091797. [Google Scholar] [CrossRef]
  130. Seephetdee, C.; Bhukhai, K.; Buasri, N.; Leelukkanaveera, P.; Lerdwattanasombat, P.; Manopwisedjaroen, S.; Phueakphud, N.; Kuhaudomlarp, S.; Olmedillas, E.; Saphire, E.O.; et al. A circular mRNA vaccine prototype producing VFLIP-X spike confers a broad neutralization of SARS-CoV-2 variants by mouse sera. Antivir. Res. 2022, 204, 105370. [Google Scholar] [CrossRef]
  131. Chen, C.K.; Cheng, R.; Demeter, J.; Chen, J.; Weingarten-Gabbay, S.; Jiang, L.H.; Snyder, M.P.; Weissman, J.S.; Segal, E.; Jackson, P.K.; et al. Structured elements drive extensive circular RNA translation. Mol. Cell 2021, 81, 4300–4318.E13. [Google Scholar] [CrossRef]
  132. Pfafenrot, C.; Schneider, T.; Muller, C.; Hung, L.H.; Schreiner, S.; Ziebuhr, J.; Bindereif, A. Inhibition of SARS-CoV-2 coronavirus proliferation by designer antisense-circRNAs. Nucleic Acids Res. 2021, 49, 12502–12516. [Google Scholar] [CrossRef]
  133. Shuai, M.; Huang, L. High Expression of hsa_circRNA_001387 in Nasopharyngeal Carcinoma and the Effect on Efficacy of Radiotherapy. Onco Targets Ther. 2020, 13, 3965–3973. [Google Scholar] [CrossRef]
  134. Liu, Q.; Shuai, M.; Xia, Y. Knockdown of EBV-encoded circRNA circRPMS1 suppresses nasopharyngeal carcinoma cell proliferation and metastasis through sponging multiple miRNAs. Cancer Manag. Res. 2019, 11, 8023–8031. [Google Scholar] [CrossRef] [Green Version]
  135. Zhao, T.; Zheng, Y.; Hao, D.; Jin, X.; Luo, Q.; Guo, Y.; Li, D.; Xi, W.; Xu, Y.; Chen, Y.; et al. Blood circRNAs as biomarkers for the diagnosis of community-acquired pneumonia. J. Cell Biochem. 2019, 120, 16483–16494. [Google Scholar] [CrossRef]
  136. Lavenniah, A.; Luu, T.D.A.; Li, Y.Q.P.; Lim, T.S.B.; Jiang, J.M.; Ackers-Johnson, M.; Foo, R.S.Y. Engineered Circular RNA Sponges Act as miRNA Inhibitors to Attenuate Pressure Overload-Induced Cardiac Hypertrophy. Mol. Ther. 2020, 28, 1506–1517. [Google Scholar] [CrossRef]
  137. Lu, D.C.; Thum, T. RNA-based diagnostic and therapeutic strategies for cardiovascular disease. Nat. Rev. Cardiol. 2019, 16, 661–674. [Google Scholar] [CrossRef]
  138. Janssen, H.L.; Reesink, H.W.; Lawitz, E.J.; Zeuzem, S.; Rodriguez-Torres, M.; Patel, K.; van der Meer, A.J.; Patick, A.K.; Chen, A.; Zhou, Y.; et al. Treatment of HCV infection by targeting microRNA. N. Engl. J. Med. 2013, 368, 1685–1694. [Google Scholar] [CrossRef] [Green Version]
  139. Yang, P.-W.; Chen, T.-T.; Zhao, W.-X.; Liu, G.-W.; Feng, X.-J.; Wang, S.-M.; Pan, Y.-C.; Wang, Q.; Zhang, S.-H. Scutellaria barbata D.Don and Oldenlandia diffusa (Willd.) Roxb crude extracts inhibit hepatitis-B-virus-associated hepatocellular carcinoma growth through regulating circRNA expression. J. Ethnopharmacol. 2021, 275, 114110. [Google Scholar] [CrossRef]
  140. Wang, Y.; Liu, J.; Ma, J.; Sun, T.; Zhou, Q.; Wang, W.; Wang, G.; Wu, P.; Wang, H.; Jiang, L.; et al. Exosomal circRNAs: Biogenesis, effect and application in human diseases. Mol. Cancer 2019, 18, 116. [Google Scholar] [CrossRef] [Green Version]
  141. Shi, X.; Wang, B.; Feng, X.; Xu, Y.; Lu, K.; Sun, M. circRNAs and Exosomes: A Mysterious Frontier for Human Cancer. Mol. Ther. Nucleic Acids 2020, 19, 384–392. [Google Scholar] [CrossRef]
  142. Wang, S.M.; Zhang, K.; Tan, S.Y.; Xin, J.Y.; Yuan, Q.Y.; Xu, H.H.; Xu, X.; Liang, Q.; Christiani, D.C.; Wang, M.L.; et al. Circular RNAs in body fluids as cancer biomarkers: The new frontier of liquid biopsies. Mol. Cancer 2021, 20, 13. [Google Scholar] [CrossRef]
  143. Song, Z.; Zhang, Q.; Zhu, J.; Yin, G.; Lin, L.; Liang, C. Identification of urinary hsa_circ_0137439 as a potential biomarker and tumor regulator of bladder cancer. Neoplasma 2020, 67, 137–146. [Google Scholar] [CrossRef]
  144. Yu, J.; Ding, W.B.; Wang, M.C.; Guo, X.G.; Xu, J.; Xu, Q.G.; Yang, Y.; Sun, S.H.; Liu, J.F.; Qin, L.X.; et al. Plasma circular RNA panel to diagnose hepatitis B virus-related hepatocellular carcinoma: A large-scale, multicenter study. Int. J. Cancer 2020, 146, 1754–1763. [Google Scholar] [CrossRef]
  145. Bahn, J.H.; Zhang, Q.; Li, F.; Chan, T.M.; Lin, X.; Kim, Y.; Wong, D.T.; Xiao, X. The landscape of microRNA, Piwi-interacting RNA, and circular RNA in human saliva. Clin. Chem. 2015, 61, 221–230. [Google Scholar] [CrossRef] [Green Version]
  146. Liu, B.; Song, F.; Yang, Q.; Zhou, Y.; Shao, C.; Shen, Y.; Zhao, Z.; Tang, Q.; Hou, Y.; Xie, J. Characterization of tissue-specific biomarkers with the expression of circRNAs in forensically relevant body fluids. Int. J. Leg. Med. 2019, 133, 1321–1331. [Google Scholar] [CrossRef]
  147. Salzman, J. Circular RNA Expression: Its Potential Regulation and Function. Trends Genet. 2016, 32, 309–316. [Google Scholar] [CrossRef] [Green Version]
  148. He, J.; Ming, Y.; MinLi, Y.; Han, Z.; Jiang, J.; Zhou, J.; Dai, B.; Lv, Y.; He, M.L.; Fang, M.; et al. hsa_circ_0006459 and hsa_circ_0015962 affect prognosis of Dengue fever. Sci. Rep. 2019, 9, 19425. [Google Scholar] [CrossRef] [Green Version]
Figure 2. Potential applications of circRNAs in antiviral therapeutics. The relatively more stable and less immunogenic nature of circRNAs compared to that of their linear cognate RNAs renders circRNAs a superior arsenal for potential antiviral therapeutics and a new biomarker for the detection of strenuously diagnosed diseases caused by viruses. (A) Circular RNA vaccines expressing the RBD of the SARS-CoV-2 spike protein demonstrated neutralizing antibody generation in mice and monkeys [52,55,130]. (B) Top panel: a circular RNA sponge containing miRNA-122 binding sites, sequestered miRNA-122, and suppressed HCV viral replication [48]; Bottom panel: antisense circRNAs targeting distinct regions of the 5′-UTR of SARS-CoV-2 that can efficiently inhibit virus replication [132]. (C) Top panel: EBV-encoded circRPMS1 and host hsa_circRNA_001387 that are significantly expressed in EBV-positive NPC tissues. Hence, they could be valuable biomarkers for the diagnosis and prognosis of NPC [133,134]; CircEAF2 inhibited EBV and large B-cell lymphoma progression via the miR-BART19-3p/APC/β-catenin axis, indicating it is a potential prognostic biomarker [99]. Bottom panel: four circular RNAs (hsa_circ_0018429, hsa_circ_0026579, hsa_circ_0125357, and hsa_circ_0099188) which were shown to be very sensitive and specific biomarkers for the diagnosis of community-acquired pneumonia (CAP) [135]. HCV: Hepatitis C virus; EBV: Epstein–Barr virus; CAP: community-acquired pneumonia. Created with BioRender.com (accessed on 4 August 2023).
Figure 2. Potential applications of circRNAs in antiviral therapeutics. The relatively more stable and less immunogenic nature of circRNAs compared to that of their linear cognate RNAs renders circRNAs a superior arsenal for potential antiviral therapeutics and a new biomarker for the detection of strenuously diagnosed diseases caused by viruses. (A) Circular RNA vaccines expressing the RBD of the SARS-CoV-2 spike protein demonstrated neutralizing antibody generation in mice and monkeys [52,55,130]. (B) Top panel: a circular RNA sponge containing miRNA-122 binding sites, sequestered miRNA-122, and suppressed HCV viral replication [48]; Bottom panel: antisense circRNAs targeting distinct regions of the 5′-UTR of SARS-CoV-2 that can efficiently inhibit virus replication [132]. (C) Top panel: EBV-encoded circRPMS1 and host hsa_circRNA_001387 that are significantly expressed in EBV-positive NPC tissues. Hence, they could be valuable biomarkers for the diagnosis and prognosis of NPC [133,134]; CircEAF2 inhibited EBV and large B-cell lymphoma progression via the miR-BART19-3p/APC/β-catenin axis, indicating it is a potential prognostic biomarker [99]. Bottom panel: four circular RNAs (hsa_circ_0018429, hsa_circ_0026579, hsa_circ_0125357, and hsa_circ_0099188) which were shown to be very sensitive and specific biomarkers for the diagnosis of community-acquired pneumonia (CAP) [135]. HCV: Hepatitis C virus; EBV: Epstein–Barr virus; CAP: community-acquired pneumonia. Created with BioRender.com (accessed on 4 August 2023).
Viruses 15 01697 g002
Table 1. List of reported virus-dysregulated circRNAs and their functions/mechanisms in regulating innate immunity or viral replication.
Table 1. List of reported virus-dysregulated circRNAs and their functions/mechanisms in regulating innate immunity or viral replication.
circRNAStimuliDifferential ExpressionFunctions/MechanismsReference
hsa_circ_0001400KSHVUpDuring KSHV de novo infections, circ_0001400 expression suppressed the expression of vital latent and lytic viral genes without significantly altering the viral genome copy number.[89]
circ_0000479HTNVUpCirc_0000479 sponged miR-149-5p and regulated RIG-I expression, thus dampening viral replication.[63]
hsa_circ_0004812HBVUpCirc_0004812 silencing enhanced the expression of IFN-α and β in HBV-infected Huh7 cells.[90]
circBACH1HBVUpCircBACH1 regulated HBV propagation through the miR-200a-3p/MAP3K2 pathway.[91]
circ-ATP5HHBVUpCirc-ATP5H boosted HBV replication by modulating the miR-138-5p/TNFAIP3 axis.[92]
circFNDC3BMERS-CoVUPThe silencing of circFNDC3B and circCNOT1 significantly suppressed the MERS-CoV viral load and its target mRNA expression, modulating various biological pathways, including the MAPK and ubiquitination pathways.[69]
circCNOT1
hsa_circ_0004445MERS-CoVUPThe knockdown of hsa_circ_0004445 inhibited MERS-CoV replication.[93]
hsa_circ_0000479SARS-CoV-2UPSARS-CoV-2 could regulate IL-6 and RIG-I activity via hsa_circ_0000479/hsa-miR-149-5p/RIG-I, IL-6axis.[64]
ssc_circ_009380 (circEZH2)TGEVDownCircEZH2 promoted the activation of NF-κB via sponging miR-22.[67]
circMerTKIAVUpCircMerTK inhibited IFN-beta production and suppressed IFN signaling, thus boosting IAV replication.[31]
circRNA AIVRIAVUpcircRNA AIVR inhibited IAV replication by predominantly absorbing miR-330-3p.[59]
circRNA_0050463IAVUpCircRNA_0050463 was found to sponge to miR-33b-5p and thereby enhanced IAV replication.[94]
circ-GATAD2AIAVUpCirc-GATAD2A promoted influenza virus multiplication by inhibiting VPS34-dependent autophagy in vitro[95]
hsa_circ_0005870IAVUPThe overexpression of these three circRNAs inhibited AIV replication and proliferation, whereas silencing these circRNAs enhanced AIV multiplication.[96]
hsa_circ_0006104
hsa_circ_0009365
circEXOSCHCVUPDepleting circEXOSC in HCV-infected cells markedly reduced viral infectivity.[97]
circTIALA significant reduction in HCV infectivity was observed after circTIAL silencing.
circPSD3HCVUPCircPSD3 promoted HCV RNA abundances at a post-translational level.
DENVUPCircPSD3 was found to reduce viral infectivity in Dengue virus-infected cells significantly.
ciTRANHIVUPHIV-1-Vpr-induced ciTRAN-sequestered SRSF1 from the HIV viral transcriptional complex to enhance viral transcription.[98]
circSIAECVB3DownCircSIAE suppressed CVB3 replication by targeting miR-331-3p and TAOK2.[65]
CircEAF2EBV DownCircEAF2 inhibited the replication of EBV and the progression of DLBCL via the miR-BART19-3p/APC/β-catenin axis.[99]
KSHV: Kaposi sarcoma-associated herpesvirus; HTNV: Hantaan virus; HBV: Hepatitis B virus; MERS-CoV: Middle East respiratory syndrome coronavirus; TGEV: Transmissible Gastroenteritis Virus; IAV: Influenza A virus; HCV: Hepatitis C virus; DENV: Dengue virus; HIV: Human immunodeficiency virus; CVB3: Coxsackievirus B3; EBV: Epstein–Barr virus.
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

Maarouf, M.; Wang, L.; Wang, Y.; Rai, K.R.; Chen, Y.; Fang, M.; Chen, J.-L. Functional Involvement of circRNAs in the Innate Immune Responses to Viral Infection. Viruses 2023, 15, 1697. https://0-doi-org.brum.beds.ac.uk/10.3390/v15081697

AMA Style

Maarouf M, Wang L, Wang Y, Rai KR, Chen Y, Fang M, Chen J-L. Functional Involvement of circRNAs in the Innate Immune Responses to Viral Infection. Viruses. 2023; 15(8):1697. https://0-doi-org.brum.beds.ac.uk/10.3390/v15081697

Chicago/Turabian Style

Maarouf, Mohamed, Lulu Wang, Yiming Wang, Kul Raj Rai, Yuhai Chen, Min Fang, and Ji-Long Chen. 2023. "Functional Involvement of circRNAs in the Innate Immune Responses to Viral Infection" Viruses 15, no. 8: 1697. https://0-doi-org.brum.beds.ac.uk/10.3390/v15081697

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