- Review
- Open access
- Published:
CircRNA as an Achilles heel of cancer: characterization, biomarker and therapeutic modalities
Journal of Translational Medicine volume 22, Article number: 752 (2024)
Abstract
Circular RNAs (circRNAs) are a class of endogenous noncoding RNAs characterized by their lack of 5′ caps and 3′ poly(A) tails. These molecules have garnered substantial attention from the scientific community. A wide range of circRNA types has been found to be expressed in various tissues of the human body, exhibiting unique characteristics such as high abundance, remarkable stability, and tissue-specific expression patterns. These attributes, along with their detectability in liquid biopsy samples such as plasma, position circRNAs an ideal choice as cancer diagnostic and prognostic biomarkers. Additionally, several studies have reported that the functions of circRNAs are associated with tumor proliferation, metastasis, and drug resistance. They achieve this through various mechanisms, including modulation of parental gene expression, regulation of gene transcription, acting as microRNA (miRNA) sponges, and encoding functional proteins. In recent years, a large number of studies have focused on synthesizing circRNAs in vitro and delivering them to tumor tissue to exert its effects in inhibit tumor progression. Herein, we briefly discuss the biogenesis, characteristics, functions, and detection of circRNAs, emphasizing their clinical potential as biomarkers for cancer diagnosis and prognosis. We also provide an overview the recent techniques for synthesizing circRNAs and delivery strategies, and outline the application of engineered circRNAs in clinical cancer therapy.
Introduction
CircRNAs are a novel class of RNAs characterized by covalently closed structures without 5′ caps and 3′ poly(A) tails. The first circRNA was discovered in viruses several decades ago but they initially received little attention [1, 2]. Although a handful of circRNAs were found afterwards, they were considered errors in RNA splicing with limited function [1, 3, 4]. With the advancements of RNA deep sequencing technology and dedicated bioinformatics algorithms, an increasing number of circRNAs have been reported in eukaryotic cells, especially in the mammalian brain [5, 6]. CircRNAs have been implicated in various diseases, including neurodegenerative diseases, atherosclerosis and cardiovascular diseases, especially cancer [7,8,9,10,11]. Studies have shown that circRNA exhibit aberrant expression patterns in multiple cancers [12,13,14]. Research has also revealed the high abundance, remarkable stability, and tissue-specific expression patterns of circRNA in vivo, which can be detected and quantified using RT‒qPCR, ddPCR, Northern Blotting, In Situ Hybridization and NanoString nCounter, making circRNAs an ideal diagnostic and prognostic biomarker for cancer [15,16,17]. Memczak et al. and Hansen et al. proposed the function of ciRS-7 as a miRNA sponge regulating gene expression in 2013 [18, 19], which garnered significant interest. Furthermore, subsequent research has uncovered additional functions of circRNAs, such as modulating parental gene expression, acting as miRNA sponges and encoding functional proteins [20, 21]. Certain circRNAs have been shown to play roles in initiation, progression and metastasis through their functions [18, 19, 22, 23]. Thus, circRNA-based cancer therapies have been extensively investigated. Linear RNAs are circularized in vitro and administered to specific tissues of patients, exerting an inhibitory effect on tumors. However, the development of engineered circRNA therapies faces challenges, including the need for highly efficient and high-throughput circRNA synthesis methods and effective delivery systems for large circRNAs. Therefore, optimizing circRNA synthesis methods and developing reliable delivery vehicles are essential for advancing the application of engineered circRNAs in cancer therapeutics.
In this review, we summarize the current understanding of circRNAs, including their biogenesis, characteristics, function and detection, as well as highlight their enormous potential as biomarkers for diagnosis and prognosis in cancer. Moreover, we offer an overview of the current knowledge regarding circRNA composition and delivery methods, and highlight the clinical application of circRNAs in cancer therapy.
Biogenesis of circRNAs
Based on their biogenesis, circRNAs can be classified into three primary categories [24]: exon-only circular RNAs (EcRNAs) [25], intron-only circular RNAs (ciRNAs) [21] and exon‒intron RNAs (EIciRNAs) [20]. Despite their diversity, circRNAs are generated from precursor mRNAs (pre-mRNAs) through back-splicing [26]. The mechanisms underlying circRNA biogenesis have not been completely elucidated. Back-splicing involves the 3’ splice donor site joining with the downstream 5′ splice site, resulting in a closed-loop structure with a 3′-5′ junction [27, 28]. Then, all or part of the introns are removed, and the rest of the sequences connect to form the three types of circRNAs. Three biogenesis models of circRNAs have been proposed: intron-pairing-driven circularization, RNA-binding protein (RBP)-induced circularization and lariat-driven circularization (Fig. 1A). In intron-pairing-driven circularization, introns on either side of exons are essential elements. The upstream introns are base-paired with the downstream introns, forming a hairpin and promoting the formation of circRNAs. These reverse complementary sequences include ALU repeats and other nonrepetitive elements [16, 28]. Additionally, certain RNA-binding proteins can interact with specific intron motifs to introns to form hairpins. In the lariat intermediate model, canonical splicing removes exons and introns from the pre-mRNA, creating a lariat structure. The intronic sequences are then removed in the internal splicing to form an EcRNA [29]. Intronic lariats that escape debranching can form ciRNAs [21]. Moreover, a small number of circRNAs originate from tRNAs that are spliced from precursor tRNAs (pretRNAs) [30]. During tRNA maturation, the tRNA splicing endonuclease complex cleaves pretRNAs, and a single enzyme called RtcB joins the exon halves and introns to form a tRNA intronic circular RNA (tricRNA) [31, 32].
Characteristic of circRNA
Stability
CircRNAs are exceptionally stable molecules with higher stability than linear transcripts [33]. With the structure of a covalently closed loop structure without the 5′ cap and 3′-poly tail, circRNAs demonstrate increased resistance to RNase-mediated cleavage [34, 35]. For instance, EcRNAs are stable in cells with a half-life of 48 h, whereas the average half-life for mRNA is 10 h [35]. High stability can result in the accumulation of circRNAs during various physiological processes [15, 36]. The enhanced stability of circRNAs facilitates their detection and quantification, often at higher levels than their linear counterparts. This makes circRNAs particularly suitable as biomarkers for disease diagnosis.
Abundance
With the extensive application of RNA sequencing technology, circRNAs have been found to be abundant and widespread [37]. Studies have shown that circRNAs are highly prevalent in human tissues, particularly in the human brain [38]. More than 25,000 different circRNAs have been found in human fibroblasts, which are ten times more abundant than their linear counterparts [16, 23, 39]. Alternative circularization between different flanking complementary introns results in a gene forming many circRNA isoforms, thereby contributing to the overall diversity and abundance of circRNAs [28].
Temporal and spatial expression specificity
The expression patterns of circRNAs are spatiotemporally specific, varying at different stages of growth, development and tumor progression. CircRNA expression patterns are strictly regulated in eukaryotic cells and correlated with neural development [6, 36]. Moreover, circRNAs play a crucial role in several hallmarks of tumor progression. For instance, at the origin of liver cancer, circ-ZNF609 enhances the stemness of hepatocellular carcinoma (HCC) cells and promotes carcinogenesis [40]. Conversely, circMEG3 is downregulated during the invasion and metastasis of HCC, reducing its inhibition of telomerase activity and promoting cell replicative immortality and tumor metastasis [41]. Similarly, circRNAs exhibit differential expression in tumor tissues and adjacent non-malignant tissues [42, 43]. An analysis was performed to illustrate the circRNA profile of prostate adenocarcinoma and normal cells, bone and osteosarcoma, colon and colorectal adenocarcinoma, and kidney and renal cell carcinomas [44]. The results showed that the total circRNA abundance was lower than that in the normal tissues. Differential expression analysis revealed 34 upregulated circRNAs with statistical significance. These distinct expression patterns underscore the potential of circRNAs as biomarkers for cancer.
The aforementioned characteristics make circRNA an ideal tumor diagnostic and prognostic markers. In clinical practice, the expression levels of specific circRNAs can be measured to determine the presence of a tumor or assess the degree of tumor progression.
Functions of circRNAs
A large number of circRNAs have been identified by genome-wide analysis, and their unique characteristics suggest that they may probably play an essential role in biological processes. To date, sevaral possible functions have been proposed that have broadened our understanding of circRNAs, such as transcriptional regulation, miRNA sponges, and translation into proteins. However, additional functions of circRNAs remain to be discovered.
Transcriptional regulator
Although most circRNAs are located in the cytoplasm, ciRNAs and EIciRNAs are restricted to the nucleus and function as transcriptional regulators [20] (Fig. 1B). For instance, EIciRNAs can interact with U1 small nuclear ribonucleoproteins to form the EIciRNA-U1snBNP complex, which associates with polymerase II at the parental gene promoters to enhance their expression. Circ-EIF3J and circ-PAIP2 enhance the expression of their parental genes in HeLa and HEK293 cells by promoting the elongating activity of polymerase II [20]. Moreover, circSEP3, located in exon 6 of SEP3, can bind to its cognate DNA locus and form an RNA‒DNA hybrid, resulting in transcriptional pausing. This pausing subsequently recruits splicing factors, leading to the formation of alternatively spliced SEP3 mRNA with exon skipping [45]. Therefore, these intron circRNAs function as transcriptional regulators of parental gene transcription.
MiRNA sponge
MicroRNA (miRNA) sponges are considered one of the most critical functions of circRNAs. MiRNAs are thought to be key regulators of gene expression by base pairing with miRNA response elements of mRNAs and regulating their translation in physiological and pathological processes [46, 47]. CircRNAs with complementary binding sites can sequester miRNAs away from their target mRNAs, thereby relieving the suppression of miRNA-underlying genes (Fig. 1C). For example, ciRS-7 is a circRNA that contains more than 60 conserved miR-7 binding sites, enabling it to stably bind to miR-7. This interaction suppression of the biological activity and function of miR-7 and elevated expression of miR-7 targets [48]. Moreover, studies have showed that there is no one-to-one correspondence between miRNAs and circRNAs. Single circRNAs can bind to multiple miRNAs and perform different functions. For example, circHIPK3 can sponge miR-558 to suppress heparanase expression in bladder cancer (BCa) and can also bind to miR-124-3p to promote lung cancer [49, 50].
Translation into proteins
CircRNAs were long considered to be untranslatable due to the absence of elements necessary for translation, such as the poly(A) tail and 5′ cap. However, endogenous circRNAs with internal ribosome entry sites (IRESs) have been found to be translated via cap-independent translation initiation mechanisms (Fig. 1D) [51]. Additionally, N6-methyladenosine (m6A) modification can also drive the translation of circRNAs [52]. In the former mechanism, the ribosome can directly bind to the IRES to initiate the translation of circRNAs. Circ-ZNF609 with IRES can undergo cap-independent translation, whereas circARHGAP35 can undergo m6A-dependent translation. Moreover, these two methods can cooperate to improve the efficiency of translation [53]. Yang et al. suggested that circRNA translation increased under heat shock conditions, indicating that circRNA-encoded proteins may play a role in stress conditions [54]. While most functions of peptides translated from circRNAs remain unknown, and some peptides lack essential functional domains, further studies are needed to elucidate the detailed mechanisms of circRNA translation and the roles of circRNA-encoded proteins.
CircRNAs can intervene at various stages of tumor development through the aforementioned functions, and through artificially modulating of circRNAs, the goal of treating tumors can be achieved. CircRNA-based therapeutics offer a potential avenue for cancer treatment, as circRNAs serving as miRNA sponges could inhibit oncogenic circRNAs. Several circRNAs have been identified as potential therapeutic targets. Overexpression of circRNAs that inhibit tumor growth or silencing of circRNAs that play a positive role during cancer development can exert opposite effects, thereby suppressing tumor progression.
Potential as a biomarker
The National Cancer Institute has defined biomarkers as “a biological molecule found in blood, other body fluids, or tissues that is a sign of a normal or abnormal process, or of a condition or disease,” outlining several essential characteristics as biomarkers (Fig. 1E). An ideal biomarker should be specific, sensitive, detectable and predictable in biofluids such as plasma, saliva, urine or even in exosomes [55]. As mentioned earlier, circRNA fulfill these criteria. Numerous studies have confirmed that many different circRNAs are aberrantly expressed in tumors compared to the surrounding normal tissue and are associated with clinical tumor characteristics, such as location, tumor size and TNM stage [56,57,58,59]. To date, many circRNAs have been identified as potential cancer diagnostic and prognostic biomarkers, both individually and synergistically with known cancer markers [60]. These circRNAs have been found to be stably present at high levels not only in tissues and cells but also in body fluids, such as plasma, serum, exosomes, and urine [61]. Their abundance facilitates their detection in detecting them in biological liquids. Because most circRNAs can be detected in biofluids, liquid biopsy has emerged as a revolutionary tool for cancer management, diagnosis and treatment guidance. Liquid biopsy is a real-time, non-invasive method for early tumor detection, comprehensive tumor characterization beyond localized malignancies, and dynamic monitoring of disease progression [62] (Fig. 2). Liquid biopsy be combined with tissue biopsy or may even replace it in the future [63]. Clinical studies have applied circRNAs as diagnostic and prognostic markers for tumors (Table 1) (https://clinicaltrials.gov/), most of which are still in the “early adopter” phase. Extensive research is needed to establish circRNA expression patterns in diverse cancer populations and identify reliable diagnostic or prognostic biomarkers.
Detection and quantification of circRNA
Studies have demonstrated the aberrant expression of circRNAs in the cytoplasm and exosomes of various cancer cells. Therefore, early detection, validation, and quantification of abnormally expressed circRNAs could enable earlier diagnosis and treatment of cancer. Here, we discuss several methods for detecting the expression levels of circRNAs and the advantages of these different methods.
RT‒qPCR
Considering their sensitivity, specificity, speed and cost-effectiveness, RT‒qPCR-based techniques are the most widely utilized methods for the detecting and quantifying circRNAs in clinical laboratories [64]. CircRNAs in tissues or biofluids can be extracted using an RNA preparation protocol. Following RNase R treatment to degrade coexpressed linear RNAs, circRNAs were enriched and isolated [65]. Because of tissue and development specificity, primers for reverse transcription were designed to target the specific circRNA of interest. During reverse transcription, complementary DNA (cDNA) is synthesized from the circRNA, containing exon-exon junction sequences that can be specifically amplified by the primers. DNA without the junction sequence is not amplified. Fluorescence signals were then detected to quantify circRNA levels. RT‒qPCR assays with RNase R play a significant role at the medium-throughput scale and can be quantified by real-time PCR (Fig. 2A). However, for the circRNAs derived from a single locus, a single primer can amplify several different amplicons. It is imperative to verify the amplicons such as the melting curves.
Nevertheless, a major source of error in RT-qPCR is the reverse transcription step [66]. Amplification errors might occur in this process, such as template switching [67] and rolling circle amplification [68], resulting in false-positive error and overestimation of circRNA concentration by 34–55% [69]. Furthermore, these errors can be avoided by reverse transcription-droplet digital polymerase chain reaction (RT-ddPCR) and hybridization techniques.
ddPCR
Drolet digital PCR is a new generation of traditional quantitative polymerase chain reaction (qPCR) and is used for the absolute quantification of target nucleic acids. To address the challenge of rolling circle amplification, Chen et al. developed a method based on RT-qPCR [70]. By employing a Poisson distribution-based algorithm, this approach eliminates the effects of rolling circle amplification. Similar to RT-qPCR, ddPCR uses the same primers, probes and reagents as qPCR. RNA was isolated and enriched from fluids, but unlike RT-qPCR, RNA was not digested by RNase R. Reverse transcription is performed, followed by degradation of the RNA strand in the DNA-RNA duplex by RNase R. Subsequently, ddPCR is conducted, where cDNAs are encapsulated within droplets, and thermal cycling is initiated. Almost all the droplets contained one or fewer copies of the DNA. After the PCR routine, concentrations were determined based on the proportion of nonfluorescent partitions by Poisson distribution [71] (Fig. 2B). This method shows better performance in the accurate quantification of trace amounts of nucleotides, with higher sensitivity compared to RT‒qPCR [72]. Its sensitivity makes it suitable for detecting targets with low copy numbers without preamplification steps. Additionally, while the qPCR method relies on fluorescent dye intensity proportional to PCR amplicons and requires a standard curve for quantification, ddPCR does not rely on standard curves, thereby avoiding associated errors [73]. Hu et al. utilized this assay to demonstrate that upregulated expression of circGSK3β was positively associated with advanced stage and poor outcome of esophageal squamous cell carcinoma (ESCC) and identified that the circGSK3β expression level in plasma is a biomarker for the detection of ESCC [58]. However, ddPCR requires automation and a cost reduction for widespread adoption. If accepted by a clinical laboratory, the future of this is promising.
Northern blotting
PCR assays using RNase R offer a straightforward and rapid methods for circRNA detection. Nevertheless, there are some limitations to PCR assays. Linear RNAs with stable structures may not be fully digested by RNase R, and large circRNAs may not be absolutely resistant to RNase R [74]. The northern blot hybridization method has the advantage of detecting circRNAs without the digestion step and reversing the transcription step. In this method, hybridization begins with labeled probes binding to the complementary back-splicing junction of circRNAs. Following hybridization, denatured RNAs are separated by agarose gel electrophoresis and transferred to a nitrocellulose membrane via capillary blotting [75]. Circular and linear RNA can be detected simultaneously because of their different migration rates in the gel. Luminescence detection of the hybridized probes enables profiling of the hybrid signals (Fig. 2C). The probe design is the most crucial step in the entire experiment. To specifically detect a circRNA, the probes designed have two options: a short probe spanning the back splice junction (BSJ) or an exonic probe directed against a region within the circRNA. Despite its advantages, the northern blotting is more laborious and sensitive, make it more suitable for low-throughput identification of putative circRNAs. Many attempts have been made to save time and improve sensitivity. For example, the digoxigenin-labeled probe method can increase sensitivity and safety and shorten time [76]. Nevertheless, northern blotting still suffers from several drawbacks, such as the requirement for good-quality RNA samples and its limited accuracy for the quantitation of circRNAs.
In situ hybridization (ISH)
ISH assays are used to detect specific RNAs within cells and localize them at the subcellular level [77]. Fixed cells and tissues were exposed to labeled DNA probes at high concentrations to form stable DNA‒RNA strands. DNA probes with fluorescence signals can then be visualized and profiled by microscopy (Fig. 2D). It is challenging to design probes that are restricted to the BSJ sites and are more difficult to detect because of the lower sensitivity, RNA degradation, competition for secondary structures and so on. There are two main types of probes used for circRNA detection: one is binding on both sides of BSJ [6] and another is a probe set covering the entire length of the RNAs [78]. Probes that complement both sides of the BSJ sites contain approximately 20 nucleotides and can detect circRNAs even in the presence of their linear counterparts [19]. BaseScope is a method that can specifically detect circRNA [79]. After binding to BSJ, the signal can be amplified, marking the localization of circRNA. This method is compassionate, but has a low-throughput. The probe sets covering the entire length of the RNAs contain 30–50 individual 20-nucleotide singly labeled probes [78] and could bind different parts of the target RNA. Multiple probes nave been designed for this study. Each probe was labeled with a unique fluorophore and pooled before detection [77]. However, this method is limited to detecting circRNAs in the absence of their linear counterparts. ISH can provide the precise spatial position of circRNAs within cells and tissues, but is limited in its ability to detect multiple circRNAs simultaneously, which restricts its use in routine clinical laboratories.
NanoString nCounter
The NanoString nCounter is an emerging technology for quantifying circRNAs. This method utilizes a combination of capture probe and reporter probe, both of which can specifically bind sequences of different RNAs, such as the BSJ of circRNAs (Fig. 2E) [80]. The identification of different transcripts depends on the arrangement and combination of a string of fluorophores in the reporter probes. In this method, probes and RNA samples are fixed together, and RNA and its corresponding probes hybridize to form a complex. After hybridization, unbound probes are washed away, and the remaining complexes are detected and quantified using a charge-coupled device (CCD) camera. This method is suitable for high-throughput detection and can detect multiple circRNAs simultaneously. It has no high requirements for sample quality, and this method can work well, although the RNAs are formalin-fixed, paraffin-embedded samples [81]. However, each probe needs to be specifically designed for different circRNAs to ensure accurate detection.
CircRNA as a diagnostic biomarker
Biomarkers, such as AFP and CEA, have long been used for the diagnosis. However, conventional biomarkers often suffer from low specificity and sensitivity, and the absence of early symptoms in some cancers can lead to diagnosis at an advanced stage, thereby missing the optimal window for treatment [82]. The aforementioned characteristics of circRNAs address these limitations and they have potential as biomarkers for early detection. For instance, Xu et al. constructed a panel of five circRNAs using a liquid biopsy assay, which could robustly distinguish between patients with early-stage pancreatic ductal adenocarcinoma (PDAC) and those without diseases [83] (Table 2). The panel achieved an AUC of 0.81 for identifying early PDAC in the validation cohort, which was the highest AUC value when compared with other gastrointestinal cancers, underscoring its specificity for PDAC. Wei et al. identified a plasma circRNA panel containing three circRNAs that demonstrated higher accuracy than AFP in distinguishing individuals with early stages HCC from healthy individuals [84]. Logistic regression analysis indicated that it was independent markers for discrimination of early HCC tissues. In non-small cell lung cancer (NSCLC), hsa_circ_0014130 in plasma is associated with clinicopathological outcome [85]. Elevated levels of hsa_circ_0014130 in tumor accurately identify NSCLC tissue and the adjacent non-cancerous tissues with the sensitivity and specificity of 87.0% and 84.8%, respectively. Moreover, Wang et al. discovered that the abnormal expression of different circRNAs can not only detect NSCLC in the early stage, but also distinguish between lung adenocarcinoma (LUAD) and squamous cell carcinoma [86]. Using qPCR validation, four circRNAs were identified as potential NSCLC biomarkers. Hsa_circ_0077837 and hsa_circ_0001821 showed diagnostic potential, with AUCs of 0.921 and 0.863, respectively, for distinguishing NSCLC from healthy tissues. Moreover, the upregulation of hsa_circ_00010739 (AUC = 0.919) and downregulation of hsa_circ_0001495 (the AUC was 0.919) could predict pathological subtyping between LUAD and squamous cell carcinoma. In osteosarcoma, hsa_circ_0081001 emerged as an independent diagnostic factor, and its expression level can dynamically monitor changes in the illness of patients [87]. Serum samples were collected from five newly diagnosed patients at each treatment stage, revealing that RNA levels in chemoresistant patients increased before surgery, decreased post-surgery, and rose again after postoperative chemotherapy, while chemosensitive patients consistently maintained low levels throughout treatment. The ROC curve indicated that hsa_circ_0081001 (AUC = 0.898) outperformed traditional markers such as ALP (AUC = 0.673) and LDH (AUC = 0.80) in diagnosing overall survival (OS). In addition, the combination of circRNAs with traditional biomarkers can increase sensitivity and specificity. In gastrointestinal tumors, circCCDC66 promotes cancer cell proliferation, migration and presents differential expression pattern in colorectal cancer tissues, showing a differential expression pattern in colorectal cancer tissues [59]. The study demonstrated that the expression level of circCCDC66 was a good predictive biomarker for colorectal cancer detection. Moreover, circRNAs in exosomes provide a new way for cancer cells to communicate with other cells [88], exosome-derived circRNAs can also be novel diagnostic biomarkers. Tang et al. found that circRNAs encapsulated in exosomes demonstrated comparable detection power for gastric cancer (GC) as plasma circRNAs [89]. CircKIAA-1244 can be a potential biomarker for cancer diagnosis, with univariate and multivariate analyses suggesting its independence as a diagnostic indicator for GC. The AUC for distinguishing GC plasma and healthy plasma is 0.748 with the sensitivity of 77.42% and specificity of 68.00%. In addition, numerous other studies have also demonstrated the diagnostic potential of circRNA (Table 2).
CircRNA as a prognostic biomarker
Accurate cancer prognostication can provide doctors with a reference for follow-up treatment and guide treatment decisions for patients, potentially improving patient outcomes. Numerous single circRNAs or circRNA panels have been proven to have superior prognostic power (Table 3) [60, 90,91,92]. Among these, ciRS-7 and circ ITCH may be the most well-known circRNA among these circRNAs and has been associated with poor prognosis in various cancers, including NSCLC [93], colorectal cancer (CRC) [60], GC [94] and other cancers. CiRS-7 is highly abundant in many cancers and functions as an oncogene via the ciRS-7/miR-7 axis [95]. MiR-7 plays a suppressive role in tumor progression and can inhibit the proliferation and migration of cancer cells through various mechanisms [96]. As ciRS-7 acts as a sponge for miR-7, its high expression indicates a poor prognosis. To validate the prognostic value of the ciRS-7, Tian et al. [90] conducted a meta-analysis involving 1714 patients from multiple nations, and their ciRS-7 expression levels were measured using RT‒qPCR. The results showed that upregulated ciRS-7 expression was significantly correlated with worse OS and higher tumor stage in NSCLC and CRC [93]. Another important circRNA that has been proven to have a prognostic role is circ ITCH [97]. Circ ITCH functions as a well-known tumor suppressor in multiple cancers via various pathways. In BCa, a positive correlation exists between tumor histological grade and circ-ITCH expression. Patients with higher circ-ITCH expression had better OS than those with lower expression levels. Circ-ITCH can inhibit BCa progression, including proliferation, invasion and tumorigenesis, in vivo and in vitro through the circ-ITCH/miR-17, miR-224/p21, and PTEN signaling regulatory networks. Luo et al. found that circ-ITCH expression in ovarian cancer was reduced than that in paired adjacent tissues [91]. Univariate Cox regression analysis indicated a positive correlation between circ-ITCH levels and prolonged OS, while multivariate analysis identified it as an independent predictor of favorable OS. Moreover, other circRNAs were also found to be a potential prognostic biomarker for the clinical outcome. In PDAC, high expression levels of circ-PDE8A and its linear counterpart were noted [98]. Upregulation of circ-PDE8A was associated with lymphatic invasion (p = 0.014) and higher TNM stages (p = 0.005), while linear-PDE8A was not significantly associated with patient survival. In gastric cancer, circPVRL3 was downregulated, and this downregulation promoted cell proliferation and migration [99]. Reduced expression of circPVRL3 was negatively correlated with survival time, suggesting it may be an ideal biomarker for predicting medical prognosis.
CircRNA-based therapeutic approach
CircRNAs play vital roles in cancer development such as proliferation, invasion and metastasis through various pathways (Fig. 3), with their expression levels dysregulated in cancer cells [100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119]. CircRNAs regulate tumorigenesis through specific signaling axes. Firstly, in the circRNA-miRNA-mRNA axis, circRNAs function as miRNA sponges, sequestering miRNAs and preventing them from repressing their target mRNAs, leading to oncogene upregulation and enhanced tumor proliferation. Secondly, the circRNA-RBP axis involves circRNAs binding to RNA-binding proteins, modulating their activity and influencing processes such as cell cycle progression. Thirdly, within the circRNA-mRNA axis, circRNAs can directly interact with mRNAs to regulate their stability and translation, affecting gene expression relevant to cancer. Lastly, the circRNA-protein axis sees circRNAs encoding functional proteins that can contribute to tumorigenesis. These diverse signaling pathways underscore the multifaceted roles of circRNAs in cancer biology, highlighting their potential as therapeutic targets and biomarkers. These findings open new avenues for treating cancer treatment, suggesting that regulating the levels of aberrantly expressed circRNAs may be an approach to cure cancer. Studies have demonstrated the potential of circRNAs as novel therapeutic agents [120, 121]. Synthetic and engineered circRNAs have been shown to have superior stability and strong translation in eukaryotic cells [122, 123]. The synthesis, purification and delivery of engineered circRNAs is attracting researchers' attention.
Synthesis of therapeutic circRNA
Chemical molecule circularization strategy
The circularization of RNA involves an intramolecular reaction between the 3′ and 5′ terminal nucleotides of linear RNA. Various molecules, such as amines and thiols, have been proven to facilitate RNA cyclization in vitro [124]. The phosphate and hydroxyl groups at the 5′ or 3′ terminus of linear RNA were replaced by other functional groups for cyclization. Ethyl-3-(3′-dimethylaminopropyl)-carbodiimide (EDC) and cyanogen bromide (BrCN) are the most commonly used reagents. However, these chemical strategies have several drawbacks. One significant issue is the formation of linear products through intermolecular ligation, which are the main byproducts. Reducing the concentration of linear RNA could reduce the occurrence of intermolecular reactions, but this approach negatively impacts the efficiency of cyclization and is unsuitable for large-scale applications. Oligonucleotide splints that can bring the two ends closer together can promote intramolecular ligation [125]. Moreover, the formation of 2′,5′-phosphodiester bonds at the ligation junction, rather than the desired 3′,5′-phosphodiester bonds, can also leads to the generation of byproducts (Fig. 4A).
Enzymatic circularization strategy
The enzymatic strategy for RNA circularization utilizes various nucleotide ligases to form phosphodiester bonds at both ends of linear RNA, thereby achieve cyclization [126, 127]. Commonly used enzymes include T4 DNA ligase, T4 RNA ligase 1 and T4 RNA ligase 2, all of which are from the bacteriophage T4 [127, 128]. For enzymatic ligation to occur, the linear RNA must have a 3′-OH and a 5′-monophosphate; otherwise, the 5' end must be phosphorylated with enzymes before the cyclization reaction can proceed. The function of T4 DNA ligase is to form connections and cyclize at the junction sites of double-stranded RNA, which requires the two ends of the substrate RNA connection to be perfectly complementary. The substrate of T4 RNA ligase 1 is single-stranded RNA, which can form 3′,5′-phosphodiester bonds in RNA and form circRNAs. The secondary structure of RNA is the key factor affecting cyclization efficiency, and T4 RNA ligase 1 requires the 3′ and 5′ termini to be in close proximity. Splints can effectively perform this function and bring the two ends of linear RNA into close proximity. T4 RNA ligase 2 is effective in cyclizing both single-stranded and double-stranded RNA, though it is more efficient at joining nicks in double-stranded RNA. Similar to chemical strategies, intermolecular reactions are the most important side reactions. The generation of byproducts can be minimized through helper oligonucleotide splints. However, enzymatic methods struggle with the ligation of large RNA molecules. In such cases, a permuted intron–exon self-splicing system (PIE strategy) may offer a more suitable approach (Fig. 4B).
Ribozyme circularization strategy
The chemical and enzymatic methods mentioned above are suitable for synthesizing smaller circRNAs. However, for larger molecules, alternative approaches are required. The ribozyme circularization strategy is effective for producing larger circRNAs [123]. The Group I Permuted Intron–Exon (PIE) method is the most commonly used ribozyme strategy. The 5′ half of the intron is seated at the tail of the exon, and the 3′ half is positioned at the head of the exon. Through self-splicing, the 5′ half of the intron is moved to the tail of the exon, and the remaining 3′ half is transferred to the head of the exon. After two transesterifications reactions at the defined splice sites, circRNAs are formed from a 3′–5′ phosphodiester bond. Group II introns, which can connect the 5′ and 3′ ends of an exon through inverse splicing reactions, can also be used for circRNA synthesis. However, group II introns form 2′,5′-phosphodiester bonds at the ligation site. Owing to the relatively simple reaction conditions and high efficiency of this method, the PIE method is the most widely applied method. However, there are still some limitations. First, the Group I intron method leaves a splicing scar in the ligation, which can induce immunogenicity [129]. Second, the RNA secondary structure could influence the efficiency of circularization and lead to the generation of byproducts (Fig. 4C).
Delivery of circRNA
The delivery of synthetic circRNAs primarily depends on viral and nonviral delivery systems (Fig. 4D). The unique terminus-free structure makes them resist degradation by exonucleases during their delivery at the cellular level. Additionally, their compact structure enhances the capacity of the delivery carriers.
Nanoparticles
Nanoparticles are nonviral delivery carriers composed of phospholipids, cholesterol and ionizable lipids. They are engineered to deliver different molecules for different applications such as imaging and therapeutic targets [130, 131]. Their sizes and compositions can be tailored for specific uses. Owing to its low production cost, it can be easily produced, transported, stored, and transported in large quantities, making them an effective carrier for circRNA [132]. In cancer therapy, nanoparticles can encapsulate circRNAs and target specific cancer cells. Upon reaching the target, the membrane ruptures, releasing the circRNA into the cytosol where it can exert its function. For example, a study demonstrated that delivering nanoparticles encapsulating siRNA targeting circDnmt1 significantly suppressed breast cancer tumor growth and extended the lifespan of mice [133]. Additional studies have shown the therapeutic potentials of nanoparticles. Nanoparticles that deliver circFoxo3 plasmids increase tumor cell apoptosis and delivered circEHMT1 plasmids inhibit lung metastasis of breast cancer [134, 135].
Exosomes
Exosomes are vesicles secreted by a range of cell types, such as T cells and cancer cells, carrying diverse molecules such as circRNAs and playing a crucial role in intercellular communication. Characterized by a lipid bilayer membrane with embedded signal receptors, exosomes isolated from cells exhibit innate bioactivity and biocompatibility, functioning effectively without further modification [132, 136]. Similar to nanoparticle, exosomes could protect their content from degradation, promote targeted cells endocytosis and release circRNAs to the cytosol by endogenous mechanism. Therefore, exosomes are recognized as ideal delivery vehicles for circRNAs. Currently, researchers have designed an exosome delivery system that could deliver mRNA vaccination [137]. As for cancers, there are almost tenfold more exosomes in tumor cells than in other cells. Thus, extracellular vesicles seem to be ideal tumor RNA carriers [138]. Wang et al. [139] found that ciRS-122 expression is a primary cause of oxaliplatin resistance in colorectal cancer (CRC), with cancer cells secreting exosomes containing ciRS-122 to spread drug resistance. Isolated exosomes from cancer cells were injected isolated exosomes into chemotherapy-sensitive mice and tumor volume did not shrink significantly after treatment with oxaliplatin. This study also used the natural carrier system to deliver siRNA targeting ciRS-122 in mice, suppressing its expression and enhancing chemotherapy sensitivity. In summary, the biocompatibility, stability in circulation, and permeability of biological barriers make exosomes ideal circRNA carriers for cancer therapy.
Viral vectors
Viral vectors, particularly adeno-associated viruses (AAV) and adenoviruses (Ads), are highly efficient in delivering circular RNA (circRNA) into various cell types, including cancer cells [140]. They enable stable and long-term expression of circRNA, with lentiviruses integrating into the host genome for prolonged expression and AAV providing stable episomal expression. Viral vectors also offer tissue-specific delivery through the selection of different serotypes or specific promoters, high transfection efficiency compared to physical or chemical methods, and precise regulation of circRNA expression using inducible elements [141, 142]. These features make viral vectors a crucial tool for delivering circRNA in basic research and therapeutic applications. In a previous study, a lentiviral vector was used to deliver circZKSCAN1 successfully. Moreover, the results showed that overexpression of circRNA repressed HCC growth in vivo and in vitro [143]. However, a disadvantage is that the vector may produce some unnecessary linear byproducts.
The antitumor application of circRNA
Gene silencing therapy
The positive roles of some circRNAs in tumors make them potential targets for cancer therapy. Several strategies have been developed to exploit this potential. One common approach is to use RNA interference (RNAi), which includes small interfering RNA (siRNA) and short hairpin RNA (shRNA), which target the back-spliced sequence of circRNA to induce cleavage. SiRNAs are long dsRNAs that target circRNAs via complementary pairing, leading to their incorporation into the RNA-induced silencing complex (RISC) for cleavage [144]. shRNAs, which have a loop and base-paired stems, can be converted into siRNAs. By specifically targeting the back-spliced sequence of circRNA, the circRNA can be knocked out, but does not involve its corresponding linear mRNA. In HCC, siRNA and shRNA targeting the specific sequence of the oncogenic circRNA Crd1as can inhibit its expression, thereby suppressing HCC cell proliferation and invasion [145]. Similarly, in lung adenocarcinoma (LUAD), Qiu et al. found that circPRKCI, which promotes tumor growth and can be silenced by E2F7, becomes a therapeutic target when transfected with small interfering circPRKCI, resulting in reduced tumor size. These findings highlight the therapeutic potential of targeting circRNAs in cancer treatment [146]. However, this strategy has some limitations, such as degradation by nucleases, lack of cell specificity, and off-target effects [131, 147]. To mitigate these issues, researchers have explored the use of circular sense RNA and its complementary linear RNA, which can reduce off-target effects due to their unique circular structure. Linear antisense RNA can form a complex with RISC, and caged circular siRNA has been designed based on this complex [148]. In a xenograft tumor model with U87-GFP cells in mice, the complex was found to induce GFP gene silencing, demonstrating its potential [149]. However, unpredictable off-target pathological effects are the main obstacle for siRNA to become a viable treatment. Since circRNAs may play roles in multiple tissues and diseases, it is crucial to understand their effects throughout the body and design siRNA specifically to minimize off-target effects before clinical application.
Gene editing therapy
The CRISPR-Cas9 system provides a precise method for genome editing, including the targeted disruption of circRNAs by interfering with the pairing of introns flanking the circularized exons [150, 151]. The CRISPR/Cas system has been proven to knockdown circRNAs with greater specificity than RNAi [152]. Zheng et al. utilized the CRISPR/Cas9 technique for the knockdown of circHIPK3 expression, resulting in suppressed cell proliferation in various human and cancer cell lines [23]. The first human trial of CRISPR-Cas9 was conducted by Cyranoski et al., where metastatic non-small cell lung cancer (NSCLC) patients who had not responded to chemotherapy, radiation, or other therapies were treated with CRISPR gene editing [153]. The trial used CRISPR-Cas9 to target the PD-1 gene in T cells, with recent results indicating the feasibility of this approach for patients with advanced lung cancer [154]. Meanwhile, clinical trials are ongoing for CRISPR-mediated PD-1 gene knockout in other cancers, including hepatocellular carcinoma (HCC) and advanced esophageal cancer [155]. In the context of HPV-related cervical cancer, where HPV E6 and E7 proteins play crucial roles in malignant transformation and maintaining the malignant phenotype, CRISPR/Cas9-mediated knockdown of these genes led to the death of HPV-16-positive cervical cancer cells and inhibited tumor growth in a nude mouse model [156].
Vaccine therapy
Recently, tumor vaccines have emerged as a novel and promising strategy for cancer therapy. Cancer vaccines can stimulate antitumor responses by expressing tumor-associated antigens, leading to the recruitment of immune cells, activation of T cells, and ultimately the destruction of tumor cells and the formation of sustained immune memory [157]. Recent studies have demonstrated that circRNAs can be translated into small peptides or proteins [158, 159]. Engineered circRNAs often exhibit consistently high levels of protein expression, resulting in prolonged antigen exposure, sustained T-cell activity, and durable immunological memory [123]. A comparison of a circRNA SARS-CoV2 RBD vaccine with an mRNA vaccine revealed that the circRNA vaccine exhibited higher and more durable expression levels [160]. Moreover, the high stability and low immunogenicity of circRNA vaccines perfectly compensate for the limitations of mRNAs, positioning them the next blockbuster of cancer therapy. Despite their promise, reports on circRNA vaccines in cancer therapy remain limited. A study published in 2022 demonstrated the positive role of circRNA vaccines in the prevention and treatment of melanoma [161]. In this study, circRNA was circularized in vitro using a Group I PIE system, purified via liquid chromatography, and encapsulated in lipid nanoparticles (LNPs) to form a stable vaccine. This vaccine was successfully delivered into cells and translated into proteins. Following intramuscular injection into mice, there was a notable increase in proinflammatory cytokines and enhanced antigen-specific T-cell responses, indicating effective induction of both innate and adaptive immune responses. In immune-desert orthotopic and metastatic melanoma models, the vaccine exhibited superior antitumor efficacy, effectively inhibiting tumor formation and blocking lung metastasis. The antitumor response induced by circRNA vaccines is mainly exerted by cytotoxic T cells rather than antibody-based B-cell responses. Therefore, to enhance therapeutic outcomes, researchers have combined circRNA vaccines with adoptive cell transfer therapy, showing improved persistence of engineered T cells and superior efficacy compared to monotherapy in targeting late-stage melanoma. This evidence underscores the potential of circRNA vaccines as a viable alternative in vaccination strategies and highlights their promising role in cancer therapy. However, the role of vaccines against other malignancies remains unclear. The function of circRNA vaccines in these cancers remains to be explored.
To date, therapies that target circRNAs have not yet been investigated in clinical studies. Several challenges must be addressed before these approaches can be applied in humans, and the development process is complex and time-consuming.
Conclusion and perspective
CircRNAs, initially regarded as mere byproducts of mRNA splicing, have emerged as a significant class of noncoding RNAs with advancements in biological and computational technologies. They are currently the subject of extensive research across various biological fields, revealing a multitude of functions including acting as miRNA sponges, regulating parental gene transcription, and serving as templates for protein translation. CircRNAs are implicated in a range of physiological and pathological processes, and recent evidence suggests their involvement in tumorigenesis, cancer cell proliferation, migration, and invasion, forming a critical network in cancer development. The stability, abundance, tissue and development specificity and detectability of circRNAs make them promising candidates for cancer diagnostics and prognostics. Therefore, many studies have applied circRNAs for the clinical diagnosis and treatment of cancer.
However, compared with other noncoding RNAs, our understanding of circRNAs is still shallow and is far from applying circRNAs in clinical practice. Future research should focus on several key areas. First, the precise mechanisms of circRNA formation, degradation and function remain unclear [162]. Further explorations are needed to fully elucidate and deepen our understanding of circRNAs. Second, we lacked a standardized methodology for identifying, validating and detecting clinical circRNAs. Reliable detection methods for circRNAs in bodily fluids are essential. Third, because most circRNAs are expressed at a low level and cannot be detected accurately, research should focus on improving the sensitivity and accuracy of circRNA detection. Finally, most circRNA studies were single-center and retrospective, with limited validation and clinical translation. Prospective validation and controlled clinical trials are necessary to establish circRNAs as reliable biomarkers.
Although the mechanisms and physiological and pathological effects of circRNAs in cancer are still controversial, circRNAs hold significant potential as diagnostic, prognostic and predictive biomarkers. This field has great translational potential in the future. However, before achieving this, we still need to overcome some knowledge challenge.
Availability of data and materials
Not applicable.
Abbreviations
- GC:
-
Gastric cancer
- NSCLC:
-
Non-small cell lung cancer
- LUAD:
-
Lung adenocarcinoma
- LUSC:
-
Lung squamous cell carcinoma
- HCC:
-
Hepatocellular carcinoma
- BCa:
-
Bladder cancer
- BC:
-
Breast cancer
- CRC:
-
Colorectal cancer
- ESCC:
-
Esophageal squamous cell carcinoma
- PDAC:
-
Pancreatic ductal adenocarcinoma
- ROC curve:
-
Receiver operating characteristic curve
- OS:
-
Overall survival
- ALP:
-
Alkaline phosphatase
- LDH:
-
Lactic dehydrogenase
- AFP:
-
Alpha-fetoprotein
- CEA:
-
Carcinoembryonic antigen
References
Sanger HL, Klotz G, Riesner D, et al. Viroids are single-stranded covalently closed circular RNA molecules existing as highly base-paired rod-like structures. Proc Natl Acad Sci USA. 1976;73(11):3852–6.
Kjems J, Garrett RA. Novel splicing mechanism for the ribosomal RNA intron in the archaebacterium Desulfurococcus mobilis. Cell. 1988;54(5):693–703.
Xu C, Zhang J. Mammalian circular RNAs result largely from splicing errors. Cell Rep. 2021;36(4): 109439.
Cocquerelle C, Mascrez B, Hétuin D, et al. Mis-splicing yields circular RNA molecules. Faseb J. 1993;7(1):155–60.
Wang PL, Bao Y, Yee MC, et al. Circular RNA is expressed across the eukaryotic tree of life. PLoS ONE. 2014;9(6): e90859.
You X, Vlatkovic I, Babic A, et al. Neural circular RNAs are derived from synaptic genes and regulated by development and plasticity. Nat Neurosci. 2015;18(4):603–10.
Junn E, Lee KW, Jeong BS, et al. Repression of alpha-synuclein expression and toxicity by microRNA-7. Proc Natl Acad Sci USA. 2009;106(31):13052–7.
Burd CE, Jeck WR, Liu Y, et al. Expression of linear and novel circular forms of an INK4/ARF-associated non-coding RNA correlates with atherosclerosis risk. PLoS Genet. 2010;6(12): e1001233.
Kumar L, Shamsuzzama, Haque R, et al. Circular RNAs: the emerging class of non-coding RNAs and their potential role in human neurodegenerative diseases. Mol Neurobiol. 2017;54(9):7224–34.
Aufiero S, Reckman YJ, Pinto YM, et al. Circular RNAs open a new chapter in cardiovascular biology. Nat Rev Cardiol. 2019;16(8):503–14.
Fang Y, Wang X, Li W, et al. Screening of circular RNAs and validation of circANKRD36 associated with inflammation in patients with type 2 diabetes mellitus. Int J Mol Med. 2018;42(4):1865–74.
Li P, Chen S, Chen H, et al. Using circular RNA as a novel type of biomarker in the screening of gastric cancer. Clin Chim Acta. 2015;444:132–6.
Qin M, Liu G, Huo X, et al. Hsa_circ_0001649: a circular RNA and potential novel biomarker for hepatocellular carcinoma. Cancer Biomark. 2016;16(1):161–9.
Chen J, Li Y, Zheng Q, et al. Circular RNA profile identifies circPVT1 as a proliferative factor and prognostic marker in gastric cancer. Cancer Lett. 2017;388:208–19.
Rybak-Wolf A, Stottmeister C, Glažar P, et al. Circular RNAs in the mammalian brain are highly abundant, conserved, and dynamically expressed. Mol Cell. 2015;58(5):870–85.
Jeck WR, Sorrentino JA, Wang K, et al. Circular RNAs are abundant, conserved, and associated with ALU repeats. RNA. 2013;19(2):141–57.
Chapman MH, Tidswell R, Dooley JS, et al. Whole genome RNA expression profiling of endoscopic biliary brushings provides data suitable for biomarker discovery in cholangiocarcinoma. J Hepatol. 2012;56(4):877–85.
Hansen TB, Jensen TI, Clausen BH, et al. Natural RNA circles function as efficient microRNA sponges. Nature. 2013;495(7441):384–8.
Memczak S, Jens M, Elefsinioti A, et al. Circular RNAs are a large class of animal RNAs with regulatory potency. Nature. 2013;495(7441):333–8.
Li Z, Huang C, Bao C, et al. Exon-intron circular RNAs regulate transcription in the nucleus. Nat Struct Mol Biol. 2015;22(3):256–64.
Zhang Y, Zhang XO, Chen T, et al. Circular intronic long noncoding RNAs. Mol Cell. 2013;51(6):792–806.
Meng S, Zhou H, Feng Z, et al. CircRNA: functions and properties of a novel potential biomarker for cancer. Mol Cancer. 2017;16(1):94.
Zheng Q, Bao C, Guo W, et al. Circular RNA profiling reveals an abundant circHIPK3 that regulates cell growth by sponging multiple miRNAs. Nat Commun. 2016;7:11215.
Wang C, Tan S, Li J, et al. CircRNAs in lung cancer - biogenesis, function and clinical implication. Cancer Lett. 2020;492:106–15.
Salzman J, Gawad C, Wang PL, et al. Circular RNAs are the predominant transcript isoform from hundreds of human genes in diverse cell types. PLoS ONE. 2012;7(2): e30733.
Chen LL, Yang L. Regulation of circRNA biogenesis. RNA Biol. 2015;12(4):381–8.
Kristensen LS, Andersen MS, Stagsted LVW, et al. The biogenesis, biology and characterization of circular RNAs. Nat Rev Genet. 2019;20(11):675–91.
Zhang XO, Wang HB, Zhang Y, et al. Complementary sequence-mediated exon circularization. Cell. 2014;159(1):134–47.
Kelly S, Greenman C, Cook PR, et al. Exon skipping is correlated with exon circularization. J Mol Biol. 2015;427(15):2414–7.
Lu Z, Filonov GS, Noto JJ, et al. Metazoan tRNA introns generate stable circular RNAs in vivo. RNA. 2015;21(9):1554–65.
Shen H, Liu B, Xu J, et al. Circular RNAs: characteristics, biogenesis, mechanisms and functions in liver cancer. J Hematol Oncol. 2021;14(1):134.
Schmidt CA, Giusto JD, Bao A, et al. Molecular determinants of metazoan tricRNA biogenesis. Nucleic Acids Res. 2019;47(12):6452–65.
Enuka Y, Lauriola M, Feldman ME, et al. Circular RNAs are long-lived and display only minimal early alterations in response to a growth factor. Nucleic Acids Res. 2016;44(3):1370–83.
Zhang Y, Xue W, Li X, et al. The biogenesis of nascent circular RNAs. Cell Rep. 2016;15(3):611–24.
Jeck WR, Sharpless NE. Detecting and characterizing circular RNAs. Nat Biotechnol. 2014;32(5):453–61.
Szabo L, Morey R, Palpant NJ, et al. Statistically based splicing detection reveals neural enrichment and tissue-specific induction of circular RNA during human fetal development. Genome Biol. 2015;16(1):126.
Li X, Yang L, Chen LL. The biogenesis, functions, and challenges of circular RNAs. Mol Cell. 2018;71(3):428–42.
Hanan M, Soreq H, Kadener S. CircRNAs in the brain. RNA Biol. 2017;14(8):1028–34.
Salzman J, Chen RE, Olsen MN, et al. Cell-type specific features of circular RNA expression. PLoS Genet. 2013;9(9): e1003777.
He Y, Huang H, Jin L, et al. CircZNF609 enhances hepatocellular carcinoma cell proliferation, metastasis, and stemness by activating the Hedgehog pathway through the regulation of miR-15a-5p/15b-5p and GLI2 expressions. Cell Death Dis. 2020;11(5):358.
Jiang X, Xing L, Chen Y, et al. CircMEG3 inhibits telomerase activity by reducing Cbf5 in human liver cancer stem cells. Mol Ther Nucleic Acids. 2021;23:310–23.
Liang D, Wilusz JE. Short intronic repeat sequences facilitate circular RNA production. Genes Dev. 2014;28(20):2233–47.
Venø MT, Hansen TB, Venø ST, et al. Spatio-temporal regulation of circular RNA expression during porcine embryonic brain development. Genome Biol. 2015;16:245.
Vo JN, Cieslik M, Zhang Y, et al. The landscape of circular RNA in cancer. Cell. 2019;176(4):869-881.e13.
Conn VM, Hugouvieux V, Nayak A, et al. A circRNA from SEPALLATA3 regulates splicing of its cognate mRNA through R-loop formation. Nat Plants. 2017;3:17053.
Ambros V. The functions of animal microRNAs. Nature. 2004;431(7006):350–5.
Gebert LFR, Macrae IJ. Regulation of microRNA function in animals. Nat Rev Mol Cell Biol. 2019;20(1):21–37.
Hansen TB, Kjems J, Damgaard CK. Circular RNA and miR-7 in cancer. Cancer Res. 2013;73(18):5609–12.
Li Y, Zheng F, Xiao X, et al. CircHIPK3 sponges miR-558 to suppress heparanase expression in bladder cancer cells. EMBO Rep. 2017;18(9):1646–59.
Chen X, Mao R, Su W, et al. Circular RNA circHIPK3 modulates autophagy via MIR124-3p-STAT3-PRKAA/AMPKα signaling in STK11 mutant lung cancer. Autophagy. 2020;16(4):659–71.
Pamudurti NR, Bartok O, Jens M, et al. Translation of CircRNAs. Mol Cell. 2017;66(1):9-21.e7.
Meyer KD, Patil DP, Zhou J, et al. 5’ UTR m(6)A promotes cap-independent translation. Cell. 2015;163(4):999–1010.
Di Timoteo G, Dattilo D, Centrón-Broco A, et al. Modulation of circRNA metabolism by m(6)A modification. Cell Rep. 2020;31(6): 107641.
Yang Y, Fan X, Mao M, et al. Extensive translation of circular RNAs driven by N(6)-methyladenosine. Cell Res. 2017;27(5):626–41.
Li Y, Zheng Q, Bao C, et al. Circular RNA is enriched and stable in exosomes: a promising biomarker for cancer diagnosis. Cell Res. 2015;25(8):981–4.
Chen Q, Zhang J, He Y, et al. hsa_circ_0061140 knockdown reverses FOXM1-mediated cell growth and metastasis in ovarian cancer through miR-370 sponge activity. Mol Ther Nucleic Acids. 2018;13:55–63.
Fang J, Hong H, Xue X, et al. A novel circular RNA, circFAT1(e2), inhibits gastric cancer progression by targeting miR-548g in the cytoplasm and interacting with YBX1 in the nucleus. Cancer Lett. 2019;442:222–32.
Hu X, Wu D, He X, et al. circGSK3β promotes metastasis in esophageal squamous cell carcinoma by augmenting β-catenin signaling. Mol Cancer. 2019;18(1):160.
Hsiao KY, Lin YC, Gupta SK, et al. Noncoding effects of circular RNA CCDC66 promote colon cancer growth and metastasis. Cancer Res. 2017;77(9):2339–50.
Weng W, Wei Q, Toden S, et al. Circular RNA ciRS-7-a promising prognostic biomarker and a potential therapeutic target in colorectal cancer. Clin Cancer Res. 2017;23(14):3918–28.
Arnaiz E, Sole C, Manterola L, et al. CircRNAs and cancer: biomarkers and master regulators. Semin Cancer Biol. 2019;58:90–9.
Heitzer E, Haque IS, Roberts CES, et al. Current and future perspectives of liquid biopsies in genomics-driven oncology. Nat Rev Genet. 2019;20(2):71–88.
Wang S, Zhang K, Tan S, et al. Circular RNAs in body fluids as cancer biomarkers: the new frontier of liquid biopsies[J]. Mol Cancer. 2021;20(1):13.
Panda AC, Gorospe M. Detection and analysis of circular RNAs by RT-PCR. Bio Protoc. 2018;8(6): e2775.
Suzuki H, Zuo Y, Wang J, et al. Characterization of RNase R-digested cellular RNA source that consists of lariat and circular RNAs from pre-mRNA splicing. Nucleic Acids Res. 2006;34(8): e63.
Nielsen AF, Bindereif A, Bozzoni I, et al. Best practice standards for circular RNA research. Nat Methods. 2022;19:1208–20.
Kulpa D, Topping R, Telesnitsky A. Determination of the site of first strand transfer during Moloney murine leukemia virus reverse transcription and identification of strand transfer-associated reverse transcriptase errors. Embo J. 1997;16(4):856–65.
Conn V, Conn SJ. SplintQuant: a method for accurately quantifying circular RNA transcript abundance without reverse transcription bias. RNA. 2019;25(9):1202–10.
Cocquet J, Chong A, Zhang G, et al. Reverse transcriptase template switching and false alternative transcripts. Genomics. 2006;88(1):127–31.
Chen D-F, Zhang L-J, Tan K, et al. Application of droplet digital PCR in quantitative detection of the cell-free circulating circRNAs. Biotechnol Biotechnol Equip. 2018;32(1):116–23.
Kojabad AA, Farzanehpour M, Galeh HEG, et al. Droplet digital PCR of viral DNA/RNA, current progress, challenges, and future perspectives. J Med Virol. 2021;93(7):4182–97.
Sillence KA, Roberts LA, Hollands HJ, et al. Fetal sex and RHD genotyping with digital PCR demonstrates greater sensitivity than real-time PCR. Clin Chem. 2015;61(11):1399–407.
Quan PL, Sauzade M, Brouzes E. dPCR: a technology review. Sensors. 2018;18(4):1271.
Schneider T, Schreiner S, Preußer C, et al. Northern blot analysis of circular RNAs. Methods Mol Biol. 2018;1724:119–33.
D’ambra E, Morlando M. Study of circular RNA expression by nonradioactive northern blot procedure. Methods Mol Biol. 2021;2348:371–83.
Wang X, Shan G. Nonradioactive northern blot of circRNAs. Methods Mol Biol. 2018;1724:135–41.
Lim AS, Lim TH. Fluorescence in situ hybridization on tissue sections. Methods Mol Biol. 2017;1541:119–25.
Raj A, Van Den Bogaard P, Rifkin SA, et al. Imaging individual mRNA molecules using multiple singly labeled probes. Nat Methods. 2008;5(10):877–9.
D’ambra E, Santini T, Vitiello E, et al. Circ-Hdgfrp3 shuttles along neurites and is trapped in aggregates formed by ALS-associated mutant FUS. iScience. 2021;24(12): 103504.
Dahl M, Daugaard I, Andersen MS, et al. Enzyme-free digital counting of endogenous circular RNA molecules in B-cell malignancies. Lab Invest. 2018;98(12):1657–69.
Kristensen LS. Profiling of circRNAs using an enzyme-free digital counting method. Methods. 2021;196:11–6.
Ju C, He J, Wang C, et al. Current advances and future perspectives on the functional roles and clinical implications of circular RNAs in esophageal squamous cell carcinoma: more influential than expected. Biomark Res. 2022;10(1):41.
Xu C, Jun E, Okugawa Y, et al. A circulating panel of circular RNA biomarkers for the noninvasive and early detection of pancreatic ductal adenocarcinoma. Gastroenterology. 2023;166(1):178–90.
Wei Y, Chen X, Liang C, et al. A noncoding regulatory RNAs network driven by Circ-CDYL acts specifically in the early stages hepatocellular carcinoma. Hepatology. 2020;71(1):130–47.
Zhang S, Zeng X, Ding T, et al. Microarray profile of circular RNAs identifies hsa_circ_0014130 as a new circular RNA biomarker in non-small cell lung cancer. Sci Rep. 2018;8(1):2878.
Wang C, Tan S, Liu WR, et al. RNA-Seq profiling of circular RNA in human lung adenocarcinoma and squamous cell carcinoma. Mol Cancer. 2019;18(1):134.
Kun-Peng Z, Chun-Lin Z, Jian-Ping H, et al. A novel circulating hsa_circ_0081001 act as a potential biomarker for diagnosis and prognosis of osteosarcoma. Int J Biol Sci. 2018;14(11):1513–20.
Lasda E, Parker R. Circular RNAs co-precipitate with extracellular vesicles: a possible mechanism for circRNA clearance. PLoS ONE. 2016;11(2): e0148407.
Tang W, Fu K, Sun H, et al. CircRNA microarray profiling identifies a novel circulating biomarker for detection of gastric cancer. Mol Cancer. 2018;17(1):137.
Tian G, Li G, Guan L, et al. Prognostic value of circular RNA ciRS-7 in various cancers: a PRISMA-compliant meta-analysis. Biomed Res Int. 2020;2020:1487609.
Luo L, Gao Y, Sun X. Circ-ITCH correlates with small tumor size, decreased FIGO stage and prolonged overall survival, and it inhibits cells proliferation while promotes cells apoptosis in epithelial ovarian cancer. Cancer Biomark. 2018;23(4):505–13.
Nanishi K, Konishi H, Shoda K, et al. Circulating circERBB2 as a potential prognostic biomarker for gastric cancer: an investigative study. Cancer Sci. 2020;111(11):4177–86.
Yan B, Zhang W, Mao XW, et al. Circular RNA ciRS-7 correlates with advance disease and poor prognosis, and its down-regulation inhibits cells proliferation while induces cells apoptosis in non-small cell lung cancer. Eur Rev Med Pharmacol Sci. 2018;22(24):8712–21.
Yu L, Gong X, Sun L, et al. The circular RNA Cdr1as act as an oncogene in hepatocellular carcinoma through targeting miR-7 expression. PLoS ONE. 2016;11(7): e0158347.
Kristensen LS, Ebbesen KK, Sokol M, et al. Spatial expression analyses of the putative oncogene ciRS-7 in cancer reshape the microRNA sponge theory. Nat Commun. 2020;11(1):4551.
Ye T, Yang M, Huang D, et al. MicroRNA-7 as a potential therapeutic target for aberrant NF-κB-driven distant metastasis of gastric cancer. J Exp Clin Cancer Res. 2019;38(1):55.
Yang C, Yuan W, Yang X, et al. Circular RNA circ-ITCH inhibits bladder cancer progression by sponging miR-17/miR-224 and regulating p21, PTEN expression. Mol Cancer. 2018;17(1):19.
Li Z, Yanfang W, Li J, et al. Tumor-released exosomal circular RNA PDE8A promotes invasive growth via the miR-338/MACC1/MET pathway in pancreatic cancer. Cancer Lett. 2018;432:237–50.
Sun HD, Xu ZP, Sun ZQ, et al. Down-regulation of circPVRL3 promotes the proliferation and migration of gastric cancer cells. Sci Rep. 2018;8(1):10111.
Yang F, Hu A, Li D, et al. Circ-HuR suppresses HuR expression and gastric cancer progression by inhibiting CNBP transactivation. Mol Cancer. 2019;18(1):158.
Zhang Y, Jiang J, Zhang J, et al. CircDIDO1 inhibits gastric cancer progression by encoding a novel DIDO1-529aa protein and regulating PRDX2 protein stability. Mol Cancer. 2021;20(1):101.
Dai X, Liu J, Guo X, et al. Circular RNA circFGD4 suppresses gastric cancer progression via modulating miR-532-3p/APC/β-catenin signalling pathway. Clin Sci. 2020;134(13):1821–39.
Pan H, Li T, Jiang Y, et al. Overexpression of circular RNA ciRS-7 abrogates the tumor suppressive effect of miR-7 on gastric cancer via PTEN/PI3K/AKT signaling pathway. J Cell Biochem. 2018;119(1):440–6.
Zheng X, Huang M, Xing L, et al. The circRNA circSEPT9 mediated by E2F1 and EIF4A3 facilitates the carcinogenesis and development of triple-negative breast cancer. Mol Cancer. 2020;19(1):73.
Wang X, Xing L, Yang R, et al. The circACTN4 interacts with FUBP1 to promote tumorigenesis and progression of breast cancer by regulating the expression of proto-oncogene MYC. Mol Cancer. 2021;20(1):91.
Liang G, Ling Y, Mehrpour M, et al. Autophagy-associated circRNA circCDYL augments autophagy and promotes breast cancer progression. Mol Cancer. 2020;19(1):65.
Wang X, Chen T, Li C, et al. CircRNA-CREIT inhibits stress granule assembly and overcomes doxorubicin resistance in TNBC by destabilizing PKR. J Hematol Oncol. 2022;15(1):122.
Liu J, Liu H, Zeng Q, et al. Circular RNA circ-MAT2B facilitates glycolysis and growth of gastric cancer through regulating the miR-515-5p/HIF-1α axis. Cancer Cell Int. 2020;20:171.
Huang XY, Zhang PF, Wei CY, et al. Circular RNA circMET drives immunosuppression and anti-PD1 therapy resistance in hepatocellular carcinoma via the miR-30-5p/snail/DPP4 axis. Mol Cancer. 2020;19(1):92.
Liu Z, Wang Q, Wang X, et al. Circular RNA cIARS regulates ferroptosis in HCC cells through interacting with RNA binding protein ALKBH5. Cell Death Discov. 2020;6:72.
Wu A, Hu Y, Xu Y, et al. Methyltransferase-like 3-mediated m6A methylation of Hsa_circ_0058493 accelerates hepatocellular carcinoma progression by binding to YTH domain-containing protein 1. Front Cell Dev Biol. 2021;9: 762588.
Li J, Fan R, Xiao H. Circ_ZFR contributes to the paclitaxel resistance and progression of non-small cell lung cancer by upregulating KPNA4 through sponging miR-195-5p. Cancer Cell Int. 2021;21(1):15.
Ma X, Yang X, Bao W, et al. Circular RNA circMAN2B2 facilitates lung cancer cell proliferation and invasion via miR-1275/FOXK1 axis. Biochem Biophys Res Commun. 2018;498(4):1009–15.
Liang Y, Wang H, Chen B, et al. circDCUN1D4 suppresses tumor metastasis and glycolysis in lung adenocarcinoma by stabilizing TXNIP expression. Mol Ther Nucleic Acids. 2021;23:355–68.
Li B, Zhu L, Lu C, et al. circNDUFB2 inhibits non-small cell lung cancer progression via destabilizing IGF2BPs and activating anti-tumor immunity. Nat Commun. 2021;12(1):295.
Guo Y, Guo Y, Chen C, et al. Circ3823 contributes to growth, metastasis and angiogenesis of colorectal cancer: involvement of miR-30c-5p/TCF7 axis. Mol Cancer. 2021;20(1):93.
Xu H, Liu Y, Cheng P, et al. CircRNA_0000392 promotes colorectal cancer progression through the miR-193a-5p/PIK3R3/AKT axis. J Exp Clin Cancer Res. 2020;39(1):283.
Zheng X, Chen L, Zhou Y, et al. A novel protein encoded by a circular RNA circPPP1R12A promotes tumor pathogenesis and metastasis of colon cancer via Hippo-YAP signaling. Mol Cancer. 2019;18(1):47.
Chen J, Wu Y, Luo X, et al. Circular RNA circRHOBTB3 represses metastasis by regulating the HuR-mediated mRNA stability of PTBP1 in colorectal cancer. Theranostics. 2021;11(15):7507–26.
Xiong DD, Dang YW, Lin P, et al. A circRNA-miRNA-mRNA network identification for exploring underlying pathogenesis and therapy strategy of hepatocellular carcinoma. J Transl Med. 2018;16(1):220.
Liu X, Abraham JM, Cheng Y, et al. Synthetic circular RNA functions as a miR-21 sponge to suppress gastric carcinoma cell proliferation. Mol Ther Nucleic Acids. 2018;13:312–21.
Costello A, Lao NT, Barron N, et al. Reinventing the wheel: synthetic circular RNAs for mammalian cell engineering. Trends Biotechnol. 2020;38(2):217–30.
Wesselhoeft RA, Kowalski PS, Anderson DG. Engineering circular RNA for potent and stable translation in eukaryotic cells. Nat Commun. 2018;9(1):2629.
Lee KH, Kim S, Lee SW. Pros and cons of in vitro methods for circular RNA preparation. Int J Mol Sci. 2022;23(21):13247.
Nakamoto K, Abe H. Chemical synthesis of circular RNAs with phosphoramidate linkages for rolling-circle translation. Curr Protoc. 2021;1(3): e43.
Cui J, Zhang L, Zhang Z, et al. A precise and efficient circular RNA synthesis system based on a ribozyme derived from Tetrahymena thermophila. Nucleic Acids Res. 2023;51(14): e78.
Petkovic S, Müller S. RNA circularization strategies in vivo and in vitro. Nucleic Acids Res. 2015;43(4):2454–65.
Chen H, Cheng K, Liu X, et al. Preferential production of RNA rings by T4 RNA ligase 2 without any splint through rational design of precursor strand. Nucleic Acids Res. 2020;48(9): e54.
Liu CX, Guo SK, Nan F, et al. RNA circles with minimized immunogenicity as potent PKR inhibitors. Mol Cell. 2022;82(2):420-434.e6.
Wolfbeis OS. An overview of nanoparticles commonly used in fluorescent bioimaging. Chem Soc Rev. 2015;44(14):4743–68.
Wang AZ, Langer R, Farokhzad OC. Nanoparticle delivery of cancer drugs. Annu Rev Med. 2012;63:185–98.
Sun YF, Pi J, Xu JF. Emerging role of exosomes in tuberculosis: from immunity regulations to vaccine and immunotherapy. Front Immunol. 2021;12: 628973.
Fang L, Du WW, Awan FM, et al. The circular RNA circ-Ccnb1 dissociates Ccnb1/Cdk1 complex suppressing cell invasion and tumorigenesis. Cancer Lett. 2019;459:216–26.
Du WW, Fang L, Yang W, et al. Induction of tumor apoptosis through a circular RNA enhancing Foxo3 activity. Cell Death Differ. 2017;24(2):357–70.
Lu M, Wu Y, Zeng B, et al. CircEHMT1 inhibits metastatic potential of breast cancer cells by modulating miR-1233-3p/KLF4/MMP2 axis. Biochem Biophys Res Commun. 2020;526(2):306–13.
Ferguson SW, Nguyen J. Exosomes as therapeutics: the implications of molecular composition and exosomal heterogeneity. J Control Release. 2016;228:179–90.
Zhang W, Jiang Y, He Y, et al. Lipid carriers for mRNA delivery. Acta Pharm Sin B. 2023;13(10):4105–26.
Xiong S, Peng H, Ding X, et al. Circular RNA expression profiling and the potential role of hsa_circ_0089172 in Hashimoto’s thyroiditis via sponging miR125a-3p. Mol Ther Nucleic Acids. 2019;17:38–48.
Wang X, Zhang H, Yang H, et al. Exosome-delivered circRNA promotes glycolysis to induce chemoresistance through the miR-122-PKM2 axis in colorectal cancer. Mol Oncol. 2020;14(3):539–55.
Bulcha JT, Wang Y, Ma H, et al. Viral vector platforms within the gene therapy landscape. Signal Transduct Target Ther. 2021;6(1):53.
Meganck RM, Borchardt EK, Castellanos Rivera RM, et al. Tissue-dependent expression and translation of circular RNAs with recombinant AAV vectors in vivo. Mol Ther Nucleic Acids. 2018;13:89–98.
Sahoo S, Kariya T, Ishikawa K. Targeted delivery of therapeutic agents to the heart. Nat Rev Cardiol. 2021;18(6):389–99.
Yao Z, Luo J, Hu K, et al. ZKSCAN1 gene and its related circular RNA (circZKSCAN1) both inhibit hepatocellular carcinoma cell growth, migration, and invasion but through different signaling pathways. Mol Oncol. 2017;11(4):422–37.
Williford JM, Wu J, Ren Y, et al. Recent advances in nanoparticle-mediated siRNA delivery. Annu Rev Biomed Eng. 2014;16:347–70.
Li R, Jiang J, Shi H, et al. CircRNA: a rising star in gastric cancer. Cell Mol Life Sci. 2020;77(9):1661–80.
Qiu M, Xia W, Chen R, et al. The circular RNA circPRKCI promotes tumor growth in lung adenocarcinoma. Cancer Res. 2018;78(11):2839–51.
Singh S, Narang AS, Mahato RI. Subcellular fate and off-target effects of siRNA, shRNA, and miRNA. Pharm Res. 2011;28(12):2996–3015.
Zhang L, Liang D, Chen C, et al. Circular siRNAs for reducing off-target effects and enhancing long-term gene silencing in cells and mice. Mol Ther Nucleic Acids. 2018;10:237–44.
Zhang L, Liang D, Wang Y, et al. Caged circular siRNAs for photomodulation of gene expression in cells and mice. Chem Sci. 2018;9(1):44–51.
Doudna JA, Charpentier E. Genome editing. The new frontier of genome engineering with CRISPR-Cas9. Science. 2014;346(6213):1258096.
Zhang H, Qin C, An C, et al. Application of the CRISPR/Cas9-based gene editing technique in basic research, diagnosis, and therapy of cancer. Mol Cancer. 2021;20(1):126.
Li S, Li X, Xue W, et al. Screening for functional circular RNAs using the CRISPR-Cas13 system. Nat Methods. 2021;18(1):51–9.
Cyranoski D. Chinese scientists to pioneer first human CRISPR trial. Nature. 2016;535(7613):476–7.
Lu Y, Xue J, Deng T, et al. Safety and feasibility of CRISPR-edited T cells in patients with refractory non-small-cell lung cancer. Nat Med. 2020;26(5):732–40.
Yi L, Li J. CRISPR-Cas9 therapeutics in cancer: promising strategies and present challenges. Biochim Biophys Acta. 2016;1866(2):197–207.
Kennedy EM, Kornepati AV, Goldstein M, et al. Inactivation of the human papillomavirus E6 or E7 gene in cervical carcinoma cells by using a bacterial CRISPR/Cas RNA-guided endonuclease. J Virol. 2014;88(20):11965–72.
Shi Y, Liu Y, Huang J, et al. Optimized mobilization of MHC class I- and II- restricted immunity by dendritic cell vaccine potentiates cancer therapy. Theranostics. 2022;12(7):3488–502.
Zhang M, Huang N, Yang X, et al. A novel protein encoded by the circular form of the SHPRH gene suppresses glioma tumorigenesis. Oncogene. 2018;37(13):1805–14.
Xia X, Li X, Li F, et al. A novel tumor suppressor protein encoded by circular AKT3 RNA inhibits glioblastoma tumorigenicity by competing with active phosphoinositide-dependent Kinase-1. Mol Cancer. 2019;18(1):131.
Qu L, Yi Z, Shen Y, et al. Circular RNA vaccines against SARS-CoV-2 and emerging variants. Cell. 2022;185(10):1728-1744.e16.
Li H, Peng K, Yang K, et al. Circular RNA cancer vaccines drive immunity in hard-to-treat malignancies. Theranostics. 2022;12(14):6422–36.
Wang J, Li X, Lu L, et al. Circular RNA hsa_circ_0000567 can be used as a promising diagnostic biomarker for human colorectal cancer. J Clin Lab Anal. 2018;32(5): e22379.
Huang M, He YR, Liang LC, et al. Circular RNA hsa_circ_0000745 may serve as a diagnostic marker for gastric cancer. World J Gastroenterol. 2017;23(34):6330–8.
Sun H, Tang W, Rong D, et al. Hsa_circ_0000520, a potential new circular RNA biomarker, is involved in gastric carcinoma. Cancer Biomark. 2018;21(2):299–306.
Zhu X, Wang X, Wei S, et al. hsa_circ_0013958: a circular RNA and potential novel biomarker for lung adenocarcinoma. Febs J. 2017;284(14):2170–82.
Zhang Y, Zhao H, Zhang L. Identification of the tumor-suppressive function of circular RNA FOXO3 in non-small cell lung cancer through sponging miR-155. Mol Med Rep. 2018;17(6):7692–700.
Yu J, Xu QG, Wang ZG, et al. Circular RNA cSMARCA5 inhibits growth and metastasis in hepatocellular carcinoma. J Hepatol. 2018;68(6):1214–27.
Zhuo F, Lin H, Chen Z, et al. The expression profile and clinical significance of circRNA0003906 in colorectal cancer. Onco Targets Ther. 2017;10:5187–93.
Hong W, Xue M, Jiang J, et al. Circular RNA circ-CPA4/ let-7 miRNA/PD-L1 axis regulates cell growth, stemness, drug resistance and immune evasion in non-small cell lung cancer (NSCLC). J Exp Clin Cancer Res. 2020;39(1):149.
Yao T, Chen Q, Shao Z, et al. Circular RNA 0068669 as a new biomarker for hepatocellular carcinoma metastasis. J Clin Lab Anal. 2018;32(8): e22572.
Li J, Hu ZQ, Yu SY, et al. CircRPN2 inhibits aerobic glycolysis and metastasis in hepatocellular carcinoma. Cancer Res. 2022;82(6):1055–69.
Wang L, Long H, Zheng Q, et al. Circular RNA circRHOT1 promotes hepatocellular carcinoma progression by initiation of NR2F6 expression. Mol Cancer. 2019;18(1):119.
Chen S, Li K, Guo J, et al. circNEIL3 inhibits tumor metastasis through recruiting the E3 ubiquitin ligase Nedd4L to degrade YBX1. Proc Natl Acad Sci USA. 2023;120(13): e2215132120.
Acknowledgements
Not applicable.
Funding
This review was supported by the National Natural Science Foundation of China (No. 82072662), Shanghai Municipal Education Commission—Gaofeng Clinical Medicine Grant Support (2019142), Shanghai Three-year Action Plan to Promote Clinical Skills and Clinical Innovation in Municipal Hospitals (SHDC2020CR4022), and the 2021 Shanghai ‘Rising Stars of Medical Talent’ Youth Development Program: Outstanding Youth Medical Talents.
Author information
Authors and Affiliations
Contributions
J.Z. and C.H. proposed concepts; J.Z. and Z.L. drafted the figures and wrote the original draft; YZ and ZQ supervised and provided important advice; J.Z, Z.L, Y.Z, M.D, Z.Q and C.H. reviewed the original draft and gave the final approval of the submitted version.
Corresponding authors
Ethics declarations
Consent for publication
All authors of this article agree to publish.
Competing interests
The authors declare that there is no conflict of interests regarding the publication of this paper.
Additional information
Publisher's Note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
Zhang, J., Luo, Z., Zheng, Y. et al. CircRNA as an Achilles heel of cancer: characterization, biomarker and therapeutic modalities. J Transl Med 22, 752 (2024). https://doi.org/10.1186/s12967-024-05562-4
Received:
Accepted:
Published:
DOI: https://doi.org/10.1186/s12967-024-05562-4