Next Article in Journal
Dopamine in Health and Disease
Next Article in Special Issue
MicroRNAs as Critical Biomarkers of Major Depressive Disorder: A Comprehensive Perspective
Previous Article in Journal
Disease-Focused Research Using Stem Cells
Previous Article in Special Issue
The Dual Role of Circular RNAs as miRNA Sponges in Breast Cancer and Colon Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Dissecting the Role of Circular RNAs in Sarcomas with Emphasis on Osteosarcomas

by
Eleftheria Lakiotaki
1,*,
Dimitrios S. Kanakoglou
1,
Andromachi Pampalou
1,
Eleni A. Karatrasoglou
1,
Christina Piperi
2,† and
Penelope Korkolopoulou
1,†
1
First Department of Pathology, Medical School, National and Kapodistrian University of Athens, 75 Mikras Asias Street, 11527 Athens, Greece
2
Department of Biological Chemistry, Medical School, National and Kapodistrian University of Athens, 75 Mikras Asias Street, 11527 Athens, Greece
*
Author to whom correspondence should be addressed.
Co-Senior Authors.
Biomedicines 2021, 9(11), 1642; https://doi.org/10.3390/biomedicines9111642
Submission received: 30 September 2021 / Revised: 4 November 2021 / Accepted: 6 November 2021 / Published: 8 November 2021

Abstract

:
Circular RNAs (circRNAs) are single-stranded RNAs generated from exons back-splicing from a single pre-mRNA, forming covalently closed loop structures which lack 5′-3′-polarity or polyadenylated tail. Ongoing research depicts that circRNAs play a pivotal role in tumorigenesis, tumor progression, metastatic potential and chemoresistance by regulating transcription, microRNA (miRNA) sponging, RNA-binding protein interactions, alternative splicing and to a lesser degree, protein coding. Sarcomas are rare malignant tumors stemming from mesenchymal cells. Due to their clinically insidious onset, they often present at advanced stage and their treatment may require aggressive chemotherapeutic or surgical options. This review is mainly focused on the regulatory functions of circRNAs on osteosarcoma progression and their potential role as biomarkers, an area which has prompted lately extensive research. The attributed oncogenic role of circRNAs on other mesenchymal tumors such as Kaposi Sarcoma (KS), Rhabdomyosarcoma (RMS) or Gastrointestinal Stromal Tumors (GISTs) is also described. The involvement of circRNAs on sarcoma oncogenesis and relevant emerging diagnostic, prognostic and therapeutic applications are expected to gain more research interest in the future.

1. Introduction

Sarcomas are rare tumors, with an incidence between 1 to 5 per 1,000,000 population, accounting for over 20% of all pediatric solid malignant tumors and less than 1% of all adult solid malignant tumors [1,2]. The most common types are reportedly leiomyosarcomas, Kaposi Sarcomas (KS), Undifferentiated Pleomorphic Sarcomas (UPS), liposarcomas and fibrosarcomas [2]. Osteosarcoma (OS) is the most frequent malignant bone tumor, occurring in all age groups, with the highest incidence been detected at 5–29 years of age and a second peak occurring above 50 years, following a bimodal distribution [2,3]. It usually affects the metaphysis of the limbs, such as the distal femur, humerus and proximal tibia [3]. Its treatment includes surgery and chemotherapy (adjuvant and neoadjuvant). Cure rates have improved from 20% to 70% for patients without metastasis [4,5]; however, patients with metastatic OS or patients who have developed chemoresistance experience poor survival rates [6].
KS is a very common tumor, of intermediate malignant potential, in patients with AIDS in the United States, and also common in patients with underlying immunodeficiency, such as organ transplant recipients [7,8], representing one of the most frequent tumors overall [2]. It is a vascular tumor driven by the Kaposi’s Sarcoma-Associated Herpesvirus (KSHV/HHV-8) which arises mainly in the skin, lymph nodes and mucous membranes, while it can also affect the majority of the visceral organs [9,10].
Rhabdomyosarcoma (RMS) comprises 5% of all pediatric tumors [11], and stems from mesenchymal cells with skeletal muscle differentiation [12]. It comprises several histological subtypes. Embryonal RMS (ERMS) is the most common soft tissue sarcoma in children and adolescents (4.5 cases/million people aged <20 years) [13]. Alveolar RMS (ARMS) is the second most common histologic subtype, occurring in a slightly older population in comparison to ERMS [14]. Pleomorphic RMS (PRMS) usually presents in adults [15], whereas Spindle cell/sclerosing RMS is the rarest histologic subtype and can affect all age groups [16]. Gastrointestinal Stromal Tumors (GISTs) can be found throughout the gastrointestinal tract, with the stomach being the most common site (54% of all GISTs), but also in extragastrointestinal locations, mainly in the mesentery, omentum and retroperitoneum. The incidence is about 10–15 cases per million per year [17]. This category of tumors is characterized by gain-of-function mutations of the KIT (KIT Proto-Oncogene, Receptor Tyrosine Kinase) or PDGFRA (Platelet Derived Growth Factor Receptor Alpha) oncogene. However, GISTs that are wild-type for KIT or PDGFRA are characterized by alterations in Succinate Dehydrogenase (SDH) subunit genes [18,19,20]. The assessment of prognosis is estimated by taking into account several parameters, namely tumor size (using as cutoff values 2 cm, 5 cm and 10 cm), mitotic count (lower or higher than 5 mitoses/5 mm2), and location (gastric vs. non-gastric) [21]. Complete surgical resection is the primary treatment for localized tumors [22], while imatinib, the tyrosine kinase inhibitor of KIT and PDGFRA receptors is considered the standard treatment for metastatic or unresectable GISTs [23,24].
Circular RNAs (circRNAs) represent a category of non-coding RNAs (ncRNAs) and constitute functional RNAs that are predetermined not to be translated, but transcribed. CircRNAs are characterized by single-stranded closed-loop structures without 5′-3′-polarity or a polyadenylated tail [25,26,27,28]. The term “circRNA” was first used by Sanger et al. when identifying the structure of viroids [29]. Their annular structure makes them more stable and resistant to degradation by exonucleases such as Ribonuclease (RNase) R, exhibiting longer half-life than linear RNA [25,27]. Upon their discovery, circRNAs were thought to be the byproducts of splicing errors and not given appropriate attention. After the advances in high-throughput RNA-sequencing technologies and bioinformatics, it has been shown that the expression of circRNAs accounts for more than 10% of gene expression [26,30,31]. Taking into account that their expression is tissue, cell or developmental stage-specific and also regulated by certain conditions such as stress, compelling evidence points towards circRNAs playing an important role in human tumorigenicity, tumor progression, metastatic spread and signaling pathway modulation [32,33,34,35].
Although the bulk of the research regarding circRNAs has been focused on carcinomas, there are reports illustrating the role of circRNAs on sarcomagenesis, with the majority of experimental work concentrating on OS. In this review, we provide an update on the current knowledge related to circRNAs’ role on sarcoma tumorigenesis, describing their implication in various tumor types.

2. Main Characteristics of CircRNAs

2.1. Categories, Biogenesis, Localization, Degradation

CircRNAs are categorized according to their synthesis by introns, exons, or both into four main categories described below:
  • Exonic circRNAs (EcircRNAs), localized to the cytoplasm [26] and consisting of a single or multiple exons, usually two or three [31,36]. A median length of 353 nucleotides is required for single-exon back-splicing.
  • Intronic circRNAs (ciRNAs), localized to the cell nucleus [37].
  • Exon-Intron circRNAs (EIcircRNAs), also localized to the cell nucleus and functionally similar to ciRNAs [38].
  • Intergenic circRNAs, generated by two ciRNAs fragments flanked by GT-AC splicing signals, acting as the splice donor and acceptor of the circular junction, and forming an integrated circRNAs [39].
CircRNAs are generated by back-splicing (Figure 1) of primary messenger RNA (mRNA) transcripts. During back-splicing, a 5′ splice site (donor) is joined to an upstream 3′ site (acceptor) by the spliceosome machinery, forming a covalently closed structure with a specific junction site [37]. This process competes with canonical splicing and can produce different circRNAs from the same sequence [30,40]. It has been reported that the more back-splicing an exon can undergo, the less it is included in the fully processed mRNA [41].
Back-splicing occurs both co-transcriptionally and post-transcriptionally and is a much less efficient process than canonical splicing. It is favored over canonical splicing up on depletion of splicing factors and can be affected by epigenetic changes within histones and gene bodies [30,42,43,44].
There are three basic models of circRNAs biogenesis. Firstly, on an exon-skipping event, lariats are formed when alternative exons are spliced out of the final mRNA product. The lariat containing the excised exons undergoes internal back-splicing and circRNA (EcircRNA or EIcircRNA) are formed [41,45]. Intron lariats that contain a ciRNA specific consensus motif consisting of an 11-Nucleotide (nt) C-rich element near the branch point and a 7-nt GU-rich element near the 5′ splice site can escape debranching and form ciRNAs according to this model [46]. Secondly, on “intron pairing-driven circularization”, inverted repeat elements (for example ALU elements) are located in the upstream and downstream introns. By base-pairing of complementary sequences of inverted repeat elements, a hairpin structure is formed and looping brings the acceptor and donor sites in proximity [36,37,41]. Thirdly, looping can be mediated by RNA-Binding Proteins (RBPs), such as Quaking, Muscleblind or FUS/TLS (Fused in Sarcoma/Translocated in Liposarcoma) protein [47,48,49]. Interestingly, not all RBPs promote circRNAs biogenesis, as ADAR1 (Adenosine Deaminase RNA-Specific Binding Protein) binds to double-stranded RNA regions and promotes the melting of stem structures to destabilize RNA pairing, thereby suppressing circRNAs formation [50].
Apart from the four main categories, there are newly emerging subtypes of circRNAs, such as tRNA Intronic circRNAs (tricRNAs), generated during pre-tRNA maturation or fusion circRNAs (f-circRNAs), created from chromosomal translocations in host genes [51,52].
Following synthesis, all circRNAs, apart from intron-containing, are transported from the nucleus to the cytoplasm in a size-dependent manner by the enzymes Adenosine Triphosphate (ATP)-dependent RNA helicase DDX39A and spliceosome RNA helicase DDX39B [53].
First reports regarding the mechanisms implicated on circRNAs degradation were based on studies of the circRNAs Cerebellar Degeneration-Related Protein 1 Transcript (CDR1as), a circRNA upregulated in OS, that described cleavage of CDR1as as mediated by Argonaute 2 (Ago2), after binding to microRNA (miRNA) miR-671 [54]. Nevertheless, whether this type of degradation is universal to circRNAs is still not clear [30]. Recent evidence points to other endonucleases functioning on circRNAs decay, upon special conditions. A portion of circRNAs that have undergone m6A (N6-Methyladenosine) modification are cleaved by the ribonuclease complex RNase P/MRP, a process mediated by the m6A reader protein YTHDF2 (YTH N6-Methyladenosine RNA Binding Protein 2) and HRSP12 (Heat-Responsive Protein 12) [55]. That could be the case for circNRIP1, a circRNA upregulated in OS. Its expression was reportedly elevated by METTL3, a methyltransferase that induces m6A modification to circNRIP1. In turn, circNRIP1 sponges miR-199a to upregulate FOXC2 expression in OS [56]. Additionally, RNase L, activated upon viral infection, can degrade all types of circRNAs. UPF1 (ATP-Dependent RNA Helicase Upstream Frameshift 1) and associated endonuclease G3BP1 (Ras Gtpase-Activating Protein-Binding Protein 1) can target and catalyze the degradation of some circRNAs [57]. Moreover, circRNAs accumulation is eliminated through exocytosis or exosome activity [58].

2.2. Biochemical Properties and Detection Methods

CircRNAs possess an exceptionally stable structure that results in their accumulation to the cytoplasm as previously mentioned, along with possible mechanisms of degradation. Most EcircRNAs’ half-life surpasses 48 h, in contrast to linear mRNA with average half-life of 10 h [26].
Another feature of circRNAs is their tissue -and developmental stage- or age-specificity. In human tissues, RNA-sequencing studies demonstrated that up to 50% of circRNAs were expressed in a highly specific pattern and showed increased levels on fetal compared to adult tissues [59].
CircRNAs are abundant in human tissues and tend to accumulate in tissues with low proliferation rate, for example cardiomyocytes. CircRNAs accumulation is also dependent on age [60,61]. Different human tissues show variable rates of circRNAs production, as in the human heart where 9% of the expressed genes produce circRNAs, whereas in the human brain the rate of production reaches 20% [62]. However, circRNA expression is not related to that of its linear isoforms, and under specific circumstances can far exceed it [30].
CircRNAs’ stability to degradation and the lack of polar structure at the end are properties that can be utilized for their detection. Their migration rate in a cross-linked gel is slower than that of long linear RNAs, and when compared with homologous gene transcription, nucleic acids show slower migration rate in weakly cross-linked gels, enabling circRNAs detection with northern blot analysis [25,63]. Subcellular location can be assessed with the Fluorescence In Situ Hybridization (FISH) technique [38,46]. Improved algorithms for high-throughput sequencing (circRNA candidate sequence boundary combination concerning different forms of exon rearrangement in comparison to sequencing data, different sequence alignment algorithms to match sequencing data to the genomic sequence and direct detection from cDNA sequences by designing multiple splice sequences) have also rendered the detection of low-abundance circRNAs feasible [64,65]. Currently, multiple circRNA-associated in silico tools and pipelines can support the de novo identification, assembly and annotation of circRNAs [66]. Depending on their implementation, circRNA identification tools are divided into three categories; BSJ-based (Back-Splicing Junction), integrated-based and machine learning-based [67]. Additionally, novel tools can perform more complex functions (e.g., alternative splicing events, expression estimation and circRNA structure prediction). In their review, Chen et al. [67] provide a comprehensive guide for hundreds of circRNA-related tools and their functionality.

2.3. Biological Functions

CircRNAs exert many important biological functions through three major mechanisms of action (Figure 2):
  • MiRNA sponging
  • Protein and mRNA interplay
  • Translation
MiRNA sponging is the most well-studied mechanism of action of circRNAs. CircRNAs contain binding sites for miRNAs and compete endogenous RNAs to sequester miRNAs via complementary base-pairing, while inhibiting miRNA from binding to their target molecules [68]. Subsequently, miRNA sponging leads to suppression of target mRNAs. The well-known paradigm of circRNA CDR1as, with more than 70 binding sites for miR-7 [69], but also other circRNAs such as circRNA-Homeodomain-Interacting Protein Kinase-2 (circHIPK2), have mechanistically confirmed sponge effect to miR124-2HG [70]. Considering their multiple binding sites, circRNAs can act both as oncogenes or tumor suppressors through binding to different miRNAs [37]. A large part of research focusing on circRNA role on sarcomas refers to circRNA-miRNA interactions and their impact on tumor growth, described later. Nevertheless, bioinformatics analysis reports that although circRNAs may have abundant binding sites, they may not exert prominent sponge effect, and under normal conditions this effect is not important on highly expressed miRNAs [31,54,71,72].
CircRNAs interactions with proteins are multifold, since they can act as protein sponges, scaffolds, enhance protein function or recruit proteins to specific subcellular compartments, or even facilitate contact between two or more proteins. It is well documented that circRNAs that carry RBP binding motifs may act as decoys or sponges to RBPs and interfere with their effects [37]. CircMbl is documented to bind both mbl and MBNL1 (Muscleblind-like Protein 1) proteins. Upon increased mbl or MBNL1 proteins, circMbl synthesis is promoted. In turn, linear mRNA splicing of the gene is favored, creating an autoregulatory loop of the gene expression [47]. In addition, RNA Polymerase II-U1 (Pol II-U1) Small Nuclear Ribonucleoproteins (snRNPs) complex binding is facilitated by interaction with circEIF3J with U1 RNA, resulting in upregulated parental gene transcription [38]. Similarly, circRNAs can bind dsRBPs [Double-Strand RNA-Binding Proteins (e.g., NF90, NF110)] [73] and create a reservoir to be used in specific conditions. Moreover, several circRNAs are reported to act as mRNA decoys or modulate mRNA stability [54,74]. Reports of proteins binding circRNAs without predicted binding sites imply that the tertiary structure of circRNAs may play an important role on protein binding, and that the tertiary structure may exhibit certain fluidity in different developmental stages or tissues [75]. The mechanisms underlying the way that tertiary structure affects circRNA function should be an object of further research.
Bioinformatics have proven that circRNAs have Open Reading Frame (ORF) and ribosome binding site, making translation possible. Translation may occur via Internal Ribosome Entry Site (IRES) elements that recruit the ribosomal 40S subunits in a cap-independent manner [76]. m6A modification may also play a role in circRNA translation, as m6A reader protein YTHDF3 interacts with translation initiation factors elF4G2 and elF3A, after binding to m6A modified circRNA [59]. The resulting peptides are truncated versions of the original proteins and, therefore, their functional relevance remains to be investigated. They might serve as dominant-negative protein variants, having been expressed under different conditions than the original protein or localized to other cellular compartments [76].
It has been mentioned above that circRNAs that localize in the nucleus (EIcircRNAs and ciRNAs) can regulate their parental gene expression by interacting with U1 snRNP complex. CircRNAs also regulate alternative splicing and transcription, given that under certain circumstances (depletion of splicing factors) circRNA biosynthesis is promoted over canonical splicing [47], exerting an important role on translation regulation by interfering with mRNA mobility and stability [54,74,77]. Other studies also show that circRNAs control the role of ribosomes in protein expression [78]. In addition, recent data demonstrate that stabilized circRNAs could form circRNA pseudogenes and be retrotranscribed and integrated into the genome [27,79], thus modifying gene expression.
Figure 2. CircRNA interactions: (a) CircRNA binds with RNA-Binding Proteins (RBPs), such as MBNL1 [80]; (b) circRNAs act as miRNA sponges; (c) EIcircRNA interacts with the U1 Small Nuclear Ribonucleoprotein (snRNP) and then binds to RNA Polymerase II (Pol II) [81]; (d) circRNA translation into proteins. The surface of the molecules MBNL1 and Pol II are modelled according to van der Waals surface (this illustration was created with BioRender.com, accessed on 3 November 2021).
Figure 2. CircRNA interactions: (a) CircRNA binds with RNA-Binding Proteins (RBPs), such as MBNL1 [80]; (b) circRNAs act as miRNA sponges; (c) EIcircRNA interacts with the U1 Small Nuclear Ribonucleoprotein (snRNP) and then binds to RNA Polymerase II (Pol II) [81]; (d) circRNA translation into proteins. The surface of the molecules MBNL1 and Pol II are modelled according to van der Waals surface (this illustration was created with BioRender.com, accessed on 3 November 2021).
Biomedicines 09 01642 g002

2.4. Roles in Homeostasis and Malignancy

CircRNAs participate in cell cycle regulation as key factors in multiple check points [82] and modulate c-MYC expression as well as other molecules downstream of c-MYC, including p21. They have also been involved in mediation of immune response [83], cell differentiation and pluripotency [84], as well as apoptosis [85].
Recent literature points to the dysregulation of circRNAs in cancer, inducing oncogenic or suppressive effects and playing a pivotal role in tumorigenesis [82]. Bioinformatics analysis has yielded interesting results on many tumor types, identifying a network of circRNAs and their target miRNAs and mRNAs as well as their interaction with important signaling pathways, such as Wnt/β-catenin, TGF-β (Transforming Growth Factor Beta) or PI3K/AKT [82]. They have also been shown to regulate Epithelial-Mesenchymal Transition (EMT), resist apoptosis, induce angiogenesis, DNA methylation and demethylation coordination as well as influence metastatic spread [68,82], whereas some studies have indicated the presence of circRNAs in exosomes, implicating their transport between cells and tissues [30,34,86]. Moreover, new evidence implies that circRNAs may affect acquired chemoresistance [87]. The above-mentioned mechanisms have been shown to be critical in sarcomagenesis [88,89,90,91,92,93,94], and a large number of circRNAs have been demonstrated to affect chemoresistance in OS, such as hsa_circ_0004674, hsa_circ_0081001 and circPVT1, among others [88,95,96,97,98].

3. CircRNAs Functions in Osteosarcoma

3.1. Upregulated CircRNAs in OS

Currently there has been an ever-increasing number of studies depicting the role of circRNAs as tumor promoters in OS. CircRNAs, by binding to multiple miRNAs and other proteins, interfere with numerous signaling pathways and promote cell proliferation, invasion, migration, chemoresistance and apoptosis inhibition. A detailed description of the current literature on upregulated circRNAs in OS is given in Table 1.

3.2. Downregulated CircRNAs in OS

The number of downregulated circRNAs in OS is much lower in comparison to upregulated circRNAs; however, there are reports of circRNAs involved in chemotherapy resensitization and tumor growth inhibition. A detailed description of the current literature on downregulated circRNAs in OS is given in Table 2.

3.3. CircRNAs of Ambiguous Significance in OS

This section provides an account of certain circRNAs detected either upregulated or downregulated in different studies, with a consequent effect on OS tumor growth and prognostic parameters. The discrepancy observed among experimental findings might be attributed to sample numbers and types (fresh, frozen) as well as to intrinsic heterogeneity between cell lines and patient tissues or to differences in methodology employed.

3.3.1. hsa_circ_0002052

The study of Wu et al. demonstrated that hsa_circ_0002052 is significantly downregulated in OS patient tissues and cell lines (143B, MG-63, HOS and U2OS). Overexpression of hsa_circ_0002052 inhibited cell proliferation, migration, and invasion in vitro and blocked cell growth in vivo [90]. hsa_circ_0002052 suppressed the activation of Wnt/β-catenin pathway by sponging miR-1205, and rescuing APC2 expression, which is a tumor suppressor often detected in many tumors [79,224,225] that regulates negatively the Wnt/β-catenin pathway [90]. Moreover, low levels of hsa_circ_0002052 were correlated with worse survival and worse PFS (Progression-Free Survival) [90].
In contrast, Zhang et al. reported that hsa_circ_0002052 is upregulated in OS tissues and is related to advanced stage, tumor size and metastasis, as well as to low survival rate. In addition, by sponging miR-382, hsa_circ_0002052 enhanced STX6 expression, resulting in Wnt/β-catenin pathway activation [214,226]. These utterly opposite results need further investigation to elucidate the role of hsa_circ_0002052 in OS.

3.3.2. circ-ITCH

circ-ITCH was reported as a tumor suppressor by the study of Ren et al., who detected downregulation of circ-ITCH on OS tissues and cell lines (MG-63, U2OS, SaOS-2) [227]. circ-ITCH overexpression sponged miR-22 and promoted apoptosis, while blocking cell viability, proliferation, migration and invasion [227]. miR-22 inhibition suppressed PTEN/PI3K/AKT and SP-1 pathways in OS [227].
Li et al. investigated circ-ITCH expression in OS cell lines SJSA-1 and U2OS and detected upregulated levels. Functional studies demonstrated that circ-ITCH participates in circ-ITCH/miR-7/EGFR pathway to promote OS cell migration, invasion and growth. Specifically, circ-ITCH acted as decoy to miR-7 and supports high levels of EGFR, while its migratory-invasive capacity was dependent to EGFR phosphorylation [228].
Zhou et al. also detected downregulation of circ-ITCH in OS tissues and cell lines. They detected a correlation of low circ-ITCH levels with promotion of OS development in OS tissues and MG-63 and KHOS cell lines, as well as with contribution to DXR (Doxorubicin) chemoresistance via the miR-524/RASSF6 axis [229].

3.3.3. circ_HIPK3

circ_HIPK3 was reported to be downregulated in OS tissues, plasma and OS cell lines (SaOS-2, HOS, KHOS, MG-63, 143B, U2OS). circ_HIPK3 levels were correlated with Enneking stage and lung metastasis [230]. Furthermore, low circ_HIPK3 levels correlated with shorter overall survival and poor prognosis. Enhancement of circ_HIPK3 blocked OS cell proliferation, migration and invasion in vitro [230].
On the other hand, Huang suggested that circ_HIPK3 can promote metastasis in OS by showing that circ_HIPK3 promotes migration, invasion and growth in OS tissues and cell lines (U2OS, SW1353) [231]. circ_HIPK3 could bind and inhibit miR-637 and further elevate STAT3 (Signal Transducer and Activator of Transcription 3) expression to exert oncogenic effects [231].
Wen et al. showed that circ_HIPK3 was upregulated in OS tissues and cell lines (HOS, MG-63, U2OS, SJSA-1 OS) and demonstrated that modulation of miR-637/HDAC4 axis promoted proliferation, migration and invasion in OS [232].

3.3.4. hsa_circ_0102049

hsa_circ_0102049 is mapped to chromosome 14 on ATL1 gene locus and has been originally found upregulated in OS [107,233]. On further studies, hsa_circ_0102049 correlated with poor prognosis, larger tumor size and pulmonary metastasis, and was found overexpressed on OS patient tissues (76 patients) and OS cell lines (U2OS, SaOS-2, MG-63, and HOS). Gain/loss of function experiments revealed accelerated cell proliferation, migration and invasion as well as attenuation of apoptosis upon hsa_circ_0102049 overexpression [233]. Additionally, this study provided evidence of hsa_circ_0102049 regulating MDM2 expression by sponging miR-1304-5p and revealed a novel hsa_circ_0102049/miR-1304-5p/MDM2 axis in OS.
On the other hand, a new study by Zhang et al. [234] found that silencing of hsa_circ_0102049 promotes proliferation, invasion, migration and cell cycle in OS cell line MG-63 by sponging miR-520g-3p and further modulating the miR-520g-3p/PLK2.Tap73 axis. Additional research is essential to elucidate hsa_circ_0102049 role on OS tumorigenesis.

3.4. Mechanism of Action and Potential Diagnostic and Prognostic Value of Important circRNAs in OS

A large number of circRNAs have been evaluated as potential biomarkers in OS tumorigenesis. However, only few of them hold promise in terms of prognostic and predictive value, as well as an effect on chemotherapy. In the first study, investigating the expression profile of circRNAs on chemoresistance to Doxorubicin (DXR) OS tissues (60 patient samples) and cell lines (MG-63, U2OS, KHOS and their (DXR) resistant pairs MG-63/DXR, U2OS/DXR and KHOS/DXR), Kun-Peng et al. detected 80 circRNAs dysregulated on chemoresistant cell lines and further studied the most upregulated circRNA, hsa_circ_0004674, which was related to poor prognosis [96]. Of the predicted targets of hsa_circ_0004674, miR-490-3p, miR-584-5p and miR-1254 are reported to have tumor suppressive roles including chemoresistance [235,236,237,238,239]. According to previous research findings of this group, chemoresistance in OS may be regulated by hsa_circ_0004674-miR-490-3p-ABCC2 and hsa_circ_0004674/miR-1254-EGFR pathways [96,240,241]. In addition, Bai et al. proposed a new mechanism of hsa_circ_0004674 promoting chemoresistance to DXR by regulating the miR-342-3p/FBN1 axis through the Wnt/β-catenin pathway [88].
On experiments aiming to elucidate the potential mechanisms of circRNAs and regulation of chemoresistance, using chemosensitive and chemoresistant paired cell lines (MG-63, U2OS, KHOS and (DXR) resistant pairs MG-63/DXR, U2OS/DXR and KHOS/DXR), hsa_circ_0081001 was detected significantly upregulated on OS tissues, cell lines and patient serum and correlated with poor survival [95]. Specifically, hsa_circ_0081001 expression was increased in higher stage patient groups, in chemoresistant patients and in lung metastasis patients. On multivariate analysis, chemoresistance, lung metastasis, Enneking stage and hsa_circ_0081001 overexpression were revealed as independent prognosticators [95]. Collectively, hsa_circ_0081001 overexpression showed a great prognostic value and may represent a biomarker in OS, superior to alkaline phosphatase and lactate dehydrogenase. Further studies by Wei et al. depicted that hsa_circ_0081001 knockdown enhanced methotrexate chemosensitivity in OS cell lines and chemoresistant pairs (U2OS-U2OS/R and HOS-HOS/R) and targeted miR-494-3p to elevate Translutaminase-2 (TGM2) levels [97].
The hsa_circ_0007534 was linked to poor prognosis in colorectal cancer [116]. Experiments on 57 OS patient tissues and OS cell lines HOS, SaOS-2, MG-63 and U2OS showed regulation of OS cell growth and apoptosis by hsa_circ_0007534 suggesting its possible prognostic value. hsa_circ_0007534 was upregulated in OS tissues and cell lines and facilitated OS tumorigenesis as well as tumor growth in xenograft mouse model. The oncogenic effects of hsa_circ_0007534 were attributed to its interaction with Phosphorylated AKT (pAKT) and Phosphorylated Glycogen Synthase Kinase-3β (pGSK-3β), as well as to the regulation of AKT/GSK-3β signaling pathway [115]. hsa_circ_0007534 high levels were correlated to tumor size and advanced histological grade, while hsa_circ_0007534 was determined as an independent prognosticator [115]. Another possible mechanism of action for hsa_circ_0007534 suggested by Zhang et al. is sponging of miR-219a-5p to upregulate Sex-Determining Region Y-box 5 (SOX-5) [242]. Recent evidence suggests that circPVT1 plays an important role in DXR and cisplatin chemoresistance. Kun-Peng conducted experiments in cell lines (MG-63, U2OS, KHOS and their (DXR) resistant pairs MG-63/DXR, U2OS/DXR and KHOS/DXR) that were also cross-resistant to cisplatin as well as tissues from 80 OS patients treated with regiments containing DXR and cisplatin [121]. circPVT1 was significantly upregulated in OS patient tissues, serum and cell lines and positively correlated to lung metastasis, advanced Enneking stage, shorter survival and chemoresistance. Furthermore, circPVT1 knockdown partly reversed DXR and cisplatin resistance in vitro and decreased ABCB1 expression which is highly expressed in drug-resistant cell lines, possibly by regulating the P-gp protein [243,244,245], further suggesting that circPVT1 inhibition mediates resensitization of OS cells to chemotherapy [121]. The effects of circPVT1 on DXR resistance were further validated in the study of Li et al. [98]. Detailed analysis indicated sponging of miR-137 by circPVT1 and subsequent regulation of TP53-regulated Inhibitor of Apoptosis 1 (TRIAP1), an inhibitor related to cisplatin sensitivity in human ovarian cancer [246].
Previous studies have shown that circ_ANKIB1 can act as an oncogene in OS [95]. miR-19b is documented to promote OS cell invasion migration and proliferation, and act as an oncogene in gliomas and colorectal carcinomas [247,248,249,250]. Mechanistic experiments conducted on OS cell lines MG-63, 143B, SaOS-2, U2OS and HOS showed that circ_ANKIB1 interacted with miR-19b and both of them were upregulated. Specifically, circ_ANKIB1 promotes miR-19b expression, as circ_ANKIB1 knockdown downregulated miR-19b expression and enhanced expression of SOCS3, respectively, with reduction in cell invasion. Collectively, it is suggested that circ_ANKIB1 and miR-19b promote OS cell invasion, while SOCS3 inhibits invasion, elucidating circ_ANKIB1 regulation on SOCS3/STAT3 pathway [125]. Another study by Zhu et al. confirmed the oncogenic role of circ_ANKIB1 in OS by regulating the miR-217/PAX3 axis [126].
Several studies have confirmed the ectopic expression of circ_001569 in OS. Zhang et al. reported circ_001569 upregulation, enhancement of cell proliferation and contribution to cisplatin resistance in OS patient tissues and cell lines (MG-63, U2OS). circ_001569 exerted its oncogenic role by activating Wnt/β-catenin signaling pathway, since increased circ_001569 upregulated phosphorylated GSK-3β and β-catenin and downregulated GSK-3β [89]. circ_001569 also correlated with distant metastasis and advanced tumor stage. Xiao et al. found that upregulated circ_001569 promotes proliferation, migration, invasion and EMT in OS by rescuing Flotillin-2 (FLOT2) expression via miR-185-5p sponging [169].
Conclusively, the abovementioned circRNAs regulate critical oncogenic pathways that justify their role as protooncogenes, and hsa_circ_0004674, hsa_circ_0081001, circPVT1 and circ_001569 were shown to exert important effect on OS prognosis, serving as potential biomarkers.

4. CircRNAs in Kaposi Sarcoma

Currently, the importance of circRNAs in immunology and significance in viral infections and host response has attracted the attention of investigators [251,252]. In 2018, Toptan et al. sequenced KSHV-infected cell lines BCBL-1 and BC-1 Primary Effusion Lymphoma (PEL) cell lines and detected constitutive expression of viral encoded circRNA circvIRF4, as well as numerous circRNAs from Polyadenylated Nuclear (PAN) RNA locus [253]. They confirmed the presence of circvIRF4 and circPAN/K7.3 isoforms in 4 out of 10 KS patient tissues, speculating that viral circRNAs participate in viral oncogenesis [253]. Tagawa et al. detected viral circRNAs in KSHV-infected primary Human Umbilical Vein Endothelial Cells (HUVECs) as well as in tissues from lymph nodes of patients suffering from KS, PEL or KSHV+ Castleman Disease [254]. This study also showed that infected cells stably expressing viral circRNAs differ in relative growth, compared to control cells in SLK cells-a KS cell line. They also identified a human circRNA, hsa_circ_0001400, induced by KSHV infection, with potential antiviral effect, and performed overrepresentation analysis that depicted enrichment of pathways involved in cancer or p53 signaling [254]. circvIRF4 and circPAN/K7.3 in KS tissues and serum were further investigated by the study of Abere et al., showing that 61/92 (66.3%) KS tissues were positive for KSHV circRNAs, namely 32/92 (34.8%) for circvIRF4, 49/92 (53.3%) for circPAN and 28/92 (30.4%) for circK7.3. A total of 5 out of 10 (50%) previously collected and stored KS patients’ sera were positive for viral circRNAs, while in fresh samples the respective percentage was 100% [255]. Moreover, Yao et al. provided an alternative mechanism of circRNA contribution to KS-driven oncogenesis, by identifying an upregulated cellular circular RNA, circARGEF1, which promotes cell invasion and angiogenesis in KS [256]. KS encodes the Viral Interferon Regulatory Factor 1 (vIRF1), which, apart from immune regulation possesses oncogenic properties, including p53 suppression and inhibition of TGF-β/Smad signaling [93,94]. Yao et al. demonstrated that vIRF1 and lymphoid enhancer binding factor 1 (Lef-1) binding induces circARGEF1 transcription, and circARGEF1 sponges miR-125a-3p to upregulate Glutaredoxin 3 (GLXR3) in vIRF1-transduced cells, KSHV-infected cells and KS tissues, promoting cell motility, proliferation and angiogenesis [256].

5. CircRNAs in Rhabdomyosarcoma

Rossi et al. utilized human primary wild-type myoblasts, ERMS RD cell line, ARMS RH4 cell line and RMS patient tissues to study the role of circZNF609 [257]. circZNF609 regulates cell cycle and immune response related genes in human primary myoblasts [76], and its silencing induces proliferation arrest. circZNF609 is upregulated in RMS tissues, RD and RRH4 cell lines, with particularly high levels in RH4 cell line. circZNF609 knockdown inhibited proliferation of RD but not of RH4 cells, and further analysis detected reduced p-Rb/Rb ratio and pAKT levels. After GO analysis, the downregulated genes were mainly associated with cell cycle progression, DNA replication and mitosis, while Gene Graph Enrichment Analysis (GGEA) detected genes involved in PI3K/AKT signaling pathway [257]. circZNF609 knockdown did not affect RH4 cells. A possible explanation is attributed to the upregulation of cell cycle-related targets of p53, that was found to be downregulated. Collectively, these results point to circZNF609 regulation of cell proliferation in RMS.

6. CircRNAs in Gastrointestinal Stromal Tumors

The study of Jia et al. used GIST-T1 and GIST-882 cell lines along with GIST patient tissues and conducted microarray analysis to detect 5770 circRNAs differentially expressed in GISTs. During expression profiling, they focused on three upregulated circRNAs (circ_0069765, circ_0084097 and circ_0079471) and their host genes (KIT, PLAT and ETV1) [258]. The importance of KIT mutation in GISTs is established as the main oncogenic event and involves ETS transcription factor ETV1. PLAT is enriched in blood vessel development involved in tissue specificity in GISTs and interacts with VEGFC, PGF and CHD7 [259]. The researchers also identified miRNAs targeted by KIT, PLAT and ETV1 and by circRNAs circ_0069765, circ_0084097 and circ_0079471, and constructed specific regulatory networks that may have important regulating roles in GISTs [258].

7. Conclusions–Perspective

CircRNAs have currently emerged as a promising research field in oncology, due to their outstanding properties and their crosstalk with multiple regulatory networks, signaling pathways, cellular processes and developmental events. Apart from their direct effect in tumorigenesis, circRNAs participate in immune modulation, which interferes with tissue microenvironment, being critical for tumor development. CircRNAs exhibit distinct expression pattern, stability and abundant detection levels in body fluids, such as blood, plasma or saliva, as well as in exosomes circulating in blood [260,261], making their function as potential biomarkers or as molecular markers to support diagnosis very attractive. The oncogenic implication of circRNAs as tumor promoters or tumor suppressors renders them a possible therapeutic target, either by the use of siRNAs complementary to the BSJ or by Antisense Oligonucleotides (ASOs) binding to the respective pre-mRNA and inhibition of oncogenic circRNAs, or by inducing expression of tumor suppressor circRNAs [262]. Modulation of circRNA levels may also be sensitive as a surrogate method to increase chemosensitivity in some tumors. Moreover, numerous studies have indicated the prognostic value of circRNAs. Future studies are highly demanded to improve detection methods as well as develop further clinical applications of these pivotal players of carcinogenesis.
To our knowledge, this is the first comprehensive literature review of circRNAs implication in all sarcoma types. Research in sarcomagenesis has been mainly focused on OS, indicating many circRNAs as possible therapeutic targets as well as prognostic biomarkers. Moreover, detection of certain circRNAs, such as hsa_circ_0081001 and circPVT1 in serum can be used for post-operative patient screening, aiding in the early detection of relapse. However, a small number of circRNAs in OS has provided ambiguous results regarding their functional role and will need further investigation. Of note, it is evident that different circRNAs play critical roles in each sarcoma type. Specifically, circRNAs present in KS are mainly of viral origin while circZNF609 detected in RMS and circ_0069765, circ_0084097 and circ_0079471 detected in GISTs have not been identified in OS. Conclusively, current data shows that circRNAs are involved in numerous significant oncogenic mechanisms that are implicated in various sarcoma types. Further research is highly demanded to investigate the pathogenic mechanisms in which circRNAs are involved in each sarcoma subtype.

Author Contributions

Conceptualization and writing—original draft preparation, E.L.; writing—original draft preparation, D.S.K.; original draft preparation, A.P.; editing, E.A.K.; review and editing, C.P.; review and supervision, P.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ADAR1Adenosine Deaminase RNA-Specific Binding Protein
Ago2Argonaute 2
APCAdenomatous Polyposis Coli Protein
ARL2ADP-Ribosylation Factor-like Protein 2
ARMSAlveolar RMS
ASOsAntisense Oligonucleotides
ATPAdenosine Triphosphate
BACH1BTB Domain and CNC Homolog 1
BSJBack-Splicing Junction
CADM1Cell Adhesion Molecule 1
CDCA4Cycle-Associated Protein 4
CDR1asCerebellar Degeneration-Related Protein 1 Transcript
CHI3L1Chitinase-3-like Protein 1
circHIPK2CircRNA-Homeodomain-Interacting Protein Kinase-2
circRNAsCircular RNAs
ciRNAsIntronic circRNAs
dsRBPsDouble-Strand RNA-Binding Proteins
DXRDoxorubicin
EcircRNAsExonic circRNAs
EIcircRNAsExon-Intron circRNAs
EMTEpithelial-Mesenchymal Transition
ERMSEmbryonal RMS
EZH2Enhancer of Zeste Homolog 2
f-circRNAsFusion circRNAs
FISHFluorescence In Situ Hybridization
FLI1Friend Leukemia Virus Integration 1
FLOT2Flotillin-2
FUS/TLSFused in Sarcoma/Translocated in Liposarcoma
G3BP1Ras Gtpase-Activating Protein-Binding Protein 1
GAB1GRB2 Associated Binding Protein 1
GGEAGene Graph Enrichment Analysis
GISTsGastrointestinal Stromal Tumors
GLXR3Glutaredoxin 3
HK2Human Glandular Kallilrein2
HRSP12Heat-Responsive Protein 12
HUVECsHuman Umbilical Vein Endothelial Cells
IL6RInterleukin 6 Receptor
IRESInternal Ribosome Entry Site
IRF2Interferon Regulatory Factor 2
JAG1Jagged1
KCNH1Potassium Voltage-Gated Channel Subfamily H Member 1
KLF10Kruppel-like Factor 10
KSKaposi Sarcoma
KSHV/HHV-8Kaposi’s Sarcoma-Associated Herpesvirus
LPAATβLysophosphatidic Acid Acyltransferase β
LSD1Lysine-Specific Histone Demethylase 1A
m6AN6-Methyladenosine
MBNL1Muscleblind-like Protein 1
miRNAmicroRNA
mRNAMessenger RNA
ncRNAsNon-coding RNAs
ntNucleotide
ORFOpen Reading Frame
OSOsteosarcoma
pAKTPhosphorylated AKT
PANPolyadenylated Nuclear
PDE7BPhosphodiesterase 7B
PELPrimary Effusion Lymphoma
PFSProgression-Free Survival
PNNPinin Desmosome Associated Protein
Pol II-U1RNA Polymerase II-U1
PRMSPleomorphic RMS
PTENPhosphate and Tensin Homolog
RBPsRNA-Binding Proteins
RMSRhabdomyosarcoma
RNaseRibonuclease
SDHSuccinate Dehydrogenase
SFRP1Secreted Frizzled Related Protein 1
siRNAsSmall Interfering RNAs
snRNPsSmall Nuclear Ribonucleoproteins
SOX-5Sex-Determining Region Y-box 5
STAT3Signal Transducer and Activator of Transcription 3
TGF-βTransforming Growth Factor Beta
TGM2Translutaminase-2
TRIAP1TP53-regulated Inhibitor of Apoptosis 1
tricRNAstRNA Intronic circRNAs
TRIM44Tripartite Motif Containing 44
TWISTTwist Family BHLH Transcription Factor
UPF1ATP-Dependent RNA Helicase Upstream Frameshift 1
UPSUndifferentiated Pleomorphic Sarcoma
VEGFVascular Endothelial Growth Factor
vIRF1Viral Interferon Regulatory Factor 1
XIAPX-linked Inhibitor of Apoptosis Protein
YTHDF2YTH N6-Methyladenosine RNA Binding Protein 2
ZNF652Zinc-Finger Protein 652

References

  1. Burningham, Z.; Hashibe, M.; Spector, L.; Schiffman, J.D. The epidemiology of sarcoma. Clin. Sarcoma Res. 2012, 2, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Ng, V.Y.; Scharschmidt, T.J.; Mayerson, J.L.; Fisher, J.L. Incidence and survival in sarcoma in the United States: A focus on musculoskeletal lesions. Anticancer Res. 2013, 33, 2597–2604. [Google Scholar] [PubMed]
  3. Bielack, S.S.; Kempf-Bielack, B.; Delling, G.; Exner, G.U.; Flege, S.; Helmke, K.; Kotz, R.; Salzer-Kuntschik, M.; Werner, M.; Winkelmann, W.; et al. Prognostic factors in high-grade osteosarcoma of the extremities or trunk: An analysis of 1702 patients treated on neoadjuvant cooperative osteosarcoma study group protocols. J. Clin. Oncol. 2002, 20, 776–790. [Google Scholar] [CrossRef] [PubMed]
  4. Jaffe, N. Progress report on high-dose methotrexate (NSC-740) with citrovorum rescue in the treatment of metastatic bone tumors. Cancer Chemother. Rep. 1974, 58, 275–280. [Google Scholar]
  5. Link, M.P.; Goorin, A.M.; Miser, A.W.; Green, A.A.; Pratt, C.B.; Belasco, J.B.; Pritchard, J.; Malpas, J.S.; Baker, A.R.; Kirkpatrick, J.A.; et al. The effect of adjuvant chemotherapy on relapse-free survival in patients with osteosarcoma of the extremity. N. Engl. J. Med. 1986, 314, 1600–1606. [Google Scholar] [CrossRef]
  6. Cao, J.; Han, X.; Qi, X.; Jin, X.; Li, X. TUG1 promotes osteosarcoma tumorigenesis by upregulating EZH2 expression via miR-144-3p. Int. J. Oncol. 2017, 51, 1115–1123. [Google Scholar] [CrossRef]
  7. Engels, E.A.; Biggar, R.J.; Hall, H.I.; Cross, H.; Crutchfield, A.; Finch, J.L.; Grigg, R.; Hylton, T.; Pawlish, K.S.; McNeel, T.S.; et al. Cancer risk in people infected with human immunodeficiency virus in the United States. Int. J. Cancer 2008, 123, 187–194. [Google Scholar] [CrossRef]
  8. Marcelin, A.G.; Calvez, V.; Dussaix, E. KSHV After an Organ Transplant: Should We Screen? In Kaposi Sarcoma Herpesvirus: New Perspectives; Boshoff, C., Weiss, R.A., Eds.; Springer: Berlin/Heidelberg, Germany, 2007; pp. 245–262. [Google Scholar]
  9. Boshoff, C.; Weiss, R. Aids-related malignancies. Nat. Rev. Cancer 2002, 2, 373–382. [Google Scholar] [CrossRef]
  10. Schwartz, R.A.; Micali, G.; Nasca, M.R.; Scuderi, L. Kaposi sarcoma: A continuing conundrum. J. Am. Acad. Dermatol. 2008, 59, 179–206. [Google Scholar] [CrossRef]
  11. Egas-Bejar, D.; Huh, W.W. Rhabdomyosarcoma in adolescent and young adult patients: Current perspectives. Adolesc. Health Med. Ther. 2014, 5, 115. [Google Scholar]
  12. The WHO Classification of Tumours Editorial Board. Soft Tissue and Bone Tumours; International Agency for Research on Cancer: Lyon, France, 2020. [Google Scholar]
  13. Skapek, S.X.; Ferrari, A.; Gupta, A.A.; Lupo, P.J.; Butler, E.; Shipley, J.; Barr, F.G.; Hawkins, D.S. Rhabdomyosarcoma. Nat. Rev. Dis. Primers 2019, 5, 1–19. [Google Scholar] [CrossRef]
  14. Rudzinski, E.R.; Anderson, J.R.; Chi, Y.Y.; Gastier-Foster, J.M.; Astbury, C.; Barr, F.G.; Skapek, S.X.; Hawkins, D.S.; Weigel, B.J.; Pappo, A. Histology, fusion status, and outcome in metastatic rhabdomyosarcoma: A report from the Children’s Oncology Group. Pediatr. Blood Cancer 2017, 64, e26645. [Google Scholar] [CrossRef]
  15. Noujaim, J.; Thway, K.; Jones, R.L.; Miah, A.; Khabra, K.; Langer, R.; Kasper, B.; Judson, I.; Benson, C.; Kollár, A. Adult pleomorphic rhabdomyosarcoma: A multicentre retrospective study. Anticancer Res. 2015, 35, 6213–6217. [Google Scholar]
  16. Agaram, N.P.; LaQuaglia, M.P.; Alaggio, R.; Zhang, L.; Fujisawa, Y.; Ladanyi, M.; Wexler, L.H.; Antonescu, C.R. MYOD1-mutant spindle cell and sclerosing rhabdomyosarcoma: An aggressive subtype irrespective of age. A reappraisal for molecular classification and risk stratification. Mod. Pathol. 2019, 32, 27–36. [Google Scholar] [CrossRef]
  17. Søreide, K.; Sandvik, O.M.; Søreide, J.A.; Giljaca, V.; Jureckova, A.; Bulusu, V.R. Global epidemiology of gastrointestinal stromal tumours (GIST): A systematic review of population-based cohort studies. Cancer Epidemiol. 2016, 40, 39–46. [Google Scholar] [CrossRef] [Green Version]
  18. Heinrich, M.C.; Corless, C.L.; Demetri, G.D.; Blanke, C.D.; Von Mehren, M.; Joensuu, H.; McGreevey, L.S.; Chen, C.J.; Van den Abbeele, A.D.; Druker, B.J. Kinase mutations and imatinib response in patients with metastatic gastrointestinal stromal tumor. J. Clin. Oncol. 2003, 21, 4342–4349. [Google Scholar] [CrossRef]
  19. Hirota, S.; Isozaki, K.; Moriyama, Y.; Hashimoto, K.; Nishida, T.; Ishiguro, S.; Kawano, K.; Hanada, M.; Kurata, A.; Takeda, M. Gain-of-function mutations of c-kit in human gastrointestinal stromal tumors. Science 1998, 279, 577–580. [Google Scholar] [CrossRef]
  20. Janeway, K.A.; Kim, S.Y.; Lodish, M.; Nosé, V.; Rustin, P.; Gaal, J.; Dahia, P.L.; Liegl, B.; Ball, E.R.; Raygada, M. Defects in succinate dehydrogenase in gastrointestinal stromal tumors lacking KIT and PDGFRA mutations. Proc. Natl. Acad. Sci. USA 2011, 108, 314–318. [Google Scholar] [CrossRef] [Green Version]
  21. Miettinen, M.; Lasota, J. Gastrointestinal stromal tumors: Pathology and prognosis at different sites. Semin. Diagn. Pathol. 2006, 23, 70–83. [Google Scholar] [CrossRef]
  22. Ho, M.Y.; Blanke, C.D. Gastrointestinal stromal tumors: Disease and treatment update. Gastroenterology 2011, 140, 1372–1376.E2. [Google Scholar] [CrossRef]
  23. Lopes, L.F.; Bacchi, C.E. Imatinib treatment for gastrointestinal stromal tumour (GIST). J. Cell Mol. Med. 2010, 14, 42–50. [Google Scholar] [CrossRef] [Green Version]
  24. Silva, M.; Veiga, I.; Ribeiro, F.R.; Vieira, J.; Pinto, C.; Pinheiro, M.; Mesquita, B.; Santos, C.; Soares, M.; Dinis, J. Chromosome copy number changes carry prognostic information independent of KIT/PDGFRA point mutations in gastrointestinal stromal tumors. BMC Med. 2010, 8, 26. [Google Scholar] [CrossRef] [Green Version]
  25. Suzuki, H.; Tsukahara, T. A view of pre-mRNA splicing from RNase R resistant RNAs. Int. J. Mol. Sci. 2014, 15, 9331–9342. [Google Scholar] [CrossRef] [Green Version]
  26. 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]
  27. Liu, J.; Liu, T.; Wang, X.; He, A. Circles reshaping the RNA world: From waste to treasure. Mol. Cancer 2017, 16, 58. [Google Scholar] [CrossRef] [Green Version]
  28. Cortés-López, M.; Miura, P. Emerging Functions of Circular RNAs. Yale J. Biol. Med. 2016, 89, 527–537. [Google Scholar]
  29. 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] [Green Version]
  30. Chen, L.L. The biogenesis and emerging roles of circular RNAs. Nat. Rev. Mol. Cell Biol. 2016, 17, 205–211. [Google Scholar] [CrossRef]
  31. Guo, J.U.; Agarwal, V.; Guo, H.; Bartel, D.P. Expanded identification and characterization of mammalian circular RNAs. Genome Biol. 2014, 15, 409. [Google Scholar] [CrossRef]
  32. Wu, J.; Qi, X.; Liu, L.; Hu, X.; Liu, J.; Yang, J.; Yang, J.; Lu, L.; Zhang, Z.; Ma, S.; et al. Emerging Epigenetic Regulation of Circular RNAs in Human Cancer. Mol. Ther. Nucleic Acids 2019, 16, 589–596. [Google Scholar] [CrossRef] [Green Version]
  33. Papatsirou, M.; Artemaki, P.I.; Scorilas, A.; Kontos, C.K. The role of circular RNAs in therapy resistance of patients with solid tumors. Pers. Med. 2020, 17, 469–490. [Google Scholar] [CrossRef] [PubMed]
  34. Shen, B.; Wang, Z.; Li, Z.; Song, H.; Ding, X. Circular RNAs: An emerging landscape in tumor metastasis. Am. J. Cancer Res. 2019, 9, 630. [Google Scholar] [PubMed]
  35. Zhou, R.; Wu, Y.; Wang, W.; Su, W.; Liu, Y.; Wang, Y.; Fan, C.; Li, X.; Li, G.; Li, Y. Circular RNAs (circRNAs) in cancer. Cancer Lett. 2018, 425, 134–142. [Google Scholar] [CrossRef] [PubMed]
  36. 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]
  37. Kristensen, L.S.; Andersen, M.S.; Stagsted, L.V.; 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]
  38. Li, Z.; Huang, C.; Bao, C.; Chen, L.; Lin, M.; Wang, X.; Zhong, G.; Yu, B.; Hu, W.; Dai, L. Exon-intron circular RNAs regulate transcription in the nucleus. Nat. Struct. Mol. Biol. 2015, 22, 256–264. [Google Scholar] [CrossRef]
  39. Gao, Y.; Wang, J.; Zhao, F. CIRI: An efficient and unbiased algorithm for de novo circular RNA identification. Genome Biol. 2015, 16, 1–16. [Google Scholar] [CrossRef] [Green Version]
  40. 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] [Green Version]
  41. Kelly, S.; Greenman, C.; Cook, P.R.; Papantonis, A. Exon skipping is correlated with exon circularization. J. Mol. Biol. 2015, 427, 2414–2417. [Google Scholar] [CrossRef]
  42. Shukla, S.; Kavak, E.; Gregory, M.; Imashimizu, M.; Shutinoski, B.; Kashlev, M.; Oberdoerffer, P.; Sandberg, R.; Oberdoerffer, S. CTCF-promoted RNA polymerase II pausing links DNA methylation to splicing. Nature 2011, 479, 74–79. [Google Scholar] [CrossRef]
  43. Bentley, D.L. Coupling mRNA processing with transcription in time and space. Nat. Rev. Genet. 2014, 15, 163–175. [Google Scholar] [CrossRef] [Green Version]
  44. Kramer, M.C.; Liang, D.; Tatomer, D.C.; Gold, B.; March, Z.M.; Cherry, S.; Wilusz, J.E. Combinatorial control of Drosophila circular RNA expression by intronic repeats, hnRNPs, and SR proteins. Genes Dev. 2015, 29, 2168–2182. [Google Scholar] [CrossRef] [Green Version]
  45. Eger, N.; Schoppe, L.; Schuster, S.; Laufs, U.; Boeckel, J.N. Circular RNA splicing. Circ. RNAs 2018, 1087, 41–52. [Google Scholar]
  46. 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] [Green Version]
  47. 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]
  48. Conn, S.J.; Pillman, K.A.; Toubia, J.; Conn, V.M.; Salmanidis, M.; Phillips, C.A.; Roslan, S.; Schreiber, A.W.; Gregory, P.A.; Goodall, G.J. The RNA binding protein quaking regulates formation of circRNAs. Cell 2015, 160, 1125–1134. [Google Scholar] [CrossRef] [Green Version]
  49. Errichelli, L.; Modigliani, S.D.; Laneve, P.; Colantoni, A.; Legnini, I.; Capauto, D.; Rosa, A.; De Santis, R.; Scarfo, R.; Peruzzi, G. FUS affects circular RNA expression in murine embryonic stem cell-derived motor neurons. Nat. Commun. 2017, 8, 14741. [Google Scholar] [CrossRef]
  50. Ivanov, A.; Memczak, S.; Wyler, E.; Torti, F.; Porath, H.T.; Orejuela, M.R.; Piechotta, M.; Levanon, E.Y.; Landthaler, M.; Dieterich, C. Analysis of intron sequences reveals hallmarks of circular RNA biogenesis in animals. Cell Rep. 2015, 10, 170–177. [Google Scholar] [CrossRef] [Green Version]
  51. Alderton, G.K. Circular RNAs from translocations. Nat. Rev. Cancer 2016, 16, 273. [Google Scholar] [CrossRef]
  52. Guarnerio, J.; Bezzi, M.; Jeong, J.C.; Paffenholz, S.V.; Berry, K.; Naldini, M.M.; Lo-Coco, F.; Tay, Y.; Beck, A.H.; Pandolfi, P.P. Oncogenic role of fusion-circRNAs derived from cancer-associated chromosomal translocations. Cell 2016, 165, 289–302. [Google Scholar] [CrossRef] [Green Version]
  53. Huang, C.; Liang, D.; Tatomer, D.C.; Wilusz, J.E. A length-dependent evolutionarily conserved pathway controls nuclear export of circular RNAs. Genes Dev. 2018, 32, 639–644. [Google Scholar] [CrossRef]
  54. 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]
  55. Park, O.H.; Ha, H.; Lee, Y.; Boo, S.H.; Kwon, D.H.; Song, H.K.; Kim, Y.K. Endoribonucleolytic cleavage of m6A-containing RNAs by RNase P/MRP complex. Mol. Cell 2019, 74, 494–507.e8. [Google Scholar] [CrossRef]
  56. Meng, Y.; Hao, D.; Huang, Y.; Jia, S.; Zhang, J.; He, X.; Liu, D.; Sun, L. Circular RNA circNRIP1 plays oncogenic roles in the progression of osteosarcoma. Mamm. Genome 2021. [Google Scholar] [CrossRef]
  57. Fischer, J.W.; Busa, V.F.; Shao, Y.; Leung, A.K. Structure-mediated RNA decay by UPF1 and G3BP1. Mol. Cell 2020, 78, 70–84. [Google Scholar] [CrossRef]
  58. Preußer, C.; Hung, L.H.; Schneider, T.; Schreiner, S.; Hardt, M.; Moebus, A.; Santoso, S.; Bindereif, A. Selective release of circRNAs in platelet-derived extracellular vesicles. J. Extracell. Vesicles 2018, 7, 1424473. [Google Scholar] [CrossRef] [Green Version]
  59. Yang, Y.; Fan, X.; Mao, M.; Song, X.; Wu, P.; Zhang, Y.; Jin, Y.; Yang, Y.; Chen, L.L.; Wang, Y.; et al. Extensive translation of circular RNAs driven by N6-methyladenosine. Cell Res. 2017, 27, 626–641. [Google Scholar] [CrossRef] [Green Version]
  60. Du, W.W.; Yang, W.; Chen, Y.; Wu, Z.K.; Foster, F.S.; Yang, Z.; Li, X.; Yang, B.B. Foxo3 circular RNA promotes cardiac senescence by modulating multiple factors associated with stress and senescence responses. Eur. Heart J. 2017, 38, 1402–1412. [Google Scholar] [CrossRef] [Green Version]
  61. Bachmayr-Heyda, A.; Reiner, A.T.; Auer, K.; Sukhbaatar, N.; Aust, S.; Bachleitner-Hofmann, T.; Mesteri, I.; Grunt, T.W.; Zeillinger, R.; Pils, D. Correlation of circular RNA abundance with proliferation—Exemplified with colorectal and ovarian cancer, idiopathic lung fibrosis and normal human tissues. Sci. Rep. 2015, 5, 8057. [Google Scholar] [CrossRef]
  62. Aufiero, S.; Reckman, Y.J.; Pinto, Y.M.; Creemers, E.E. Circular RNAs open a new chapter in cardiovascular biology. Nat. Rev. Cardiol. 2019, 16, 503–514. [Google Scholar] [CrossRef]
  63. Tabak, H.F.; Van der Horst, G.; Smit, J.; Winter, A.J.; Mul, Y.; Groot Koerkamp, M.J. Discrimination between RNA circles, interlocked RNA circles and lariats using two-dimensional polyacrylamide gel electrophoresis. Nucleic Acids Res. 1988, 16, 6597–6605. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. 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]
  65. Hoffmann, S.; Otto, C.; Doose, G.; Tanzer, A.; Langenberger, D.; Christ, S.; Kunz, M.; Holdt, L.M.; Teupser, D.; Hackermüller, J.; et al. A multi-split mapping algorithm for circular RNA, splicing, trans-splicing and fusion detection. Genome Biol. 2014, 15, R34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Zeng, X.; Lin, W.; Guo, M.; Zou, Q. A comprehensive overview and evaluation of circular RNA detection tools. PLoS Comput. Biol. 2017, 13, e1005420. [Google Scholar] [CrossRef]
  67. Chen, L.; Wang, C.; Sun, H.; Wang, J.; Liang, Y.; Wang, Y.; Wong, G. The bioinformatics toolbox for circRNA discovery and analysis. Brief. Bioinform. 2021, 22, 1706–1728. [Google Scholar] [CrossRef]
  68. Ng, W.L.; Mohd Mohidin, T.B.; Shukla, K. Functional role of circular RNAs in cancer development and progression. RNA Biol. 2018, 15, 995–1005. [Google Scholar] [CrossRef] [Green Version]
  69. Hansen, T.B.; Jensen, T.I.; Clausen, B.H.; Bramsen, J.B.; Finsen, B.; Damgaard, C.K.; Kjems, J. Natural RNA circles function as efficient microRNA sponges. Nature 2013, 495, 384–388. [Google Scholar] [CrossRef]
  70. Huang, R.; Zhang, Y.; Han, B.; Bai, Y.; Zhou, R.; Gan, G.; Chao, J.; Hu, G.; Yao, H. Circular RNA HIPK2 regulates astrocyte activation via cooperation of autophagy and ER stress by targeting MIR124–2HG. Autophagy 2017, 13, 1722–1741. [Google Scholar] [CrossRef] [Green Version]
  71. You, X.; Vlatkovic, I.; Babic, A.; Will, T.; Epstein, I.; Tushev, G.; Akbalik, G.; Wang, M.; Glock, C.; Quedenau, C.; et al. Neural circular RNAs are derived from synaptic genes and regulated by development and plasticity. Nat. Neurosci. 2015, 18, 603–610. [Google Scholar] [CrossRef] [Green Version]
  72. Thomson, D.W.; Dinger, M.E. Endogenous microRNA sponges: Evidence and controversy. Nat. Rev. Genet. 2016, 17, 272–283. [Google Scholar] [CrossRef]
  73. 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]
  74. Jeck, W.R.; Sharpless, N.E. Detecting and characterizing circular RNAs. Nat. Biotechnol. 2014, 32, 453–461. [Google Scholar] [CrossRef]
  75. 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]
  76. Legnini, I.; Di Timoteo, G.; Rossi, F.; Morlando, M.; Briganti, F.; Sthandier, O.; Fatica, A.; Santini, T.; Andronache, A.; Wade, M. Circ-ZNF609 is a circular RNA that can be translated and functions in myogenesis. Mol. Cell 2017, 66, 22–37.e29. [Google Scholar] [CrossRef] [Green Version]
  77. 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]
  78. Holdt, L.M.; Stahringer, A.; Sass, K.; Pichler, G.; Kulak, N.A.; Wilfert, W.; Kohlmaier, A.; Herbst, A.; Northoff, B.H.; Nicolaou, A.; et al. Circular non-coding RNA ANRIL modulates ribosomal RNA maturation and atherosclerosis in humans. Nat. Commun. 2016, 7, 12429. [Google Scholar] [CrossRef] [Green Version]
  79. Xu, S.; Zhou, L.; Ponnusamy, M.; Zhang, L.; Dong, Y.; Zhang, Y.; Wang, Q.; Liu, J.; Wang, K. A comprehensive review of circRNA: From purification and identification to disease marker potential. PeerJ 2018, 6, e5503. [Google Scholar] [CrossRef]
  80. Park, S.; Phukan, P.D.; Zeeb, M.; Martinez-Yamout, M.A.; Dyson, H.J.; Wright, P.E. Structural Basis for Interaction of the Tandem Zinc Finger Domains of Human Muscleblind with Cognate RNA from Human Cardiac Troponin T. Biochemistry 2017, 56, 4154–4168. [Google Scholar] [CrossRef] [Green Version]
  81. Jishage, M.; Yu, X.; Shi, Y.; Ganesan, S.J.; Chen, W.Y.; Sali, A.; Chait, B.T.; Asturias, F.J.; Roeder, R.G. Architecture of Pol II(G) and molecular mechanism of transcription regulation by Gdown1. Nat. Struct. Mol. Biol. 2018, 25, 859–867. [Google Scholar] [CrossRef]
  82. Li, J.; Sun, D.; Pu, W.; Wang, J.; Peng, Y. Circular RNAs in Cancer: Biogenesis, Function, and Clinical Significance. Trends Cancer 2020, 6, 319–336. [Google Scholar] [CrossRef]
  83. 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.e225. [Google Scholar] [CrossRef] [Green Version]
  84. Yu, C.Y.; Li, T.C.; Wu, Y.Y.; Yeh, C.H.; Chiang, W.; Chuang, C.Y.; Kuo, H.C. The circular RNA circBIRC6 participates in the molecular circuitry controlling human pluripotency. Nat. Commun. 2017, 8, 1149. [Google Scholar] [CrossRef]
  85. Zhang, H.; Wang, G.; Ding, C.; Liu, P.; Wang, R.; Ding, W.; Tong, D.; Wu, D.; Li, C.; Wei, Q.; et al. Increased circular RNA UBAP2 acts as a sponge of miR-143 to promote osteosarcoma progression. Oncotarget 2017, 8, 61687–61697. [Google Scholar] [CrossRef]
  86. 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]
  87. Kartal-Yandim, M.; Adan-Gokbulut, A.; Baran, Y. Molecular mechanisms of drug resistance and its reversal in cancer. Crit. Rev. Biotechnol. 2016, 36, 716–726. [Google Scholar] [CrossRef] [Green Version]
  88. Bai, Y.; Li, Y.; Bai, J.; Zhang, Y. Hsa_circ_0004674 promotes osteosarcoma doxorubicin resistance by regulating the miR-342-3p/FBN1 axis. J. Orthop. Surg. Res. 2021, 16, 510. [Google Scholar] [CrossRef]
  89. Zhang, H.; Yan, J.; Lang, X.; Zhuang, Y. Expression of circ_001569 is upregulated in osteosarcoma and promotes cell proliferation and cisplatin resistance by activating the Wnt/β-catenin signaling pathway. Oncol. Lett. 2018, 16, 5856–5862. [Google Scholar] [CrossRef] [Green Version]
  90. Wu, Z.; Shi, W.; Jiang, C. Overexpressing circular RNA hsa_circ_0002052 impairs osteosarcoma progression via inhibiting Wnt/β-catenin pathway by regulating miR-1205/APC2 axis. Biochem. Biophys. Res. Commun. 2018, 502, 465–471. [Google Scholar] [CrossRef]
  91. Wang, L.; Du, Z.G.; Huang, H.; Li, F.S.; Li, G.S.; Xu, S.N. Circ-0003998 promotes cell proliferative ability and invasiveness by binding to miR-197-3p in osteosarcoma. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 10638–10646. [Google Scholar] [CrossRef]
  92. Zhang, Z.; Pu, F.; Wang, B.; Wu, Q.; Liu, J.; Shao, Z. Hsa_circ_0000285 functions as a competitive endogenous RNA to promote osteosarcoma progression by sponging hsa-miRNA-599. Gene Ther. 2020, 27, 186–195. [Google Scholar] [CrossRef]
  93. Seo, T.; Park, J.; Lee, D.; Hwang Sun, G.; Choe, J. Viral Interferon Regulatory Factor 1 of Kaposi’s Sarcoma-Associated Herpesvirus Binds to p53 and Represses p53-Dependent Transcription and Apoptosis. J. Virol. 2001, 75, 6193–6198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Shin Young, C.; Nakamura, H.; Liang, X.; Feng, P.; Chang, H.; Kowalik Timothy, F.; Jung Jae, U. Inhibition of the ATM/p53 Signal Transduction Pathway by Kaposi’s Sarcoma-Associated Herpesvirus Interferon Regulatory Factor 1. J. Virol. 2006, 80, 2257–2266. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Zhu, K.-P.; Zhang, C.-L.; Hu, J.-P.; Zhang, L. A novel circulating hsa_circ_0081001 act as a potential biomarker for diagnosis and prognosis of osteosarcoma. Int. J. Biol. Sci. 2018, 14, 1513–1520. [Google Scholar] [CrossRef] [Green Version]
  96. Zhu, K.-P.; Ma, X.-L.; Zhang, L.; Zhang, C.-L.; Hu, J.-P.; Zhan, T.-C. Screening circular RNA related to chemotherapeutic resistance in osteosarcoma by RNA sequencing. Epigenomics 2018, 10, 1327–1346. [Google Scholar] [CrossRef]
  97. Wei, W.; Ji, L.; Duan, W.; Zhu, J. Circular RNA circ_0081001 knockdown enhances methotrexate sensitivity in osteosarcoma cells by regulating miR-494-3p/TGM2 axis. J. Orthop. Surg. Res. 2021, 16, 50. [Google Scholar] [CrossRef]
  98. Li, D.; Huang, Y.; Wang, G. Circular RNA circPVT1 Contributes to Doxorubicin (DXR) Resistance of Osteosarcoma Cells by Regulating TRIAP1 via miR-137. Biomed Res. Int. 2021, 2021, 7463867. [Google Scholar] [CrossRef]
  99. Pan, G.; Hu, T.; Chen, X.; Zhang, C. Upregulation Of circMMP9 Promotes Osteosarcoma Progression Via Targeting miR-1265/CHI3L1 Axis. Cancer Manag. Res. 2019, 11, 9225–9231. [Google Scholar] [CrossRef] [Green Version]
  100. Wang, R.; Zhang, S.; Chen, X.; Li, N.; Li, J.; Jia, R.; Pan, Y.; Liang, H. EIF4A3-induced circular RNA MMP9 (circMMP9) acts as a sponge of miR-124 and promotes glioblastoma multiforme cell tumorigenesis. Mol. Cancer 2018, 17, 166. [Google Scholar] [CrossRef]
  101. Jin, H.; Jin, X.; Zhang, H.; Wang, W. Circular RNA hsa-circ-0016347 promotes proliferation, invasion and metastasis of osteosarcoma cells. Oncotarget 2017, 8, 25571–25581. [Google Scholar] [CrossRef] [Green Version]
  102. Li, Z.; Fu, Y.; Ouyang, W.; He, M.; Wang, Y.; Wang, X.; Tan, W. Circ_0016347 Promotes Osteosarcoma Progression by Regulating miR-1225-3p/KCNH1 Axis. Cancer Biother. Radiopharm. 2021, 21, 166. [Google Scholar] [CrossRef]
  103. Li, J.F.; Song, Y.Z. Circular RNA GLI2 promotes osteosarcoma cell proliferation, migration, and invasion by targeting miR-125b-5p. Tumour. Biol. 2017, 39, 1010428317709991. [Google Scholar] [CrossRef] [Green Version]
  104. Deng, N.q.; Li, L.; Gao, J.; Zhou, J.; Wang, Y.; Wang, C.; Liu, Y. Hsa_circ_0009910 promotes carcinogenesis by promoting the expression of miR-449a target IL6R in osteosarcoma. Biochem. Biophys. Res. Commun. 2018, 495, 189–196. [Google Scholar] [CrossRef]
  105. Song, Y.Z.; Li, J.F. Circular RNA hsa_circ_0001564 regulates osteosarcoma proliferation and apoptosis by acting miRNA sponge. Biochem. Biophys. Res. Commun. 2018, 495, 2369–2375. [Google Scholar] [CrossRef]
  106. Jin, J.; Chen, A.; Qiu, W.; Chen, Y.; Li, Q.; Zhou, X.; Jin, D. Dysregulated circRNA_100876 suppresses proliferation of osteosarcoma cancer cells by targeting microRNA-136. J. Cell Biochem. 2019, 120, 15678–15687. [Google Scholar] [CrossRef]
  107. Yao, J.T.; Zhao, S.H.; Liu, Q.P.; Lv, M.Q.; Zhou, D.X.; Liao, Z.J.; Nan, K.J. Over-expression of CircRNA_100876 in non-small cell lung cancer and its prognostic value. Pathol. Res. Pract. 2017, 213, 453–456. [Google Scholar] [CrossRef]
  108. Liu, G.; Huang, K.; Jie, Z.; Wu, Y.; Chen, J.; Chen, Z.; Fang, X.; Shen, S. CircFAT1 sponges miR-375 to promote the expression of Yes-associated protein 1 in osteosarcoma cells. Mol. Cancer 2018, 17, 170. [Google Scholar] [CrossRef]
  109. Gu, H.; Cheng, X.; Xu, J.; Zhou, K.; Bian, C.; Chen, G.; Yin, X. Circular RNA circFAT1(e2) Promotes Osteosarcoma Progression and Metastasis by Sponging miR-181b and Regulating HK2 Expression. Biomed Res. Int. 2020, 2020, 3589871. [Google Scholar] [CrossRef]
  110. Fang, J.; Hong, H.; Xue, X.; Zhu, X.; Jiang, L.; Qin, M.; Liang, H.; Gao, L. 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, 222–232. [Google Scholar] [CrossRef]
  111. Xi, Y.; Fowdur, M.; Liu, Y.; Wu, H.; He, M.; Zhao, J. Differential expression and bioinformatics analysis of circRNA in osteosarcoma. Biosci. Rep. 2019, 39, BSR20181514. [Google Scholar] [CrossRef] [Green Version]
  112. Zhang, M.; Yu, G.Y.; Liu, G.; Liu, W.D. Circular RNA circ_0002137 regulated the progression of osteosarcoma through regulating miR-433-3p/IGF1R axis. J. Cell Mol. Med. 2021. [Google Scholar] [CrossRef]
  113. Zhou, X.; Natino, D.; Qin, Z.; Wang, D.; Tian, Z.; Cai, X.; Wang, B.; He, X. Identification and functional characterization of circRNA-0008717 as an oncogene in osteosarcoma through sponging miR-203. Oncotarget 2017, 9, 22288–22300. [Google Scholar] [CrossRef] [Green Version]
  114. Li, L.; Guo, L.; Yin, G.; Yu, G.; Zhao, Y.; Pan, Y. Upregulation of circular RNA circ_0001721 predicts unfavorable prognosis in osteosarcoma and facilitates cell progression via sponging miR-569 and miR-599. Biomed. Pharmacother. 2019, 109, 226–232. [Google Scholar] [CrossRef]
  115. Li, B.; Li, X. Overexpression of hsa_circ_0007534 predicts unfavorable prognosis for osteosarcoma and regulates cell growth and apoptosis by affecting AKT/GSK-3β signaling pathway. Biomed. Pharmacother. 2018, 107, 860–866. [Google Scholar] [CrossRef]
  116. Zhang, R.; Xu, J.; Zhao, J.; Wang, X. Silencing of hsa_circ_0007534 suppresses proliferation and induces apoptosis in colorectal cancer cells. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 118–126. [Google Scholar]
  117. Hao, L.; Rong, W.; Bai, L.; Cui, H.; Zhang, S.; Li, Y.; Chen, D.; Meng, X. Upregulated circular RNA circ_0007534 indicates an unfavorable prognosis in pancreatic ductal adenocarcinoma and regulates cell proliferation, apoptosis, and invasion by sponging miR-625 and miR-892b. J. Cell. Biochem. 2019, 120, 3780–3789. [Google Scholar] [CrossRef]
  118. Li, G.F.; Li, L.; Yao, Z.Q.; Zhuang, S.J. Hsa_circ_0007534/miR-761/ZIC5 regulatory loop modulates the proliferation and migration of glioma cells. Biochem. Biophys. Res. Commun. 2018, 499, 765–771. [Google Scholar] [CrossRef]
  119. Song, L.; Xiao, Y. Downregulation of hsa_circ_0007534 suppresses breast cancer cell proliferation and invasion by targeting miR-593/MUC19 signal pathway. Biochem. Biophys. Res. Commun. 2018, 503, 2603–2610. [Google Scholar] [CrossRef]
  120. Huang, L.; Chen, M.; Pan, J.; Yu, W. Circular RNA circNASP modulates the malignant behaviors in osteosarcoma via miR-1253/FOXF1 pathway. Biochem. Biophys. Res. Commun. 2018, 500, 511–517. [Google Scholar] [CrossRef]
  121. Zhu, K.-P.; Ma, X.-L.; Zhang, C.-L. Overexpressed circPVT1, a potential new circular RNA biomarker, contributes to doxorubicin and cisplatin resistance of osteosarcoma cells by regulating ABCB1. Int. J. Biol. Sci. 2018, 14, 321–330. [Google Scholar] [CrossRef]
  122. Chen, J.; Li, Y.; Zheng, Q.; Bao, C.; He, J.; Chen, B.; Lyu, D.; Zheng, B.; Xu, Y.; Long, Z.; et al. Circular RNA profile identifies circPVT1 as a proliferative factor and prognostic marker in gastric cancer. Cancer Lett. 2017, 388, 208–219. [Google Scholar] [CrossRef]
  123. Zhu, K.; Niu, L.; Wang, J.; Wang, Y.; Zhou, J.; Wang, F.; Cheng, Y.; Zhang, Q.; Li, H. Circular RNA hsa_circ_0000885 Levels are Increased in Tissue and Serum Samples from Patients with Osteosarcoma. Med. Sci. Monit. 2019, 25, 1499–1505. [Google Scholar] [CrossRef] [PubMed]
  124. Li, S.; Pei, Y.; Wang, W.; Liu, F.; Zheng, K.; Zhang, X. Circular RNA 0001785 regulates the pathogenesis of osteosarcoma as a ceRNA by sponging miR-1200 to upregulate HOXB2. Cell Cycle 2019, 18, 1281–1291. [Google Scholar] [CrossRef] [PubMed]
  125. Du, Y.X.; Guo, L.X.; Pan, H.S.; Liang, Y.M.; Li, X. Circ_ANKIB1 stabilizes the regulation of miR-19b on SOCS3/STAT3 pathway to promote osteosarcoma cell growth and invasion. Hum. Cell 2020, 33, 252–260. [Google Scholar] [CrossRef]
  126. Zhu, X.; Liu, C.; Shi, J.; Zhou, Z.; Chen, S.; Jami, S.A. Circular RNA circANKIB1 promotes the progression of osteosarcoma by regulating miR-217/PAX3 axis. J. Bone Oncol. 2021, 27, 100347. [Google Scholar] [CrossRef] [PubMed]
  127. Chen, J.; Liu, G.; Wu, Y.; Ma, J.; Wu, H.; Xie, Z.; Chen, S.; Yang, Y.; Wang, S.; Shen, P.; et al. CircMYO10 promotes osteosarcoma progression by regulating miR-370-3p/RUVBL1 axis to enhance the transcriptional activity of β-catenin/LEF1 complex via effects on chromatin remodeling. Mol. Cancer 2019, 18, 150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Wu, Y.; Xie, Z.; Chen, J.; Chen, J.; Ni, W.; Ma, Y.; Huang, K.; Wang, G.; Wang, J.; Ma, J.; et al. Circular RNA circTADA2A promotes osteosarcoma progression and metastasis by sponging miR-203a-3p and regulating CREB3 expression. Mol. Cancer 2019, 18, 73. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Yu, W.; Jiang, H.; Zhang, H.; Li, J. Hsa_circ_0003998 promotes cell proliferation and invasion by targeting miR-326 in non-small cell lung cancer. Onco. Targets Ther. 2018, 11, 5569–5577. [Google Scholar] [CrossRef] [Green Version]
  130. Wang, L.; Wang, P.; Su, X.; Zhao, B. Circ_0001658 promotes the proliferation and metastasis of osteosarcoma cells via regulating miR-382-5p/YB-1 axis. Cell Biochem. Funct. 2020, 38, 77–86. [Google Scholar] [CrossRef]
  131. Li, X.; Sun, X.H.; Xu, H.Y.; Pan, H.S.; Liu, Y.; He, L. Circ_ORC2 enhances the regulatory effect of miR-19a on its target gene PTEN to affect osteosarcoma cell growth. Biochem. Biophys. Res. Commun. 2019, 514, 1172–1178. [Google Scholar] [CrossRef]
  132. Yanbin, Z.; Jing, Z. CircSAMD4A accelerates cell proliferation of osteosarcoma by sponging miR-1244 and regulating MDM2 mRNA expression. Biochem. Biophys. Res. Commun. 2019, 516, 102–111. [Google Scholar] [CrossRef]
  133. Wang, J.; Ni, J.; Song, D.; Ding, M.; Huang, J.; Li, W.; He, G. The regulatory effect of has-circ-0001146/miR-26a-5p/MNAT1 network on the proliferation and invasion of osteosarcoma. Biosci. Rep. 2020, 40, BSR20201232. [Google Scholar] [CrossRef]
  134. Fang, C.; Wang, X.; Guo, D.; Fang, R.; Zhu, T. Circular RNA CircITGA7 Promotes Tumorigenesis of Osteosarcoma via miR-370/PIM1 Axis. Comput. Math. Methods Med. 2020, 2020, 1367576. [Google Scholar] [CrossRef]
  135. Yang, G.; Zhang, T.; Ye, J.; Yang, J.; Chen, C.; Cai, S.; Ma, J. Circ-ITGA7 sponges miR-3187-3p to upregulate ASXL1, suppressing colorectal cancer proliferation. Cancer Manag. Res. 2019, 11, 6499–6509. [Google Scholar] [CrossRef] [Green Version]
  136. Li, S.; Yang, J.; Liu, X.; Guo, R.; Zhang, R. circITGA7 Functions as an Oncogene by Sponging miR-198 and Upregulating FGFR1 Expression in Thyroid Cancer. BioMed Res. Int. 2020, 2020, 8084028. [Google Scholar] [CrossRef]
  137. Li, L.; Kong, X.-a.; Zang, M.; Dong, J.; Feng, Y.; Gui, B.; Hu, Y. Hsa_circ_0003732 promotes osteosarcoma cells proliferation via miR-545/CCNA2 axis. Biosci. Rep. 2020, 40, BSR20191552. [Google Scholar] [CrossRef]
  138. Lin, E.; Liu, S.; Xiang, W.; Zhang, H.; Xie, C. CircEIF4G2 Promotes Tumorigenesis and Progression of Osteosarcoma by Sponging miR-218. BioMed Res. Int. 2020, 2020, 8386936. [Google Scholar] [CrossRef]
  139. Mao, Y.; Zhang, L.; Li, Y. circEIF4G2 modulates the malignant features of cervical cancer via the miR-218/HOXA1 pathway. Mol. Med. Rep. 2019, 19, 3714–3722. [Google Scholar] [CrossRef] [Green Version]
  140. Zhang, H.; Meng, F.; Dong, S. circSMARCA5 Promoted Osteosarcoma Cell Proliferation, Adhesion, Migration, and Invasion through a Competing Endogenous RNA Network. BioMed Res. Int. 2020, 2020, 2539150. [Google Scholar] [CrossRef]
  141. Kong, Z.; Wan, X.; Zhang, Y.; Zhang, P.; Zhang, Y.; Zhang, X.; Qi, X.; Wu, H.; Huang, J.; Li, Y. Androgen-responsive circular RNA circSMARCA5 is up-regulated and promotes cell proliferation in prostate cancer. Biochem. Biophys. Res. Commun. 2017, 493, 1217–1223. [Google Scholar] [CrossRef]
  142. Li, Z.; Zhou, Y.; Yang, G.; He, S.; Qiu, X.; Zhang, L.; Deng, Q.; Zheng, F. Using circular RNA SMARCA5 as a potential novel biomarker for hepatocellular carcinoma. Clin. Chim. Acta 2019, 492, 37–44. [Google Scholar] [CrossRef]
  143. Tian, J.D.C.; Liang, L. Involvement of circular RNA SMARCA5/microRNA-620 axis in the regulation of cervical cancer cell proliferation, invasion and migration. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 8589–8598. [Google Scholar] [CrossRef] [PubMed]
  144. Barbagallo, D.; Caponnetto, A.; Cirnigliaro, M.; Brex, D.; Barbagallo, C.; D’Angeli, F.; Morrone, A.; Caltabiano, R.; Barbagallo, G.M.; Ragusa, M.; et al. CircSMARCA5 Inhibits Migration of Glioblastoma Multiforme Cells by Regulating a Molecular Axis Involving Splicing Factors SRSF1/SRSF3/PTB. Int. J. Mol. Sci. 2018, 19, 480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Lu, Q.; Fang, T. Circular RNA SMARCA5 correlates with favorable clinical tumor features and prognosis, and increases chemotherapy sensitivity in intrahepatic cholangiocarcinoma. J. Clin. Lab. Anal. 2020, 34, e23138. [Google Scholar] [CrossRef] [PubMed]
  146. Ji, X.; Shan, L.; Shen, P.; He, M. Circular RNA circ_001621 promotes osteosarcoma cells proliferation and migration by sponging miR-578 and regulating VEGF expression. Cell Death Dis. 2020, 11, 18. [Google Scholar] [CrossRef] [PubMed]
  147. Li, H.; He, L.; Tuo, Y.; Huang, Y.; Qian, B. Circular RNA hsa_circ_0000282 contributes to osteosarcoma cell proliferation by regulating miR-192/XIAP axis. BMC Cancer 2020, 20, 1026. [Google Scholar] [CrossRef] [PubMed]
  148. Tan, X.; Tan, D.; Li, H.; Lin, Y.; Wen, Z.; Zeng, C. circEPSTI1 Acts as a ceRNA to Regulate the Progression of Osteosarcoma. Curr. Cancer Drug Targets 2020, 20, 288–294. [Google Scholar] [CrossRef] [PubMed]
  149. Chen, B.; Wei, W.; Huang, X.; Xie, X.; Kong, Y.; Dai, D.; Yang, L.; Wang, J.; Tang, H.; Xie, X. circEPSTI1 as a Prognostic Marker and Mediator of Triple-Negative Breast Cancer Progression. Theranostics 2018, 8, 4003–4015. [Google Scholar] [CrossRef]
  150. Li, X.; Liu, Y.; Zhang, X.; Shen, J.; Xu, R.; Liu, Y.; Yu, X. Circular RNA hsa_circ_0000073 contributes to osteosarcoma cell proliferation, migration, invasion and methotrexate resistance by sponging miR-145-5p and miR-151-3p and upregulating NRAS. Aging 2020, 12, 14157–14173. [Google Scholar] [CrossRef]
  151. Zhang, C.; Zhou, H.; Yuan, K.; Xie, R.; Chen, C. Overexpression of hsa_circ_0136666 predicts poor prognosis and initiates osteosarcoma tumorigenesis through miR-593-3p/ZEB2 pathway. Aging 2020, 12, 10488–10496. [Google Scholar] [CrossRef]
  152. Jin, C.; Wang, A.; Liu, L.; Wang, G.; Li, G. Hsa_circ_0136666 promotes the proliferation and invasion of colorectal cancer through miR-136/SH2B1 axis. J. Cell. Physiol. 2019, 234, 7247–7256. [Google Scholar] [CrossRef]
  153. Liu, L.H.; Tian, Q.Q.; Liu, J.; Zhou, Y.; Yong, H. Upregulation of hsa_circ_0136666 contributes to breast cancer progression by sponging miR-1299 and targeting CDK6. J. Cell Biochem. 2019, 120, 12684–12693. [Google Scholar] [CrossRef]
  154. Liu, Y.D.; Liu, L.P. Circ100284 promotes invasion and migration of osteosarcoma cells by down-regulating PTEN and EMP1. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 6540–6550. [Google Scholar] [CrossRef]
  155. Cao, J.; Liu, X.S. Circular RNA 0060428 sponges miR-375 to promote osteosarcoma cell proliferation by upregulating the expression of RPBJ. Gene 2020, 740, 144520. [Google Scholar] [CrossRef]
  156. Li, J.; Zhang, F.; Li, H.; Peng, F.; Wang, Z.; Peng, H.; He, J.; Li, Y.; He, L.; Wei, L. Circ_0010220-mediated miR-503-5p/CDCA4 axis contributes to osteosarcoma progression tumorigenesis. Gene 2020, 763, 145068. [Google Scholar] [CrossRef]
  157. Lu, Z.; Wang, C.; Lv, X.; Dai, W. Hsa_circ_0010220 regulates miR-198/Syntaxin 6 axis to promote osteosarcoma progression. J. Bone Oncol. 2021, 28, 100360. [Google Scholar] [CrossRef]
  158. Jiang, Y.; Hou, J.; Zhang, X.; Xu, G.; Wang, Y.; Shen, L.; Wu, Y.; Li, Y.; Yao, L. Circ-XPO1 upregulates XPO1 expression by sponging multiple miRNAs to facilitate osteosarcoma cell progression. Exp. Mol. Pathol. 2020, 117, 104553. [Google Scholar] [CrossRef]
  159. Chen, G.; Wang, Q.; Yang, Q.; Li, Z.; Du, Z.; Ren, M.; Zhao, H.; Song, Y.; Zhang, G. Circular RNAs hsa_circ_0032462, hsa_circ_0028173, hsa_circ_0005909 are predicted to promote CADM1 expression by functioning as miRNAs sponge in human osteosarcoma. PLoS ONE 2018, 13, e0202896. [Google Scholar] [CrossRef] [Green Version]
  160. Xu, B.; Yang, T.; Wang, Z.; Zhang, Y.; Liu, S.; Shen, M. CircRNA CDR1as/miR-7 signals promote tumor growth of osteosarcoma with a potential therapeutic and diagnostic value. Cancer Manag. Res. 2018, 10, 4871–4880. [Google Scholar] [CrossRef] [Green Version]
  161. Xu, L.; Zhang, M.; Zheng, X.; Yi, P.; Lan, C.; Xu, M. The circular RNA ciRS-7 (Cdr1as) acts as a risk factor of hepatic microvascular invasion in hepatocellular carcinoma. J. Cancer Res. Clin. Oncol. 2017, 143, 17–27. [Google Scholar] [CrossRef]
  162. Yu, L.; Gong, X.; Sun, L.; Zhou, Q.; Lu, B.; Zhu, L. The Circular RNA Cdr1as Act as an Oncogene in Hepatocellular Carcinoma through Targeting miR-7 Expression. PLoS ONE 2016, 11, e0158347. [Google Scholar] [CrossRef] [Green Version]
  163. Barbagallo, D.; Condorelli, A.; Ragusa, M.; Salito, L.; Sammito, M.; Banelli, B.; Caltabiano, R.; Barbagallo, G.; Zappalà, A.; Battaglia, R.; et al. Dysregulated miR-671-5p/CDR1-AS/CDR1/VSNL1 axis is involved in glioblastoma multiforme. Oncotarget 2016, 7, 4746–4759. [Google Scholar] [CrossRef]
  164. Uhr, K.; Sieuwerts, A.M.; de Weerd, V.; Smid, M.; Hammerl, D.; Foekens, J.A.; Martens, J.W.M. Association of microRNA-7 and its binding partner CDR1-AS with the prognosis and prediction of 1st-line tamoxifen therapy in breast cancer. Sci. Rep. 2018, 8, 9657. [Google Scholar] [CrossRef]
  165. Zhao, Z.; Ji, M.; Wang, Q.; He, N.; Li, Y. Circular RNA Cdr1as Upregulates SCAI to Suppress Cisplatin Resistance in Ovarian Cancer via miR-1270 Suppression. Mol. Ther. Nucleic Acids 2019, 18, 24–33. [Google Scholar] [CrossRef] [Green Version]
  166. Yang, Y.; Kong, F.; Cai, Y.; Ding, Q.; Tang, B. Enhanced CDR1as Promotes the Development of Bladder Urothelial Carcinoma. Clin. Lab. 2020, 66. [Google Scholar] [CrossRef]
  167. Li, C.; Li, M.; Xue, Y. Downregulation of CircRNA CDR1as specifically triggered low-dose Diosbulbin-B induced gastric cancer cell death by regulating miR-7-5p/REGγ axis. Biomed. Pharmacother. 2019, 120, 109462. [Google Scholar] [CrossRef]
  168. Li, Y.; Zhang, J.; Pan, S.; Zhou, J.; Diao, X.; Liu, S. CircRNA CDR1as knockdown inhibits progression of non-small-cell lung cancer by regulating miR-219a-5p/SOX5 axis. Thorac. Cancer 2020, 11, 537–548. [Google Scholar] [CrossRef] [Green Version]
  169. Xiao, B.; Zhang, X.; Li, X.; Zhao, Z. Circ_001569 regulates FLOT2 expression to promote the proliferation, migration, invasion and EMT of osteosarcoma cells through sponging miR-185-5p. Open Life Sci. 2020, 15, 476–487. [Google Scholar] [CrossRef]
  170. Xie, H.; Ren, X.; Xin, S.; Lan, X.; Lu, G.; Lin, Y.; Yang, S.; Zeng, Z.; Liao, W.; Ding, Y.Q.; et al. Emerging roles of circRNA_001569 targeting miR-145 in the proliferation and invasion of colorectal cancer. Oncotarget 2016, 7, 26680–26691. [Google Scholar] [CrossRef]
  171. Qi, H.; Sun, Y.; Jiang, Y.; Li, X. Upregulation of circular RNA circ_0000502 predicts unfavorable prognosis in osteosarcoma and facilitates cell progression via sponging miR-1238. J. Cell Biochem. 2018. [Google Scholar] [CrossRef]
  172. Zheng, S.; Qian, Z.; Jiang, F.; Ge, D.; Tang, J.; Chen, H.; Yang, J.; Yao, Y.; Yan, J.; Zhao, L.; et al. CircRNA LRP6 promotes the development of osteosarcoma via negatively regulating KLF2 and APC levels. Am. J. Transl. Res. 2019, 11, 4126–4138. [Google Scholar]
  173. Mao, X.; Guo, S.; Gao, L.; Li, G. Circ-XPR1 promotes osteosarcoma proliferation through regulating the miR-214-5p/DDX5 axis. Hum. Cell 2021, 34, 122–131. [Google Scholar] [CrossRef] [PubMed]
  174. Ma, W.; Xue, N.; Zhang, J.; Wang, D.; Yao, X.; Lin, L.; Xu, Q. circUBAP2 regulates osteosarcoma progression via the miR-204-3p/HMGA2 axis. Int. J. Oncol. 2021, 58, 298–311. [Google Scholar] [CrossRef] [PubMed]
  175. Sun, J.; Yin, A.; Zhang, W.; Lv, J.; Liang, Y.; Li, H.; Li, Y.; Li, X. CircUBAP2 Inhibits Proliferation and Metastasis of Clear Cell Renal Cell Carcinoma via Targeting miR-148a-3p/FOXK2 Pathway. Cell Transpl. 2020, 29, 963689720925751. [Google Scholar] [CrossRef] [PubMed]
  176. Wu, Y.; Zhi, L.; Zhao, Y.; Yang, L.; Cai, F. Knockdown of circular RNA UBAP2 inhibits the malignant behaviours of esophageal squamous cell carcinoma by microRNA-422a/Rab10 axis. Clin. Exp. Pharmacol. Physiol. 2020, 47, 1283–1290. [Google Scholar] [CrossRef]
  177. Wang, S.; Li, Q.; Wang, Y.; Li, X.; Wang, R.; Kang, Y.; Xue, X.; Meng, R.; Wei, Q.; Feng, X. Upregulation of circ-UBAP2 predicts poor prognosis and promotes triple-negative breast cancer progression through the miR-661/MTA1 pathway. Biochem. Biophys. Res. Commun. 2018, 505, 996–1002. [Google Scholar] [CrossRef]
  178. Sheng, M.; Wei, N.; Yang, H.Y.; Yan, M.; Zhao, Q.X.; Jing, L.J. CircRNA UBAP2 promotes the progression of ovarian cancer by sponging microRNA-144. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 7283–7294. [Google Scholar] [CrossRef]
  179. Gao, A.M.; Yuan, C.; Hu, A.X.; Liu, X.S. circ_ARF3 regulates the pathogenesis of osteosarcoma by sponging miR-1299 to maintain CDK6 expression. Cell Signal 2020, 72, 109622. [Google Scholar] [CrossRef]
  180. Liu, X.; Zhong, Y.; Li, J.; Shan, A. Circular RNA circ-NT5C2 acts as an oncogene in osteosarcoma proliferation and metastasis through targeting miR-448. Oncotarget 2017, 8, 114829–114838. [Google Scholar] [CrossRef] [Green Version]
  181. Liu, W.; Zhang, J.; Zou, C.; Xie, X.; Wang, Y.; Wang, B.; Zhao, Z.; Tu, J.; Wang, X.; Li, H. Microarray expression profile and functional analysis of circular RNAs in osteosarcoma. Cell. Physiol. Biochem. 2017, 43, 969–985. [Google Scholar] [CrossRef]
  182. Li, Z.Q.; Wang, Z.; Zhang, Y.; Lu, C.; Ding, Q.L.; Ren, R.; Cheng, B.B.; Lou, L.X. CircRNA_103801 accelerates proliferation of osteosarcoma cells by sponging miR-338-3p and regulating HIF-1/Rap1/PI3K-Akt pathway. J. Biol. Regul. Homeost. Agents 2021, 35, 1021–1028. [Google Scholar] [CrossRef]
  183. Qin, G.; Wu, X. Hsa_circ_0032463 acts as the tumor promoter in osteosarcoma by regulating the miR-330-3p/PNN axis. Int. J. Mol. Med. 2021, 47, 92. [Google Scholar] [CrossRef]
  184. Qin, G.; Wu, X. Circular RNA hsa_circ_0032463 Acts as the Tumor Promoter in Osteosarcoma by Regulating the MicroRNA 498/LEF1 Axis. Mol. Cell Biol. 2021, 41, e0010021. [Google Scholar] [CrossRef]
  185. Chen, Z.; Xu, W.; Zhang, D.; Chu, J.; Shen, S.; Ma, Y.; Wang, Q.; Liu, G.; Yao, T.; Huang, Y.; et al. circCAMSAP1 promotes osteosarcoma progression and metastasis by sponging miR-145-5p and regulating FLI1 expression. Mol. Ther. Nucleic Acids 2021, 23, 1120–1135. [Google Scholar] [CrossRef]
  186. Liu, Y.; Jiang, Y.; Xu, L.; Qu, C.; Zhang, L.; Xiao, X.; Chen, W.; Li, K.; Liang, Q.; Wu, H. circ-NRIP1 Promotes Glycolysis and Tumor Progression by Regulating miR-186-5p/MYH9 Axis in Gastric Cancer. Cancer Manag. Res. 2020, 12, 5945–5956. [Google Scholar] [CrossRef]
  187. Dong, Z.; Liu, Y.; Wang, Q.; Wang, H.; Ji, J.; Huang, T.; Khanal, A.; Niu, H.; Cao, Y. The circular RNA-NRIP1 plays oncogenic roles by targeting microRNA-505 in the renal carcinoma cell lines. J. Cell Biochem. 2020, 121, 2236–2246. [Google Scholar] [CrossRef]
  188. Li, X.; Ma, N.; Zhang, Y.; Wei, H.; Zhang, H.; Pang, X.; Li, X.; Wu, D.; Wang, D.; Yang, Z.; et al. Circular RNA circNRIP1 promotes migration and invasion in cervical cancer by sponging miR-629-3p and regulating the PTP4A1/ERK1/2 pathway. Cell Death Dis. 2020, 11, 399. [Google Scholar] [CrossRef]
  189. Xu, Y.; Yao, T.; Ni, H.; Zheng, R.; Huang, K.; Huang, Y.; Gao, J.; Qiao, D.; Shen, S.; Ma, J. Circular RNA circSIPA1L1 Contributes to Osteosarcoma Progression Through the miR-411-5p/RAB9A Signaling Pathway. Front. Cell Dev. Biol. 2021, 9, 642605. [Google Scholar] [CrossRef]
  190. Yuan, J.; Liu, Y.; Zhang, Q.; Ren, Z.; Li, G.; Tian, R. CircPRDM2 Contributes to Doxorubicin Resistance of Osteosarcoma by Elevating EZH2 via Sponging miR-760. Cancer Manag. Res. 2021, 13, 4433–4445. [Google Scholar] [CrossRef]
  191. Tang, G.; Liu, L.; Xiao, Z.; Wen, S.; Chen, L.; Yang, P. CircRAB3IP upregulates twist family BHLH transcription factor (TWIST1) to promote osteosarcoma progression by sponging miR-580-3p. Bioengineered 2021, 12, 3385–3397. [Google Scholar] [CrossRef]
  192. Wu, P.F.; Tang, X.M.; Sun, H.L.; Jiang, H.T.; Ma, J.; Kong, Y.H. Depletion of circRNA circ_CDK14 inhibits osteosarcoma progression by regulating the miR-520a-3p/GAB1 axis. Neoplasma 2021, 68, 798–809. [Google Scholar] [CrossRef]
  193. Zhang, Z.; Zhou, Q.; Luo, F.; Zhou, R.; Xu, J.; Xiao, J.; Dai, F.; Song, L. Circular RNA circ-CHI3L1.2 modulates cisplatin resistance of osteosarcoma cells via the miR-340-5p/LPAATβ axis. Hum. Cell 2021, 34, 1558–1568. [Google Scholar] [CrossRef]
  194. Fang, Y.; Long, F. Circular RNA circ_0000337 contributes to osteosarcoma via the miR-4458/BACH1 pathway. Cancer Biomark 2020, 28, 411–419. [Google Scholar] [CrossRef]
  195. Song, H.; Xu, D.; Shi, P.; He, B.; Li, Z.; Ji, Y.; Agbeko, C.K.; Wang, J. Upregulated circ RNA hsa_circ_0000337 promotes cell proliferation, migration, and invasion of esophageal squamous cell carcinoma. Cancer Manag. Res. 2019, 11, 1997–2006. [Google Scholar] [CrossRef] [Green Version]
  196. Wu, X.; Yan, L.; Liu, Y.; Shang, L. Circ_0000527 promotes osteosarcoma cell progression through modulating miR-646/ARL2 axis. Aging 2021, 13, 6091–6102. [Google Scholar] [CrossRef]
  197. Zhang, L.; Wu, J.; Li, Y.; Jiang, Y.; Wang, L.; Chen, Y.; Lv, Y.; Zou, Y.; Ding, X. Circ_0000527 promotes the progression of retinoblastoma by regulating miR-646/LRP6 axis. Cancer Cell Int. 2020, 20, 301. [Google Scholar] [CrossRef]
  198. Yang, B.; Li, L.; Tong, G.; Zeng, Z.; Tan, J.; Su, Z.; Liu, Z.; Lin, J.; Gao, W.; Chen, J.; et al. Circular RNA circ_001422 promotes the progression and metastasis of osteosarcoma via the miR-195-5p/FGF2/PI3K/Akt axis. J. Exp. Clin. Cancer Res. 2021, 40, 235. [Google Scholar] [CrossRef]
  199. Zhang, Z.; Zhao, M.; Wang, G. Hsa_circ_0051079 functions as an oncogene by regulating miR-26a-5p/TGF-β1 in osteosarcoma. Cell Biosci. 2019, 9, 94. [Google Scholar] [CrossRef]
  200. Huo, S.; Dou, D. Circ_0056285 Regulates Proliferation, Apoptosis and Glycolysis of Osteosarcoma Cells via miR-1244/TRIM44 Axis. Cancer Manag. Res. 2021, 13, 1257–1270. [Google Scholar] [CrossRef]
  201. Gao, P.; Zhao, X.; Yu, K.; Zhu, Z. Circ_0084582 Facilitates Cell Growth, Migration, Invasion, and Angiopoiesis in Osteosarcoma via Mediating the miR-485-3p/JAG1 Axis. Front. Genet. 2021, 12, 690956. [Google Scholar] [CrossRef]
  202. Guarnerio, J.; Zhang, Y.; Cheloni, G.; Panella, R.; Mae Katon, J.; Simpson, M.; Matsumoto, A.; Papa, A.; Loretelli, C.; Petri, A.; et al. Intragenic antagonistic roles of protein and circRNA in tumorigenesis. Cell Res. 2019, 29, 628–640. [Google Scholar] [CrossRef]
  203. Hu, Y.; Gu, J.; Shen, H.; Shao, T.; Li, S.; Wang, W.; Yu, Z. Circular RNA LARP4 correlates with decreased Enneking stage, better histological response, and prolonged survival profiles, and it elevates chemosensitivity to cisplatin and doxorubicin via sponging microRNA-424 in osteosarcoma. J. Clin. Lab. Anal. 2020, 34, e23045. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Zhang, J.; Liu, H.; Hou, L.; Wang, G.; Zhang, R.; Huang, Y.; Chen, X.; Zhu, J. Circular RNA_LARP4 inhibits cell proliferation and invasion of gastric cancer by sponging miR-424-5p and regulating LATS1 expression. Mol. Cancer 2017, 16, 151. [Google Scholar] [CrossRef] [PubMed]
  205. Chen, Z.; Zuo, X.; Pu, L.; Zhang, Y.; Han, G.; Zhang, L.; Wu, J.; Wang, X. circLARP4 induces cellular senescence through regulating miR-761/RUNX3/p53/p21 signaling in hepatocellular carcinoma. Cancer Sci. 2019, 110, 568–581. [Google Scholar] [CrossRef] [PubMed]
  206. Zou, T.; Wang, P.L.; Gao, Y.; Liang, W.T. Circular RNA_LARP4 is lower expressed and serves as a potential biomarker of ovarian cancer prognosis. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 7178–7182. [Google Scholar] [CrossRef]
  207. Wang, L.; Zhang, G.C.; Kang, F.B.; Zhang, L.; Zhang, Y.Z. hsa_circ0021347 as a Potential Target Regulated by B7-H3 in Modulating the Malignant Characteristics of Osteosarcoma. BioMed Res. Int. 2019, 2019, 9301989. [Google Scholar] [CrossRef] [Green Version]
  208. Zhu, K.P.; Zhang, C.L.; Ma, X.L.; Hu, J.P.; Cai, T.; Zhang, L. Analyzing the Interactions of mRNAs and ncRNAs to Predict Competing Endogenous RNA Networks in Osteosarcoma Chemo-Resistance. Mol. Ther. 2019, 27, 518–530. [Google Scholar] [CrossRef] [Green Version]
  209. Li, S.; Pei, Y.; Wang, W.; Liu, F.; Zheng, K.; Zhang, X. Extracellular nanovesicles-transmitted circular RNA has_circ_0000190 suppresses osteosarcoma progression. J. Cell Mol. Med. 2020, 24, 2202–2214. [Google Scholar] [CrossRef]
  210. Chen, S.; Li, T.; Zhao, Q.; Xiao, B.; Guo, J. Using circular RNA hsa_circ_0000190 as a new biomarker in the diagnosis of gastric cancer. Clin. Chim. Acta 2017, 466, 167–171. [Google Scholar] [CrossRef]
  211. Feng, Y.; Zhang, L.; Wu, J.; Khadka, B.; Fang, Z.; Gu, J.; Tang, B.; Xiao, R.; Pan, G.; Liu, J. CircRNA circ_0000190 inhibits the progression of multiple myeloma through modulating miR-767-5p/MAPK4 pathway. J. Exp. Clin. Cancer Res. 2019, 38, 54. [Google Scholar] [CrossRef]
  212. Du, R.; Fu, B.; Sun, G.; Ma, B.; Deng, M.; Zhu, X.; Kong, D. Circular RNA circ_0046264 Suppresses Osteosarcoma Progression via microRNA-940/Secreted Frizzled Related Protein 1 Axis. Tohoku J. Exp. Med. 2021, 254, 189–197. [Google Scholar] [CrossRef]
  213. Yang, L.; Wang, J.; Fan, Y.; Yu, K.; Jiao, B.; Su, X. Hsa_circ_0046264 up-regulated BRCA2 to suppress lung cancer through targeting hsa-miR-1245. Respir Res. 2018, 19, 115. [Google Scholar] [CrossRef]
  214. Yang, J.; Han, Q.; Li, C.; Yang, H.; Chen, X.; Wang, X. Circular RNA circ_0001105 Inhibits Progression and Metastasis of Osteosarcoma by Sponging miR-766 and Activating YTHDF2 Expression. Onco. Targets Ther. 2020, 13, 1723–1736. [Google Scholar] [CrossRef] [Green Version]
  215. Jiang, X.; Chen, D. Circular RNA hsa_circ_0000658 inhibits osteosarcoma cell proliferation and migration via the miR-1227/IRF2 axis. J. Cell Mol. Med. 2021, 25, 510–520. [Google Scholar] [CrossRef]
  216. Liu, D.Y.; Li, Z.; Zhang, K.; Jiao, N.; Lu, D.G.; Zhou, D.W.; Meng, Y.B.; Sun, L. Circular RNA CircMTO1 suppressed proliferation and metastasis of osteosarcoma through miR-630/KLF6 axis. Eur. Rev. Med. Pharmacol. Sci. 2021, 25, 86–93. [Google Scholar] [CrossRef]
  217. Liu, Y.; Dong, Y.; Zhao, L.; Su, L.; Luo, J. Circular RNA-MTO1 suppresses breast cancer cell viability and reverses monastrol resistance through regulating the TRAF4/Eg5 axis. Int. J. Oncol. 2018, 53, 1752–1762. [Google Scholar] [CrossRef]
  218. Chen, M.; Ai, G.; Zhou, J.; Mao, W.; Li, H.; Guo, J. circMTO1 promotes tumorigenesis and chemoresistance of cervical cancer via regulating miR-6893. Biomed. Pharmacother. 2019, 117, 109064. [Google Scholar] [CrossRef]
  219. Han, D.; Li, J.; Wang, H.; Su, X.; Hou, J.; Gu, Y.; Qian, C.; Lin, Y.; Liu, X.; Huang, M.; et al. Circular RNA circMTO1 acts as the sponge of microRNA-9 to suppress hepatocellular carcinoma progression. Hepatology 2017, 66, 1151–1164. [Google Scholar] [CrossRef] [Green Version]
  220. Cheng, C.; Zhang, H.; Dai, Z.; Zheng, J. Circular RNA circVRK1 suppresses the proliferation, migration and invasion of osteosarcoma cells by regulating zinc finger protein ZNF652 expression via microRNA miR-337-3p. Bioengineered 2021, 12, 5411–5427. [Google Scholar] [CrossRef]
  221. Yan, N.; Xu, H.; Zhang, J.; Xu, L.; Zhang, Y.; Zhang, L.; Xu, Y.; Zhang, F. Circular RNA profile indicates circular RNA VRK1 is negatively related with breast cancer stem cells. Oncotarget 2017, 8, 95704–95718. [Google Scholar] [CrossRef] [Green Version]
  222. He, Y.; Mingyan, E.; Wang, C.; Liu, G.; Shi, M.; Liu, S. CircVRK1 regulates tumor progression and radioresistance in esophageal squamous cell carcinoma by regulating miR-624-3p/PTEN/PI3K/AKT signaling pathway. Int. J. Biol. Macromol. 2019, 125, 116–123. [Google Scholar] [CrossRef]
  223. Liu, S.; Zhang, J.; Zheng, T.; Mou, X.; Xin, W. Circ_WWC3 overexpression decelerates the progression of osteosarcoma by regulating miR-421/PDE7B axis. Open Life Sci. 2021, 16, 229–241. [Google Scholar] [CrossRef]
  224. Xu, G.; Zhang, Z.; Zhang, L.; Chen, Y.; Li, N.; Lv, Y.; Li, Y.; Xu, X. miR-4326 promotes lung cancer cell proliferation through targeting tumor suppressor APC2. Mol. Cell. Biochem. 2018, 443, 151–157. [Google Scholar] [CrossRef]
  225. Ying, X.; Qi, L.-Y.; Zhou, F.; Wang, Y.; Liu, J.-H. MiR-939 promotes the proliferation of human ovarian cancer cells by repressing APC2 expression. Biomed. Pharmacother. 2015, 71, 64–69. [Google Scholar] [CrossRef]
  226. Zhang, P.R.; Ren, J.; Wan, J.S.; Sun, R.; Li, Y. Circular RNA hsa_circ_0002052 promotes osteosarcoma via modulating miR-382/STX6 axis. Hum. Cell 2020, 33, 810–818. [Google Scholar] [CrossRef]
  227. Ren, C.; Liu, J.; Zheng, B.; Yan, P.; Sun, Y.; Yue, B. The circular RNA circ-ITCH acts as a tumour suppressor in osteosarcoma via regulating miR-22. Artif. Cells Nanomed. Biotechnol. 2019, 47, 3359–3367. [Google Scholar] [CrossRef] [Green Version]
  228. Li, H.; Lan, M.; Liao, X.; Tang, Z.; Yang, C. Circular RNA cir-ITCH Promotes Osteosarcoma Migration and Invasion through cir-ITCH/miR-7/EGFR Pathway. Technol. Cancer Res. Treat. 2020, 19, 1533033819898728. [Google Scholar] [CrossRef] [Green Version]
  229. Zhou, W.; Liu, Y.; Wu, X. Down-regulation of circITCH promotes osteosarcoma development and resistance to doxorubicin via the miR-524/RASSF6 axis. J. Gene Med. 2021, 23, e3373. [Google Scholar] [CrossRef]
  230. Ma, X.-L.; Zhu, K.-P.; Zhang, C.-L. Circular RNA circ_HIPK3 is down-regulated and suppresses cell proliferation, migration and invasion in osteosarcoma. J. Cancer 2018, 9, 1856–1862. [Google Scholar] [CrossRef]
  231. Huang, Z.; Yuan, C.; Gu, H.; Cheng, X.; Zhou, K.; Xu, J.; Yin, X.; Xia, J. Circular RNA circHIPK3 Promotes Cell Metastasis through miR-637/STAT3 Axis in Osteosarcoma. BioMed Res. Int. 2020, 2020, 2727060. [Google Scholar] [CrossRef]
  232. Wen, Y.; Li, B.; He, M.; Teng, S.; Sun, Y.; Wang, G. circHIPK3 promotes proliferation and migration and invasion via regulation of miR-637/HDAC4 signaling in osteosarcoma cells. Oncol. Rep. 2021, 45, 169–179. [Google Scholar] [CrossRef]
  233. Jin, Y.; Li, L.; Zhu, T.; Liu, G. Circular RNA circ_0102049 promotes cell progression as ceRNA to target MDM2 via sponging miR-1304-5p in osteosarcoma. Pathol. Res. Pract. 2019, 215, 152688. [Google Scholar] [CrossRef] [PubMed]
  234. Zhang, X.; Hu, Z.; Li, W.; Liu, Z.; Li, J.; Wang, Z.; Martin, V.T.; Yan, B.; Yu, B. Circular RNA 0102049 suppresses the progression of osteosarcoma through modulating miR-520g-3p/PLK2 axis. Bioengineered 2021, 12, 2022–2032. [Google Scholar] [CrossRef] [PubMed]
  235. Tian, J.; Xu, Y.Y.; Li, L.; Hao, Q. MiR-490-3p sensitizes ovarian cancer cells to cisplatin by directly targeting ABCC2. Am. J. Transl. Res. 2017, 9, 1127–1138. [Google Scholar] [PubMed]
  236. Chen, S.; Chen, X.; Xiu, Y.L.; Sun, K.X.; Zong, Z.H.; Zhao, Y. microRNA 490-3P enhances the drug-resistance of human ovarian cancer cells. J. Ovarian Res. 2014, 7, 84. [Google Scholar] [CrossRef]
  237. Xiang, X.; Mei, H.; Qu, H.; Zhao, X.; Li, D.; Song, H.; Jiao, W.; Pu, J.; Huang, K.; Zheng, L.; et al. miRNA-584-5p exerts tumor suppressive functions in human neuroblastoma through repressing transcription of matrix metalloproteinase 14. Biochim. Biophys. Acta 2015, 1852, 1743–1754. [Google Scholar] [CrossRef] [Green Version]
  238. Li, Q.; Li, Z.; Wei, S.; Wang, W.; Chen, Z.; Zhang, L.; Chen, L.; Li, B.; Sun, G.; Xu, J.; et al. Overexpression of miR-584-5p inhibits proliferation and induces apoptosis by targeting WW domain-containing E3 ubiquitin protein ligase 1 in gastric cancer. J. Exp. Clin. Cancer Res. 2017, 36, 59. [Google Scholar] [CrossRef]
  239. Li, G.; Wu, X.; Qian, W.; Cai, H.; Sun, X.; Zhang, W.; Tan, S.; Wu, Z.; Qian, P.; Ding, K.; et al. CCAR1 5′ UTR as a natural miRancer of miR-1254 overrides tamoxifen resistance. Cell Res. 2016, 26, 655–673. [Google Scholar] [CrossRef] [Green Version]
  240. Zhang, C.; Zhao, Y.; Zeng, B. Enhanced chemosensitivity by simultaneously inhibiting cell cycle progression and promoting apoptosis of drug-resistant osteosarcoma MG63/DXR cells by targeting Cyclin D1 and Bcl-2. Cancer Biomark 2012, 12, 155–167. [Google Scholar] [CrossRef]
  241. Zhao, Y.; Zhang, C.L.; Zeng, B.F.; Wu, X.S.; Gao, T.T.; Oda, Y. Enhanced chemosensitivity of drug-resistant osteosarcoma cells by lentivirus-mediated Bcl-2 silencing. Biochem. Biophys. Res. Commun. 2009, 390, 642–647. [Google Scholar] [CrossRef]
  242. Zhang, P.; Li, J. Down-regulation of circular RNA hsa_circ_0007534 suppresses cell growth by regulating miR-219a-5p/SOX5 axis in osteosarcoma. J. Bone Oncol. 2021, 27, 100349. [Google Scholar] [CrossRef]
  243. Sui, H.; Fan, Z.; Li, Q. Signal transduction pathways and transcriptional mechanisms of ABCB1/Pgp-mediated multiple drug resistance in human cancer cells. J. Int. Med. Res. 2012, 40, 426–435. [Google Scholar] [CrossRef] [Green Version]
  244. Hee Choi, Y.; Yu, A.M. ABC transporters in multidrug resistance and pharmacokinetics, and strategies for drug development. Curr. Pharm. Des. 2014, 20, 793–807. [Google Scholar] [CrossRef]
  245. Bruhn, O.; Cascorbi, I. Polymorphisms of the drug transporters ABCB1, ABCG2, ABCC2 and ABCC3 and their impact on drug bioavailability and clinical relevance. Expert Opin. Drug Metab. Toxicol. 2014, 10, 1337–1354. [Google Scholar] [CrossRef]
  246. Zhang, T.-M. TRIAP1 inhibition activates the cytochrome c/Apaf-1/caspase-9 signaling pathway to enhance human ovarian cancer sensitivity to cisplatin. Chemotherapy 2019, 64, 119–128. [Google Scholar] [CrossRef]
  247. Li, X.; Yang, H.; Tian, Q.; Liu, Y.; Weng, Y. Upregulation of microRNA-17-92 cluster associates with tumor progression and prognosis in osteosarcoma. Neoplasma 2014, 61, 453–460. [Google Scholar] [CrossRef] [Green Version]
  248. Li, X.; Wang, F.; Wu, Z.; Lin, J.; Lan, W.; Lin, J. MicroRNA-19b targets Mfn1 to inhibit Mfn1-induced apoptosis in osteosarcoma cells. Neoplasma 2014, 61, 265–273. [Google Scholar] [CrossRef]
  249. Jia, Z.; Wang, K.; Zhang, A.; Wang, G.; Kang, C.; Han, L.; Pu, P. miR-19a and miR-19b overexpression in gliomas. Pathol. Oncol. Res. 2013, 19, 847–853. [Google Scholar] [CrossRef]
  250. Jiang, T.; Ye, L.; Han, Z.; Liu, Y.; Yang, Y.; Peng, Z.; Fan, J. miR-19b-3p promotes colon cancer proliferation and oxaliplatin-based chemoresistance by targeting SMAD4: Validation by bioinformatics and experimental analyses. J. Exp. Clin. Cancer Res. 2017, 36, 131. [Google Scholar] [CrossRef] [Green Version]
  251. 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]
  252. Huang, A.; Zheng, H.; Wu, Z.; Chen, M.; Huang, Y. Circular RNA-protein interactions: Functions, mechanisms, and identification. Theranostics 2020, 10, 3503–3517. [Google Scholar] [CrossRef]
  253. Toptan, T.; Abere, B.; Nalesnik, M.A.; Swerdlow, S.H.; Ranganathan, S.; Lee, N.; Shair, K.H.; Moore, P.S.; Chang, Y. Circular DNA tumor viruses make circular RNAs. Proc. Natl. Acad. Sci. USA 2018, 115, E8737–E8745. [Google Scholar] [CrossRef] [Green Version]
  254. 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]
  255. Abere, B.; Li, J.; Zhou, H.; Toptan, T.; Moore Patrick, S.; Chang, Y.; Damania, B. 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]
  256. Yao, S.; Jia, X.; Wang, F.; Sheng, L.; Song, P.; Cao, Y.; Shi, H.; Fan, W.; Ding, X.; Gao, S.J.; et al. CircRNA ARFGEF1 functions as a ceRNA to promote oncogenic KSHV-encoded viral interferon regulatory factor induction of cell invasion and angiogenesis by upregulating glutaredoxin 3. PLoS Pathog. 2021, 17, e1009294. [Google Scholar] [CrossRef]
  257. Rossi, F.; Legnini, I.; Megiorni, F.; Colantoni, A.; Santini, T.; Morlando, M.; Di Timoteo, G.; Dattilo, D.; Dominici, C.; Bozzoni, I. Circ-ZNF609 regulates G1-S progression in rhabdomyosarcoma. Oncogene 2019, 38, 3843–3854. [Google Scholar] [CrossRef] [Green Version]
  258. Jia, N.; Tong, H.; Zhang, Y.; Katayama, H.; Wang, Y.; Lu, W.; Zhang, S.; Wang, J. CeRNA Expression Profiling Identifies KIT-Related circRNA-miRNA-mRNA Networks in Gastrointestinal Stromal Tumour. Front. Genet. 2019, 10, 825. [Google Scholar] [CrossRef]
  259. Ma, N.; Xu, H.; Zhou, Y.; Liu, M.; Wei Zhou, J.; Jie Wang, C. Analyzing the molecular mechanism of the tissue specificity of gastrointestinal stromal tumors by using bioinformatics approaches. J. Buon 2018, 23, 1149–1155. [Google Scholar]
  260. 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]
  261. Dou, Y.; Cha, D.J.; Franklin, J.L.; Higginbotham, J.N.; Jeppesen, D.K.; Weaver, A.M.; Prasad, N.; Levy, S.; Coffey, R.J.; Patton, J.G. Circular RNAs are down-regulated in KRAS mutant colon cancer cells and can be transferred to exosomes. Sci. Rep. 2016, 6, 37982. [Google Scholar] [CrossRef]
  262. Zhao, W.; Dong, M.; Pan, J.; Wang, Y.; Zhou, J.; Ma, J.; Liu, S. Circular RNAs: A novel target among non-coding RNAs with potential roles in malignant tumors (Review). Mol. Med. Rep. 2019, 20, 3463–3474. [Google Scholar] [CrossRef]
Figure 1. Biogenesis and diversity of circRNAs. CircRNAs are produced by exon-skipping or another non-canonical event (like back-splicing) that can initiate internal splicing. CircRNAs are present in blood, body fluids and tissues, and can serve as potential disease biomarkers. They are non-polyadenylated, unlike mRNAs, and they are covalently closed. CircRNAs are resistant to exonucleases digestion due to their lack of 5′ or 3′ end. CircRNAs can be exonic, intronic or a combination of both (this illustration was created with BioRender.com, accessed on 3 November 2021).
Figure 1. Biogenesis and diversity of circRNAs. CircRNAs are produced by exon-skipping or another non-canonical event (like back-splicing) that can initiate internal splicing. CircRNAs are present in blood, body fluids and tissues, and can serve as potential disease biomarkers. They are non-polyadenylated, unlike mRNAs, and they are covalently closed. CircRNAs are resistant to exonucleases digestion due to their lack of 5′ or 3′ end. CircRNAs can be exonic, intronic or a combination of both (this illustration was created with BioRender.com, accessed on 3 November 2021).
Biomedicines 09 01642 g001
Table 1. Upregulated circRNAs in OS.
Table 1. Upregulated circRNAs in OS.
CircRNASample TypesTargeted Molecules/AxisInvolved Signaling PathwayPrognostic ImplicationsRole in OS TumorigenesisOther Tumor Types Tested
circMMP9 [99]MG-63, SaOS-2, HOS, U2OS, hFOB1.19, patient tissuesmiR1265/CHI3L1 1-Advanced tumor stage, adverse prognosisCell proliferation, invasion, migration, apoptosis inhibitionGlioblastoma [100]
hsa-circ-0016347 [101,102]MG-63, SaOS-2, hFOB1.19, patient tissues, mouse xenograft modelmiR-214/Caspase-1--Cell proliferation, invasion, metastasis-
OS cell lines, mouse xenograft modelmiR-1225-3p/KCNH1 2Tumor progression
cir-GLI2 [103]MG-63, SaOS-2, HOS, U2OS, hFOB1.19, HEK293, patient tissuesmiR-125b-5p--Cell proliferation, invasion, migration-
hsa_circ_0009910 [104]MG-63, SaOS-2, U2OS, hFOB1.19, patient tissuesmiR-449a/IL6R 3JAK1/STAT3-Cell proliferation, cell cycle arrest, apoptosis inhibition-
hsa_circ_0001564 [105]MG-63, SaOS-2, HOS, U2OS, patient tissuesmiR-29c-3p--Cell proliferation, cell cycle arrest, apoptosis inhibition-
hsa_circ_0004674 [88,96]KHOS, KHOS/DXR, U2OS, U2OS/DXR, MG-63, MG-63/DXR, patient tissuesmiR-490-3p-ABCC2,
miR-1254-EGFR
-Adverse prognosisDoxorubicin chemoresistance-
KHOS, KHOS/DXR, U2OS, U2OS/DXR, patient tissues, mouse xenograft modelmiR-342-3p/FBN1Wnt/β-catenin-
circRNA_100876 [106]MG-63, OS-732, SaOS-2, HOS, U2OS, 143B, patient tissues, mouse xenograft modelmiR-136-Large tumor size, poor tumor differentiation degreeCell proliferation, cell cycle progression, apoptosis inhibitionNon-small cell lung cancer [107]
hsa_circ_0081001 [95,97]MG-63, MG-63/DXR KHOS, KHOS/DXR, U2OS, U2OS/DXR, patient tissues and serum--Advanced tumor stage, lung metastasis, advanced Enneking stage, independent prognosticatorTumor growth, methotrexate chemoresistance-
HOS, HOS/MTXR, U2OS, U2OS/MTXR, patient tissues, mouse xenograft modelmiR-494-3p/TGM2 4--
circFAT1 [108,109]SJSA-1, MG-63, SaOS-2, HOS, U2OS, 143B, hFOB1.19, patient tissues, mouse xenograft modelmiR-375/YAP1Hippo-Cell proliferation, migration, apoptosis inhibitionGastric cancer [110]
SW1353, MG-63, hFOB1.19miR-181b/HK2 5-Cell growth, migration, metastasis
circ_2137, circ_20403 [111,112]Patient tissues-Phosphatidyl-inositol signaling pathway, Inositol phosphate metabolism---
MG-63, U2OS, 143B, G292, hFOB1.19, patient tissues, mouse xenograft modelcirc_2137/miR-433-3p/IGF1R-Invasion, cell cycle progression, apoptosis inhibition
circ_0001721 [113,114]MG-63, SaOS-2, HOS, U2OS, hFOB1.19, patient tissues, mouse xenograft modelmiR-569, miR-599-Large tumor size, high differentiation grade, adverse 5-year prognosisCell proliferation-
hsa_circ_0007534 [115,116]MG-63, SaOS-2, HOS, U2OS, hFOB1.19, patient tissues, mouse xenograft model-AKT/GSK-3βLarge tumor size, poor differentiation grade, independent prognosticatorTumor growth, apoptosis inhibitionPancreatic ductal carcinoma [117], glioma [118], colorectal cancer [116], breast cancer [119]
MG-63, HOS, U2OS, 143B, hFOB1.19, patient tissues, mouse xenograft modelmiR-219a-5p/SOX-5 6--Cell proliferation, colony formation, migration, invasion
circNASP [120]MG-63, 143B, patient tissuesmiR-1253/FOXF1-Large tumor size, lung metastasisCell proliferation, cell cycle progression, invasion-
circPVT1 [98,121]MG-63, SaOS-2, KHOS, U2OS, patient tissues and serumP-gp/ABCB1-Lung metastasis, advanced Enneking stage, poor survivalTumor growth, doxorubicin chemoresistanceGastric cancer [122]
KHOS, U2OS, 293T, patient tissues, mouse xenograft modelmiR-137/TRIAP1 7--
hsa_circ_0000885 [123]patient tissues and serum--Advanced Enneking stage, lung metastasis--
circ-0001785 [124]MG-63, SJSA-1, SaOS-2, HOS, U2OS, hFOB1.19, mouse xenograft modelmiR-1200/HOXB2Bcl-2 family, PI3K/AKT/mTOR-Cell proliferation, apoptosis inhibition-
circ_ANKIB1 [95,125,126]MG-63, SaOS-2, HOS, U2OS, 143B, hFOB1.19miR-19b/SOCS3SOCS3/STAT3-Cell proliferation, invasion-
HOS, U2OS, hFOB1.19, patient tissues, mouse xenograft modelmiR-217/PAX3-Cell proliferation, migration, invasion, tumor growth, apoptosis inhibition
circMYO10 [127]SJSA-1, MG-63, HOS, U2OS, 143B, hFOB1.19, patient tissues, mouse xenograft modelmiR-370-3p/RUVBL1β-catenin/LEF1/c-MYC-Cell proliferation, EMT-
circTADA2A [128]HEK-293, SJSA-1, MG-63, HOS, U2OS, 143B, patient tissues, mouse xenograft modelmiR-203a-3p/CREB3--Cell proliferation, migration, invasion-
circ-0003998 [91]MG-63, SaOS-2, U2OS, 143B, hFOB1.19, patient tissuesmiR-197-3p/KLF10 8TGF-βPoor overall survivalCell proliferation, invasionNon-small cell lung cancer [129]
circ_0001658 [130]MG-63, SaOS-2, HOS, U2OS, 143B, hFOB1.19, patient tissuesmiR-382-5p/YB-1-Early relapseCell proliferation, migration, metastasis, apoptosis inhibition-
circ_ORC2 [131]MG-63, SaOS-2, HOS, U2OS, 143B, hFOB1.19miR-19a/PTEN 9PI3K/AKT-Cell proliferation, invasion, apoptosis inhibition-
circSAMD4A [132]MG-63, HOS, U2OS, 143B, hFOB1.19, patient tissues, mouse xenograft modelmiR-1244/MDM2-Poor overall survivalCell proliferation, metastasis-
has-circ-0001146 [133]MG-63, 143B, hFOB1.19, patient tissues, mouse xenograft modelmiR-26a-5p/MNAT1--Cell proliferation, viability, invasion, apoptosis inhibition-
circITGA7 [134]SW1353, MG-63, HOS, U2OS, patient tissuesmiR-370/PIM1--Cell proliferation, migration, metastasisColorectal cancer [135], thyroid cancer [136]
hsa_circ_0003732 [137]MG-63, SaOS-2, HOS, U2OS, hFOB1.19, patient tissuesmiR-545/CCNA2-Poor prognosisCell proliferation-
circEIF4G [138]MG-63, HOS, patient tissuesmiR-218EGFR, PI3K/AKT, ErbB-Cell proliferation, migration, invasionCervical cancer [139]
circSMARCA5 [140]MG-63, HOS---Cell proliferation, cell cycle progression, adhesion, migration, metastasisProstate cancer [141], hepatocellular carcinoma [142], cervical cancer [143], glioma [144], intrahepatic cholangiocarcinoma [145]
circ_001621 [146]MG-63, U2OS, hFOB1.19, patient tissues, mouse xenograft modelmiR-578/VEGF 10, CDK4, MMP9-Advanced tumor stage, poor survival, metastasisCell proliferation, migration-
hsa_circ_0000282 [147]SOSP-9607, MG-63, U2OS, 143B, hFOB1.19, patient tissuesmiR-192/XIAP 11-High tumor differentiation grade, advanced Enneking stageCell proliferation, apoptosis inhibition-
circEPSTI1 [148]MG-63, U2OS, mouse xenograft modelmiR-892b/MCL1--Cell proliferation, migration, invasionBreast cancer [149]
hsa_circ_0000073 [150]MG-63, MG-63/DXR, HOS, U2OS, U2OS/DXR, 143B, patient tissuesmiR-145-5p/NRAS, miR-151-3p/NRAS--Cell proliferation, migration, invasion, methotrexate chemoresistance-
hsa_circ_0136666 [151]OS cell lines, hFOB1.19, patient tissues, mouse xenograft modelmiR-593-3p/ZEB2-Large tumor size, adnanved tumor stage, poor survivalCell proliferation, migration, invasion, apoptosis inhibitionColorectal cancer [152], breast cancer [153]
circ_100284 [154]MG-63, SaOS-2, U2OS, hFOB1.19, patient tissuesLSD1 12-EZH2 13/PTEN-Large tumor size, lung metastasis, poor survivalCell viability, invasion, cell cycle progression, apoptosis inhibition-
circ-0060428 [155]SaOS-2, HOS, U2OS, 143B, hFOB1.19miR-375/RPBJ/Bax-bcl-2-cleaved-caspace-3--Cell proliferation, apoptosis inhibition-
circ_0010220 [156,157]MG-63, SaOS-2, HOS, U2OS, 143B, hFOB1.19miR-503-5p/CDCA4 14--Tumor growth, migration, invasion-
HOS, U2OS, hFOB1.19 patient tissues, mouse xenograft modelmiR-198/Syntaxin_6Poor survivalCell proliferation, migration, invasion,
circ-XPO1 [158]MG-63, SaOS-2, HOS, U2OS, hFOB1.19, patient tissuesmiR-23a-3p/XPO1, miR-23b-3p/XPO1,
miR-130a-5p/XPO1,
miR-23c/XPO1
-Poor survivalTumor growth, invasion, apoptosis inhibition-
hsa_circ_0032462, hsa_circ_0028173, hsa_circ_0005909 [159]MG-63, HOS, U2OS, 143B, hFOB1.19has-miR-338-3/CADM1 15,
has-miR-142-5p/CADM1
Cell cycle pathway, cell adhesion molecules pathway, p53 signaling, oxidative phosphorylation pathway, cytokine-cytokine receptor interaction pathway---
CDR1as [160]MG-63, SaOS-2, HOS, U2OS, 143B, hFOB1.19, patient tissues, mouse xenograft modelmiR-7/EGFRA, miR-7/CCNE1, miR-7/PI3KCD, miR-7/RAF1, N-cadherin, E-cadherin-Large tumor size, advanced Enneking stage, distant metastasisEMT, cell migration, tumor growthHepatocelular carcinoma [161,162], glioblastoma [163], breast cancer [164], ovarian cancer [165], urothelial cancer [166], gastric cancer [167], non-small cell lung cancer [168]
circ_001569 [89,169]MG-63, U2OS, hFOB1.19, patient tissuesGSK-3β, β-cateninWnt/β-cateninAdvanced tumor stage, distant metastasisCell proliferation, cisplatin chemoresistanceColorectal cancer [170]
MG-63, SaOS-2, HOS, hFOB1.19, patient tissues, mouse xenograft modelmiR-185-5p/FLOT2 16--Cell proliferation, migration, invasion, EMT
circ_0000502 [171]MG-63, SaOS-2, HOS, U2OS, patient tissues, mouse xenograft modelmiR-1238--Cell proliferation, migration, invasion, apoptosis inhibition-
circLRP6 [172]MG-63, HOS, U2OS, SaOS-2, patient tissuesLSD1, EZH2-Shorter disease-free survival and overall survivalCell proliferation, migration, invasion-
hsa_circ_0000285 [92]MG-63, SaOS-2, HOS, U2OS, hFOB1.19, mouse xenograft modelhsa-miR-599/TGF-β2--Cell proliferation and migration-
circ-XPR1 [173]MG-63, U2OS, patient tissuesmiR-214-5p/DDX5-Poor overall survival and disease-free survivalCell proliferation-
circUBAP2 [85,174]MG-63, U2OS, hFOB1.19, patient tissues, mouse xenograft modelmiR-143/bcl-2-Advanced tumor stage, poor survival, poor prognosisTumor growth, apoptosis inhibitionRenal cancer [175] esophageal squamous carcinoma [176], triple-negative breast cancer [177], ovarian cancer [178]
MG-63, HOS, SaOS-2, U2OS, patient tissuesmiR-204-3p/HMGA2-Poor survivalCell proliferation, migration, invasion, apoptosis inhibition
circ_ARF3 [179]MG-63, SaOS-2, U2OS, patient tissues, mouse xenograft modelmiR-1299/CDK6--Cell growth, cell cycle progression-
circ-NT5C2SOSP-9607, MG-63, SaOS-2, U2OS, patient tissues, mouse xenograft modelmiR-448 [180]-Advanced Enneking stage, lung metastasisCell proliferation, invasion-
circRNA-0008717 [113]SW1353, MG-63, SaOS-2, HOS, U2OS, hFOB1.19, patient tissuesmiR-203/Bmi-1-Poor overall survival and progression-free survivalCell proliferation, invasion-
hsa_circRNA_103801 [181,182]MG-63, HOS, U2OS, U2OS/MTX300, ZOS, ZOS-M, 143B, hFOB1.19, patient tissues-HIF-1, VEGF, angiogenesis pathway, Rap1 signaling pathway, PI3K/AKT signaling pathway---
OS cell lines, patient tissuesmiR-338-3p/HIF1-Rap1PI3K/AKT-Cell proliferation, migration, invasion
hsa_circ_0032463 [183,184]SOSP-9607, HOS, U2OS, SaOS-2, patient tissuesmiR-330-3p/PNN 17--Cell proliferation, viability, invasion, adhesion, apoptosis inhibition-
SOSP-9607, SW1353, SaOS-2, HOS, U2OS, hFOB1.19, patient tissuesmiR-498/LEF1--Tumor proliferation, migration, apoptosis inhibition
circCAMSAP1 [185]HOS, U2OS, 143B, hFOB1.19, patient tissues, mouse xenograft modelmiR-145-5p/FLI1 18--Cell growth, apoptosis inhibition, migration, invasion-
circNRIP1 [56]MG-63, U2OS, patient tissuesMETTL3/circNRIP1/miR-199a/FOXC2--Cell proliferation, migration, apoptosisGastric cancer [186], renal cancer [187], cervical cancer [188]
circSIPA1L1 [189]MG-63, SJSA-1, HOS, U2OS, 143B, hFOB1.19, patient tissues, mouse xenograft modelmiR-411-5p/RAB9A--Cell vitality, invasion, migration, proliferation-
circPRDM2 [190]MG-63, MG-63/DXR, KHOS, KHOS/DXR, hFOB1.19, patient tissues, mouse xenograft modelmiR-760/EZH2--Cell migration, invasion, colony formation, doxorubicin chemoresistance, apoptosis inhibition-
circRAB3IP [191]MG-63, HOS, U2OS, 143B, patient tissues, mouse xenograft modelmiR-580-3p/TWIST 19-Advanced tumor stage, distant metastasisCell proliferation, migration, invasion-
circ_CDK14 [192]OS cell lines, patient tissuesmiR-520a-3p/GAB1 20--Cell proliferation, metastasis, tumorigenesis, apoptosis-
circ-CHI3L1.2 [193]OS cell linesmiR-340-5p/LPAATβ 21--Cell migration, invasion, EMT, cisplatin chemoresistance-
circ_0000337 [194]MG-63, HOS, U2OS, 143B, hFOB1.19miR-4458/BACH1--Cell growth, migrationEsophageal squamous carcinoma [195]
circ_0000527 [196]MG-63, SaOS-2, HOS, U2OS, hFOB1.19, patient tissuesmiR-646/ARL2--Cell growth, cell cycle, inflammationRetinoblastoma [197]
circ_001422 [198]MNNG, MG-63, SaOS-2, U2OS, 143B, hFOB1.19, patient tissues, mouse xenograft modelmiR-195-5p/FGF2PI3K/AKTAdvanced tumor stage, large tumor size, distant metastasis, poor overall survivalCell proliferation, metastasis, apoptosis inhibition-
hsa_circ_0051079 [199]MG-63, SaOS-2, HOS, KHOS, U2OS, 143B, patient tissues, mouse xenograft modelmiR-26a-5p/TGF-β1--Tumor proliferation, metastasis-
circ_0056285 [200]MG-63, HOS, U2OS, 143B, hFOB1.19, patient tissues and serum, mouse xenograft modelmiR-1244/TRIM44 22--Cell proliferation, glycolysis, apoptosis inhibition-
circ_0084582 [201]MG-63, U2OS, hFOB1.19, patient tissues, mouse xenograft modelmiR-485-3p/JAG1 23Notch pathway-Cell proliferation, cell cycle progression, migration, invasion, angiogenesis-
circPOK [202]-ILF2/3 complex--Tumorigenic-
1 Chitinase-3-like Protein 1; 2 Potassium Voltage-Gated Channel Subfamily H Member 1; 3 Interleukin 6 Receptor; 4 Translutaminase-2; 5 Human Glandular Kallilrein2; 6 Sex-Determining Region Y-box 5; 7 TP53-regulated Inhibitor of Apoptosis 1; 8 Kruppel-like Factor 10; 9 Phosphate and Tensin Homolog; 10 Vascular Endothelial Growth Factor; 11 X-linked Inhibitor of Apoptosis Protein; 12 Lysine-Specific Histone Demethylase 1A; 13 Enhancer of Zeste Homolog 2; 14 Cycle-Associated Protein 4; 15 Cell Adhesion Molecule 1; 16 Flotillin-2, 17 Pinin Desmosome Associated Protein; 18 Friend Leukemia Virus Integration 1, 19 Twist Family BHLH Transcription Factor; 20 GRB2 Associated Binding Protein 1; 21 Lysophosphatidic Acid Acyltransferase β; 22 Tripartite Motif Containing 44; 23 Jagged1.
Table 2. Downregulated circRNAs in OS.
Table 2. Downregulated circRNAs in OS.
CircRNASample TypesTargeted Molecules/AxisInvolved Signaling PathwayPrognostic RelevanceRole in OS TumorigenesisOther Tumor Types Tested
circ-LARP4 [203]MG-63, SaOS-2, patient tissuesmiR-424-Low Enneking stage, tumor cell necrosis after adjuvant chemotherapy, prolonged disease-free survival and overall survivalIncreasing chemosensitivity to cisplatin and doxorubicinGastric cancer [204], hepatocellular carcinoma [205], ovarian cancer [206]
hsa_circ0021347 [207]MG-63, patient tissuesB7-H3-Low tumor stage, low Enneking stage, prolonged survival--
hsa_circ_0001258 [208]MG-63, MG-63/DXR KHOS, KHOS/DXR, U2OS, U2OS/DXR, patient tissuesmiR-744-3p/GSTM2--Cell viability, increasing chemosensitivity to doxorubicin-
has_circ_0000190 [209]MG-63, SaOS-2, KHOS, SJSA1, patient tissues, mouse xenograft modelmiR-76-5p/TET1--Cell proliferation, migration, and invasion inhibitionGastric cancer [210], plasma cell myeloma [211]
circ_0046264 [212]MG-63, HOS, U2OS, 143B, patient tissuesmiR-940/SFRP1 1-Low tumor size and Ki67 proliferation indexCell proliferation, migration, and invasion inhibitionLung cancer [213]
circ_0001105 [214]MG-63, U2OS, 143B, patient tissues, mouse xenograft modelmiR-766/YTHDF2-Prolonged survivalCell viability, migration, and invasion inhibition-
circ_32279, circ_24831 [111]Patient tissues-Phosphatidylinositol signaling pathway, inositol phosphate metabolism---
hsa_circRNA_104980 [181]MG-63, HOS, U2OS, U2OS/MTX300, ZOS, ZOS-M, 143B, hFOB1.19, patient tissueshsa-miR-1298-3p, hsa-miR-660-3pPhosphoric ester hydrolase activity, carbohydrate derivative binding---
hsa_circ_0000658 [215]MG-63, SJSA-1, SaOS-2, HOS, U2OS, hFOB1.19, patient tissues, mouse xenograft modelmiR-1227/IRF2 2--Cell cycle, proliferation, invasion, and migration inhibition-
circMTO1 [216]MG-63, SaOS-2, HOS, U2OS, hFOB1.19, patient tissuesmiR-630/KLF6-Low Enneking stage, prolonged overall survivalProliferation, migration and invasion inhibition, apoptosis inductionBreast cancer [217], cervical cancer [218], hepatocellular carcinoma [219]
circVRK1 [220]MG-63, SaOS-2, HOS, U2OS, 143B, hFOB1.19, patient tissues, mouse xenograft modelmiR-337/ZNF652 3-Low levels correlate to poor prognosis and distant metastasisGrowth, migration, invasion inhibitionBreast cancer [221], esophageal cancer [222]
circ_WWC3 [223]MG-63, SaOS-2, HOS, U2OS, hFOB1.19, patient tissues, mouse xenograft modelmiR-421/PDE7B 4--Cell growth, migration, invasion inhibition, apoptosis induction-
1 Secreted Frizzled Related Protein 1; 2 Interferon Regulatory Factor 2; 3 Zinc-Finger Protein 652; 4 Phosphodiesterase 7B.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lakiotaki, E.; Kanakoglou, D.S.; Pampalou, A.; Karatrasoglou, E.A.; Piperi, C.; Korkolopoulou, P. Dissecting the Role of Circular RNAs in Sarcomas with Emphasis on Osteosarcomas. Biomedicines 2021, 9, 1642. https://doi.org/10.3390/biomedicines9111642

AMA Style

Lakiotaki E, Kanakoglou DS, Pampalou A, Karatrasoglou EA, Piperi C, Korkolopoulou P. Dissecting the Role of Circular RNAs in Sarcomas with Emphasis on Osteosarcomas. Biomedicines. 2021; 9(11):1642. https://doi.org/10.3390/biomedicines9111642

Chicago/Turabian Style

Lakiotaki, Eleftheria, Dimitrios S. Kanakoglou, Andromachi Pampalou, Eleni A. Karatrasoglou, Christina Piperi, and Penelope Korkolopoulou. 2021. "Dissecting the Role of Circular RNAs in Sarcomas with Emphasis on Osteosarcomas" Biomedicines 9, no. 11: 1642. https://doi.org/10.3390/biomedicines9111642

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