Next Article in Journal
NAGS, CPS1, and SLC25A13 (Citrin) at the Crossroads of Arginine and Pyrimidines Metabolism in Tumor Cells
Previous Article in Journal
A Guide to Perform 3D Histology of Biological Tissues with Fluorescence Microscopy
Previous Article in Special Issue
The Role of microRNAs in Inflammation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Micro RNAs and Circular RNAs in Different Forms of Otitis Media

Department of Pediatric Otolaryngology, Poznan University of Medical Sciences, 60-572 Poznan, Poland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(7), 6752; https://doi.org/10.3390/ijms24076752
Submission received: 22 February 2023 / Revised: 25 March 2023 / Accepted: 30 March 2023 / Published: 4 April 2023
(This article belongs to the Special Issue Non-coding RNA and Inflammation)

Abstract

:
The aim of this comprehensive review was to present the current knowledge on the role of microRNAs (miRNAs) in acute, recurrent, and chronic forms of otitis media. Special attention was focused on cholesteatoma of the middle ear. MicroRNAs modulate gene expression, which, in turn, influences the development and likelihood of the recurrence of acute and aggressive chronic middle ear inflammatory processes. Moreover, this study discusses the modulating role of a specific subgroup of noncoding RNA, circular RNA (circRNA). Recognizing the precise potential pathways and the mechanisms of their function may contribute to a better understanding of the molecular bases of middle ear diseases and identifying novel methods for treating this demanding pathology. Articles published between 2009 and 2022 were used in this analysis. In this review, we provide a complete overview of the latest progress in identifying the role and mechanisms of particular miRNAs and circRNAs in acute, recurrent and chronic forms of otitis media.

1. Introduction

Acute otitis media (AOM) is an inflammatory process of the middle ear that is caused by viral or bacterial infection. It is a common disease, especially amongst children, whose middle ear anatomy is different from adults, making them more susceptible to this condition [1]. For the same reason, in this group of patients the recurrence is more often observed and recurrent otitis media (rAOM) with the persistent presence of ear effusion eventually can lead to chronic otitis media (COM) development [2]. Otitis media with effusion (OME) is characterized by a presence of nonpurulent fluid behind the tympanic membrane without any symptoms of acute infection and often is a consequence of AOM or rAOM. Effusion, if it persists for longer periods, may lead to damage to the middle ear and cause conductive hearing loss which is especially important in pre-school children where it can induce speech development delay [3]. The evolution of COM is mediated by cytokines such as IL-1 and TNF-α that are secreted in response to bacterial infection during an AOM episode and trigger an inflammation response which in effect leads to epithelium remodeling [4]. However, the exact molecular mechanisms of COM formation are still not fully understood. Perhaps, miRNAs that control gene expression at the post-transcriptional stage might be involved in this process.
Acquired middle ear cholesteatoma (AMEC) may be a consequence of OME and due to the recurrence rates its treatment remains a real challenge. AMEC ia a solid mass composed of keratinizing squamous epithelium in the middle ear but with the potential to expand to the surrounding structures. Although considered benign, it is characterized by expanding growth that can destroy the bony structures of the middle ear and eventually lead to many serious conditions such as hearing impairment, vestibular dysfunction, facial nerve palsy, or severe intracranial complications [5]. The pathogenesis of chronic otitis media with cholesteatoma (COMC) is multifactorial and still not fully understood. There are four main theories for explaining cholesteatoma formation, which are based on one of the following: invagination, immigration, squamous metaplasia, or basal cell hyperplasia. The invagination theory includes the accumulation of hyperproliferating epithelial keratinocytes with inhibited apoptosis in a progressive retraction pocket, leading to the formation of cholesteatoma [6]. However, the exact mechanisms underlying cholesteatoma development remain unknown [7]. Surgical resection remains the only method of treatment [8]. This provides a strong incentive for researchers to seek an explanation for cholesteatoma formation at the molecular level, which could lead to a paradigm shift in treatment.
MicroRNAs (miRNAs) are small, noncoding molecules that modulate gene expression at the post-transcriptional stage; however, their exact role in the pathogenesis of human diseases remains unknown. Their influence on the human genome is multifactorial—one miRNA can modulate the expression of many genes, and a single gene can be affected by multiple miRNAs [9]. What is known is that miRNAs regulate around 60% of the protein genes in the human genome, making them the most promising group of interest for molecular research [10].
Many studies have investigated the role of miRNAs in the pathogenesis of various diseases, demonstrating that miRNAs actively partake in many vital processes, such as proliferation, maturation, differentiation, apoptosis, and angiogenesis [11,12]. As almost half of miRNA transcripts are located in regions that are affected during carcinogenesis, it is clear that they may play a pivotal role in the development or transformation to malignancy [13]. MiRNAs whose alteration leads to cancer development are commonly called oncomirs; they were first described in cases of chronic lymphocytic leukemia (CLL) where the expression of two miRNAs (miR-15a and miR-16-1) was upregulated, affecting the expression of antiapoptotic B-cell lymphoma 2 (Bcl2) protein [14]. MicroRNAs can function as oncogenes or tumor suppressors; they can also perform both functions simultaneously, as with miR-26, which is an oncogene in cases of glioma and glioblastoma, inhibiting PTEN (a protein regulating the Akt pathway), in addition to acting as a suppressor in other cancers, such as breast cancer, thyroid cancer, and hepatocellular cancer [15,16].
Recent evidence shows that miRNAs have great diagnostic and therapeutic potential, as they are shown to be more stable than mRNAs and can be easily collected and analyzed. They are present in human plasma, serum, saliva, and urine, which not only makes them excellent non-invasive candidates to be used as biomarkers for diagnosis, prognosis, and disease monitoring, but also shows their potential to become future targets for pharmacological treatment [17,18]. The discovery of microRNA-regulated carcinogenesis has led to the development of anti-miRNAs called antagomirs, which may directly exert a therapeutic effect by blocking oncomirs [19].
Some studies have reported that miRNAs might function as ligands transmitting signals to the cell surface receptors, such as in the case of miR-21, which binds to TLRs (Toll-like receptors) in the immune cells, activating the inflammation response but also—in some cases—tumor growth [20]. Inducing the immune response via microRNA-regulated pathways, as later presented in this paper, plays an important role in the development of middle ear pathologies. The most recent studies have focused on a specific subgroup of noncoding RNA—circular RNA (circRNA). These molecules may play a specific modulating role acting as a competing endogenous RNA [21,22].
The aim of this paper is to present a complete overview of the latest progress on the role of miRNA and circRNA in AOM, rAOM, OME, and AMEC, based on a comprehensive literature review.

2. Literature Review

This review identifies and summarizes the literature concerning the role of miRNA and circRNA in middle ear pathologies that were published between 2009 and 2022 in the PubMed, Scopus, and Web of Science databases. Each database was searched using the following combination of terms: ‘miRNA’ and ‘otitis media’, ‘circRNA’ and ‘otitis media’, ‘miRNA’ and ‘cholesteatoma’, and ‘circRNA’ and ‘cholesteatoma’. The initial search results were identified and a comparison between the search results in different databases was conducted. The same article in different databases was treated as one position. The inclusion criteria regarding the search results were established as English language original research with the full text available. The exclusion criteria included as follows: studies conducted on non-human tissues, non-English language papers, papers that did not contain original research, or those for which only an abstract was available. The inclusion and exclusion criteria were all defined prior to the search. All articles which met the aforementioned criteria were screened for their applicability based on the full text content. No studies were excluded on the basis of poor-quality methods or nonsubstantiated results. The final research material consisting of 27 articles was analysed and presented in this study (Table 1).
Scientific evidence supporting the involvement of several miRNAs and circRNAs in AOM, rAOM, COME, and cholesteatoma formation are summarized below. Their most important potential regulatory mechanisms, pathways, and clinical outcomes are presented in Table 2 (organized by the type of RNA).

2.1. Noncoding RNA in Acute, Recurrent, and Chronic Otitis Media with Effusion

There is a paucity of studies focusing on the presence of miRNAs in the middle ear and the assessment of the impact in the development of acute and chronic inflammation.
A study conducted by Song et al. was the first to describe the presence of miRNA in human middle ear epithelial cells (HMECs) after a 2 h exposition to lipopolysaccharide (LPS) [43]. In that study, they performed a microarray analysis and compared this to the control, and found that 15 miRNAs were dysregulated in the LPS exposed group. As LPS is a component of the outer membrane of Gram-negative bacteria, they hypothesized that the HMECs cells sensitization by LPS may result in differentially expressed miRNAs patterns. The study was the first to demonstrate the difference in miRNAs expression in HMECs after an inflammatory stimulus. In the predicted target genes of the altered miRNAs were the genes responsible for the innate inflammatory response, cell growth, cell differentiation, communication and adhesion or complement activation, so the authors of the study hypothesized that altered miRNAs may take part in triggering the acute inflammation response [43].
Another study presented a link between miRNA expression and Toll-like receptor (TLR) signaling activation in response to an inflammatory stimulus [33]. TLR has a well-documented role in OM development due to bacterial exposition, however, whether miRNAs were part of this signaling pathway was never elucidated [44]. Bacteria, such as Haemophilus influenzae, Streptococcus pneumoniae, and Moraxella catarrhalis are responsible for most cases of acute OM [45]. TLRs are triggered by these pathogens and play an important role in activating an inflammation response, as they evoke changes in the cytokines gene expression, resulting in mucosal hyperplasia and leukotic infiltration in the middle ear [46]. Samuels et al. demonstrated that proinflammatory cytokines, such as tumor necrosis factor-alpha (TNF-α) and interleukin-1 beta (IL-1 β) induced miRNA-146 expression in HMECs [33]. Moreover, increased miRNA-146 expression in OM patients correlated with increased middle ear epithelial thickness when compared to the control, showing miRNA-146 might be involved in stimulating mucosal hyperplasia [33]. Other studies supported this thesis, showing the potential of miRNA-146 in triggering acute inflammation via enhancing the TLR signaling pathway [47].
A study conducted by Val et al. was the first to confirm the presence of miRNAs in the exosomes in the middle ear fluid of patients suffering from otitis media with effusion (OME) [37]. From 800 miRNAs probed in the study, the expression of 29 was significantly altered, 17 of them were unique for middle ear exosomes and miRNA-223-3p expression was the most upregulated. The target genes analysis was then conducted and showed that miRNAs found in the middle ear exosomes influence IL-8 production which, among others, takes part in neutrophil degranulation [38]. This finding supports the hypothesis that the progression from acute to chronic OM might depend on the proinflammatory mediators released after the bacterial stimulation of the middle ear epithelium [48]. Moreover, Val et al. investigated miRNAs expression in HMEECs secretions after exposition to Nontypeable Haemophilus influenzae (NTHi) lysate [37]. They found five miRNAs altered by the NTHi lysate and the target gene analysis showed those miRNAs potential in initiating an innate immune response. NTHi is a common pathogen in AOM etiology that activates cytokines secretion in HMECs through TLR-2 and MAPK (mitogen-activated protein kinase) pathways [39]. Perhaps a cascade of events after NTHi exposition, initiated and regulated by altered miRNA expression, may cause middle ear epithelium remodeling, enabling it to produce mucins, which consequently is responsible for the progression from AOM to the chronic form [4].
In 2021, Zhang et al. made an attempt to identify a role of miR-210 in OME. They revealed significantly lower levels of miR-210 in the serum and middle ear effusions of patients with OME compared to healthy individuals [36]. The expression levels of IL-1β, IL-6, IL-8 and TNF-a were significantly higher both in the serum and in the middle ear effusion in children with OME, whereas the levels of the IL-10 and IL-5 were lower. Moreover, nitric oxide (NO) and vascular endothelial growth factor (VEGF) levels were also higher in the OME group. Western blot confirmed the increased phosphorylation of NF-κB p65 in middle ear effusion samples from patients with OME. In vitro tests proved that overexpression of miR-210 significantly improved cell viability in LPS-treated HMEEC cells. Additionally, the reduced cell apoptosis ratio in LPS-treated HMEEC cells compared with miR-210 NC overexpression was noticed. The authors assumed and then proved that hypoxia-inducible factor-1 alpha (HIF-1α) may be a potential target gene of miR-210 [36]. Acting as an oxygen-dependent transcriptional activator, HIF-1α affects the innate immunity, proinflammatory gene expression, antibacterial activities, and cell migration [49,50].
A nuclear-enriched abundant transcript 1 (NEAT1) is a lncRNA known from various functions in different physiological and pathological processes [40]. Hu et al. decided to explore the potential role of NEAT1 in COME. The authors revealed that miR-495 was a target of NEAT1 and identified the activation of the p38 MAPK signaling pathway. This particular lncRNA could promote inflammation by influencing the cytokine levels, cell proliferation, and apoptosis [40].

2.2. Noncoding RNA and Cholesteatoma

2.2.1. Research Directions

The first attempts to identify the underlying molecular mechanisms of cholesteatoma were directed at the mechanisms dependent on growth factors and inflammatory mediators [51,52,53,54]. It was also agreed that AMEC may develop secondary to infection and chronic inflammation and be maintained by persistent cytokine production [5]. Keratinocytes were found to secrete proinflammatory cytokines such as interleukin (IL)-1b, IL-6, and IL-8; these cytokines stimulate perimatrix fibroblasts to release other cytokines and proinflammatory factors, causing persistent inflammation in the middle ear, which might eventually lead to keratinocyte hyperproliferation and cholesteatoma development [55,56].
Epigenetic regulatory mechanisms have been a dominant theme of molecular studies in recent years. The dysregulation of small noncoding RNAs (miRNAs) has been revealed to be crucial in AMEC formation and development [9,10,11,13,14].
Many researchers have aimed to identify the mRNA targets of miRNAs in order to explain their regulatory mechanisms [28,57]. Nevertheless, the biological phenotype of cholesteatoma cells is difficult to predict, as an individual miRNA is able to control the expression of multiple mRNAs, and a particular mRNA can be regulated by numerous miRNAs [44]. Moreover, the level of mRNA expression is not strictly related to the level of protein expression [58].
The most recent studies have been dominated by a specific subgroup of noncoding RNA, circular RNA (circRNA). CircRNA has a covalescent closed-loop structure and represents a group of noncoding RNAs. Its unique structure is thought to be responsible for its significantly higher stability compared with linear transcripts, preventing circRNAs from undergoing exonucleolytic decomposition [59,60,61]. There are numerous miRNA binding sites—called miRNA response elements (MREs)—on circRNAs, which serve as competitive endogenous RNAs (ceRNAs) [13,14]. All RNA transcripts that share common MREs (circRNAs, long noncoding RNAs, pseudogenes, and protein-coding genes) can function as ceRNAs [29,59,62]. CeRNAs are responsible for ‘sponging’, which is the process of modulating miRNA activity. It is accomplished by the capture of miRNA molecules, thus, limiting their ability to influence mRNA [63]. This attribute of ceRNA is also called a ‘molecular sponge effect’, and it affects the whole network of interactions composed of ceRNAs, miRNAs, and mRNAs and, ultimately, the expression of multiple genes [63].

2.2.2. miRNA

The first study to suggest the possible role of miRNAs in cholesteatoma formation was published in 2009 by Friedland et al. [28]. They found that miRNA-21 expression is upregulated in cholesteatoma tissue compared with normal skin, and that this correlates with decreased levels of PTEN (phosphatase and tensin homolog). PTEN downregulation was previously implicated in carcinogenesis, where it is also negatively regulated by miRNA-21, among other miRNAs [64].
In the abovementioned study, a model was proposed in which the lipopolysaccharide receptor was activated during middle ear infections. Its activation stimulated IL-6 secretion, which in turn resulted in the JAK/STAT/MAPK pathway activation, eventually leading to STAT3 activation and increased miR-21 expression. In effect, miR-21 upregulation results in the inhibited translation of PTEN and PDCD4 (programmed cell death 4), which are important regulators of cell division, in effect causing constant keratinocyte proliferation and cholesteatoma invasion [28].
The results of another study, conducted by Chen at al. [23], are in accordance with the findings of Friedland et al. [28]. They compared the miR-21 expression levels of cholesteatoma tissue between adults and children, observing significantly higher levels of miR-21 and decreased levels of the abovementioned targets—PTEN and PDCD4—in pediatric patients rather than in adults. This reflects the clinical observation that cholesteatoma seems more aggressive and grows faster and more expansively in children [23]. They also found that another miRNA, miR-let-7a, is upregulated in cholesteatoma cells. MiR-let-7a is a known tumor suppressor that has been shown to be deregulated in many human cancers. Its low expression might be correlated with poor treatment outcomes in head and neck squamous cell carcinoma [65]. MiR-let-7a targets HMGA2 (high-mobility group AT-hook 2 protein), an oncogene that controls proliferation, invasion, and cell apoptosis and has recently been described as a prognostic factor of tumor grade in some cancers [66].
A study conducted by Zhang et al. seems to confirm the role of miR-let-7a in the development of cholesteatoma [24]. This study documented that miR-let-7a inhibits keratinocyte proliferation through two separate mechanisms: enhancing apoptosis and causing cell cycle arrest. Moreover, their study demonstrated the relationship between miR-21 and miR-let-7a, showing that miR-let-7a inhibits the expression of miR-21 and, therefore, its target genes. As a result, miR-let-7a upregulation may increase the expression of the miR-21 target genes PTEN and PDCD4 as part of a mechanism for inhibiting keratinocyte proliferation and cholesteatoma growth [24].
This theory is supported by Chen et al., who also hypothesized that there is a relationship between the expression of miR-21 and miR-let-7a and indicated that increased levels of miR-let-7a induce miR-21 downregulation [57].
Yao et al. analyzed the role of microRNA-199a (miR-199a), a cancer-promoting microRNA, as a probable factor influencing the development of cholesteatoma [34]. They revealed that miR-199a is upregulated in cholesteatoma tissues, which is thought to promote keratinocyte proliferation, migration, and invasion in this pathology. PNRC1 was found to be a direct target of miR-199a, with a negative regulatory interrelationship [34].
Another miRNA shown to be involved in cholesteatoma pathogenesis is miR-802. According to a study conducted by Li et al., miR-802 targets more than twenty-two genes in seventeen types of tumors through at least five different pathways [41]. Although downregulated in many neoplastic tumors, miR-802 is upregulated in cholesteatoma. Its overexpression stimulates keratinocyte proliferation and cholesteatoma growth. MiR-802 targets PTEN; its upregulation is correlated with PTEN downregulation, which alleviates the inhibitory effect of PTEN on the PI3K/AKT pathway [41]. Moreover, miR-802 expression is induced by NF-kB (nuclear factor kappa-light-chain-enhancer of activated B cells), a protein complex that controls the cytokine production in response to inflammatory stimuli. This finding supports the proposal that cholesteatoma growth might be induced by cytokines secreted during the inflammation response, and that the NF-kb/miR-802/PTEN signaling pathway might play a role in expansion [67].
Zang et al. revealed that miR-203a influences keratinocyte hyperproliferation [35]. The authors demonstrated that miR-203a is downregulated in cholesteatoma keratinocytes and its target gene—Bmi1—is upregulated. Bmi1 is a complex protein that has previously been described to play a role in carcinogenesis. Bmi1 enhances the expression of p-Akt, which contributes to cell hyperproliferation and invasion and, in effect, plays a role in tumor growth and metastasis [68]. Therefore, another signaling pathway model was proposed by the authors, suggesting that the miR-203a/Bmi1/p-Akt axis might participate in the development of cholesteatoma and become a future target in pharmacological treatment [68].
A study by Sui et al. confirmed the negative correlation between miR-142-5p and cyclin-dependent kinase 5 (CDK5) in terms of expression and regulation [32]. CDK5 was previously reported to contribute to increasing inflammatory responses [69,70]. Sui et al. observed CDK5 overexpression in cholesteatoma tissue samples, which was found to be positively correlated with levels of the cytokines TNF-α, TGF-β1, IL-5, IL-6, and IL-17A [32]. It is commonly known that TNF-α, IL-6, and IL-17A affect the development and proliferation of epidermal cells [71]. Therefore, it was suspected that the formation of a specific inflammatory microenvironment in cholesteatoma may result from the upregulation of IL-5 in AMEC [32]. Moreover, Sui et al. hypothesized that the CDK5/MAPK/NF-κB pathway is used by miR-142-5p to regulate the secretion of inflammatory cytokines [32].
Previous studies have highlighted the role of new blood vessel formation in the development of tissue masses such as cholesteatoma. Increased angiogenesis is essential in tumor growth and expansion, as it enables nutrient delivery and oxygen exchange [72]. Li et al. investigated the presence of miRNAs in exosomes derived from cholesteatoma perimatrix fibroblasts (hCPFs-Exo) and found that miR-106b-5p is downregulated [31]. As miR-106b-5p inhibits the expression of angiopoietin 2 (Angpt2), a growth factor that regulates new vessel formation, they hypothesized that its downregulation promotes angiogenesis by disinhibiting Angpt2 [31].
The results of a microarray analysis of miRNAs expressed in cholesteatoma demonstrated that the expression of 219 miRNAs is altered (22 upregulated and 175 downregulated) when compared with normal skin. The results for some miRNAs were similar to those of previous studies, such as in regard to the upregulation of miR-21-3p, whereas they were divergent for others. Surprisingly, in that study, the expression of members of the miR-let-7a family did not differ in cholesteatoma compared with normal skin. The GO and KEGG pathway target gene analyses were performed on five selected miRNAs and showed that these miRNAs take part in proliferation, differentiation, the cell cycle, and apoptosis [26].
Recent evidence shows that miRNAs are also involved in osteoclast formation in cholesteatoma tissue. Gong et al. established that miR-17 is secreted by keratinocytes and—through exosomes—can interact with fibroblasts, favoring their differentiation into osteoclasts by upregulating the RANKL levels [27]. RANKL, a protein that is a member of the TNF (tumor necrosis factor) family, has been reported to participate in osteoclast activation [73]. The authors hypothesized that, upon osteoclast hyperactivation, the balance between bone formation and resorption is lost, which may support further cholesteatoma development.
Taken together, the results of recent studies imply that miRNAs might be crucial regulators of post-translational processes, leading to a series of events that, in effect, result in cholesteatoma progression. There is evidence that they participate in almost every stage of cholesteatoma formation, as they are engaged in keratinocyte hyperproliferation, angiogenesis, and fibroblast-to-osteoclast transformation.

2.2.3. circRNA

In 2018, a study on cholesteatoma formation conducted by Gao et al. revealed that long noncoding RNAs can serve as ceRNAs [29]. However, the real breakthrough was their study published in 2020, using microarray analysis and confirmation by qRT-PCR [25]. The authors determined that the profiles of circRNAs in AMEC were divergent from those derived from normal skin samples. Using a bioinformatics approach, the authors constructed the network composed of circRNA–lncRNA–miRNA–mRNA and investigated the ‘sponging’ potential of circRNAs [25]. Two circRNAs sharing common MREs were selected (hsa-circRNA-101458 and hsa-circRNA-102747). CircRNA-102747 was found to interact with miR-21-3p [25]. The role of the miR-21 family in the formation and spread of cholesteatoma has been previously demonstrated [28,57]. The link between circRNA-101458 and miR-let-7a-3p has also been confirmed [25]. The upregulation of miR-let-7 from the same family was thought to promote the benign nature of AMEC by supporting antiproliferation [23,74]. The authors hypothesized the opposite regulatory influence on mRNAs of different circRNA–lncRNA–miRNA networks [25]. Thus, circRNA-101458/miR-let-7a-3p and circRNA-102747/lncRNA-uc001kfc.1/miR-21-3p may be, respectively, regarded as benign and malignant regulators. This suggests they may be considered as potential targets for the conventional treatment of AMEC [25].
Knowing that miR-125a-5p [75,76,77] and miR-22-3p [78,79,80] exhibit antitumor activity in some types of malignant neoplasms, Hu et al. identified that circ_0074491 plays the role of decoy for miR-125a-5p and miR-22-3p [42]. They suspected its regulative effect on AMEC progression involves miR-125a-5p and miR-22-3p. The results confirmed that the PI3K/Akt pathway is involved in the regulatory influence of circ_0074491 on the adsorption of miR-125a-5p and miR-22-3p in cholesteatoma keratinocytes [42]. Circ_0074491 upregulation was found to be negatively correlated with the malignant behaviors of AMEC keratinocytes [25]. Xie et al. demonstrated that the expression of hsa_circRNA_000319 was significantly decreased whereas hsa_circRNA_104327 and hsa_circRNA_404655 were expressed significantly higher in cholesteatoma samples [81]. The authors speculated that hsa_circRNA_404655 may play a role of the sponge for miRNA-3664-3p, influencing the proliferation and local invasion of AMEC [81]. The lower expression of miRNA-3664-3p may lead to an increased expression of p53 via growth differentiation factor 15 (GDF15) [82].
Another study presenting the circRNA–miRNA–target mRNA axis was published by Liu et al. in 2021 [6]. The authors selected miR-508-3p and hsa_circ_0000007 and identified significantly higher expression of miR-508-3p in cholesteatoma samples than in HaCaT cells or retroauricular skin. In addition, has_circ_0000007 negatively influenced the expression of miR-508-3p. Their results revealed that miR-508-3p directly targets PTEN and influences the phenotype of AMEC. The authors concluded that, when present in cholesteatoma cells miR-508-3p is able to inhibit apoptosis and promote proliferation via the PTEN/PI3K/Akt pathway [6].

2.2.4. Practical Implications

An experimental study using nanotechnology conducted by Zheng et al. represents an important step forward in the practical use of miRNAs in treatment [30]. Based on AMEC tissue, the authors prepared and delivered rubine, a regulator of miR-34a, into cells. Their experiments revealed that miR-34a may target EGFR and, via this mechanism, suppress the growth and development of tumors. Overexpression of miR-34a can promote apoptosis and prevent cell proliferation and migration. The study thus confirmed that the expression of miR-34 may be enhanced by nanoparticles and thereby influence CyclinD1, Bcl-2, and Cdk6 [30].

3. Conclusions

In this comprehensive review, we summarized the contemporary molecular biology achievements analyzing the modulating role of noncoding RNAs and their potential mechanisms in the pathogenesis of different forms of otits media and especially chronic otits media with cholesteatoma. The recent progress that has been made in biomolecular research indicates the direction for future studies, wherein the circRNA–lncRNA–miRNA–mRNA axis is likely to be key. Both in vitro and in vivo studies on the regulatory influence of ceRNA on circRNAs are necessary for progress in elucidating the pathogenesis of AOM, rAOM, COME and AMEC. This may ultimately result in a paradigm shift in cholesteatoma treatment modalities, evolving into pharmacotherapy based on the modulation of transcriptional control.

Author Contributions

Conceptualization, M.K. and P.A.; methodology, M.K. and P.A.; investigation, M.K. and P.A.; resources, M.K. and P.A.; data curation, M.K. and P.A.; writing—original draft preparation, M.K. and P.A.; writing—review and editing, M.K.; visualization, M.K.; supervision, J.S. 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.

Data Availability Statement

Data available from the corresponding author on request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Toll, E.C.; Nunez, D.A. Diagnosis and treatment of acute otitis media: Review. J. Laryngol. Otol. 2012, 126, 976–983. [Google Scholar] [CrossRef]
  2. Granath, A. Recurrent Acute Otitis Media: What Are the Options for Treatment and Prevention? Curr. Otorhinolaryngol. Rep. 2017, 5, 93–100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Simon, F.; Haggard, M.; Rosenfeld, R.; Jia, H.; Peer, S.; Calmels, M.-N.; Couloigner, V.; Teissier, N. International consensus (ICON) on management of otitis media with effusion in children. Eur. Ann. Otorhinolaryngol. Head Neck Dis. 2018, 135, S33–S39. [Google Scholar] [CrossRef] [PubMed]
  4. Juhn, S.K.; Jung, M.-K.; Hoffman, M.D.; Drew, B.R.; Preciado, D.A.; Sausen, N.J.; Jung, T.T.; Kim, B.H.; Park, S.-Y.; Lin, J.; et al. The Role of Inflammatory Mediators in the Pathogenesis of Otitis Media and Sequelae. Clin. Exp. Otorhinolaryngol. 2008, 1, 117–138. [Google Scholar] [CrossRef] [Green Version]
  5. Semaan, M.T.; Megerian, C.A. The Pathophysiology of Cholesteatoma. Otolaryngol. Clin. N. Am. 2006, 39, 1143–1159. [Google Scholar] [CrossRef]
  6. Liu, D.; Ma, X. MiR-508-3p promotes proliferation and inhibits apoptosis of middle ear cholesteatoma cells by targeting PTEN/PI3K/AKT pathway. Int. J. Med. Sci. 2021, 18, 3224–3235. [Google Scholar] [CrossRef]
  7. Kuo, C.-L. Etiopathogenesis of acquired cholesteatoma: Prominent theories and recent advances in biomolecular research. Laryngoscope 2014, 125, 234–240. [Google Scholar] [CrossRef]
  8. Kuo, C.-L.; Liao, W.-H.; Shiao, A.-S. A review of current progress in acquired cholesteatoma management. Eur. Arch. Oto-Rhino-Laryngol. 2014, 272, 3601–3609. [Google Scholar] [CrossRef] [PubMed]
  9. Kozomara, A.; Birgaoanu, M.; Griffiths-Jones, S. miRBase: From microRNA sequences to function. Nucleic Acids Res. 2019, 47, D155–D162. [Google Scholar] [CrossRef] [PubMed]
  10. Friedman, R.C.; Farh, K.K.-H.; Burge, C.B.; Bartel, D.P. Most mammalian mRNAs are conserved targets of microRNAs. Genome Res. 2009, 19, 92–105. [Google Scholar] [CrossRef] [Green Version]
  11. Bouchie, A. First microRNA mimic enters clinic. Nat. Biotechnol. 2013, 31, 577. [Google Scholar] [CrossRef]
  12. Zhu, Z.-J.; Huang, P.; Chong, Y.-X.; Kang, L.-X.; Huang, X.; Nie, L. MicroRNA-181a promotes proliferation and inhibits apoptosis by suppressing CFIm25 in osteosarcoma. Mol. Med. Rep. 2016, 14, 4271–4278. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Calin, G.A.; Dumitru, C.D.; Shimizu, M.; Bichi, R.; Zupo, S.; Noch, E.; Aldler, H.; Rattan, S.; Keating, M.; Rai, K.; et al. Frequent deletions and down-regulation of micro- RNA genes miR15 and miR16 at 13q14 in chronic lymphocytic leukemia. Proc. Natl. Acad. Sci. USA 2002, 99, 15524–15529. [Google Scholar] [CrossRef] [Green Version]
  14. Calin, G.A.; Sevignani, C.; Dumitru, C.D.; Hyslop, T.; Noch, E.; Yendamuri, S.; Shimizu, M.; Rattan, S.; Bullrich, F.; Negrini, M.; et al. Human microRNA genes are frequently located at fragile sites and genomic regions involved in cancers. Proc. Natl. Acad. Sci. USA 2004, 101, 2999–3004. [Google Scholar] [CrossRef] [Green Version]
  15. Huse, J.T.; Brennan, C.; Hambardzumyan, D.; Wee, B.; Pena, J.; Rouhanifard, S.H.; Sohn-Lee, C.; le Sage, C.; Agami, R.; Tuschl, T.; et al. The PTEN-regulating microRNA miR-26a is amplified in high-grade glioma and facilitates gliomagenesis in vivo. Genes Dev. 2009, 23, 1327–1337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Gao, J.; Liu, Q.-G. The role of miR-26 in tumors and normal tissues (Review). Oncol. Lett. 2011, 2, 1019–1023. [Google Scholar] [CrossRef] [Green Version]
  17. Lu, J.; Getz, G.; Miska, E.A.; Alvarez-Saavedra, E.; Lamb, J.; Peck, D.; Sweet-Cordero, A.; Ebert, B.L.; Mak, R.H.; Ferrando, A.A.; et al. MicroRNA expression profiles classify human cancers. Nature 2005, 435, 834–838. [Google Scholar] [CrossRef]
  18. Ghai, V.; Wang, K. Recent progress toward the use of circulating microRNAs as clinical biomarkers. Arch. Toxicol. 2016, 90, 2959–2978. [Google Scholar] [CrossRef] [PubMed]
  19. Rossbach, M. Small Non-Coding RNAs as Novel Therapeutics. Curr. Mol. Med. 2010, 10, 361–368. [Google Scholar] [CrossRef]
  20. Fabbri, M.; Paone, A.; Calore, F.; Galli, R.; Gaudio, E.; Santhanam, R.; Lovat, F.; Fadda, P.; Mao, C.; Nuovo, G.J.; et al. MicroRNAs bind to Toll-like receptors to induce prometastatic inflammatory response. Proc. Natl. Acad. Sci. USA 2012, 109, E2110–E2116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Lei, K.; Bai, H.; Wei, Z.; Xie, C.; Wang, J.; Li, J.; Chen, Q. The mechanism and function of circular RNAs in human diseases. Exp. Cell Res. 2018, 368, 147–158. [Google Scholar] [CrossRef]
  22. Kristensen, L.S.; Andersen, M.S.; Stagsted, L.V.W.; Ebbesen, K.K.; Hansen, T.B.; Kjems, J. The biogenesis, biology and characterization of circular RNAs. Nat. Rev. Genet. 2019, 20, 675–691. [Google Scholar] [CrossRef] [PubMed]
  23. Chen, X.; Qin, Z. Post-Transcriptional Regulation by MicroRNA-21 and let-7a MicroRNA in Paediatric Cholesteatoma. J. Int. Med. Res. 2011, 39, 2110–2118. [Google Scholar] [CrossRef]
  24. Zhang, W.; Qin, Z.; Chen, X. MicroRNA let-7a suppresses the growth and invasion of cholesteatoma keratinocytes. Mol. Med. Rep. 2014, 11, 2097–2103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Gao, J.; Tang, Q.; Xue, R.; Zhu, X.; Wang, S.; Zhang, Y.; Liu, W.; Gao, Z.; Yang, H. Comprehensive circular RNA expression profiling with associated ceRNA network reveals their therapeutic potential in cholesteatoma. Oncol. Rep. 2020, 43, 1234–1244. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Xie, S.; Liu, X.; Pan, Z.; Chen, X.; Peng, A.; Yin, T.; Ren, J.; Liu, W. Microarray Analysis of Differentially-expressed MicroRNAs in Acquired Middle Ear Cholesteatoma. Int. J. Med. Sci. 2018, 15, 1547–1554. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Gong, N.; Zhu, W.; Xu, R.; Teng, Z.; Deng, C.; Zhou, H.; Xia, M.; Zhao, M. Keratinocytes-derived exosomal miRNA regulates osteoclast differentiation in middle ear cholesteatoma. Biochem. Biophys. Res. Commun. 2020, 525, 341–347. [Google Scholar] [CrossRef]
  28. Friedland, D.R.; Eernisse, R.; Erbe, C.; Gupta, N.; Cioffi, J.A. Cholesteatoma Growth and Proliferation. Otol. Neurotol. 2009, 30, 998–1005. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Gao, J.; Tang, Q.; Zhu, X.; Wang, S.; Zhang, Y.; Liu, W.; Gao, Z.; Yang, H. Long noncoding RNAs show differential expression profiles and display ceRNA potential in cholesteatoma pathogenesis. Oncol. Rep. 2018, 39, 2091–2100. [Google Scholar] [CrossRef] [Green Version]
  30. Zheng, H.; Wang, W.; Li, S.; Han, L. The Effect of Zbxz23ir-21 NANO(nanomaterials) Delivery Vector on Apoptosis and PTEN(phosphatase and tensin homolog deleted on chromosome ten)/PI3K(Intracellular phosphatidylinositol kinase)/AKT(related to the A and C kinase) in Children with CHOLESTEATOMA in Middle Ear. Bioengineered 2021, 12, 8809–8821. [Google Scholar] [CrossRef]
  31. Li, Y.; Liang, J.; Hu, J.; Ren, X.; Sheng, Y. Down-regulation of exosomal miR-106b-5p derived from cholesteatoma perimatrix fibroblasts promotes angiogenesis in endothelial cells by overexpression of Angiopoietin 2. Cell Biol. Int. 2018, 42, 1300–1310. [Google Scholar] [CrossRef] [PubMed]
  32. Sui, R.; Shi, W.; Han, S.; Fan, X.; Zhang, X.; Wang, N.; Zhang, H.; Xu, A.; Liu, C. MiR-142-5p directly targets cyclin-dependent kinase 5-mediated upregulation of the inflammatory process in acquired middle ear cholesteatoma. Mol. Immunol. 2022, 141, 236–245. [Google Scholar] [CrossRef] [PubMed]
  33. Samuels, T.L.; Yan, J.; Khampang, P.; MacKinnon, A.; Hong, W.; Johnston, N.; Kerschner, J.E. Association of microRNA 146 with middle ear hyperplasia in pediatric otitis media. Int. J. Pediatr. Otorhinolaryngol. 2016, 88, 104–108. [Google Scholar] [CrossRef] [Green Version]
  34. Yao, L.; Zhang, W.; Zheng, J.; Lu, X.; Zhang, F. miR-199a Targeting PNRC1 to Promote Keratinocyte Proliferation and Invasion in Cholesteatoma. BioMed Res. Int. 2021, 2021, 1–10. [Google Scholar] [CrossRef]
  35. Zang, J.; Hui, L.; Yang, N.; Yang, B.; Jiang, X. Downregulation of MiR-203a Disinhibits Bmi1 and Promotes Growth and Proliferation of Keratinocytes in Cholesteatoma. Int. J. Med. Sci. 2018, 15, 447–455. [Google Scholar] [CrossRef] [Green Version]
  36. Zhang, J.; He, J.; Luo, Y.; Liu, Y.; Fan, X. miR-210 regulates the inflammation of otitis media with effusion by inhibiting the expression of hypoxia-inducible factor (HIF)-1a. Biochem. Biophys. Res. Commun. 2020, 534, 401–407. [Google Scholar] [CrossRef]
  37. Val, S.; Krueger, A.; Poley, M.; Cohen, A.; Brown, K.; Panigrahi, A.; Preciado, D. Nontypeable Haemophilus influenzae lysates increase heterogeneous nuclear ribonucleoprotein secretion and exosome release in human middle-ear epithelial cells. FASEB J. 2018, 32, 1855–1867. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Johnson, M.; Leonard, G.; Kreutzer, D.L. Murine Model of Interleukin-8-Induced Otitis Media. Laryngoscope 1997, 107, 1405–1408. [Google Scholar] [CrossRef] [PubMed]
  39. Lee, H.-Y.; Takeshita, T.; Shimada, J.; Akopyan, A.; Woo, J.-I.; Pan, H.; Moon, S.K.; Andalibi, A.; Park, R.-K.; Kang, S.-H.; et al. Induction of beta defensin 2 by NTHi requires TLR2 mediated MyD88 and IRAK-TRAF6-p38MAPK signaling pathway in human middle ear epithelial cells. BMC Infect. Dis. 2008, 8, 87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Hu, Y.; Dong, H.; Huang, J.; Huang, J.; Tao, D.; Huang, C.; Hu, L.; Xu, H.; Sun, Y. Long non-coding RNA (lncRNA) nuclear enriched abundant transcript 1 (NEAT1) promotes the inflammation and apoptosis of otitis media with effusion through targeting microRNA (miR)-495 and activation of p38 MAPK signaling pathway. Bioengineered 2021, 12, 8080–8088. [Google Scholar] [CrossRef] [PubMed]
  41. Li, N.; Qin, Z.-B. Inflammation-induced miR-802 promotes cell proliferation in cholesteatoma. Biotechnol. Lett. 2014, 36, 1753–1759. [Google Scholar] [CrossRef]
  42. Hu, Y.M.; Qian, X. Hsa_circ_0074491 regulates the malignance of cholesteatoma keratinocytes by modulating the PI3K/Akt pathway by binding to miR-22-3p and miR-125a-5p. Medicine 2021, 100, e27122. [Google Scholar] [CrossRef]
  43. Song, J.-J.; Kwon, S.K.; Cho, C.G.; Park, S.-W.; Chae, S.-W. Microarray analysis of microRNA expression in LPS induced inflammation of human middle ear epithelial cells (HMEECs). Int. J. Pediatr. Otorhinolaryngol. 2011, 75, 648–651. [Google Scholar] [CrossRef] [PubMed]
  44. Emonts, M.; Veenhoven, R.H.; Wiertsema, S.P.; Houwing-Duistermaat, J.J.; Walraven, V.; de Groot, R.; Hermans, P.W.; Sanders, E.A. Genetic Polymorphisms in Immunoresponse Genes TNFA, IL6, IL10, and TLR4 Are Associated With Recurrent Acute Otitis Media. Pediatrics 2007, 120, 814–823. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Leung, A.K. Acute Otitis Media in Children. Recent Pat. Inflamm. Allergy Drug Discov. 2017, 11, 32–40. [Google Scholar] [CrossRef]
  46. Jotic, A.; Jesic, S.; Zivkovic, M.; Tomanovic, N.; Kuveljic, J.; Stankovic, A. Polymorphisms in Toll-like receptors 2 and 4 genes and their expression in chronic suppurative otitis media. Auris Nasus Larynx 2015, 42, 431–437. [Google Scholar] [CrossRef] [PubMed]
  47. Quinn, S.R.; O’Neill, L.A. A trio of microRNAs that control Toll-like receptor signalling. Int. Immunol. 2011, 23, 421–425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Lin, J.; Tsuprun, V.; Kawano, H.; Paparella, M.M.; Zhang, Z.; Anway, R.; Ho, S.B. Characterization of mucins in human middle ear and Eustachian tube. Am. J. Physiol. Cell. Mol. Physiol. 2001, 280, L1157–L1167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Bui, B.P.; Nguyen, P.L.; Lee, K.; Cho, J. Hypoxia-Inducible Factor-1: A Novel Therapeutic Target for the Management of Cancer, Drug Resistance, and Cancer-Related Pain. Cancers 2022, 14, 6054. [Google Scholar] [CrossRef]
  50. Infantino, V.; Santarsiero, A.; Convertini, P.; Todisco, S.; Iacobazzi, V. Cancer Cell Metabolism in Hypoxia: Role of HIF-1 as Key Regulator and Therapeutic Target. Int. J. Mol. Sci. 2021, 22, 5703. [Google Scholar] [CrossRef]
  51. Kupper, T.S. The Activated Keratinocyte: A Model for Inducible Cytokine Production by Non-Bone-Marrow-Derived Cells in Cutaneous Inflammatory and Immune Responses. J. Investig. Dermatol. 1990, 94, s146–s150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Kojima, H.; Shiwa, M.; Kamide, Y.; Moriyama, H. Expression and Localization of mRNA for Epidermal Growth Factor and Epidermal Growth Factor Receptor in Human Cholesteatoma. Acta Oto-Laryngol. 1994, 114, 423–429. [Google Scholar] [CrossRef]
  53. Ergün, S.; Zheng, X.; Carlsöö, B. Expression of transforming growth factor-alpha and epidermal growth factor receptor in middle ear cholesteatoma. Am. J. Otol. 1996, 17, 393–396. [Google Scholar]
  54. Olszewska, E.; Wagner, M.; Bernal-Sprekelsen, M.; Dazert, S.; Hildmann, H.; Sudhoff, H. Etiopathogenesis of cholesteatoma. Eur. Arch. Oto-Rhino-Laryngol. 2003, 261, 6–24. [Google Scholar] [CrossRef] [PubMed]
  55. Cheshire, I.M.; Blight, A.; Ratcliffe, W.A.; Proops, D.W.; A Heath, D. Production of parathyroid-hormone-related protein by cholesteatoma cells in culture. Lancet 1991, 338, 1041–1043. [Google Scholar] [CrossRef]
  56. Yetiser, S.; Satar, B.; Aydin, N. Expression of Epidermal Growth Factor, Tumor Necrosis Factor-α, and Interleukin-1α in Chronic Otitis Media with or without Cholesteatoma. Otol. Neurotol. 2002, 23, 647–652. [Google Scholar] [CrossRef] [PubMed]
  57. Chen, X.; Li, X.; Qin, Z. MicroRNA-21 promotes the proliferation and invasion of cholesteatoma keratinocytes. Acta Oto-Laryngol. 2016, 136, 1261–1266. [Google Scholar] [CrossRef] [PubMed]
  58. Chen, G.; Gharib, T.G.; Huang, C.-C.; Taylor, J.M.G.; Misek, D.E.; Kardia, S.L.R.; Giordano, T.J.; Iannettoni, M.D.; Orringer, M.B.; Hanash, S.M.; et al. Discordant Protein and mRNA Expression in Lung Adenocarcinomas. Mol. Cell. Proteom. 2002, 1, 304–313. [Google Scholar] [CrossRef] [Green Version]
  59. 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] [PubMed] [Green Version]
  60. Starke, S.; Jost, I.; Rossbach, O.; Schneider, T.; Schreiner, S.; Hung, L.-H.; Bindereif, A. Exon Circularization Requires Canonical Splice Signals. Cell Rep. 2015, 10, 103–111. [Google Scholar] [CrossRef] [Green Version]
  61. Guo, J.U.; Agarwal, V.; Guo, H.; Bartel, D.P. Expanded identification and characterization of mammalian circular RNAs. Genome Biol. 2014, 15, 1–14. [Google Scholar] [CrossRef] [PubMed]
  62. Salmena, L.; Poliseno, L.; Tay, Y.; Kats, L.; Pandolfi, P.P. A ceRNA Hypothesis: The Rosetta Stone of a Hidden RNA Language? Cell 2011, 146, 353–358. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Mitra, A.; Pfeifer, K.; Park, K.-S. Circular RNAs and competing endogenous RNA (ceRNA) networks. Transl. Cancer Res. 2018, 7, S624–S628. [Google Scholar] [CrossRef]
  64. Song, M.S.; Salmena, L.; Pandolfi, P.P. The functions and regulation of the PTEN tumour suppressor. Nat. Rev. Mol. Cell Biol. 2012, 13, 283–296. [Google Scholar] [CrossRef]
  65. Fedele, M.; Battista, S.; Kenyon, L.; Baldassarre, G.; Fidanza, V.; Klein-Szanto, A.J.P.; Parlow, A.F.; Visone, R.; Pierantoni, G.M.; Outwater, E.; et al. Overexpression of the HMGA2 gene in transgenic mice leads to the onset of pituitary adenomas. Oncogene 2002, 21, 3190–3198. [Google Scholar] [CrossRef] [Green Version]
  66. Motoyama, K.; Inoue, H.; Nakamura, Y.; Uetake, H.; Sugihara, K.; Mori, M. Clinical Significance of High Mobility Group A2 in Human Gastric Cancer and Its Relationship to let-7 MicroRNA Family. Clin. Cancer Res. 2008, 14, 2334–2340. [Google Scholar] [CrossRef] [Green Version]
  67. Si, Y.; Bin Chen, Y.; Chen, S.J.; Zheng, Y.Q.; Liu, X.; Liu, Y.; Jiang, H.L.; Xu, G.; Li, Z.H.; Huang, Q.H.; et al. TLR4 drives the pathogenesis of acquired cholesteatoma by promoting local inflammation and bone destruction. Sci. Rep. 2015, 5, 16683. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Wang, M.-C.; Jiao, M.; Wu, T.; Jing, L.; Cui, J.; Guo, H.; Tian, T.; Ruan, Z.-P.; Wei, Y.-C.; Jiang, L.-L.; et al. Polycomb complex protein BMI-1 promotes invasion and metastasis of pancreatic cancer stem cells by activating PI3K/AKT signaling, an ex vivo, in vitro, and in vivo study. Oncotarget 2016, 7, 9586–9599. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Sun, H.; Cai, J.; Shen, S.; Ren, X. Hypothermia treatment ameliorated cyclin-dependent kinase 5-mediated inflammation in ischemic stroke and improved outcomes in ischemic stroke patients. Clinics 2019, 74, e938. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Pfänder, P.; Fidan, M.; Burret, U.; Lipinski, L.; Vettorazzi, S. Cdk5 Deletion Enhances the Anti-inflammatory Potential of GC-Mediated GR Activation During Inflammation. Front. Immunol. 2019, 10, 1554. [Google Scholar] [CrossRef] [Green Version]
  71. Yu, J.; Xiao, Z.; Zhao, R.; Lu, C.; Zhang, Y. Paeoniflorin suppressed IL-22 via p38 MAPK pathway and exerts anti-psoriatic effect. Life Sci. 2017, 180, 17–22. [Google Scholar] [CrossRef] [PubMed]
  72. Carmeliet, P.; Jain, R.K. Molecular mechanisms and clinical applications of angiogenesis. Nature 2011, 473, 298–307. [Google Scholar] [CrossRef] [Green Version]
  73. Yoon, W.-J.; Kim, K.-N.; Heo, S.-J.; Han, S.-C.; Kim, J.; Ko, Y.-J.; Kang, H.-K.; Yoo, E.-S. Sargachromanol G inhibits osteoclastogenesis by suppressing the activation NF-κB and MAPKs in RANKL-induced RAW 264.7 cells. Biochem. Biophys. Res. Commun. 2013, 434, 892–897. [Google Scholar] [CrossRef] [PubMed]
  74. Johnson, C.D.; Esquela-Kerscher, A.; Stefani, G.; Byrom, M.; Kelnar, K.; Ovcharenko, D.; Wilson, M.; Wang, X.; Shelton, J.; Shingara, J.; et al. The let-7 MicroRNA Represses Cell Proliferation Pathways in Human Cells. Cancer Res. 2007, 67, 7713–7722. [Google Scholar] [CrossRef] [Green Version]
  75. Vo, D.T.; Karanam, N.K.; Ding, L.; Saha, D.; Yordy, J.S.; Giri, U.; Heymach, J.V.; Story, M.D. miR-125a-5p Functions as Tumor Suppressor microRNA And Is a Marker of Locoregional Recurrence And Poor prognosis in Head And Neck Cancer. Neoplasia 2019, 21, 849–862. [Google Scholar] [CrossRef]
  76. Chen, L.-Y.; Wang, L.; Ren, Y.-X.; Pang, Z.; Liu, Y.; Sun, X.-D.; Tu, J.; Zhi, Z.; Qin, Y.; Sun, L.-N.; et al. The circular RNA circ-ERBIN promotes growth and metastasis of colorectal cancer by miR-125a-5p and miR-138-5p/4EBP-1 mediated cap-independent HIF-1α translation. Mol. Cancer 2020, 19, 1–17. [Google Scholar] [CrossRef]
  77. Cao, Q.; Wang, N.; Ren, L.; Tian, J.; Yang, S.; Cheng, H. miR-125a-5p post-transcriptionally suppresses GALNT7 to inhibit proliferation and invasion in cervical cancer cells via the EGFR/PI3K/AKT pathway. Cancer Cell Int. 2020, 20, 117. [Google Scholar] [CrossRef] [PubMed]
  78. Hussein, N.A.E.M.; El Kholy, Z.A.; Anwar, M.M.; Ahmad, M.A.; Ahmad, S.M. Plasma miR-22-3p, miR-642b-3p and miR-885-5p as diagnostic biomarkers for pancreatic cancer. J. Cancer Res. Clin. Oncol. 2016, 143, 83–93. [Google Scholar] [CrossRef] [PubMed]
  79. Dong, H.; Wang, R.; Jin, X.; Zeng, J.; Pan, J. LncRNA DGCR5 promotes lung adenocarcinoma (LUAD) progression via inhibiting hsa-mir-22-3p. J. Cell. Physiol. 2017, 233, 4126–4136. [Google Scholar] [CrossRef]
  80. Gan, L.; Lv, L.; Liao, S. Long non-coding RNA H19 regulates cell growth and metastasis via the miR-22-3p/Snail1 axis in gastric cancer. Int. J. Oncol. 2019, 54, 2157–2168. [Google Scholar] [CrossRef] [PubMed]
  81. Xie, S.; Jin, L.; Yin, T.; Ren, J.; Liu, W. Microarray analysis and functional prediction of differentially expressed circular RNAs in acquired middle ear cholesteatoma. Biomed. Eng. Online 2021, 20, 1–13. [Google Scholar] [CrossRef] [PubMed]
  82. An, T.; Fan, T.; Zhang, X.Q.; Liu, Y.-F.; Huang, J.; Liang, C.; Lv, B.-H.; Wang, Y.-Q.; Zhao, X.-G.; Liu, J.-X.; et al. Comparison of Alterations in miRNA Expression in Matched Tissue and Blood Samples during Spinal Cord Glioma Progression. Sci. Rep. 2019, 9, 9169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Table 1. Identification of publications meeting the inclusion criteria based on the results of searching the PubMed, Web of Science and Scopus databases.
Table 1. Identification of publications meeting the inclusion criteria based on the results of searching the PubMed, Web of Science and Scopus databases.
DatabasePubMedScopusWeb of Science
Searched Terms
‘otits media’ and ‘miRNA’444
‘otits media’ and ‘circRNA’000
‘cholesteatoma’ and ‘miRNA’161311
‘cholesteatoma’ and ‘circRNA’345
Total articles identified *211918
Total articles included *27
* duplications removed.
Table 2. Selected noncoding RNAs and their potential role in AOM, rAOM, COME and cholesteatoma.
Table 2. Selected noncoding RNAs and their potential role in AOM, rAOM, COME and cholesteatoma.
Noncoding RNARegulatory Mechanism or Signaling PathwayInfluenceReferences
miR-let-7aHMAG2proliferation of kerationocytes[23,24,25]
miR-10a-5pPI3K/Akt hyperproliferation of cholesteatoma[26]
miR-16-1-3pPI3K/Akt hyperproliferation of cholesteatoma[26]
miR-17RANKLdifferentiation of fibroblasts into osteoclasts, osteoclasts activation[27]
miR-21PTEN, PDCD4proliferation of kerationocytes[23,25,28,29]
miR-34aPTEN/PI3K/Akt/EGFRproliferation, apoptosis[30]
miR-106b-5pAngpt2angiogenesis[31]
miR-142-5pCDK5/MAPK/NF-κB positive correlation with: TNF-α, TGF-β1, IL-5, IL-6, and IL-17A[32]
miR-146(TNF-α) (IL-1 β)/TLRmucosal hyperplasia[33]
miR-199aPNRC1keratinocyte proliferation, migration, and invasion[34]
miR-203aBmi1/p-Aktproliferation, migration and antiapoptotic abilities[35]
miR-210NF-κB/HIF-1αcell viability, apoptosis[36]
miRNA-223-3pIL-8neutrophil degranulation[37,38]
miR-378a-3p + miR-378i,
miR-200a-3p,
miR-378g,
miR30d-5p,
miR-222-3p miRNAs
TLR-2 and MAPKepithelium remodeling[37,39]
NEAT1/miR-495p38 MAPKcell proliferation,
levels of inflammatory cytokines,
cell apoptosis
[40]
miR-802PTEN/PI3K/Akt proliferation of kerationocytes [41]
circ-0000007/
miR-508-3p
PTEN/PI3K/Aktproliferation apoptosis[20]
circ-0074491/
miR-22-3p
PI3K/Akt proliferation, colony formation, apoptosis of keratinocytes[42]
circ-0074491/
miR-125a-5p
PI3K/Akt proliferation, colony formation, apoptosis of keratinocytes[42]
circRNA-101458/
miR-let-7a
PI3K/Aktproliferation of kerationocytes[25]
circ-102747/
miR-21-3p
PI3K/Aktformation and invasion of cholesteatoma[25]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kotowski, M.; Adamczyk, P.; Szydlowski, J. Micro RNAs and Circular RNAs in Different Forms of Otitis Media. Int. J. Mol. Sci. 2023, 24, 6752. https://doi.org/10.3390/ijms24076752

AMA Style

Kotowski M, Adamczyk P, Szydlowski J. Micro RNAs and Circular RNAs in Different Forms of Otitis Media. International Journal of Molecular Sciences. 2023; 24(7):6752. https://doi.org/10.3390/ijms24076752

Chicago/Turabian Style

Kotowski, Michal, Paulina Adamczyk, and Jaroslaw Szydlowski. 2023. "Micro RNAs and Circular RNAs in Different Forms of Otitis Media" International Journal of Molecular Sciences 24, no. 7: 6752. https://doi.org/10.3390/ijms24076752

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