Next Article in Journal
Phosphoproteome Reveals Extracellular Regulated Protein Kinase Phosphorylation Mediated by Mitogen-Activated Protein Kinase Kinase-Regulating Granulosa Cell Apoptosis in Broody Geese
Next Article in Special Issue
Circulating miR-206 and miR-1246 as Markers in the Early Diagnosis of Lung Cancer in Patients with Chronic Obstructive Pulmonary Disease
Previous Article in Journal
Guar Gum as an Eco-Friendly Corrosion Inhibitor for N80 Carbon Steel under Sweet Environment in Saline Solution: Electrochemical, Surface, and Spectroscopic Studies
Previous Article in Special Issue
Obese Asthma Phenotype Is Associated with hsa-miR-26a-1-3p and hsa-miR-376a-3p Modulating the IGF Axis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Expression and Regulatory Mechanisms of MicroRNA in Cholesteatoma: A Systematic Review

by
Karolina Dżaman
1,
Katarzyna Czerwaty
1,
Torsten E. Reichert
2,
Mirosław J. Szczepański
1,3,* and
Nils Ludwig
2
1
Department of Otolaryngology, The Medical Centre of Postgraduate Education, 01-813 Warsaw, Poland
2
Department of Oral and Maxillofacial Surgery, University Hospital Regensburg, 93053 Regensburg, Germany
3
Department of Biochemistry, Medical University of Warsaw, 02-097 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(15), 12277; https://doi.org/10.3390/ijms241512277
Submission received: 8 July 2023 / Revised: 29 July 2023 / Accepted: 30 July 2023 / Published: 31 July 2023

Abstract

:
Cholesteatoma is a temporal bone disease characterized by dysfunctions of keratinocytes. MicroRNAs (miRNAs) are evolutionary conserved noncoding RNAs that regulate mRNA expression. They can be packaged into exosomes and transported to target cells that can be used in the future therapy of cholesteatoma. This study aimed to collect knowledge on the role of miRNAs and exosomal miRNAs in cholesteatoma and was conducted according to the PRISMA guidelines for systematic reviews. Four databases were screened: Pubmed/MEDLINE, Web of Science, Scopus, and the Cochrane Library. The last search was run on the 6th of June 2023. We included full-text original studies written in English, which examined miRNAs in cholesteatoma. The risk of bias was assessed using the Office of Health Assessment and Translation (OHAT) Risk of Bias Rating Tool, modified for the needs of this review. We identified 118 records and included 18 articles. Analyses revealed the downregulation of exosomal miR-17 as well as miR-10a-5p, miR-125b, miR-142-5p, miR34a, miR-203a, and miR-152-5p and the overexpression of exosomal miR-106b-5p as well as miR-1297, miR-26a-5p, miR-199a, miR-508-3p, miR-21-3p, miR-584-5p, and miR-16-1-3p in cholesteatoma. The role of differentially expressed miRNAs in cholesteatoma, including cell proliferation, apoptosis, the cell cycle, differentiation, bone resorption, and the remodeling process, was confirmed, making them a potential therapeutic target in this disease.

1. Introduction

Cholesteatoma is a well-demarcated non-neoplastic cystic formation composed of stratified keratinizing squamous epithelium, developing in the temporal bone. In cholesteatoma, the hyperproliferation, migration, and altered differentiation of keratinocytes are observed, which are associated with invasive and destructive growth [1].
The incidence rate of cholesteatoma is estimated at 5–9.2 per 100,000 per year [2,3,4,5,6]. The familial occurrence of cholesteatoma is the basis for investigating the genetic background of the disease [7,8], although a family history in this disease is nevertheless quite rare and can only explain a limited number of cases [9]. To date, the higher incidence of cholesteatoma has been associated with various genetic syndromes characterized by craniofacial anomalies, such as Down syndrome [10,11] or Turner syndrome [12,13]. There is a classic clinical classification of cholesteatoma into two phenotypes: congenital and acquired [14]. Middle ear cholesteatomas can be also classified according to the growth pattern into anterior epitympanic, posterior epitympanic, posterior mesotympanic, two routes (both the pars flaccida and the pars tensa are involved), and undetermined [15].
Four theories of the origin of acquired cholesteatoma were presented, including the theories of invagination, immigration, squamous metaplasia, and basal cell hyperplasia [16]. The growth and proliferation of cholesteatoma are closely linked to the upregulation of growth factors, such as epidermal growth factor (EGF) and keratinocyte growth factor (KGF) and its receptors [17,18,19], and cytokines, including IL-1, IL-6, and tumor necrosis factor α (TNF-α) [20]. Cholesteatoma growth can cause both sensorineural and conductive hearing loss, as well as other symptoms, such as dizziness or facial nerve paralysis, and can lead to other complications that include meningitis, mastoiditis, bacterial labyrinthitis, sigmoid sinus thrombophlebitis, or brain abscess.
Currently, the only effective method of treating cholesteatoma is the complete removal and repair of the damaged middle ear structures, and various surgical accesses are used for this purpose [21,22,23]. Unfortunately, recurrence after surgery is still common, and risk factors for recurrence include young age, cholesteatoma localized to the mastoid, stapes, or incus erosion [3]. It was also observed that surgical outcomes are different when comparing rural and urban cohorts [24]. Therefore, there is an urgent need to develop nonsurgical, alternative treatments based on molecular mechanisms. For this reason, it is very important to explore the cellular and molecular mechanisms underlying the pathogenesis of middle ear cholesteatoma.
MicroRNAs (miRNAs) are among small, evolutionary conserved, noncoding RNAs whose function is to regulate the activity of mRNA expression, thus affecting various biological processes. The importance of miRNA-directed gene regulation is becoming increasingly apparent as more miRNAs and their regulatory targets and functions are discovered. A single miRNA can regulate many targets [25]. MiRNAs can be packaged into exosomes and transported to different cells. This process allows for the transfer of genetic information between cells and can have significant functional implications; therefore, they may be used in the future therapy of cholesteatoma.
The role of certain miRNAs in differentiation, development, apoptosis, and oncogenesis in various cancers has been recognized. However, miRNAs have also been shown to play an important role in the development of various otological disorders, such as progressive sensorineural hearing loss, age-related hearing loss, noise-induced hearing loss, cholesteatoma, schwannomas, and inner ear inflammation [26].
In this systematic review, we collect and analyze the results of all studies on the role of miRNAs and exosomal miRNAs in cholesteatoma pathogenesis and their potential usefulness in the nonsurgical treatment of cholesteatoma.

2. Materials and Methods

This systematic review was conducted and written in accordance with the updated 2020 Preferred Reporting Items for Systematic Reviews and MetaAnalyses Statement (PRISMA) [27,28].
The PubMed/MEDLINE, Web of Science, Scopus, and Cochrane Library databases were searched systematically. The last search was initiated on the 6th of June 2023 with the language limited to English. No publication time limits were imposed. The detailed search strategy is presented in Table 1. Subsequently, the automatic duplicate identifier function built in EndNote X9 (Clarivate Analytics, London, UK) was used.
Both types of studies, observational and experimental, were included and assessed. Two reviewers (KD and KC) independently screened the studies by title and abstract. Following this selection, full texts were checked. Studies that met the selection criteria were included (Table 2).
Only original papers according to the study design were eligible for inclusion. Any apparent discrepancies resulting from the selection process were discussed and resolved by the co-authors (MS and NL or by discussion). Figure 1 provides an overview of the selection process summarized in the PRISMA flowchart.
After assessing the eligibility of all included studies, data for each study were extracted individually by two reviewers. The following results are presented in this review: first author, year of publication, country of origin, the purpose of the study, the miRNAs investigated, and finally the main results on the role of miRNAs in cholesteatoma.
The risk of bias was assessed independently by two authors (KD and KC) using the Office of Health Assessment and Translation (OHAT) Risk of Bias Rating Tool for Human and Animal Studies, modified for the needs of this review [29]. Any disagreement was discussed and resolved by the fourth and fifth co-authors (MS and NL). In general, the studies included were of high to moderate quality. The data are presented in Supplementary Table S2.

3. Results and Discussion

3.1. Search Results, Study Characteristics, and Study Quality

In total, 118 articles were retrieved through the database search. After applying the search strategy described above, a total of 18 articles were identified that met all inclusion criteria [30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47]. A flow diagram of the detailed study selection process is presented in Figure 1.
The articles were analyzed for their basic data and the main results are described in Table S1. The included studies were published between 2009 and 2023, more than half of them in the last five years, reflecting the rapid development in miRNA research in recent years. Almost all of the included studies were conducted in China, except for one study by Friedland et al. [33], which originated in the USA. All included studies used cholesteatoma tissues for analysis. Additionally, the majority of studies included in vitro experiments on cell lines [30,32,34,35,36,37,38,39,42,43,44,46,47]. In the Zheng study [46], a mouse model was also used for in vivo testing.
The quality of the included studies was estimated using the OHAT Risk of Bias Rating Tool for Human and Animal Studies, modified for the needs of this review [29], and it showed that the quality of most studies was intermediate or high. The results are presented in Table S2.

3.2. Differentially Expressed MicroRNAs in Acquired Middle Ear Cholesteatoma

In the included studies, miRNA expression profiling analysis was performed between acquired middle ear cholesteatoma and normal skin to identify miRNAs that may be involved in the aetiopathogenesis of middle ear cholesteatoma. Selected miRNAs with significantly increased or decreased expression in cholesteatoma tissue compared to normal skin are shown in Table 3. In Xie’s study [40], the miRNA microarray technology revealed 44 miRNAs with increased expressions (miR-21-3p, miR-584-5p, miR-16-1-3p, etc.) and 175 miRNAs with decreased expressions (miR-10a-5p, miR-152-5p, miR-203b-5p, etc.) in cholesteatoma tissues with a 2-fold change compared to normal skin. Subsequently, qRT-PCR validation showed that miR-21-3p and miR-16-1-3p had significantly higher expressions, while miR-10a-5p showed a markedly reduced expression in middle ear cholesteatoma tissues compared to normal skin. Finally, the Gene Ontology (GO) and Kyoto Encyclopedia Genes and Genomes (KEGG) pathway analyses provided clues that these differentially expressed miRNAs may play an important role in the aetiopathogenesis of middle ear cholesteatoma, including cell proliferation, apoptosis, the cell cycle, differentiation, bone resorption, and the remodeling process.
Differences in miRNA expression may have a potential role in the pathogenesis of cholesteatoma and be a future therapeutic target. The role of some miRNAs has been identified in cholesteatoma and the results of the available studies are presented below. More cholesteatoma research is needed to extend these early discovered relationships.

3.3. Role of Exosomal MicroRNAs in Cholesteatoma

3.3.1. MicroRNA-17 Carried by Small Extracellular Vesicles

It was identified that miRNAs may regulate osteoclasts in bone diseases, such as bone loss, metastasis, and osteosarcoma [48]. Furthermore, the role of osteoclasts in bone resorption in cholesteatoma has been confirmed [49,50]. The balance between bone resorption by osteoclasts and bone formation by osteoblasts and the communication between these cells are crucial for bone homeostasis [51,52].
Gong et al. [34] demonstrated that miRNA-17 carried by small extracellular vesicles (sEVs) isolated from keratinocytes of patients with middle ear cholesteatoma can increase receptor activator of nuclear factor-κB ligand (RANKL) expression in fibroblasts and induce osteoclast differentiation. Hereby, RANKL is known to induce osteoclast differentiation [52,53]. Using an in vitro co-culture system consisting of keratinocytes, fibroblasts, and osteoclast precursors, keratinocytes have been shown to stimulate osteoclast differentiation through the induction of RANKL in fibroblasts [52]. Gong et al. [34] observed that fibroblasts treated with keratinocyte-derived sEVs (Ker-Exo) from cholesteatoma patients increased RANKL expression and promoted osteoclast differentiation, and this was dependent on the sEV component miRNA-17, which was downregulated in Ker-Exo in cholesteatoma. These new findings provide a foundation for further research into the effect of miRNAs carried by sEVs on osteoclast differentiation in cholesteatoma, which may be used in the future to treat cholesteatoma and reduce bone destruction in this disease.

3.3.2. MicroRNA-106b-5p Carried by Small Extracellular Vesicles

Proliferating tissue in cholesteatoma requires an increased blood supply; therefore, angiogenesis appears to be a necessary condition for cholesteatoma expansion. The presence and distribution of blood vessels in cholesteatoma were investigated and revealed the presence of increased vascularization in the subepithelial connective tissue of cholesteatoma (perimatrix), which probably plays a role in maintaining continued abnormal growth [54,55,56]. Analysis of the angiogenic growth factors in middle ear cholesteatoma revealed their altered expression compared to middle ear mucosa or auditory meatal skin [55,57,58,59,60,61,62,63]. Therefore, the possibility of using antiangiogenic molecules in the adjuvant treatment of cholesteatoma surgery has been proposed [62].
Moreover, it was confirmed that miRNAs carried by sEVs can participate in stimulating angiogenesis [64,65,66]. Therefore, miRNAs carried by sEVs were one of the factors studied in cholesteatoma angiogenesis. sEVs (size range of 30–150 nm) released into the extracellular space carry cell-specific cargos of proteins, lipids, and nucleic acids (including miRNAs) that can be transferred and participate in intercellular communication [67]. One source of miRNAs in human cholesteatoma is perimatrix fibroblasts (hCPFs), which stimulate microvascular endothelial cells to proliferate and migrate and alter their gene expression patterns by releasing various angiogenic factors [37]. Li et al. [37] were the first to study the function of hCPF-derived sEVs (hCPF-Exo). Their results indicate that hCPF-Exo transports low-expressed miR-106b-5p into endothelial cells and promotes angiogenesis by overexpressing angiopoietin 2.
In conclusion, to develop effective antiangiogenic therapies, it is necessary to understand the mechanisms underlying abnormal angiogenesis.

3.4. Other MicroRNAs in Cholesteatoma

3.4.1. MicroRNA-21

The main miRNA investigated in cholesteatoma is miR-21. Its expression is altered in many cancers and other pathologies, including cardiomyopathies [68]. The upregulation of miR-21 in colorectal cancer was shown to be associated with the downregulation of phosphatase tension homolog (PTEN) protein expression, which affects proliferation, antiapoptosis, cell cycle progression, and the invasion of colorectal cancer cells [69]. Furthermore, the tumor suppressor PTEN modulates cell cycle progression and cell survival and its expression was found to be significantly lower in cholesteatoma [70,71], which may account for the impaired inhibition in cholesteatoma. Moreover, programmed cell death 4 (PDCD4) plays a role in the pathogenesis of many cancers, including skin cancer [72,73,74]. However, IL-6 and signal transducers and activators of transcription 3 (STAT3) are upstream activators of miR-21, whose role in cholesteatoma epithelial hyperproliferation has also been confirmed [71,75].
Some investigators [33,76] observed a more than 4-fold higher expression of miR-21 and a reduction in the downstream targets of miR-21, PTEN, and PDCD4 in cholesteatoma compared to normal skin. Chen et al. [31] confirmed these results and additionally reported that miR-21 was particularly elevated in pediatric patients, potentially leading to greater tumor cell proliferation and cholesteatoma invasion, which is reflected in clinical practice, where the disease process is more aggressive and invasive in children than adults [77]. Moreover, PTEN, PDCD4, and high-mobility group AT-hook 2 (HMGA2) protein levels were significantly decreased in pediatric versus adult cholesteatoma patients [31].
Others [30] provided pieces of evidence that miR-21 promotes the proliferation and invasion of cholesteatoma keratinocytes. The number of proliferative cholesteatoma keratinocytes increased after transfection with miR-21 mimics, compared to miRNA controls and miR-21 inhibitors, indicating that miR-21 can promote the growth of cholesteatoma keratinocytes. The inhibitory effect of miR-21 on the apoptosis of cholesteatoma keratinocytes was shown by changing the percentage of apoptotic cells—decreasing it after transfection with miR-21 mimetics and increasing it after transfection with miR-21 inhibitors. The percentage of migrated cholesteatoma keratinocytes transfected with miR-21 mimics was 4.82 times higher when compared to cells transfected with miR-21 inhibitors, indicating that miR-21 promotes the migration and invasion of cholesteatoma keratinocytes.
In a recent study, Chen et al. [32] demonstrated that miR-21 inhibition to some extent slows keratinocyte proliferation and induces the apoptosis of cholesteatoma keratinocytes by inducing cell cycle arrest in the G0/G1 phase through a mechanism associated with the negative regulation of PTEN and PDCD4 expression. Specifically, it was sequentially demonstrated that the proliferation of CK cells (cultured keratinocytes) in the miR-21 inhibition group was significantly lower than in the negative and blind control group, the percentage of CK cells in the G0/G1 phase in the miR-21 inhibition group was significantly higher than in the negative and blind control groups, and protein and mRNA expression levels of PTEN and PDCD4 in CK in the miR-21 group were significantly higher than in the negative and blind control groups.
To sum up, the results of the above studies on cholesteatoma showed a higher level of miR-21 expression, which was associated with the downregulation of two cholesteatoma suppressors, PTEN and PDCD4, which was particularly noticeable for pediatric cholesteatoma. Therefore, the researchers point to the potential use of miR-21 as a therapeutic target in the disease.

3.4.2. MicroRNA-508-3p and Hsa_circ_0000007

Circular RNAs (circRNAs) are covalently closed RNA molecules, which are stable, conserved, and expressed in a tissue-specific manner [78]. CircRNA has been discovered to play a role in the pathogenesis of many cancers [78,79], along with other diseases, such as cardiovascular [80], renal [81], and skin diseases [82], diabetes mellitus [83], and neurological and neuropsychiatric disorders [84,85]. The circRNA expression profile was examined in cholesteatoma and revealed 101 upregulated and 254 downregulated circRNAs in cholesteatoma [86]. CircRNAs can regulate gene expression (transcription and translation) and act as miRNA sponges [79].
It was demonstrated that the increased expression of miR-508-3p in cholesteatoma tissues and cells is inversely correlated with hsa_circ_0000007 expression [38]. Hsa_circ_0000007 expression was significantly lower in cholesteatoma than in normal skin. miR-508-3p appeared to be the targeted miRNA downstream of hsa_circ_0000007; it was upregulated in cholesteatoma, and there was a statistically negative correlation between miR-508-30 and hsa_circ_0000007 in cholesteatoma tissue. The miR-508-3p inhibitor elevated the level of pro-apoptotic factor B-cell lymphoma-2 (Bcl-2), increased PTEN expression, and impeded class I phosphoinositide 3-kinase (PIK3) and protein kinase B (Akt) expressions. Therefore, the miR-508-3p mimic decreased PTEN expression, increased the expression of PIK3 and phosphorylated Akt (p-Akt), and promoted the proliferation of middle ear cholesteatoma cells. The PIK3/Akt signaling pathway plays a key role in cell growth, survival, proliferation, metabolism, and motility in various cancer phenotypes and PIK3/Akt signaling pathway inhibitors are a highly effective treatment strategy [87,88].
In conclusion, the findings point to the targeted regulation of the PTEN/PI3K/Akt signaling pathway by miR-508-3p in the cholesteatoma pathogenesis, while miR-508-3p overexpression in cholesteatoma is probably mediated by the regulation of upstream hsa_circ_0000007.

3.4.3. Let-7a MicroRNA

Let-7a miRNA is categorized as a tumor suppressor and targets oncogenes, including the HMGA2 [89,90]. However, in some rare cases, let-7a miRNA can also act as an oncogene, increasing cancer invasion and chemoresistance [91]. Furthermore, its expression is also altered in cardiovascular diseases and inflammatory processes [92].
The level of let-7a miRNA was significantly elevated in cholesteatoma tissue compared to normal skin, particularly in pediatric patients [31]. HMGA2 protein levels were significantly decreased in pediatric versus adult cholesteatoma patients, which may correspond to an increased apoptosis of keratinocytes and decreased proliferation of cholesteatoma cells [31].
It was observed that let-7a miRNA inhibited the growth of cholesteatoma keratinocytes by reducing keratinocyte proliferation by promoting cell cycle arrest in the G0/G1 phase and inducing cell apoptosis. The authors demonstrated also that let-7a miRNA induces the cell apoptosis, migration, and invasion of cholesteatoma keratinocytes. Finally, let-7a miRNA downregulated miR-21 expression in cholesteatoma keratinocytes, which was also confirmed by the upregulation of miR-21 in the cholesteatoma keratinocytes transfected with the let-7a inhibitor.
To sum up, the results of those studies indicate that let-7a plays a positive role in inhibiting cholesteatoma and can control the proliferation and apoptosis of cholesteatoma keratinocytes through both miR-21 and HMGA2 downregulation.

3.4.4. MicroRNA-125 and Circ_0074491

Another miRNA whose role has been explored in cholesteatoma pathogenesis is miE-125. It has been linked to the regulation of tumorigenesis and tumor development, and its downstream targets include transcription factors such as STAT3, cytokines such as IL-6 and TGFβ, and suppressor protein p53 [93,94]. MiR-125b was found to be significantly downregulated in psoriatic skin, which may contribute to the extensive proliferation and inappropriate differentiation of keratinocytes observed in this disease [95,96]. STAT3 target genes are involved in proliferation, survival, self-renewal, invasion, and angiogenesis [97,98]. The blocking of the JAK2/STAT3 signaling pathway inhibits IL-17-induced vascular endothelial growth factor (VEGF) expression, leading to the inhibition of angiogenesis in inflammatory skin diseases [99]. The expression of STAT3 was found to be significantly higher in the cholesteatoma epithelium than in the normal epithelium of the external auditory canal or retroauricular skin, indicating its potential role in the mechanisms of hyperproliferation and growth in cholesteatoma [100,101].
Zang et al. [44] observed downregulation of miR-125b and upregulation of STAT3, cyclin D1, survivin, and VEGF in cholesteatoma tissues, which is consistent with the results of previous studies [17,57,102,103]. These reports imply that the inhibition of miR-125b expression in cholesteatoma may contribute to high proliferation and low apoptosis through an increased expression of STAT3, which could be used as a potential therapeutic target for intratympanic pharmacotherapy of cholesteatoma [17].
Hu et al. [35] presented the results of a study in which circ_0074491 was defined as a decoy for miR-22-3p and miR-125a-5p in cholesteatoma keratinocytes and was downregulated in cholesteatoma specimens. Furthermore, circ_0074491 knockdown reduced cell cycle arrest and apoptosis; facilitated cell proliferation, migration, colony formation, and the invasion of cholesteatoma keratinocytes; and activated the PIK3/Akt pathway via miR-22-3p and miR-125a-5p in cholesteatoma keratinocytes. Moreover, both miR-22-3p and miR-125a-5p inhibitors reversed the impacts of circ_0074491 silencing on the proliferation, apoptosis, migration, and invasion of cholesteatoma keratinocytes. In conclusion, circ_0074491 was shown to regulate cholesteatoma progression through the PIK3/Akt pathway and binding to miR-22-3p and miR-125a-5p.
Finally, most researchers confirmed the protective role of miR-125 in cholesteatoma, especially through the impact on the apoptosis pathway, and found it to be significantly downregulated in patients with this disease.

3.4.5. MicroRNA-10a-5p

MiR-10a-5p is another miRNA that plays a role in cancer and was studied in cholesteatoma. A lower expression of miR-10a-5p has been linked to the progression of hepatocellular carcinoma [104], melanoma [105], and ovarian cancer [106]. However, miR-10a-5p was also found to be upregulated in the affected skin of atopic dermatitis patients and in proliferating keratinocytes [107]. Xie et al. [40] identified the phosphatidylinositol-4,5-bisphosphonate 3-kinase catalytic subunit α (PIK3CA) as an important miR-10a-5p target gene. PIK3CA is a catalytic subunit of PIK3 that plays a role in the EGFR/PIK3/Akt/cyclinD1 signaling pathway, which is active in cholesteatoma and may have a key function in epithelial hyperproliferation [17].
In another study [41], middle ear cholesteatoma tissues showed significantly decreased miR-10a-5p expression and increased PIK3CA expression compared to normal posterior ear skin tissues (p < 0.05). The results of target gene prediction revealed that PIK3CA may be an miR-10a-5p target gene. Additionally, miR-10a-5p and PIK3CA expression levels were significantly negatively correlated in middle ear cholesteatoma tissues (r = −0.926, p < 0.001). These results indicate that miR-10a-5p can inhibit cholesteatoma proliferation and differentiation through the negative regulation of its target gene, PIK3CA.

3.4.6. MicroRNA-802

MiR-802 is another miRNA that has been studied in cholesteatoma. Previously, its role was confirmed in mediating the pathogenesis of a variety of tumors [108], obesity [109,110], osteoporosis [111], and inflammatory bowel disease [112].
Li et al. [36], for the first time, demonstrated the role of the NF-κB/miR-802/PTEN signaling pathway in cholesteatoma. In this study, cholesteatoma tissues showed a high activation of NF-κB and upregulation of proinflammatory cytokines such as TNFα, IL-1b, and IL-6. The researchers observed the upregulation of miR-802 in keratinocytes treated with TNFα, IL-1b, or IL-6 and revealed that the miR-802 promoter contained a functional NF-κB/P65 binding site. The authors demonstrated that miR-802 can promote keratinocyte proliferation and cell cycle progression. They also noticed that keratinocytes transfected with an has-miR-802 mimetic showed a significantly increased proliferation rate, a significantly reduced percentage of cells in the G1/G0 phase, and an increased percentage of cells in the S phase. Additionally, the inhibition of miR-802 reduced the above effects. Finally, the authors performed a computational analysis using miRWalk software and found that the gene encoding PTEN harbored an miR-802 binding site and suggested that miR-802 could directly inhibit PTEN expression by targeting its 30-untranslated region (UTR).

3.4.7. MicroRNA-1297, MicroRNA-26a-5p, and MicroRNA-203a

MiR-1297 and miR-26a-5p are other miRNAs that have been studied in cholesteatoma and are widely investigated in oncology. The inhibitory effect of miR-1297 on cancer progression has been observed in various malignant entities [113,114,115,116,117,118,119,120,121,122,123,124]. In addition to its role in tumorigenesis, miR-26a-5p similarly plays an important role in pulmonary fibrosis [125] and Alzheimer’s disease [126]. However, the included studies provided new information on the role of miR-1297 and miR-26a-5p concerning the pathogenesis of cholesteatoma. miR-1297 and miR-26a-5p were shown to inhibit the progression of cholesteatoma keratinocytes by targeting the B-cell-specific Moloney murine leukemia virus insertion site 1 (BMI1) [47]. BMI1 is a component of the polycomb 70 repressive complex 1, which is a transcriptional repressor that increases cell survival [127]. The role of BMI1 has been identified in various cancers [128,129,130,131,132,133] and was shown to be linked to poor prognosis. Additionally, BMI1 levels were found to be elevated in transformed keratinocytes, skin tumors, and psoriasis and were shown to play a role in keratinocyte survival [127,134]. It was noted that BMI1 is expressed mainly in the basal and suprabasal layers of the epithelium in normal skin [43,134,135], but in cholesteatoma it was found in almost all layers of the epithelium, and the intensity of staining was comparably stronger [43].
A significantly elevated mRNA expression of BMI1 and protein level of BMI1 were demonstrated in cholesteatoma tumor tissues relative to normal retroauricular skin specimens [43,47]. The downregulation of BMI1 was shown to significantly inhibit the proliferation of human immortalized keratinocytes, indicating a pro-oncogenic role for BMI1 in cholesteatomas, which is consistent with observations in other malignant entities [47]. The prediction result obtained with DianaTools software showed that BMI1 directly interacts with miR-1297 and miR-26a-5p in human immortalized keratinocytes.
Another miRNA targeting BMI1 is miR-203a. The role of miR-203 as a tumor suppressor has been demonstrated in a number of cancers [136,137,138]. It is known that miR-203 promotes epidermal differentiation by limiting the proliferative potential [139] and is upregulated in psoriatic lesions [95], while the upregulation of miR-203 expression by oleic acid treatment accelerates human keratinocyte differentiation [140]. Therefore, its role in the growth and proliferation of cholesteatoma was investigated [43], leading to the observation that miR-203a is downregulated and BMI1 upregulated in cholesteatoma. Subsequently, the authors demonstrated in the dual-luciferase reporter assay (DLRA) that BMI1 is a direct target of miR-203a. Furthermore, reduced miR-203a expression increased BMI1 expression; promoted the proliferation, colony formation, and migration of HaCaT cells (a spontaneously transformed aneuploid immortal keratinocyte cell line from adult human skin); and inhibited apoptosis. Moreover, p-Akt was significantly increased in cholesteatoma tissues and was positively correlated with BMI1, which is consistent with other studies [141,142]. Finally, BMI1 suppression decreased p-Akt expression in HaCaT cells and subsequent miR-203a inhibition reversed this effect [43].
Taken together, all the studies confirmed that all three miRs, miR-1297, miR-26a-5p, and miR-203a, inhibit the progression of cholesteatoma keratinocytes by targeting BMI1. The new reports reveal the need for further research to explore specific therapeutic methods for cholesteatoma patients through a combination of nanotechnology and molecular drugs targeting those miRNAs or BMI1.

3.4.8. MicroRNA-199a

MiR-199a was shown to be involved in the proliferation, migration, invasion, and apoptosis of cancer cells from different primary tumors [143]. Its role has also been confirmed in other noncancerous diseases, such as generalized epilepsy [144] or osteoporosis [145].
Yao et al. [42] reported that miR-199a was significantly upregulated in cholesteatoma tissues compared to the normal retroauricular skin tissues, which facilitated the proliferation, migration, and invasion of cholesteatoma keratinocytes in vitro, while miR-199a downregulation caused the opposite effects. The authors demonstrated a negative regulatory relationship between miR-199a and PNRC1 and found miR-199a binding to the 3′-UTR of PNRC1.
To sum up, the upregulation of miR-199a in cholesteatoma potentially plays a role in the development of cholesteatoma and could be used as a biomarker or therapeutic target.

3.4.9. MicroRNA-142-5p

Another study focused on the role of miR-142 in cholesteatoma pathogenesis. miR-142 is expressed in many tissues and was found to play an important role in inflammatory and immune responses [146,147,148,149]. The differential expression patterns of miR-142-5p were observed in cancer tissues [150,151,152,153,154]. A downstream target gene of miR-142-5p is cyclin-dependent kinase 5 (CDK5) [39], whose role has been found in the development and progression of many cancers [155,156,157] and neurodegenerative [158,159] and inflammatory disorders [160,161,162]. Sui et al. [39] demonstrated that miR-142-5p directly inhibits the CDK5-mediated upregulation of inflammatory cytokines in acquired middle ear cholesteatoma, making it a potential therapeutic target in this disease.

3.4.10. Micro-RNA-34

MiR-34 is characterized by antitumor properties, is induced by the tumor suppressor p53, and can be used to control tumor progression [163,164]. Thus, one of the included studies revealed the function of miR-34 in cholesteatoma [42]. The researchers conducted an experimental study using nanoparticles in which therapeutic targets were achieved by regulating miR-34a expression. They showed that free rubone and rubon-containing drug nanoparticles (RC NPs) can significantly inhibit the proliferation and migration ability of cholesteatoma cells in children’s middle ear cholesteatoma. The authors demonstrated that drug-loaded nanoparticles could deliver rubone into cells and upregulate miR-34a levels and downregulate mRNA expression levels of Bcl-2, CDK6, and cyclin D1 in cells. The expression level of miR-34a in the free rubone group was slightly higher than that in the RC NPs group, which was due to the slow and controlled release of nanoparticles. The study indicated the potential of these novel forms of treatment in both cholesteatoma and also other oncologic diseases.

3.5. Strengths and Limitations of the Included Studies

Research on the association between cholesteatoma and microRNAs is still in the exploratory phase. The main advantage of the included studies is that they provide discoveries in the field of understanding the etiopathogenesis of cholesteatoma, which may open up the possibility of developing new microparticle-based treatment strategies for this disease. It should be noted that some of the studies included cholesteatoma samples collected in the pediatric population, which also provide better insight into the pathogenesis of this more aggressive form of the disease [31,32,46]. Moreover, new studies involving miRNAs carried by sEVs were also among the included studies, which provided new insight into the role of these nanoparticles in cholesteatoma [34,37].
However, many of the included studies analyzed a small number of clinical samples (most studies include cholesteatoma tissues from less than 30 patients), and the population from which tissues were collected was not homogeneous in terms of age or other parameters. In some studies, it was not specified precisely whether the samples were only from acquired cholesteatoma patients. Another limitation is the lack of information on the number of people from whom the tissues analyzed in some studies were collected [32,33,34,37]. Future studies should use a larger number of samples and ensure the homogeneity of the groups to avoid factors that could affect the results of the study. The pathogenesis of acquired and congenital cholesteatoma can differ, so it is important to emphasize which form of this disease the study is concerned with. Larger, more homogenized cohorts would provide more robust and generalizable data on differentially expressed miRNAs. Another problem is the limited validation in the included studies. Many findings are based on initial microarray screens that have not been extensively replicated in independent datasets. Further validation is needed to confirm key microRNAs and their roles. Unfortunately, microRNA mechanisms of action in promoting or inhibiting cholesteatoma progression are still unclear, and further work is needed to clarify how they affect keratinocyte behaviors.
Retroauricular skin has been used as a control material in many of the included studies [31,33,35,36,37,38,40,41,42,43,44,45,47]. There are reports that the selection of control group material in cholesteatoma studies can be important for the outcome of the studies, with deep meatal skin specimens being more similar in protein profile to cholesteatoma than retroauricular skin [165,166]. Moreover, canal skin is exposed to infection, which can stimulate cytokine expression and changes in protein expression, which can also affect research results [33].
Most studies show only correlational analyses established between miRNA levels and target mRNAs/proteins. Experimental manipulations are often lacking to demonstrate causality.
The study by Zhu et al. [47] provides new insights into the involvement of miR-1297/BMI1 and miR-26a-5p/BMI1 regulatory mechanisms in the pathogenesis of cholesteatoma, but only in vitro experiments were performed, which have some limitations. Animal studies are needed to confirm the relationships discovered.
The study by Li et al. [37] has its limitations because it is an in vitro study in which miR-106b-5p carried by sEVs derived from cholesteatoma hCPFs is just one of the sEV miRNA cargo components that promote angiogenesis in middle ear cholesteatoma, and the possible influence of other pro-angiogenic factors of the sEV cargo cannot be ignored. Therefore, further studies should be performed in an animal model of middle ear cholesteatoma to investigate the suppressive effect of sEV-associated miR-106b-5p on cholesteatoma growth in vivo. Animal models could help determine the true significance of microRNAs in cholesteatoma pathogenesis.
Recent studies suggest that abnormal miRNA expression may contribute to cholesteatoma pathogenesis, though causal links remain unclear based on preliminary evidence. Targeting relevant microRNAs could potentially offer adjunctive treatments to supplement current surgical treatments like type 1 tympanoplasty [167,168]. However, microRNA-based therapies for cholesteatoma remain in the early stages of research, with limited validation in large cohorts and animal models. Further rigorous investigation is needed before these approaches could mature sufficiently to realistically improve outcomes of procedures like type 1 tympanoplasty in clinical practice.

4. Conclusions

Cholesteatoma is a disease that significantly impairs function, and its complications can be life-threatening. The pathogenesis of this disease is still under investigation, and knowledge in this field has recently been supplemented by studying the role of certain miRNAs and circRNAs. These are very important discoveries as they might contribute to the identification of pharmacological targets for the treatment of cholesteatoma, but there is a need for further validation and causative evidence. The presence of miRNAs in exosomes suggests a mechanism by which cells can communicate and modulate gene expression in a paracrine or endocrine manner. However, further large-scale validation studies, the experimental manipulation of microRNA levels, the elucidation of mechanisms, and in vivo testing are needed to substantiate these initial correlations and guide potential therapeutic applications. Comprehensive microRNA profiling in larger cohorts to identify robust disease-associated miRNAs are needed. Furthermore, future research should include integrated analyses of miRNAs, target mRNAs, and proteins to map regulatory networks. Moreover, the specific mechanisms by which miRNAs regulate cholesteatoma development remain largely unknown. Future work should contain functional studies manipulating microRNA levels and assessing impacts on keratinocyte behaviors and clarify how miRNAs influence keratinocyte proliferation, apoptosis, differentiation, etc. Additionally, there is limited functional testing in vitro and a lack of in vivo animal models to determine the true significance of miRNAs in disease pathogenesis. Investigations of miRNA roles in growth factor and cytokine signaling dysregulation are required. Finally, future work should also consider the development and testing of miRNA-based therapies in animal models. Hopefully, in the future, RNA- and protein-based therapies using nanoparticles will be possible for the nonsurgical or adjunctive treatment of cholesteatoma.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms241512277/s1.

Author Contributions

Conceptualization, K.D. and M.J.S.; methodology, K.D. and K.C.; investigation, K.D. and K.C.; data curation, K.D. and K.C.; writing—original draft preparation, K.D., K.C., T.E.R., M.J.S. and N.L.; writing—review and editing, M.J.S. and N.L.; visualization, K.D. and K.C.; supervision, M.J.S.; project administration, M.J.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Science Centre, Poland, grant 2017/25/B/NZ5/02949 to MJS.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kim, H.J.; Tinling, S.P.; Chole, R.A. Expression patterns of cytokeratins in cholesteatomas: Evidence of increased migration and proliferation. J. Korean Med. Sci. 2002, 17, 381–388. [Google Scholar] [CrossRef] [Green Version]
  2. Li, J.; Jufas, N.; Forer, M.; Patel, N. Incidence and trends of middle ear cholesteatoma surgery and mastoidectomy in Australia-A national hospital morbidity database analysis. Laryngoscope Investig. Otolaryngol. 2022, 7, 210–218. [Google Scholar] [CrossRef]
  3. Britze, A.; Møller, M.L.; Ovesen, T. Incidence, 10-year recidivism rate and prognostic factors for cholesteatoma. J. Laryngol. Otol. 2017, 131, 319–328. [Google Scholar] [CrossRef]
  4. Shibata, S.; Murakami, K.; Umeno, Y.; Komune, S. Epidemiological study of cholesteatoma in Fukuoka City. J. Laryngol. Otol. 2015, 129 (Suppl. 2), S6–S11. [Google Scholar] [CrossRef]
  5. Kemppainen, H.O.; Puhakka, H.J.; Laippala, P.J.; Sipilä, M.M.; Manninen, M.P.; Karma, P.H. Epidemiology and aetiology of middle ear cholesteatoma. Acta Otolaryngol. 1999, 119, 568–572. [Google Scholar] [CrossRef] [PubMed]
  6. Homøe, P.; Rosborg, J. Family cluster of cholesteatoma. J. Laryngol. Otol. 2007, 121, 65–67. [Google Scholar] [CrossRef] [PubMed]
  7. Jennings, B.A.; Prinsley, P.; Philpott, C.; Willis, G.; Bhutta, M.F. The genetics of cholesteatoma. A systematic review using narrative synthesis. Clin. Otolaryngol. 2018, 43, 55–67. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Collins, R.; Ta, N.H.; Jennings, B.A.; Prinsley, P.; Philpott, C.M.; Steel, N.; Clark, A. Cholesteatoma and family history: An international survey. Clin. Otolaryngol. 2020, 45, 500–505. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Bonnard, Å.; Engmér Berglin, C.; Wincent, J.; Eriksson, P.O.; Westman, E.; Feychting, M.; Mogensen, H. The Risk of Cholesteatoma in Individuals With First-degree Relatives Surgically Treated for the Disease. JAMA Otolaryngol. Head Neck Surg. 2023, 149, 390–396. [Google Scholar] [CrossRef]
  10. Poliner, A.; Mahomva, C.; Williams, C.; Alfonso, K.; Anne, S.; Musso, M.; Liu, Y.C. Prevalence and surgical management of cholesteatoma in Down Syndrome children. Int. J. Pediatr. Otorhinolaryngol. 2022, 157, 111126. [Google Scholar] [CrossRef]
  11. Spinner, A.; Munjuluru, A.; Wootten, C.T. Prevalence of Cholesteatoma in Children With Down Syndrome Receiving Treatment at Pediatric Health Care Facilities. JAMA Otolaryngol. Head Neck Surg. 2020, 146, 864–865. [Google Scholar] [CrossRef] [PubMed]
  12. Zanetti, D.; Di Lella, F.; Negri, M.; Vincenti, V. Surgical management of middle ear cholesteatoma in children with Turner syndrome: A multicenter experience. Acta Biomed. 2018, 89, 382–388. [Google Scholar] [CrossRef] [PubMed]
  13. Dorney, I.; Otteson, T.; Kaelber, D.C. Middle ear cholesteatoma prevalence in over 3,600 children with Turner Syndrome. Int. J. Pediatr. Otorhinolaryngol. 2022, 161, 111289. [Google Scholar] [CrossRef] [PubMed]
  14. Meyerhoff, W.L.; Truelson, J. Cholesteatoma staging. Laryngoscope 1986, 96, 935–939. [Google Scholar] [CrossRef]
  15. Rosito, L.S.; Netto, L.F.; Teixeira, A.R.; da Costa, S.S. Classification of Cholesteatoma According to Growth Patterns. JAMA Otolaryngol. Head Neck Surg. 2016, 142, 168–172. [Google Scholar] [CrossRef] [Green Version]
  16. Kuo, C.L. Etiopathogenesis of Acquired Cholesteatoma: Prominent Theories and Recent Advances in Biomolecular Research. Laryngoscope 2015, 125, 234–240. [Google Scholar] [CrossRef]
  17. Liu, W.; Ren, H.; Ren, J.; Yin, T.; Hu, B.; Xie, S.; Dai, Y.; Wu, W.; Xiao, Z.; Yang, X.; et al. The role of EGFR/PI3K/Akt/cyclinD1 signaling pathway in acquired middle ear cholesteatoma. Mediat. Inflamm. 2013, 2013, 651207. [Google Scholar] [CrossRef] [Green Version]
  18. Yamamoto-Fukuda, T.; Takahashi, H.; Koji, T. Expression of keratinocyte growth factor (KGF) and its receptor in a middle-ear cavity problem. Int. J. Pediatr. Otorhinolaryngol. 2012, 76, 76–81. [Google Scholar] [CrossRef] [Green Version]
  19. Yamamoto-Fukuda, T.; Akiyama, N. Keratinocyte growth factor signaling promotes stem/progenitor cell proliferation under p63 expression during middle ear cholesteatoma formation. Curr. Opin. Otolaryngol. Head Neck Surg. 2020, 28, 291–295. [Google Scholar] [CrossRef]
  20. Kuczkowski, J.; Sakowicz-Burkiewicz, M.; Iżycka-Świeszewska, E.; Mikaszewski, B.; Pawełczyk, T. Expression of tumor necrosis factor-α, interleukin-1α, interleukin-6 and interleukin-10 in chronic otitis media with bone osteolysis. ORL J. Otorhinolaryngol. Relat. Spec. 2011, 73, 93–99. [Google Scholar] [CrossRef]
  21. Mulazimoglu, S.; Meco, C. Endoscopic diving technique for hearing preservation in managing labyrinth-invading cholesteatomas. Eur. Arch. Otorhinolaryngol. 2023, 280, 1639–1646. [Google Scholar] [CrossRef]
  22. Salem, J.; Bakundukize, J.; Milinis, K.; Sharma, S.D. Mastoid obliteration versus canal wall down or canal wall up mastoidectomy for cholesteatoma: Systematic review and meta-analysis. Am. J. Otolaryngol. 2023, 44, 103751. [Google Scholar] [CrossRef] [PubMed]
  23. Shakya, D.; Nepal, A. Transcanal Endoscopic Retrograde Mastoidectomy for Cholesteatoma: A Prospective Study. Ear Nose Throat J. 2023, 102, Np269–Np276. [Google Scholar] [CrossRef] [PubMed]
  24. Kennedy, K.L.; Connolly, K.M.; Albert, C.L.; Goldman, J.L.; Cash, E.D.; Severtson, M.A. Postoperative Recurrent Cholesteatoma in Rural Versus Urban Populations. Otol. Neurotol. 2021, 42, e459–e463. [Google Scholar] [CrossRef] [PubMed]
  25. Bartel, D.P. MicroRNAs: Genomics, biogenesis, mechanism, and function. Cell 2004, 116, 281–297. [Google Scholar] [CrossRef] [Green Version]
  26. Mahmoudian-Sani, M.R.; Mehri-Ghahfarrokhi, A.; Ahmadinejad, F.; Hashemzadeh-Chaleshtori, M.; Saidijam, M.; Jami, M.S. MicroRNAs: Effective elements in ear-related diseases and hearing loss. Eur. Arch. Otorhinolaryngol. 2017, 274, 2373–2380. [Google Scholar] [CrossRef]
  27. Page, M.J.; McKenzie, J.E.; Bossuyt, P.M.; Boutron, I.; Hoffmann, T.C.; Mulrow, C.D.; Shamseer, L.; Tetzlaff, J.M.; Akl, E.A.; Brennan, S.E.; et al. The PRISMA 2020 statement: An updated guideline for reporting systematic reviews. J. Clin. Epidemiol. 2021, 134, 178–189. [Google Scholar] [CrossRef] [PubMed]
  28. Page, M.J.; Moher, D.; Bossuyt, P.M.; Boutron, I.; Hoffmann, T.C.; Mulrow, C.D.; Shamseer, L.; Tetzlaff, J.M.; Akl, E.A.; Brennan, S.E.; et al. PRISMA 2020 explanation and elaboration: Updated guidance and exemplars for reporting systematic reviews. BMJ 2021, 372, n160. [Google Scholar] [CrossRef] [PubMed]
  29. Eick, S.M.; Goin, D.E.; Chartres, N.; Lam, J.; Woodruff, T.J. Assessing risk of bias in human environmental epidemiology studies using three tools: Different conclusions from different tools. Syst. Rev. 2020, 9, 249. [Google Scholar] [CrossRef]
  30. 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]
  31. 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]
  32. Chen, X.; Xiao, F.; Dong, S.; Chen, Y.; Wang, J. MiR-21 inhibits the proliferation of childhood cholesteatoma glioma cells by negatively regulating the expressions of PTEN and PDCD4. Trop. J. Pharm. Res. 2021, 20, 1119–1124. [Google Scholar] [CrossRef]
  33. Friedland, D.R.; Eernisse, R.; Erbe, C.; Gupta, N.; Cioffi, J.A. Cholesteatoma growth and proliferation: Posttranscriptional regulation by microRNA-21. Otol. Neurotol. 2009, 30, 998–1005. [Google Scholar] [CrossRef] [Green Version]
  34. 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]
  35. Hu, Y.; 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: An observational study. Medicine 2021, 100, e27122. [Google Scholar] [CrossRef]
  36. Li, N.; Qin, Z.B. Inflammation-induced miR-802 promotes cell proliferation in cholesteatoma. Biotechnol. Lett. 2014, 36, 1753–1759. [Google Scholar] [CrossRef] [PubMed]
  37. 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]
  38. 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] [PubMed]
  39. 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]
  40. 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] [Green Version]
  41. Yang, J.; Yan, W.; Tang, S.; Huang, Z.; Ye, M.; Lu, Z.; Liu, Q. Expression and Correlation Research of MicroRNA10a-5p and PIK3CA in Middle Ear Cholesteatoma. J. Int. Adv. Otol. 2023, 19, 212–216. [Google Scholar] [CrossRef]
  42. 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, 1442093. [Google Scholar] [CrossRef]
  43. 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]
  44. Zang, J.; Yang, B.; Feng, S.; Jiang, X. Low expression of microRNA-125b enhances the expression of STAT3 and contributes to cholesteatoma growth. Arch. Med. Sci. 2022, 18, 1596–1606. [Google Scholar] [CrossRef] [PubMed]
  45. Zhang, W.; Chen, X.; Qin, Z. MicroRNA let-7a suppresses the growth and invasion of cholesteatoma keratinocytes. Mol. Med. Rep. 2015, 11, 2097–2103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. 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]
  47. Zhu, X.; Ye, F.; Hao, S.; Yu, Q.; Wang, Y.; Lou, W.; Zhao, K.; Li, H. MiR-1297 and MiR-26a-5p Inhibit Cell Progression of Keratinocytes in Cholesteatoma Depending on the Regulation of BMI1. Biotechnol. Bioprocess Eng. 2022, 27, 79–88. [Google Scholar] [CrossRef]
  48. Doghish, A.S.; Elballal, M.S.; Elazazy, O.; Elesawy, A.E.; Shahin, R.K.; Midan, H.M.; Sallam, A.M.; Elbadry, A.M.M.; Mohamed, A.K.I.; Ishak, N.W.; et al. miRNAs as potential game-changers in bone diseases: Future medicinal and clinical uses. Pathol. Res. Pract. 2023, 245, 154440. [Google Scholar] [CrossRef]
  49. Kuczkowski, J.; Brzoznowski, W.; Nowicki, T. Bone Damage in Chronic Otitis Media. Ear Nose Throat J. 2022, 101, 428–429. [Google Scholar] [CrossRef]
  50. Imai, R.; Sato, T.; Iwamoto, Y.; Hanada, Y.; Terao, M.; Ohta, Y.; Osaki, Y.; Imai, T.; Morihana, T.; Okazaki, S.; et al. Osteoclasts Modulate Bone Erosion in Cholesteatoma via RANKL Signaling. J. Assoc. Res. Otolaryngol. 2019, 20, 449–459. [Google Scholar] [CrossRef]
  51. Chen, X.; Wang, Z.; Duan, N.; Zhu, G.; Schwarz, E.M.; Xie, C. Osteoblast-osteoclast interactions. Connect Tissue Res. 2018, 59, 99–107. [Google Scholar] [CrossRef]
  52. Iwamoto, Y.; Nishikawa, K.; Imai, R.; Furuya, M.; Uenaka, M.; Ohta, Y.; Morihana, T.; Itoi-Ochi, S.; Penninger, J.M.; Katayama, I.; et al. Intercellular Communication between Keratinocytes and Fibroblasts Induces Local Osteoclast Differentiation: A Mechanism Underlying Cholesteatoma-Induced Bone Destruction. Mol. Cell Biol. 2016, 36, 1610–1620. [Google Scholar] [CrossRef] [Green Version]
  53. 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]
  54. Bujía, J.; Holly, A.; Stammberger, M.; Sudhoff, H. Angiogenesis in cholesteatoma of the middle ear. Acta Otorrinolaringol. Esp. 1996, 47, 187–192. [Google Scholar] [PubMed]
  55. Sudhoff, H.; Dazert, S.; Gonzales, A.M.; Borkowski, G.; Park, S.Y.; Baird, A.; Hildmann, H.; Ryan, A.F. Angiogenesis and angiogenic growth factors in middle ear cholesteatoma. Am. J. Otol. 2000, 21, 793–798. [Google Scholar]
  56. Olszewska, E.; Chodynicki, S.; Chyczewski, L. Apoptosis in the pathogenesis of cholesteatoma in adults. Eur. Arch. Oto-Rhino-Laryngol. 2006, 263, 409–413. [Google Scholar] [CrossRef]
  57. Fukudome, S.; Wang, C.; Hamajima, Y.; Ye, S.; Zheng, Y.; Narita, N.; Sunaga, H.; Fujieda, S.; Hu, X.; Feng, L.; et al. Regulation of the angiogenesis of acquired middle ear cholesteatomas by inhibitor of DNA binding transcription factor. JAMA Otolaryngol. Head Neck Surg. 2013, 139, 273–278. [Google Scholar] [CrossRef] [Green Version]
  58. Starke, R.D.; Ferraro, F.; Paschalaki, K.E.; Dryden, N.H.; McKinnon, T.A.; Sutton, R.E.; Payne, E.M.; Haskard, D.O.; Hughes, A.D.; Cutler, D.F.; et al. Endothelial von Willebrand factor regulates angiogenesis. Blood 2011, 117, 1071–1080. [Google Scholar] [CrossRef]
  59. Samuelson Bannow, B.; Recht, M.; Négrier, C.; Hermans, C.; Berntorp, E.; Eichler, H.; Mancuso, M.E.; Klamroth, R.; O’Hara, J.; Santagostino, E.; et al. Factor VIII: Long-established role in haemophilia A and emerging evidence beyond haemostasis. Blood Rev. 2019, 35, 43–50. [Google Scholar] [CrossRef]
  60. Acharya, S.S.; Kaplan, R.N.; Macdonald, D.; Fabiyi, O.T.; DiMichele, D.; Lyden, D. Neoangiogenesis contributes to the development of hemophilic synovitis. Blood 2011, 117, 2484–2493. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Costa, J.R.; Rego, A.R.; Soares, T.; Sousa, C.A.E.; Coutinho, M.B. Changes in Coagulation Study and Risk of Developing Cholesteatoma: Is There a Link? J. Audiol. Otol. 2023, 27, 30–36. [Google Scholar] [CrossRef] [PubMed]
  62. Reis Rego, Â.; Santos, M.; Coutinho, M.; Feliciano, T.; Almeida e Sousa, C. Is von Willebrand disease linked to cholesteatoma aetiology? Med. Hypotheses 2017, 100, 43–45. [Google Scholar] [CrossRef]
  63. Yoshikawa, M.; Kojima, H.; Yaguchi, Y.; Okada, N.; Saito, H.; Moriyama, H. Cholesteatoma Fibroblasts Promote Epithelial Cell Proliferation through Overexpression of Epiregulin. PLoS ONE 2013, 8, e66725. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. van Balkom, B.W.; de Jong, O.G.; Smits, M.; Brummelman, J.; den Ouden, K.; de Bree, P.M.; van Eijndhoven, M.A.; Pegtel, D.M.; Stoorvogel, W.; Würdinger, T.; et al. Endothelial cells require miR-214 to secrete exosomes that suppress senescence and induce angiogenesis in human and mouse endothelial cells. Blood 2013, 121, 3997–4006. [Google Scholar] [CrossRef] [Green Version]
  65. Qu, Q.; Liu, L.; Cui, Y.; Liu, H.; Yi, J.; Bing, W.; Liu, C.; Jiang, D.; Bi, Y. miR-126-3p containing exosomes derived from human umbilical cord mesenchymal stem cells promote angiogenesis and attenuate ovarian granulosa cell apoptosis in a preclinical rat model of premature ovarian failure. Stem Cell Res. Ther. 2022, 13, 352. [Google Scholar] [CrossRef]
  66. Cabello, P.; Torres-Ruiz, S.; Adam-Artigues, A.; Forés-Martos, J.; Martínez, M.T.; Hernando, C.; Zazo, S.; Madoz-Gúrpide, J.; Rovira, A.; Burgués, O.; et al. miR-146a-5p Promotes Angiogenesis and Confers Trastuzumab Resistance in HER2+ Breast Cancer. Cancers 2023, 15, 2138. [Google Scholar] [CrossRef]
  67. Dżaman, K.; Czerwaty, K. Roles of Exosomes in Chronic Rhinosinusitis: A Systematic Review. Int. J. Mol. Sci. 2022, 23, 1284. [Google Scholar] [CrossRef]
  68. Surina, S.; Fontanella, R.A.; Scisciola, L.; Marfella, R.; Paolisso, G.; Barbieri, M. miR-21 in Human Cardiomyopathies. Front. Cardiovasc. Med. 2021, 8, 767064. [Google Scholar] [CrossRef]
  69. Wu, Y.; Song, Y.; Xiong, Y.; Wang, X.; Xu, K.; Han, B.; Bai, Y.; Li, L.; Zhang, Y.; Zhou, L. MicroRNA-21 (Mir-21) Promotes Cell Growth and Invasion by Repressing Tumor Suppressor PTEN in Colorectal Cancer. Cell Physiol. Biochem. 2017, 43, 945–958. [Google Scholar] [CrossRef]
  70. Yune, T.Y.; Byun, J.Y. Expression of PTEN and phosphorylated Akt in human cholesteatoma epithelium. Acta Otolaryngol. 2009, 129, 501–506. [Google Scholar] [CrossRef] [PubMed]
  71. Liu, D.; Ma, X. Expression and significance of PTEN, P-ERK and P-AKT in the middle ear cholesteatoma. Lin Chuang Er Bi Yan Hou Tou Jing Wai Ke Za Zhi 2014, 28, 238–242, 245. [Google Scholar]
  72. Cai, Q.; Yang, H.S.; Li, Y.C.; Zhu, J. Dissecting the Roles of PDCD4 in Breast Cancer. Front. Oncol. 2022, 12, 855807. [Google Scholar] [CrossRef] [PubMed]
  73. Mao, X.H.; Chen, M.; Wang, Y.; Cui, P.G.; Liu, S.B.; Xu, Z.Y. MicroRNA-21 regulates the ERK/NF-κB signaling pathway to affect the proliferation, migration, and apoptosis of human melanoma A375 cells by targeting SPRY1, PDCD4, and PTEN. Mol. Carcinog. 2017, 56, 886–894. [Google Scholar] [CrossRef]
  74. Kim, J.Y.; Lee, H.; Kim, E.K.; Lee, W.M.; Hong, Y.O.; Hong, S.A. Low PDCD4 Expression Is Associated With Poor Prognosis of Colorectal Carcinoma. Appl. Immunohistochem. Mol. Morphol. 2021, 29, 685–692. [Google Scholar] [CrossRef] [PubMed]
  75. Xie, S.; Xiang, Y.; Wang, X.; Ren, H.; Yin, T.; Ren, J.; Liu, W. Acquired cholesteatoma epithelial hyperproliferation: Roles of cell proliferation signal pathways. Laryngoscope 2016, 126, 1923–1930. [Google Scholar] [CrossRef]
  76. Friedland, D.R.; Eernisse, R.; Erbe, C.; Gupta, N.; Cioffi, J.A. MicroRNA regulation of cholesteatoma growth microrna regulation of cholesteatoma growth. Laryngoscope 2009, 119, S114. [Google Scholar] [CrossRef]
  77. Orobello, N.; Harrington, C.; Reilly, B.K. Updates in paediatric cholesteatoma. Curr. Opin. Otolaryngol. Head Neck Surg. 2022, 30, 422–425. [Google Scholar] [CrossRef] [PubMed]
  78. 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]
  79. Li, W.; Liu, J.Q.; Chen, M.; Xu, J.; Zhu, D. Circular RNA in cancer development and immune regulation. J. Cell Mol. Med. 2022, 26, 1785–1798. [Google Scholar] [CrossRef]
  80. Altesha, M.A.; Ni, T.; Khan, A.; Liu, K.; Zheng, X. Circular RNA in cardiovascular disease. J. Cell Physiol. 2019, 234, 5588–5600. [Google Scholar] [CrossRef]
  81. Jin, J.; Sun, H.; Shi, C.; Yang, H.; Wu, Y.; Li, W.; Dong, Y.H.; Cai, L.; Meng, X.M. Circular RNA in renal diseases. J. Cell Mol. Med. 2020, 24, 6523–6533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Wu, X.; Xiao, Y.; Ma, J.; Wang, A. Circular RNA: A novel potential biomarker for skin diseases. Pharmacol. Res. 2020, 158, 104841. [Google Scholar] [CrossRef] [PubMed]
  83. Zhang, Y.L.; Zhang, S.Y.; Hu, J.; Zhang, H.H. Circular RNA in Diabetes and its Complications. Zhongguo Yi Xue Ke Xue Yuan Xue Bao 2022, 44, 521–528. [Google Scholar] [CrossRef]
  84. Li, T.R.; Jia, Y.J.; Wang, Q.; Shao, X.Q.; Lv, R.J. Circular RNA: A new star in neurological diseases. Int. J. Neurosci. 2017, 127, 726–734. [Google Scholar] [CrossRef]
  85. Singh, M.; Dwibedy, S.L.L.; Biswal, S.R.; Muthuswamy, S.; Kumar, A.; Kumar, S. Circular RNA: A novel and potential regulator in pathophysiology of schizophrenia. Metab. Brain Dis. 2022, 37, 1309–1316. [Google Scholar] [CrossRef]
  86. 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] [Green Version]
  87. He, Y.; Sun, M.M.; Zhang, G.G.; Yang, J.; Chen, K.S.; Xu, W.W.; Li, B. Targeting PI3K/Akt signal transduction for cancer therapy. Signal Transduct Target Ther. 2021, 6, 425. [Google Scholar] [CrossRef] [PubMed]
  88. Stanciu, S.; Ionita-Radu, F.; Stefani, C.; Miricescu, D.; Stanescu, S., II; Greabu, M.; Ripszky Totan, A.; Jinga, M. Targeting PI3K/AKT/mTOR Signaling Pathway in Pancreatic Cancer: From Molecular to Clinical Aspects. Int. J. Mol. Sci. 2022, 23, 132. [Google Scholar] [CrossRef]
  89. Mao, X.B.; Sheng, T.; Zhuang, L.P.; Wu, C.K.; Zhao, G.G. Expression of miRNA let-7a and HMGA2 and Diagnostic Value of Serum miRNA let-7a Level in Pancreatic Cancer. Sichuan Da Xue Xue Bao Yi Xue Ban 2020, 51, 540–545. [Google Scholar] [CrossRef]
  90. 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]
  91. Chirshev, E.; Oberg, K.C.; Ioffe, Y.J.; Unternaehrer, J.J. Let-7 as biomarker, prognostic indicator, and therapy for precision medicine in cancer. Clin. Transl. Med. 2019, 8, 24. [Google Scholar] [CrossRef] [Green Version]
  92. Bernstein, D.L.; Jiang, X.; Rom, S. let-7 microRNAs: Their Role in Cerebral and Cardiovascular Diseases, Inflammation, Cancer, and Their Regulation. Biomedicines 2021, 9, 606. [Google Scholar] [CrossRef]
  93. Yin, H.; Sun, Y.; Wang, X.; Park, J.; Zhang, Y.; Li, M.; Yin, J.; Liu, Q.; Wei, M. Progress on the relationship between miR-125 family and tumorigenesis. Exp. Cell Res. 2015, 339, 252–260. [Google Scholar] [CrossRef] [PubMed]
  94. Wang, Q.W.; Sun, Y.N.; Tan, L.J.; Zhao, J.N.; Zhou, X.J.; Yu, T.J.; Liu, J.T. MiR-125 family improves the radiosensitivity of head and neck squamous cell carcinoma. Mol. Biol. Rep. 2023, 50, 5307–5317. [Google Scholar] [CrossRef] [PubMed]
  95. Wang, M.J.; Xu, Y.Y.; Huang, R.Y.; Chen, X.M.; Chen, H.M.; Han, L.; Yan, Y.H.; Lu, C.J. Role of an imbalanced miRNAs axis in pathogenesis of psoriasis: Novel perspectives based on review of the literature. Oncotarget 2017, 8, 5498–5507. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Xu, N.; Brodin, P.; Wei, T.; Meisgen, F.; Eidsmo, L.; Nagy, N.; Kemeny, L.; Ståhle, M.; Sonkoly, E.; Pivarcsi, A. MiR-125b, a microRNA downregulated in psoriasis, modulates keratinocyte proliferation by targeting FGFR2. J. Investig. Dermatol. 2011, 131, 1521–1529. [Google Scholar] [CrossRef] [PubMed]
  97. Chan, K.S.; Sano, S.; Kiguchi, K.; Anders, J.; Komazawa, N.; Takeda, J.; DiGiovanni, J. Disruption of Stat3 reveals a critical role in both the initiation and the promotion stages of epithelial carcinogenesis. J. Clin. Investig. 2004, 114, 720–728. [Google Scholar] [CrossRef] [Green Version]
  98. Frank, D.A. STAT3 as a central mediator of neoplastic cellular transformation. Cancer Lett. 2007, 251, 199–210. [Google Scholar] [CrossRef]
  99. Xu, Y.; Xu, X.; Gao, X.; Chen, H.; Geng, L. Shikonin suppresses IL-17-induced VEGF expression via blockage of JAK2/STAT3 pathway. Int. Immunopharmacol. 2014, 19, 327–333. [Google Scholar] [CrossRef]
  100. Ho, K.Y.; Huang, H.H.; Hung, K.F.; Chen, J.C.; Chai, C.Y.; Chen, W.T.; Tsai, S.M.; Chien, C.Y.; Wang, H.M.; Wu, Y.J. Cholesteatoma growth and proliferation: Relevance with serpin B3. Laryngoscope 2012, 122, 2818–2823. [Google Scholar] [CrossRef]
  101. Liu, W.; Xie, S.; Chen, X.; Rao, X.; Ren, H.; Hu, B.; Yin, T.; Xiang, Y.; Ren, J. Activation of the IL-6/JAK/STAT3 signaling pathway in human middle ear cholesteatoma epithelium. Int. J. Clin. Exp. Pathol. 2014, 7, 709–715. [Google Scholar] [PubMed]
  102. Park, H.R.; Min, S.K.; Min, K.; Jun, S.Y.; Seo, J.; Kim, H.J. Increased expression of p63 and survivin in cholesteatomas. Acta Otolaryngol. 2009, 129, 268–272. [Google Scholar] [CrossRef] [PubMed]
  103. Hamajima, Y.; Komori, M.; Preciado, D.A.; Choo, D.I.; Moribe, K.; Murakami, S.; Ondrey, F.G.; Lin, J. The role of inhibitor of DNA-binding (Id1) in hyperproliferation of keratinocytes: The pathological basis for middle ear cholesteatoma from chronic otitis media. Cell Prolif. 2010, 43, 457–463. [Google Scholar] [CrossRef] [PubMed]
  104. Shen, D.; Zhao, H.Y.; Gu, A.D.; Wu, Y.W.; Weng, Y.H.; Li, S.J.; Song, J.Y.; Gu, X.F.; Qiu, J.; Zhao, W. miRNA-10a-5p inhibits cell metastasis in hepatocellular carcinoma via targeting SKA1. Kaohsiung J. Med. Sci. 2021, 37, 784–794. [Google Scholar] [CrossRef]
  105. Zhu, H.; Kang, M.; Bai, X. TCF21 regulates miR-10a-5p/LIN28B signaling to block the proliferation and invasion of melanoma cells. PLoS ONE 2021, 16, e0255971. [Google Scholar] [CrossRef] [PubMed]
  106. Liu, L.J.; Sun, X.Y.; Yang, C.X.; Zou, X.Y. MiR-10a-5p restrains the aggressive phenotypes of ovarian cancer cells by inhibiting HOXA1. Kaohsiung J. Med. Sci. 2021, 37, 276–285. [Google Scholar] [CrossRef]
  107. Vaher, H.; Runnel, T.; Urgard, E.; Aab, A.; Carreras Badosa, G.; Maslovskaja, J.; Abram, K.; Raam, L.; Kaldvee, B.; Annilo, T.; et al. miR-10a-5p is increased in atopic dermatitis and has capacity to inhibit keratinocyte proliferation. Allergy 2019, 74, 2146–2156. [Google Scholar] [CrossRef]
  108. Gao, T.; Zou, M.; Shen, T.; Duan, S. Dysfunction of miR-802 in tumors. J. Clin. Lab. Anal. 2021, 35, e23989. [Google Scholar] [CrossRef]
  109. Sun, D.; Chen, J.; Wu, W.; Tang, J.; Luo, L.; Zhang, K.; Jin, L.; Lin, S.; Gao, Y.; Yan, X.; et al. MiR-802 causes nephropathy by suppressing NF-κB-repressing factor in obese mice and human. J. Cell Mol. Med. 2019, 23, 2863–2871. [Google Scholar] [CrossRef] [Green Version]
  110. Zhang, F.; Ma, D.; Zhao, W.; Wang, D.; Liu, T.; Liu, Y.; Yang, Y.; Liu, Y.; Mu, J.; Li, B.; et al. Obesity-induced overexpression of miR-802 impairs insulin transcription and secretion. Nat. Commun. 2020, 11, 1822. [Google Scholar] [CrossRef] [Green Version]
  111. Liu, W.; Li, G.; Li, J.; Chen, W. Long noncoding RNA TRG-AS1 protects against glucocorticoid-induced osteoporosis in a rat model by regulating miR-802-mediated CAB39/AMPK/SIRT-1/NF-κB axis. Hum. Cell 2022, 35, 1424–1439. [Google Scholar] [CrossRef]
  112. Yao, J.; Gao, R.; Luo, M.; Li, D.; Guo, L.; Yu, Z.; Xiong, F.; Wei, C.; Wu, B.; Xu, Z.; et al. miR-802 participates in the inflammatory process of inflammatory bowel disease by suppressing SOCS5. Biosci. Rep. 2020, 40, BSR20192257. [Google Scholar] [CrossRef] [Green Version]
  113. Wang, Y.; Liu, X.; Wang, L.; Zhang, Z.; Li, Z.; Li, M. Circ_PGPEP1 Serves as a Sponge of miR-1297 to Promote Gastric Cancer Progression via Regulating E2F3. Dig. Dis. Sci. 2021, 66, 4302–4313. [Google Scholar] [CrossRef]
  114. Zhang, X.; Zhang, M.; Guo, Q.; Hu, X.; Zhao, Z.; Ni, L.; Liu, L.; Wang, X.; Wang, Z.; Tong, D.; et al. MicroRNA-1297 inhibits proliferation and promotes apoptosis in gastric cancer cells by downregulating CDC6 expression. Anticancer Drugs 2019, 30, 803–811. [Google Scholar] [CrossRef]
  115. Liang, L.; Feng, L.; Wei, B. microRNA-1297 involves in the progression of oral squamous cell carcinoma through PTEN. Saudi J. Biol. Sci. 2018, 25, 923–927. [Google Scholar] [CrossRef] [PubMed]
  116. Liu, F.; He, Y.; Shu, R.; Wang, S. MicroRNA-1297 regulates hepatocellular carcinoma cell proliferation and apoptosis by targeting EZH2. Int. J. Clin. Exp. Pathol. 2015, 8, 4972–4980. [Google Scholar]
  117. Wang, Y.; Xue, J.; Kuang, H.; Zhou, X.; Liao, L.; Yin, F. microRNA-1297 Inhibits the Growth and Metastasis of Colorectal Cancer by Suppressing Cyclin D2 Expression. DNA Cell Biol. 2017, 36, 991–999. [Google Scholar] [CrossRef]
  118. Park, C.R.; Lee, M.; Lee, S.Y.; Kang, D.; Park, S.J.; Lee, D.C.; Koo, H.; Park, Y.G.; Yu, S.L.; Jeong, I.B.; et al. Regulating POLR3G by MicroRNA-26a-5p as a promising therapeutic target of lung cancer stemness and chemosensitivity. Noncoding RNA Res. 2023, 8, 273–281. [Google Scholar] [CrossRef]
  119. Li, M.; Xiao, Y.; Liu, M.; Ning, Q.; Xiang, Z.; Zheng, X.; Tang, S.; Mo, Z. MiR-26a-5p regulates proliferation, apoptosis, migration and invasion via inhibiting hydroxysteroid dehydrogenase like-2 in cervical cancer cell. BMC Cancer 2022, 22, 876. [Google Scholar] [CrossRef]
  120. Chen, X.; Wu, G.; Qing, J.; Li, C.; Chen, X.; Shen, J. LINC00240 knockdown inhibits nasopharyngeal carcinoma progress by targeting miR-26a-5p. J. Clin. Lab. Anal. 2022, 36, e24424. [Google Scholar] [CrossRef] [PubMed]
  121. Cai, B.; Qu, X.; Kan, D.; Luo, Y. miR-26a-5p suppresses nasopharyngeal carcinoma progression by inhibiting PTGS2 expression. Cell Cycle 2022, 21, 618–629. [Google Scholar] [CrossRef] [PubMed]
  122. Chung, Y.H.; Cheng, Y.T.; Kao, Y.H.; Tsai, W.C.; Huang, G.K.; Chen, Y.T.; Shen, Y.C.; Tai, M.H.; Chiang, P.H. MiR-26a-5p as a useful therapeutic target for upper tract urothelial carcinoma by regulating WNT5A/β-catenin signaling. Sci. Rep. 2022, 12, 6955. [Google Scholar] [CrossRef] [PubMed]
  123. Wang, Z.; He, S.; Guo, P.; Guo, X.; Zheng, J. MicroRNA-1297 inhibits metastasis and epithelial-mesenchymal transition by targeting AEG-1 in cervical cancer. Oncol. Rep. 2017, 38, 3121–3129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Pan, X.; Li, H.; Tan, J.; Weng, X.; Zhou, L.; Weng, Y.; Cao, X. miR-1297 Suppresses Osteosarcoma Proliferation and Aerobic Glycolysis by Regulating PFKFB2. Onco. Targets Ther. 2020, 13, 11265–11275. [Google Scholar] [CrossRef]
  125. Zhao, J.; Jiang, Q.; Xu, C.; Jia, Q.; Wang, H.; Xue, W.; Wang, Y.; Zhu, Z.; Tian, L. MiR-26a-5p from HucMSC-derived extracellular vesicles inhibits epithelial mesenchymal transition by targeting Adam17 in silica-induced lung fibrosis. Ecotoxicol. Environ. Saf. 2023, 257, 114950. [Google Scholar] [CrossRef]
  126. Xie, T.; Pei, Y.; Shan, P.; Xiao, Q.; Zhou, F.; Huang, L.; Wang, S. Identification of miRNA-mRNA Pairs in the Alzheimer’s Disease Expression Profile and Explore the Effect of miR-26a-5p/PTGS2 on Amyloid-β Induced Neurotoxicity in Alzheimer’s Disease Cell Model. Front. Aging Neurosci. 2022, 14, 909222. [Google Scholar] [CrossRef]
  127. Balasubramanian, S.; Adhikary, G.; Eckert, R.L. The Bmi-1 polycomb protein antagonizes the (-)-epigallocatechin-3-gallate-dependent suppression of skin cancer cell survival. Carcinogenesis 2010, 31, 496–503. [Google Scholar] [CrossRef] [Green Version]
  128. Ma, N.; Zhao, S.; Yang, W.; Wang, Y. B-cell-specific Moloney murine leukemia virus integration site 1 knockdown impairs adriamycin resistance of gastric cancer cells. Arab J. Gastroenterol. 2023. [Google Scholar] [CrossRef]
  129. Liu, J.Y.; Jiang, Y.N.; Huang, H.; Xu, J.F.; Wu, Y.H.; Wang, Q.; Zhu, Y.; Zheng, B.; Shen, C.; Qian, W.F.; et al. BMI-1 promotes breast cancer proliferation and metastasis through different mechanisms in different subtypes. Cancer Sci. 2023, 114, 449–462. [Google Scholar] [CrossRef]
  130. Liu, S.; Yang, Y.; Chen, L.; Liu, D.; Dong, H. MicroRNA-154 functions as a tumor suppressor in non-small cell lung cancer through directly targeting B-cell-specific Moloney murine leukemia virus insertion site 1. Oncol. Lett. 2018, 15, 10098–10104. [Google Scholar] [CrossRef] [Green Version]
  131. Li, J.; Wang, Y.; Ge, J.; Li, W.; Yin, L.; Zhao, Z.; Liu, S.; Qin, H.; Yang, J.; Wang, L.; et al. Doublecortin-Like Kinase 1 (DCLK1) Regulates B Cell-Specific Moloney Murine Leukemia Virus Insertion Site 1 (Bmi-1) and is Associated with Metastasis and Prognosis in Pancreatic Cancer. Cell Physiol. Biochem. 2018, 51, 262–277. [Google Scholar] [CrossRef] [PubMed]
  132. Espersen, M.L.; Linnemann, D.; Christensen, I.J.; Alamili, M.; Troelsen, J.T.; Høgdall, E. The prognostic value of polycomb group protein B-cell-specific moloney murine leukemia virus insertion site 1 in stage II colon cancer patients. Apmis 2016, 124, 541–546. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Abobaker, S.; Kulbe, H.; Taube, E.T.; Darb-Esfahani, S.; Richter, R.; Denkert, C.; Jank, P.; Sehouli, J.; Braicu, E.I. Polycomb Protein BMI-1 as a Potential Therapeutic Target in Mucinous Ovarian Cancer. Anticancer Res. 2022, 42, 1739–1747. [Google Scholar] [CrossRef] [PubMed]
  134. Lee, K.; Adhikary, G.; Balasubramanian, S.; Gopalakrishnan, R.; McCormick, T.; Dimri, G.P.; Eckert, R.L.; Rorke, E.A. Expression of Bmi-1 in epidermis enhances cell survival by altering cell cycle regulatory protein expression and inhibiting apoptosis. J. Invest. Dermatol. 2008, 128, 9–17. [Google Scholar] [CrossRef] [Green Version]
  135. Reinisch, C.M.; Uthman, A.; Erovic, B.M.; Pammer, J. Expression of BMI-1 in normal skin and inflammatory and neoplastic skin lesions. J. Cutan Pathol. 2007, 34, 174–180. [Google Scholar] [CrossRef]
  136. Wang, S.L.; Dong, X.W.; Zhao, F.; Li, C.X. MiR-203 inhibits cell proliferation, invasion, and migration of ovarian cancer through regulating RGS17. J. Biol. Regul. Homeost. Agents 2021, 35, 1109–1115. [Google Scholar] [CrossRef]
  137. Song, S.; Johnson, K.S.; Lujan, H.; Pradhan, S.H.; Sayes, C.M.; Taube, J.H. Nanoliposomal Delivery of MicroRNA-203 Suppresses Migration of Triple-Negative Breast Cancer through Distinct Target Suppression. Noncoding. RNA 2021, 7, 45. [Google Scholar] [CrossRef]
  138. Altan, Z.; Sahin, Y. miR-203 suppresses pancreatic cancer cell proliferation and migration by modulating DUSP5 expression. Mol. Cell Probes. 2022, 66, 101866. [Google Scholar] [CrossRef]
  139. Yi, R.; Poy, M.N.; Stoffel, M.; Fuchs, E. A skin microRNA promotes differentiation by repressing ‘stemness’. Nature 2008, 452, 225–229. [Google Scholar] [CrossRef] [Green Version]
  140. Kim, Y.J.; Lee, S.B.; Lee, H.B. Oleic acid enhances keratinocytes differentiation via the upregulation of miR-203 in human epidermal keratinocytes. J. Cosmet. Dermatol. 2019, 18, 383–389. [Google Scholar] [CrossRef] [Green Version]
  141. Liu, W.; Yin, T.; Ren, J.; Li, L.; Xiao, Z.; Chen, X.; Xie, D. Activation of the EGFR/Akt/NF-κB/cyclinD1 survival signaling pathway in human cholesteatoma epithelium. Eur. Arch. Otorhinolaryngol. 2014, 271, 265–273. [Google Scholar] [CrossRef]
  142. Huisman, M.A.; De Heer, E.; Grote, J.J. Survival signaling and terminal differentiation in cholesteatoma epithelium. Acta Otolaryngol. 2007, 127, 424–429. [Google Scholar] [CrossRef]
  143. Wang, Q.; Ye, B.; Wang, P.; Yao, F.; Zhang, C.; Yu, G. Overview of microRNA-199a Regulation in Cancer. Cancer Manag. Res. 2019, 11, 10327–10335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Zahra, M.A.; Kamha, E.S.; Abdelaziz, H.K.; Nounou, H.A.; Deeb, H.M.E. Aberrant Expression of Serum MicroRNA-153 and -199a in Generalized Epilepsy and its Correlation with Drug Resistance. Ann. Neurosci. 2022, 29, 203–208. [Google Scholar] [CrossRef] [PubMed]
  145. Tang, J.; Yu, H.; Wang, Y.; Duan, G.; Wang, B.; Li, W.; Zhu, Z. microRNA-199a counteracts glucocorticoid inhibition of bone marrow mesenchymal stem cell osteogenic differentiation through regulation of Klotho expression in vitro. Cell Biol. Int. 2020, 44, 2532–2540. [Google Scholar] [CrossRef]
  146. Sharma, S. Immunomodulation: A definitive role of microRNA-142. Dev. Comp. Immunol. 2017, 77, 150–156. [Google Scholar] [CrossRef]
  147. Berrien-Elliott, M.M.; Sun, Y.; Neal, C.; Ireland, A.; Trissal, M.C.; Sullivan, R.P.; Wagner, J.A.; Leong, J.W.; Wong, P.; Mah-Som, A.Y.; et al. MicroRNA-142 Is Critical for the Homeostasis and Function of Type 1 Innate Lymphoid Cells. Immunity 2019, 51, 479–490.e476. [Google Scholar] [CrossRef] [PubMed]
  148. Zhou, C.; Wang, P.; Lei, L.; Huang, Y.; Wu, Y. Overexpression of miR-142-5p inhibits the progression of nonalcoholic steatohepatitis by targeting TSLP and inhibiting JAK-STAT signaling pathway. Aging 2020, 12, 9066–9084. [Google Scholar] [CrossRef]
  149. Shrestha, A.; Mukhametshina, R.T.; Taghizadeh, S.; Vásquez-Pacheco, E.; Cabrera-Fuentes, H.; Rizvanov, A.; Mari, B.; Carraro, G.; Bellusci, S. MicroRNA-142 is a multifaceted regulator in organogenesis, homeostasis, and disease. Dev. Dyn. 2017, 246, 285–290. [Google Scholar] [CrossRef] [Green Version]
  150. Yao, R.; Xu, L.; Wei, B.; Qian, Z.; Wang, J.; Hui, H.; Sun, Y. miR-142-5p regulates pancreatic cancer cell proliferation and apoptosis by regulation of RAP1A. Pathol. Res. Pract. 2019, 215, 152416. [Google Scholar] [CrossRef]
  151. Yu, W.; Li, D.; Zhang, Y.; Li, C.; Zhang, C.; Wang, L. MiR-142-5p Acts as a Significant Regulator Through Promoting Proliferation, Invasion, and Migration in Breast Cancer Modulated by Targeting SORBS1. Technol. Cancer Res. Treat 2019, 18, 1533033819892264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Yan, J.; Yang, B.; Lin, S.; Xing, R.; Lu, Y. Downregulation of miR-142-5p promotes tumor metastasis through directly regulating CYR61 expression in gastric cancer. Gastric. Cancer 2019, 22, 302–313. [Google Scholar] [CrossRef] [Green Version]
  153. Cheng, D.; Li, J.; Zhang, L.; Hu, L. miR-142-5p suppresses proliferation and promotes apoptosis of human osteosarcoma cell line, HOS, by targeting PLA2G16 through the ERK1/2 signaling pathway. Oncol. Lett. 2019, 17, 1363–1371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Wei, G.; Yuan, Y.; He, X.; Jin, L.; Jin, D. Enhanced plasma miR-142-5p promotes the progression of intrahepatic cholangiocarcinoma via targeting PTEN. Exp. Ther. Med. 2019, 17, 4190–4196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. NavaneethaKrishnan, S.; Rosales, J.L.; Lee, K.Y. Loss of Cdk5 in breast cancer cells promotes ROS-mediated cell death through dysregulation of the mitochondrial permeability transition pore. Oncogene 2018, 37, 1788–1804. [Google Scholar] [CrossRef] [Green Version]
  156. Wang, Q.; Yin, R.; Chen, X.; Hu, B.; Jiang, B.; Tang, W.; Zhang, X.; Jin, X.; Ying, M.; Fu, J. Higher Levels of Tumour-Infiltrating Lymphocytes (TILs) are Associated with a Better Prognosis, While CDK5 Plays a Different Role Between Nonmetastatic and Metastatic Colonic Carcinoma. Cancer Control 2023, 30, 10732748231169396. [Google Scholar] [CrossRef]
  157. Ling, R.; Sheng, Y.; Hu, Y.; Wang, D.; Zhou, Y.; Shu, Y. Comprehensive analysis of CDK5 as a novel biomarker for progression in esophageal cancer. Esophagus 2023, 20, 502–514. [Google Scholar] [CrossRef]
  158. Shao, X.; Yang, Y.; Chen, J.; Zhao, R.; Xu, L.; Guo, X.; Feng, Y.; Qin, L. Identification of Two CDK5R1-Related Subtypes and Characterization of Immune Infiltrates in Alzheimer’s Disease Based on an Integrated Bioinformatics Analysis. Comput. Math Methods Med. 2022, 2022, 6766460. [Google Scholar] [CrossRef]
  159. Batra, S.; Jahan, S.; Ashraf, A.; Alharby, B.; Jawaid, T.; Islam, A.; Hassan, I. A review on cyclin-dependent kinase 5: An emerging drug target for neurodegenerative diseases. Int. J. Biol. Macromol. 2023, 230, 123259. [Google Scholar] [CrossRef]
  160. Liu, C.C.; Zhang, H.L.; Zhi, L.L.; Jin, P.; Zhao, L.; Li, T.; Zhou, X.M.; Sun, D.S.; Cheng, G.H.; Xin, Q.; et al. CDK5 Regulates PD-L1 Expression and Cell Maturation in Dendritic Cells of CRSwNP. Inflammation 2019, 42, 135–144. [Google Scholar] [CrossRef]
  161. Zhang, D.; Lee, H.; Wang, X.; Groot, M.; Sharma, L.; Dela Cruz, C.S.; Jin, Y. A potential role of microvesicle-containing miR-223/142 in lung inflammation. Thorax 2019, 74, 865–874. [Google Scholar] [CrossRef] [PubMed]
  162. Ji, Y.; Fang, Q.Y.; Wang, S.N.; Zhang, Z.W.; Hou, Z.J.; Li, J.N.; Fu, S.Q. Lnc-RNA BLACAT1 regulates differentiation of bone marrow stromal stem cells by targeting miR-142-5p in osteoarthritis. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 2893–2901. [Google Scholar] [CrossRef] [PubMed]
  163. Rokavec, M.; Huang, Z.; Hermeking, H. Meta-analysis of miR-34 target mRNAs using an integrative online application. Comput. Struct. Biotechnol. J. 2023, 21, 267–274. [Google Scholar] [CrossRef] [PubMed]
  164. Li, S.; Wei, X.; He, J.; Cao, Q.; Du, D.; Zhan, X.; Zeng, Y.; Yuan, S.; Sun, L. The comprehensive landscape of miR-34a in cancer research. Cancer Metastasis Rev. 2021, 40, 925–948. [Google Scholar] [CrossRef] [PubMed]
  165. Mehta, D.; Daudia, A.; Birchall, J.P.; Banerjee, A.R. The localization of matrix metalloproteinases-8 and -13 in cholesteatoma, deep-meatal and post-auricular skin: A comparative analysis. Acta Otolaryngol. 2007, 127, 138–142. [Google Scholar] [CrossRef]
  166. Lee, R.J.; Mackenzie, I.C.; Hall, B.K.; Gantz, B.J. The nature of the epithelium in acquired cholesteatoma. Clin. Otolaryngol. Allied Sci. 1991, 16, 168–173. [Google Scholar] [CrossRef]
  167. Ferlito, S.; Fadda, G.; Lechien, J.R.; Cammaroto, G.; Bartel, R.; Borello, A.; Cavallo, G.; Piccinini, F.; La Mantia, I.; Cocuzza, S.; et al. Type 1 Tympanoplasty Outcomes between Cartilage and Temporal Fascia Grafts: A Long-Term Retrospective Study. J. Clin. Med. 2022, 11, 7000. [Google Scholar] [CrossRef]
  168. Ferlito, S.; La Mantia, I.; Merlino, F.; Cocuzza, S.; Di Stadio, A.; Cammaroto, G.; Bartel, R.; Fadda, G.; Iannella, G.; Mat, Q.; et al. Long-Term Anatomical and Hearing Outcomes of Canal Wall down Tympanoplasty for Tympano-Mastoid Cholesteatoma: A 20-Year Retrospective Study. Life 2022, 12, 1745. [Google Scholar] [CrossRef]
Figure 1. The PRISMA 2020 flow diagram for new systematic reviews, which included searches of databases and registers only.
Figure 1. The PRISMA 2020 flow diagram for new systematic reviews, which included searches of databases and registers only.
Ijms 24 12277 g001
Table 1. Databases used and corresponding search lines.
Table 1. Databases used and corresponding search lines.
DatabasesNumber of HitsSearch Lines
Scopus83TITLE-ABS-KEY ((microRNA) OR (miRNA)) AND (cholesteatoma)
Limited to
1. Document type: Article
2. Language: English
Web of Science18(AB = ((microRNA) OR (miRNA))) AND (AB = (cholesteatoma))
Pubmed/Medline17(“Cholesteatoma” [Mesh]) AND “MicroRNAs” [Mesh]
Cochrane0#1 “cholesteatoma” [Mesh]
#2 “microRNAs” [Mesh]
#1 AND #2
Table 2. Inclusion and exclusion criteria for the studies included in the review.
Table 2. Inclusion and exclusion criteria for the studies included in the review.
Inclusion CriteriaExclusion Criteria
TopicStudies concerning miRNA/cholesteatomaStudies not related to miRNA/cholesteatoma
Study typeOriginal articlesReviews, case reports, book chapters, expert opinions, letters to the editor, conference reports, posters
Study statusCompleted, publishedUnfinished, unpublished
LanguageFull text available in EnglishLanguage other than English or only abstract available in English
QualityGood-quality research studiesPoor-quality research studies
Table 3. Differentially expressed miRNAs in cholesteatoma tissue.
Table 3. Differentially expressed miRNAs in cholesteatoma tissue.
StudymiRNAExpression in
Cholesteatoma Tissue
Yang, J. [41]miR-10a-5pdownregulated
Zhu, X. [47]miR-1297,
miR-26a-5p
downregulated
Zang, J. [44]miR-125bdownregulated
Sui, R. [39]miR-142-5pdownregulated
Yao, L. [42]miR-199aupregulated
Liu, D. [38]miR-508-3pupregulated
Hu, Y. [35]miR-22-3p,
miR-125a-5p
upregulated
Gong, N. [34]exosomal miR-17downregulated
Zang, J. [43]miR-203adownregulated
Xie, S. [40]miR-21-3p,
miR-584-5p,
miR-16-1-3p,
miR-338-5p,
miR-320b,
miR-181a-3p,
miR-181a-5p,
miR-181b-5p,
miR-335-3p,
miR-155-5p,
miR-224-3p, etc.
upregulated
miR-10a-5p,
miR-152-5p,
miR-203b-5p,
miR-30a-5p,
miR-1297,
miR-539-3p,
miR-9-3p,
miR-769-3p, etc.
downregulated
Li, Y. [37]exosomal miR-106b-5pdownregulated
Chen, X. [31]miR-21, let-7a miRNAupregulated
Friedland, D.R. [33]miR-21upregulated
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

Dżaman, K.; Czerwaty, K.; Reichert, T.E.; Szczepański, M.J.; Ludwig, N. Expression and Regulatory Mechanisms of MicroRNA in Cholesteatoma: A Systematic Review. Int. J. Mol. Sci. 2023, 24, 12277. https://doi.org/10.3390/ijms241512277

AMA Style

Dżaman K, Czerwaty K, Reichert TE, Szczepański MJ, Ludwig N. Expression and Regulatory Mechanisms of MicroRNA in Cholesteatoma: A Systematic Review. International Journal of Molecular Sciences. 2023; 24(15):12277. https://doi.org/10.3390/ijms241512277

Chicago/Turabian Style

Dżaman, Karolina, Katarzyna Czerwaty, Torsten E. Reichert, Mirosław J. Szczepański, and Nils Ludwig. 2023. "Expression and Regulatory Mechanisms of MicroRNA in Cholesteatoma: A Systematic Review" International Journal of Molecular Sciences 24, no. 15: 12277. https://doi.org/10.3390/ijms241512277

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