Next Article in Journal
Liquid Biopsy with Detection of NRASQ61K Mutation in Cerebrospinal Fluid: An Alternative Tool for the Diagnosis of Primary Pediatric Leptomeningeal Melanoma
Next Article in Special Issue
Integrated Approaches to Identify miRNA Biomarkers Associated with Cognitive Dysfunction in Multiple Sclerosis Using Text Mining, Gene Expression, Pathways, and GWAS
Previous Article in Journal
AI-Based Chest CT Analysis for Rapid COVID-19 Diagnosis and Prognosis: A Practical Tool to Flag High-Risk Patients and Lower Healthcare Costs
Previous Article in Special Issue
Genes of the Glutamatergic System and Tardive Dyskinesia in Patients with Schizophrenia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

MiRNAs in Lung Cancer: Diagnostic, Prognostic, and Therapeutic Potential

1
Department of Biochemistry, Government Medical College (GMC-Srinagar), Karan Nagar, Srinagar 190010, J&K, India
2
Clinical Drug Development of Herbal Medicine, College of Pharmacy, Taipei Medical University, Taipei 11031, Taiwan
3
Department of Clinical Pharmacy, College of Pharmacy, King Saud University, Riyadh 11451, Saudi Arabia
4
Department of Pharmacology & Toxicology, College of Pharmacy Girls Section, Prince Sattam Bin Abdulaziz University, P.O. Box-173, Al-Kharj 11942, Saudi Arabia
5
Department of Pulmonary Medicine, Government Medical College (GMC-Srinagar), Karan Nagar, Srinagar 190010, J&K, India
6
Department of Pharmaceutics, College of Pharmacy, King Saud University, Riyadh 11451, Saudi Arabia
7
Department of Pharmacy Practice, College of Pharmacy, AlMaarefa University, Ad Diriyah, Riyadh 13713, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Diagnostics 2022, 12(7), 1610; https://doi.org/10.3390/diagnostics12071610
Submission received: 7 March 2022 / Revised: 13 April 2022 / Accepted: 17 April 2022 / Published: 1 July 2022
(This article belongs to the Special Issue MicroRNA in Diagnosis and Prognosis)

Abstract

:
Lung cancer is the dominant emerging factor in cancer-related mortality around the globe. Therapeutic interventions for lung cancer are not up to par, mainly due to reoccurrence/relapse, chemoresistance, and late diagnosis. People are currently interested in miRNAs, which are small double-stranded (20–24 ribonucleotides) structures that regulate molecular targets (tumor suppressors, oncogenes) involved in tumorigeneses such as cell proliferation, apoptosis, metastasis, and angiogenesis via post-transcriptional regulation of mRNA. Many studies suggest the emerging role of miRNAs in lung cancer diagnostics, prognostics, and therapeutics. Therefore, it is necessary to intensely explore the miRNOME expression of lung tumors and the development of anti-cancer strategies. The current review focuses on the therapeutic, diagnostic, and prognostic potential of numerous miRNAs in lung cancer.

1. Introduction

Lung cancer is the second most prevalent cancer with an incidence rate of 11.4% (22,06,771 cases) [1,2] and is a leading cause of cancer-related mortalities worldwide with a rate of 18% (17,96,144 cases) [1,2,3]. The age-adjusted incidence rate per 1000 in south-central Asian countries is 9.4 in males and 3.4 in females [4]. Lung cancer progression is a complex and multistep process that leads to the sequential development of genetic and molecular defects. It begins with 9p and 3p chromosome loss and ends with cyclin D1 and E over-expression [5]. Small cell lung cancer (SCLC) is less prevalent (16.8%) but more lethal than non-small cell lung cancer (NSCLC), which is more prevalent (80.4%) and heterogeneous among the population [6]. Chemotherapy using cytotoxic drugs and radiotherapy is the standard treatment option for lung cancer. Later, precise therapeutic drugs are employed to target specific genetic aberrations in lung tumors, guided by tumor genomic profiling. For instance, epidermal growth factor receptor (EGFR) mutation was the first genetic aberration targeted by tyrosine kinase inhibitors (TKIs), such as erlotinib and gefitinib [7,8]. The other strategy is immunotherapy, which prevents immune system compromise to the tumor (immunological anergy) and stimulates the immune system against the tumor. The immune checkpoint proteins that cause immunological anergy include programmed cell death protein1 (PD1), programmed death ligand1 (PDL1), cytotoxic T lymphocyte-associated protein4 (CTLA-4), etc. For instance, the monoclonal antibody Ipilimumab is used against CTLA-4, enhancing T-cell activation against tumors [9].
The current diagnostic methods for lung cancer include radiologic investigations, such as chest X-ray (CXR), low dose computer tomography (LDCT), positron emission tomography and computer tomography (PET-CT), and pathological investigations like lung cancer biopsy, bronchoscopy, etc. [10,11]. However, several drawbacks are associated with existing diagnostic regimens, such as low CXR sensitivity and overdiagnosis of lung cancer associated with LDCT, which results in psychological stress and unnecessary treatments [12]. Due to the physiological variations in (18F)-fluoro-2-deoxy-D-glucose uptake, malignancies that are PET negative, such as SCLC, and carcinoid tumors [13], and non-malignant conditions that respond to PET-positive, like infection, and inflammation [14], confound PET-CT interpretation. Also, technical artefacts and quantitative errors cause misinterpretation of PET-CT results [15,16,17]. Lung bronchoscopy is an invasive diagnostic method, and recently individuals have been predisposed to respiratory infection after undergoing bronchoscopy for lung cancer confirmation [18].
The use of integrative therapeutic strategies for lung cancer, especially immunotherapy and molecular-targeted therapy, has significantly improved the survival of specific populations over the past 20 years. Nevertheless, the efficacy of existing lung cancer therapies is inadequate owing to the lack of effective diagnostic techniques and the development of drug resistance, which, together, restrict the increase in survival after treatment. Thus, there is a need to understand lung cancer biology and explore new effective biomarkers and treatment approaches to lower the burden of lung cancer [19]. MiRNAs are small 20–22 nucleotide single-stranded non-coding RNA molecules, discovered in 1993 [20]. Later, in 2004, Takamizawa et al. highlighted the relationship between miRNA expression and lung cancer [21]. The primary function of miRNA is to control gene expression, mostly at the post-transcriptional stage [22]. In lung cancer cells and tissue, miRNAs exhibit an altered pattern of expression [23,24]. MiRNAs may be pro-oncogenic or anti-oncogenic and can regulate cell invasion, proliferation, migration [25,26], cell viability [25], epithelial-to-mesenchymal transition (EMT) [27], metastasis [28], and therapeutic resistance [29] in lung cancer. For example, miRNAs such as miRNA-195 and miRNA-497 inhibit the progression and formation of colonies in lung cancer by increasing the expression of TGF-β [30]. At the same time, miRNA-196b-59 promotes the progression and formation of colonies in lung cancer by decreasing FAS expression [31]. MiRNA-143-3p promotes metastasis of lung cancer cells to the brain by stimulating N6-methyladenosine [32], while miRNA-192-5p prevents metastasis of lung cancer cells to bones by negatively regulating TRIM44 [33].
The currently emerging clinical investigation method for lung cancer is miRNA expression analysis in tumor cells and biofluids like blood, due to its minimally invasive nature and high potential for early lung cancer detection [24]. Since a single miRNA targets multiple gene products or pathways, i.e., showing pleiotropic property, it possesses more therapeutic intervention potential than a drug targeting a single protein-coding gene. It also seems that single target therapeutic intervention is more prone to acquired chemo-resistance [34]. Specific miRNAs, such as miRNA-432-5p, promote resistance against EGFR inhibitors [29], while miRNA-27b can cause lung cancer chemo-resistance to retreat by down-regulating Snail1-mediated suppression of EMT [35]. As a result, miRNAs may serve as potential biomarkers for lung cancer diagnosis, prognosis, and treatment [36,37]. Thus, miRNAs have the potential to be therapeutic targets in the future, and should be explored as such.

2. MiRNA Deregulation in Lung Cancer

Fluctuations in miRNA expression are the dominant consequence of miRNA deregulation in cancer [38]. The expression of different miRNAs has been found to be altered in various lung cancer types, such as NSCLC [39,40], lung carcinoids [41,42,43], lung adenocarcinoma, large cell neuroendocrine carcinoma, and squamous cell carcinoma [44] (Table 1).
The miRNA expression within a cell is controlled by genetic and epigenetic mechanisms (DNA methylation) that regulate the transcription of pri-miRNA, its biosynthesis machinery, or specific transcription factors related to its biosynthesis [147]. The transcription factors of miRNA-processing enzymes (DROSHA and DICER) either promote or repress their transcription, affecting miRNAs’ expressions. For example, transcription of the DROSHA gene was either activated by MYC or repressed by ADARB1 transcription factors [148,149]. MYC binding to DROSHA gene promoter accelerated DROSHA transcription, causing elevated miRNA processing in A549 lung cancer cells [148]. TAP63 transcription factor binding to the DICER promoter facilitates its transcription, and dysfunctional TAP63 was found in many cancers [150]. For example, reduced DICER expression was associated with the development of lung cancer [151]. In addition to DROSHA, dysfunctional miRNA silencing complex (miRISC) constituents, like AGO2 and TARBP2, also affect the gene silencing mechanism.
Unfortunately, a reasonable number of miRNA genes (50%) are present at fragile sites prone to deletion, amplification, or translocation in cancer, which becomes an essential factor for dysregulation of target mRNA that can initiate cancer progression as lung cancer [152,153]. Mutations of miRNA binding sites in the 3’UTR of oncogenic mRNAs or single nucleotide polymorphism (SNP) in the seed sequence of mature miRNAs also increased cancer risk, due to decreased target control. For example, genetic tumor profiling revealed that people with KRAS who have a SNP in the 3’ UTR are more likely to develop NSCLC. This may be because KRAS is now not controlled by the miRNA [154]. Hypoxic conditions within cancer cells stimulate EGFR to phosphorylate AGO2 at Tyr393, resulting in decreased AGO2 function that causes decreased DICER–AGO2 interaction, leading to decreased miRNA maturation and function [155]. Epigenetic modifications of chromatin within cancer cells also lead to miRNA dysregulation. MiRNA genes are controlled by epigenetic modulators like DNA methylation and histone modifications [156]. Epigenetic modification, like methylation of the miR-126 promoter sequence, decreases its expression, promoting lung carcinogenesis [157,158]. For example, it is found that promoters of miRNA genes containing CpG islands are heavily methylated in lung tumors, as in the case of miR-34a and miR-34b-miR-34c transcribed from chromosome 1p36 and chromosome 11q23, respectively [159].
Similarly, promoter methylation has been linked with downregulation of miR-200c that stimulates poor differentiation, poor E-cadherin expression, and lymph node metastasis in NSCLC [160]. Davalos et al. (2012) investigated the methylation status of CpG islands present in the regulatory sequences of miR-200c/141 and miR-200b/200a/429 and observed that they are heavily methylated in transformed lung cells, compared to healthy cells, which encourages EMT [161].
ZEB1and ZEB2 transcription activators promote gene expression of the mesenchymal phenotype and also discourage known EMT [87]. Histone modifications also have a vital role in miRNA dysregulation via miRNA gene transcription change, and therefore some miRNA genes were downregulated independently of hypermethylation. For example, decreased expression of miR-212 has been frequently linked with H3K9me2 and H3K27me3 histone modification in lung cancer cells [162]. In addition to epigenetic modifications, SNPs in promoter region and RNA editing affect miRNA biogenesis and function in NSCLC [163,164,165].

3. Tumor Enhancer miRNA

Tumor enhancer miRNA levels are generally elevated in tumors like conventional oncogenes. They are also referred to as oncomiRs, whose deregulation, usually over-expression, triggers carcinogenesis since they negatively regulate tumor suppressor genes, with or without affecting cell differentiation or the apoptosis transcriptome. Most solid tumors experience deregulated miRNA levels [166]. Here we have discussed some well-known oncomiRs (Table 2).

3.1. MiR-21

Mature miR-21 is encoded from chromosome 17q23.2, a conserved single gene locus [174,175]. Lung cancer patients demonstrated significantly higher miR-21 expression than normal controls [176]. The possible reasons for frequent over-expression of miRNA-21 force us to think about activating mutations in its regulatory sequence or amplifying its genomic locus. MiR-21 is a crucial protagonist in cell growth, proliferation, migration, invasion, and metastasis. The upregulation of miR-21, commonly in solid tumors [177,178,179], suppresses the expression of tumor suppresser genes like RECK, NFIB, TIMP3, TPM1, STAT3, etc. [60,61,62]. MiR-21 also has diverse targets of negative regulators of the cell growth (Spry1, Spry2, Btg2, and Pdcd4) pathway and apoptosis (Apaf1, Faslg, RhoB, and Pdcd4) pathway [63].
Anti-miR-21 treatment of the A549 lung cancer cell line inhibits proliferation [180]. It may discourage malignant NSCLC spreading to distant sites by releasing the PTEN tumor suppressor gene product, a direct target of miR-21 [181]. Interestingly, miR-21 over-expression was exaggerated in EGFR gene mutant cases and suppressed by treating EGFR tyrosine kinase inhibitors in never-smoking lung cancer patients [182]. The lung adenocarcinoma-derived cell line (pc-9) becomes Gefitinib-resistant when miR-21 is forcedly overexpressed [183], and miRNA upregulation is associated with the overall survival of lung cancer patients [184,185]. Another study found that miR-21 interferes with EMT by negatively regulating the expression of SMAD7, which is a crucial inhibitor of TNFα [186]. MiR-21 deletion/knockout sensitizes cells to DNA-damaging chemotherapy, which suggests inhibition of miR-21 by anti-miRNA agents improves chemotherapeutic action in lung cancer [63]. MiR-21 is an essential anti-apoptotic miRNA, positively regulated by EGFR cell signaling. Interestingly, miRNA controls protein-coding mRNA, and long non-coding RNA (LncRNA), like GAS5 controls the miRNA expression. The in-vitro cell culture and xenograft mouse model study conducted by Ziqiang Zhang et al. observed a double-negative feedback loop in which GAS5 negatively regulates the expression of miR-21 [187]. Further investigation found that forced under-expression of GAS5 increases chemoresistance of the NSCLC cell line (H157, H460) cisplatin (DDP), since GAS5 sponges the miR-21, relieves PTEN mRNA from inhibition, and improves its expression within the cell [188].
Lung cancer patients exhibit increased levels of miRNA-21-5p in their blood circulation [54]. MiRNA-21-5p inhibits the expression of SMAD7, which may enhance lung cancer cell proliferation, migration, and invasion [189]. Inhibiting miRNA-21-5p can effectively prevent the progression of lung cancer [190]. Thus, miR-21 can be utilized as an effective non-invasive biomarker for the diagnosis and prognosis of lung cancer [176].

3.2. MiR-17-92

Since the miR-17-92 cluster is found to have a well-documented contribution to the development of malignant diseases and is involved in the development of multiple organs in mammals, it has been extensively studied [191]. The miR-17-92 cluster is found on chromosome 13(q31.3) within the third intron of the C13 or f25 gene [192], which was found to be a first tumor enhancer miRNA [193]. The miR-17-92 cluster is transcribed as a polycistronic precursor transcript containing tandem stem-loop structures trimmed by dicer enzyme to generate mature miR-17-92 family members: miR-17, miR-18a, miR-19a, miR-19b-1, miR-20a and miR-92a [194]. The leading oncogenic player of miR-19-72 was miR-19, which directly targets the PTEN gene, and thus membrane-bound PI 3,4,5 triphosphate is not dephosphorylated, leading to uncontrolled proliferative AKT signaling [195]. About 30 targets of the miR-7-92 cluster were found, especially tumor suppressor genes like PTEN, RB1, P53, IRF2 SPRY4, etc., validated by the luciferase enzyme assay in lung cancer. The over-expression of miR-17-5p and miR-18a are also correlated with shorter survival in lung cancer individuals [102,111,112,113].
Mir-17-5p shows substantial upregulation in patients with NSCLC compared to normal controls, signifying this miRNA’s tremendous clinical importance in NSCLC diagnosis [196]. In lung cancer patients, miR-18a levels were found to be positively correlated with clinicopathological characteristics, such as TNM (lung cancer staging), while being negatively correlated with a therapeutic response (radiotherapy) [197]. A p53-induced lncRNA called TP53TG1 is under-expressed in lung cancer cells and tissues. Further experiments found that it decreased miR-18a expression and increased cisplatin sensitivity in the lung cancer cell line(A549) [198]. MiR-19a is a critical oncogenic family member of the miR-17-92 cluster since it promotes proliferative EGFR signaling in lung cancer [108]. The forced over-expression of miR-19a promotes EMT [199], reduces survival in NSCLC patients [145], and increases gefitinib-resistance in NSCLC cells via directly targeting the c-MET gene, both in in vivo and in vitro experimental systems [109]. MiR-20a and E2F1 form a negative feedback loop augmented by c-MYC [200,201,202]. Qin et al. discovered that miR-20a is over-expressed in malignant lung cells and tissues while RUNX3, a downstream effector of the TNF-α pathway, is downregulated. MiR-20a inhibits TNF-α pathway activation by directly targeting RUNX3 [169]. The introduction of miR-92a into NSCLC cell lines has the same effect as the knockdown of the PTEN tumor suppressor gene. Further research found PTEN and RGS3 are the direct targets of miR-92a and miR-92a-enhanced cell survival, caspase-3 activity, cell proliferation, and metastasis of lung cancer cells [170,203]. Aarati et al. found that over-expressed miR-92a-2 was associated with poor prognosis and resistance to chemotherapy drugs and miR-147 and miR-574-5p [204]. The miR-17-92 complex might be a potential biomarker for diagnosing NSCLC [205].

3.3. MiRNA-221

MiR-221/222 is transcribed from the X chromosome (Xp11.3). Its target selectivity remained highly conserved in vertebrates during evolution as it possesses the same seed sequence in humans, mice, and rats [206]. When the April/TRAIL sensitized human lung cancer cell line (H460) is transfected with miR-221/222, it becomes resistant [92]. Culturing TRAIL-resistant cells in the presence of miR-221/222 inhibitors improved TRAIL sensitivity [92]. MiR-221/222 reduces the expression of cyclin-dependent kinase inhibitors, such as p27Kip1. A reduced expression of p27kip1 in NSCLC cells may be a reason for decreased sensitivity towards TRAIL-induced apoptosis. The coincidence of over-expression of miR-221/222 with TRAIL resistance in epithelial cancers [206,207,208] found additional targets of this miRNA, linked to cell proliferation and apoptosis, such as TIMP3 [207], PUMA [93], and PTEN that promotes proliferation in A549 cells [209]. MiR-221 and miR-222 are subjected to activation by MET through the c-JUN transcription factor [130]. Interestingly, miR-130a improves TRAIL sensitivity by downregulating miR-221/222, directly targeting MET gene expression [210]. MiR-221 expression and quantity in patient tissue or serum predict the extent of lymph node metastasis and disease progression in non-small cell lung cancer [211]. However, Sun et al. found that miR-221 negatively regulates the expression of HOTAIR LncRNA and promotes apoptosis of NSCLC cells [212].
MiR-221 expression has been elevated in patients with benign metastasizing leiomyoma (BML) lungs [213]. In NSCLC patients, miR-221 levels increase significantly in plasma, suggesting its potential as a therapeutic target or non-invasive marker for early diagnosis and screening of NSCLC [214]. Wang et al. (2020) also suggested miR-221 as one of the biomarkers for early detection of NSCLC [215].

4. Tumor Suppressor miRNA

On analyzing miRNA of malignant cells, some miRNAs are downregulated. These types of miRNAs are considered tumor suppressor genes (Table 3). They discourage the process of cancer progression by specifically targeting oncogenes and, in other cases, genes that play a role in cell differentiation or apoptosis. Several miRNAs are currently considered tumor suppressor genes, for example, miRNA let-7. Here, we discuss a few of them.

4.1. MiR-Let-7 Family

Reinhart first discovered the 21 nucleotide non-coding RNA (let 7) when he studied developmental timing in C. elegans [228]. Later, it was found that a low level of let-7 correlated with the shorter survival rate of lung cancer patients [21]. It has been found that the Let-7 family constitutes 11 members, among which six members are located in genetically fragile regions of the genome that have a high propensity to genetic alteration [152]. The causes of let-7 downregulation may be a direct epigenetic or genetic aberration in its gene or be due to the activity of repressors of its transcription [229]. Let-7 has a prominent role in diminishing the effects of proliferative, inflammatory, and anti-apoptotic pathways via downregulating their downstream effectors, such as KRAS, c-MYC, CDK6, HOXA9, TGFBR1, BCL-XL, and MAP4K3, thereby creating an anti-malignant environment within the cell [97]. Several Let-7 members are transcribed from genomic regions frequently found missing in lung cancer individuals, such as let-7a, let-7c, and let-7g [230]. MiR-let-7 acts as a tumor suppressor miRNA that downregulates KRAS expression [231].MiR-let-7 mimics injected into the genetically engineered NSCLC mouse model dramatically reduced its tumor area, size, and metastasis compared to the placebo group, suggesting it could be a specific therapeutic tool for lung cancer [232], and is associated with better prognostic value in lung cancer [233]. The transfection of Let-7c and miR-200c in A549 cells improved erlotinib sensitivity, but the same results were obtained when treated with Hh (hedgehog) signaling inhibitor (GDC-0449) [234]. The lowered expression of Let-7 and miR-17 is associated with self-renewal and proliferation by targeting MYC and CDKN1A, which lead to gefitinib-resistance in NSCLC patients [235]. It is found through both in vivo and invitro studies that any imbalance in LIN28 gene expression and Let-7 miRNA expression leads to chemoresistance or radio-resistance in NSCLC, since they interact in a double-negative feedback mechanism [136,236].
The Let-7 family shows potential as a non-invasive marker for cancer diagnosis [237]. Lower serum miR-let-7a expression has been strongly associated with poor prognosis and efficacy of radiotherapy in lung cancer brain metastasis [238]. Furthermore, miR-let-7a regulates DICER1 and may be a critical predictive marker of lung cancer brain metastasis [238]. MiR-let-7e is an essential constituent of the lncRNA SNHG4/let-7e/KDM3A/p21 pathway, which has been related to NSCLC development and is possibly one of the vital therapeutic targets for NSCLC [239].

4.2. MiR-200 Family

Human miRNA-200 is composed of five members, three of them (miR-200a, miR-200b, miR-429) are transcribed from chr12p13 as a polycistronic mRNA, and the rest of them (miR-200c, miR-141) are transcribed from the chr1p36 region [240]. The miR-200 family is commonly recognized as being negative regulators of EMT, since they directly target negative regulators of E-cadherin transcription, such as ZEB1 and ZEB2(zinc finger E-box-binding homeobox) [88,163,221]. Interestingly, miR-200 and ZEB1 form a double-negative feedback loop in which ZEB1 also acts as a transcriptional repressor of the miR-200 family and maintains homeostasis in cell migration, invasion, and EMT [163,241,242]. MiR-200 inhibits angiogenesis, which is required for tumor survival, by lowering vascular endothelial growth factor (VEGF) and vascular endothelial growth factor-receptor1 (VEGF-R1) mRNA levels [243,244]. MiR-200c upregulation may improve radiosensitivity in lung cancer individuals, since it directly targets oxidative stress response genes such as PRDX2, GAPB/Nrf2, and SESN1. This suggests that miR-200c can improve radiotherapy in lung cancer patients [245]. The miR-200/ZEB loop may serve as a prognostic factor for nintedanib sensitivity in malignant lung cells since nintedanib sensitivity is associated with upregulated miR-200a/b, miR-141, and E-cadherin levels and block TGF-β1-induced EMT [246]. Another study found Decitabine discourages TGF-β1-mediated abnormal methylation of the miR-200 gene, which decreases tumor cell migration [247]. A recent study conducted by Kim et al. found that miR-200 directly interacts with QKI (quaking homolog), a kind of STAR protein, and reduces its expression in malignant cells of the lung. MiR-200 and QKI interact in a so-called adverse feedback loop in which QKI knockdown reduces miR-200 expression [248].
Recently miR-200 has been reported to exhibit solid diagnostic ability in liquid biopsies of lung carcinoma [249]. The miR-200 family strongly suppresses the metastasis of lung adenocarcinoma. Decreased expression of miR-200 has been observed, especially in mouse lung adenocarcinoma metastasis, which is strongly related to poor survival of the patient. In a mouse model with lung adenocarcinoma (KrasLSL-G12D/+; Trp53flox/flox), metastasis has been reported drastically promoted by miR-200c/141 deletion, resulting in a desmoplastic tumor stroma that remarkably resembles human metastatic lung carcinoma. Deficiency of miR-200 in lung cancer cells activates neighboring cancer-associated fibroblasts and promotes their proliferation, which increases the capability of cancer cells for metastasis [250].

4.3. MiR-206

The genes of human miR-206 (hsa-miR-206) and miR-133 are transcribed from Chr 6p12.2 genomic landmarks adjacent to one another [67]. Hsa-mir-206 is a myomir, which means it is usually expressed in muscles, especially skeletal muscle tissue, and has a prominent role in myogenesis in humans [251]. MiR-206 is found downregulated in individuals with advanced lung cancer. Cell transfection studies have proven that miR-206 has a tumor-suppressive role, i.e., proapoptotic, antimetastatic, and antiangiogenic. MiR-206 directly interacts with the 3UTR of c-MET, EGFR, and Bcl2 in NSCLC cell lines, promoting apoptosis, but discouraging cell proliferation [216,217,218]. MiR-206 prevents the suppression of tumor angiogenesis, both in vivo and in vitro, by inhibiting the 14-3-3z/STAT3/HIF-1α/VEGF signaling pathway [252] and preventing normal fibroblast to cancer-associated fibroblast conversion by downregulating VEGFA expression [253]. MiR-206 restoration improved the cisplatin sensitivity in NSCLC cell lines and discouraged the potential of EMT, invasion, and migration by silencing MET gene expression, and repressing the PI3k/Akt/mTOR signaling pathway [254]. Further research found pentose pathway genes (G6PD, PGD, TKT, GPD2) to be direct targets of miR-206, which is the reason for suppression of the growth of H1437 and A549 cell lines. MiR-206 expression is decreased by over-expressing NRF2, leading to increased pentose phosphate pathway gene activity and redirection of carbon flux towards the pentose phosphate pathway and tricarboxylic acid cycle [255].
Ke-gang Jia et al. found that HK2 is the direct target of miR-206 and prevents explicit cancer cell proliferation by neutralizing the Warburg effect [256]. It is found that an LncRNA, called SNHG14, downregulates miR-206 by directly binding it, and a higher expression of SNHG14 promotes cell proliferation, invasion, and migration in NSCLC. Thus, the expression of SNHG14 indirectly encourages the expression of G6PD [257]. Further research found that miR-206 downregulates COROIC protein, a key player in proliferation, metastasis, and invasion of malignant lung cells (A549) and the A549 xenograft model. The same study also observed restoration of the miR-206 inhibitory effect by CORO1C gene knockdown [258].
Lower miR-206 expression decreased patient survival in NSCLC [259]. The LncRNA/WTI-AS/miR-206/NAMPT cluster, with miR-206 as an important component, could be a new key marker for lung adenocarcinoma diagnosis and prognosis [259]. In EGFR mutant lung cancer cells, miR-206 was shown to reduce HGF-induced gefitinib resistance [260]. MiR-206 affects EVI1 expression and activates the Akt/JNK pathway in SCLC to regulate stem cell proliferation and division [261]. MiR-206 is a metastatic tumor suppressor, and it may be used as a therapeutic target in the clinical treatment of NSCLC [257].

4.4. MiR-146 Family

The miR-146 family consists of two miRNA members (miR-146a-5p and miR-146b-5p). MiR-146a is transcribed from a long non-coding RNA gene located at chromosome 5(5q33.3), and miR-146b is transcribed from an intron located at chromosome 10(10q24.32) [262]. MiR-146a regulates various immune responses, such as antiviral, inflammatory, and innate immune responses [263]. The results of many studies suggest miRNA-146a is a potent anti-inflammatory miRNA that directly targets COX-2 [45] and FLAP protein [264]. Knockout of the miR-146a locus in mice makes them hypersensitive to bacterial attack [265]. Cell line-based studies found miR-146a mimic transfection results suggest their interaction with an IRAK1 target either enhanced (regulatory T-cell) or suppressed their expression (Breast cancer cell line) [46]. Many functional polymorphisms in the miR-146a gene have been found in human cancers, increasing the risk of cancer development [156,266,267]. A study conducted on Chinese non-smoker women found that miR-146a rs2910164(CG/GG/CC) polymorphism decreased the risk for lung cancer, since the target binding preference of miR-146a has changed and has now acquired the ability to target the 3UTR of the TRAF6 oncogene [47].
MiR-146a has a well-recognized prognostic value, as its expression has been linked to improved overall survival and response to chemotherapy (EGFR-TKI) and has antiproliferative, antimetastatic, and pro-apoptotic properties in NSCLC [86]. Later, Chaohui Wu et al. found that higher serum miRNA-19b and lower levels of miR-146a are associated with poor overall survival, chemoresistance, and advanced TNM in NSCLC individuals [52]. MiR-146a mimic transfection improved cisplatin sensitivity, promoted apoptosis, and inhibited metastasis in A549/DDP via stimulating the JNK2-p53-Bcl2 axis [268]. Another study found that over-expression of miR-146a, or knockdown of cyclin J (CCNJ), produces the same effect: improved sensitivity to cisplatin mediated through inhibition of cell cycle, cell viability, and motility, and promotion of apoptosis in cisplatin-resistant NSCLC cell lines (A549, SPC-A1). This suggests that miR-146a directly interacts with the 3UTR region and decreases the CCJN mRNA level. The absence of this interaction may be the reason for drug resistance to cisplatin in NSCLC. Thus, treating A549/DPP cells with miR-146a mimic could be a novel strategy for solving the cisplatin resistance in NSCLC patients [269].
In lung cancer cells, miR-146-p regulates the expression of claudin-12, which promotes cell survival, migration, and invasion, inhibiting apoptosis and activating signaling pathways, such as Wnt/β-catenin and PI3K/AKT/MAPK [270]. The expression of miR-146a in lung cancer changes and it may serve as biomarker and therapeutic target [271].

5. Therapeutic Potential of miRNAs in Lung Cancer

MiRNAs have multiple mRNA targets, and their functional abnormality may lead to pleiotropic effects and a syndrome of disorders, including cancer. Their clinical use as biomarkers and in diagnostics is rapidly emerging. Compared to small drug molecules and protein-based drugs, they have a broader range of targets and traverse the cell plasma membrane, while monoclonal antibodies-based drugs can only target cell surface receptors and circulating proteins. MiRNAs can fine-tune the expression of virtually any gene and its mRNA transcripts [272].

5.1. MiRNAs as Therapeutic Agents

MiRNAs have a well-recognized role in cancer, and many studies have proved their significance in therapeutics and chemoresistance in cancer. However, the first human trial was conducted on siRNA, which is similar to miRNA, in 2004, and in 2018 the first siRNA drug was approved [273]. The deregulation of miRNAs in cells can be managed by directly introducing miRNAs (restoration strategy) or modulating miRNA expression by therapeutic agents. In the restoration strategy, synthetic dsRNA structures, called miRNA mimics, are used to replace and restore the function of diminished tumor-suppressive miRNA. A synthetic oligonucleotide mimic of miR-34a packaged into liposomal nanoparticles was the first miRNA-based therapy preferentially utilized for cancer [274]. For instance, a combination of Let-7 and miR-34 mimic delivered to a Kras-Trp53 (Kras mutation-p53 deletion) NSCLC mouse model showed promising results, which improved further when complemented with EGFR inhibitor (erlotinib) [67,68]. With the power of genetic engineering and nanotechnology, novel, effective strategies have been developed to deliver miRNA mimics. For instance, a study conducted by Talekar et al. (2016) observed that delivery of wild-type p53 along with miR-125b mimics by dual CD44/EGFR-targeted hyaluronic acid (HA)-based nanoparticles promoted significant macrophage repolarization and stimulated apoptosis in a KP (Kras mutation-p53 deletion) mouse model and SK-LU-1 cells [275]. One study used cationic liposome/CL-pVAX-miR-143 complex (CL-pVAX-miR-143) to deliver miR-143 mimics, which inhibited tumor metastasis in an NSCLC mouse model [276]. Recently miRNA mimics have been delivered by human-derived extracellular vesicles like exosomes not to mouse models but to 3D microfluidic lung cancer models. For instance, an miR-497 mimic was delivered to a 3D lung cancer model [277] (Figure 1).

5.2. MiRNAs as Targets for Therapy

Nucleic acid-based biomedicines, such as oligonucleotides and miRNA sponges, use microRNAs as targets directly or indirectly to achieve therapeutic responses in lung cancer. MiRNA sponges are similar to LncRNA present within the cells, containing multiple binding sites for a specific miRNA. They function by trapping miRNA and preventing binding to its endogenous targets [278]. For instance, long non-coding RNA TUG1 (taurine upregulated gene 1) promotes chemosensitivity of platinum-based chemotherapy by blocking miR-211 function and relieving PTEN from inhibition [279]. Some other examples are anti-miRs antagomirs, LNA (locked nucleic acids), etc. They are designed to block the specific function of miRNA, and their backbone is modified (O2-C4 bridge) or tagged with specific functional groups (2-methoxyethyl) [280]. These modifications make them nuclease-proof and improve thermostability and target specificity [281]. A study conducted by Fu-Gang Dua et al. found that knockdown of miR-421, both in in vivo and in vitro experimental models, by AMO (antisense morpholino) improved paclitaxel sensitivity significantly [282]. The miRNA repression is also achieved by specific chemical agents that target miRNA biogenesis or discourage miRNA-target interaction. Some small inhibitory molecules may target and block miRNA interaction with the RISC complex and possess anti-tumor properties [278]. A cell line-based study conducted by Xigan He et al. found that Docetaxel (a semisynthetic analogue of paclitaxel) discourages proliferation through upregulation of miR-7 in NSCLC cell lines. However, the underlying mechanism is not known [283]. Small molecules also directly target microRNA secondary structure. They directly interact with miRNA precursors to stop Drosha or Dicer cleavage. It is found that aminoglycosides directly interact with RNA secondary structures. Recent findings regarding curcumin found it to have anti-lung cancer properties via downregulation of miR-21 and upregulation of onco-suppressive miR-192-5p and miR-215 [284,285]. Another similar study found that curcumin suppresses the metastasis of NSCLC by stimulating miR-206 expression and discouraging the mTOR signaling pathway [286].

6. MiRNAs as Potential Lung Cancer Biomarkers

MiRNAs may develop as biomarkers in coming years since they possess remarkable stability in various specimen types [287] and are resilient to extremes of pH and temperature, and they exhibit high specificity, reproducibility, and robustness in expression patterns [288]. MiRNAs may now be added to the panel of possible biomarkers due to their reduced size and the availability of assays that can accurately assess their level, such as qRT-PCR, microarray, and others [289].

6.1. MiRNAs as Diagnostic Biomarkers

Lung cancer diagnosis and prognosis can be precisely determined by analyzing several miRNA expressions levels simultaneously (miRNA signatures) [290,291]. Many studies on lung cancer patients unraveled many unique miRNA signatures that are useful for diagnosis, possessing better sensitivity and specificity [292,293,294]. Most studies conducted on serum or plasma found that circulating cell-free miRNA (cfmiRNAs) may be an ideal screening agent for early lung cancer diagnosis [295].
MiRNA expression studies may be exploited to recognize the unknown origin of metastatic tissue. For example, Rosenfeld conducted a miRNA analysis of 22 most common solid tumors and developed a 48-miRNA classifier to identify the origin of unknown primary cancers with 81% sensitivity [296]. MiRNA dysregulation may be perceived at any stage, starting from initiation to progression, allowing us to observe dynamic changes in real-time [297]. These discoveries raise hopes for minimally invasive and early lung cancer diagnosis by exploiting cell-free miRNA expression behavior (cfmiRNA).
Over-expression of miRNA-21 in sputum, a well-recognized EDGF-regulated anti-apoptotic factor, easily distinguished NSCLC individuals from cancer-free individuals [298]. Several unique miRNA signatures, having diagnostic or prognostic importance, were observed in sputum [299]. A metanalysis study conducted by Jipei Liao et al. on miRNA-based lung cancer diagnosis found an integrated panel of biomarkers consisting of both plasma miRNA and sputum miRNA significantly increase the sensitivity and specificity of a lung cancer diagnosis. On further investigation, the performance of the integrated panel of biomarkers was found to be independent of histology and stage of NSCLC, and patients’ age, sex, and ethnicity [300].
Since lung cancer is highly heterogenous and the miRNA species produced by each type are different, these factors can also help classify lung cancer subtypes. MiRNA profile was found to be specific to a histological subtype of cancer, and this property is exploited to discriminate among different histological subtypes of cancer. Many studies on miRNA proved robust in discriminating among lung cancer subtypes [24,301,302]. These miRNA-based clinical studies were of good diagnostic value since squamous cell carcinoma (SqCC) and adenocarcinoma (AD) originating from different cells of lung tissue require different treatment [303]. MiRNAs could also serve as a marker to differentiate primary lung tumors from lung metastases originating from other locations by analyzing the expression of some specific miRNAs. MiR-182 was most significantly over-expressed in primary lung tumors, while miR-126 was over-expressed in lung metastases originating from other tissues of the body [304] (Figure 1).
MiRNAs such as miR-23a and miR-let7i might be clinically valuable biomarkers for the diagnosis of NSCLC [305]. Similarly, miR-21-5p, miR-150, miR-210, and miR-1290 can be utilized as useful early diagnostic and prognostic biomarkers in NSCLC [306]. The panel of miRNAs, including miR-30a-3p, miR-30b-5p, miR-30c-5p, miR-34a-5p, and miR-4286a, might serve as novel biomarkers for the diagnosis and prognosis of lung cancer [307]. One more recent study reported 5 miRNA-based panels (hsa-miR-31, hsa-miR-34c, hsa-miR-196b, hsa-miR-653, and hsa-miR-891a) as potential biomarkers for the diagnosis and prognosis of lung cancer [308]. Likewise, Yu-Long Zhao et al. (2022) reported serum miR-205-5p as a new and useful diagnostic biomarker for lung cancer [167]. ]. MiR-3182 can be utilized as a potential diagnostic biomarker for lung cancer [309].
MiRNA signature utilization for cancer screening has reached the pre-clinical and clinical testing stages. The mir-Test is a pre-clinical effort utilizing miRNA for cancer diagnosis [310]. This test aims to formulate a sensitive, non-invasive method of detecting early lung cancer in a high-risk group (heavy smokers greater than 50 years of age). To establish ideal miRNA biomarkers for clinical use, such as early lung cancer diagnosis, a heavy bulk of miRNA-associated data, generated from miRNA studies conducted on a much larger sample size and a standard operating protocol with standardized platforms and data analysis methods, is essential. Markers established under such conditions may have a higher success rate for clinical adaptation (Figure 2).

Exosomal miRNAs as Diagnostic Biomarkers

Pan and Johnstone first discovered exosomes in 1983 during their sheep reticulocyte-related experiment [311]. Exosomes are the smallest vesicles among the extracellular vesicle group, whose diameter may range from 0.03 µm to 0.1 µm [312]. Exosomes act as cargo, transferring their biologically active components [313], such as miRNAs, mRNA, DNA fragments, etc. into targeted cells, playing the role of a cell-to-cell communicator, [314,315,316] like tumorigenesis, carcinogenesis, metastasis, and drug resistance, in cancer. Exosomes form a large proportion of circulating vesicles, from which cancer-associated exosomal miRNAs are purified with high sensitivity and specificity [317]. Tumor-derived exosomal miRNAs are novel diagnostic and predictive biomarkers for overall survival, cancer relapse, and drug resistance. Exosomal miRNAs are considered better biomarkers than non-exosomal miRNAs. Exosomes released by cancer cells harbor cancer-specific miRNAs, RNA, DNA fragments, and membrane proteins, promoting their cancer specificity [318]. Besides, exosomal miRNAs have excellent stability and anti-degradation ability [39].
Exosomal miRNAs are better diagnostic and prognostic biomarkers in lung cancer than non-exosomal circulating miRNAs. Exosomal miRNAs have been found significantly elevated in lung carcinoma patients compared to benign pulmonary diseases. For instance, the circulating blood exosomal miRNAs (miRNA-361-3p, miRNA-625) help in discriminating malignant lung lesions from benign lung lesions [79]. Another study for diagnosing and screening lung cancer found serum exosomal miRNAs (miRNA-200b-5p, miRNA-378, miRNA-502-5p, miRNA-629, miRNA-17, and miRNA-100) were significantly lower in pulmonary granuloma and healthy smokers compared to lung adenocarcinoma subjects [319]. Apart from this, a study by Munagala et al. found serum exosomal miRNAs for predicting lung cancer recurrence or relapse. Based on in vitro cell culture and animal models, 77 exosomal miRNAs were found dysregulated. Of these, 47 were upregulated, and 30 were downregulated. Mirna-21 and miRNA-155 showed significant upregulation in recurrent tumors compared to primary tumors [320]. Peripheral blood exosomes were shown to have 30 specific molecular markers. Thus, exosomes and associated molecules may provide a theoretical basis for determining biomarkers for diagnosing lung cancer at an early stage.

7. MiRNAs as Lung Cancer Prognostic Biomarkers

MiRNA analysis of tumors by microarray and qRT-PCR has been found helpful in predicting clinical outcomes, such as response to cancer treatment, cancer relapse, and overall survival [289]. ] Takamizawa et al. discovered for the first time that a reduced level of Let-7 is significantly correlated with worsened prognosis after curative resection in lung cancer subjects [21].

7.1. MiRNAs as Biomarkers of Survival

A meta-analysis by S. R. Lamichhane found that miR-21, miR- 155, miR-148a, miR-148b, and miR-let-7 are consistently up or downregulated in NSCLC and show significant prognostic potential in the diagnosis, treatment, and follow-up of NSCLC [321]. Another meta-analysis by Wendi Xiao et al. found downregulation of miRNA-26b, miRNA-381, miRNA-146α, miRNA-148α, miRNA-204, miRNA-374α, miRNA-638 and miRNA-148b) and upregulation of miRNA-125b, miRNA-21, miRNA-141, miRNA-200c, miRNA-197, miRNA-41, miRNA-370, miRNA-376α, miRNA-192 and miRNA-662 are consistently linked to poorer overall survival of lung cancer patients [322]. Further investigation revealed their significant correlation with the overall survival of patients [323]. A similar study on SCLC individuals found a triplet miRNA signature (miR-194, miR-608, and miR-9) expression profile in serum that can easily predict overall survival [324].

7.2. MiRNAs as Biomarkers of Response to Treatment

Circulating miRNAs, including miR-21, mir-126, and miR-513a, may act as predictive markers for platinum therapy response in NSCLC besides having potential for diagnosis [325]. Zhu et al. (2022) reported a higher expression of miR-1274a associated with poor prognosis, which might be utilized as a potential prognostic biomarker in NSCLC [326]. We can predict which individuals would respond better to targeted therapy or chemotherapy, and develop resistance to treatment, and so modify therapeutic strategy accordingly based on miRNA expression analysis. Resistance of cancers to radiotherapy treatment can also be predicted based on miRNA expression in cancer. When Ma et al. treated four malignant lung cells with increased doses of radiation, they found a 2.5-fold upregulation of miR-95 in the most radioresistant cell line [327]. A study conducted on advanced NSCLC individuals found a specific group of patients that responded well to high-dose radiation therapy. On further investigation, this group possesses a unique miRNA expression pattern compared to poor respondents [328] (Figure 2)

8. Challenges in Use of miRNAs as Theragnostic Agents

Many pre-analytical and analytical variables significantly affect the diagnostic and prognostic potential of miRNA expression in lung cancer. Some of them are inherent to miRNAs, such as the addition or deletion of nucleotides, SNPs, isomers, and GC content of miRNAs, which affect miRNA recovery, and change miRNA sequence, and, thus, significantly affect miRNA profiling. Other factors are method used for isolation, storage conditions, and type of profiling method used. It should be noted that cell-free miRNAs (cf-miRNAs) have a non-homogenous origin, i.e., released miRNAs originate from residing cells and endothelial cells. This effect masks the number and level of miRNAs liberated by tumor-derived cells in biofluids. Also, visceral organs (lungs, liver, and kidney) that experience an elevated blood hydrostatic pressure may be a possible source of liberation [329,330,331]. The heterogeneity of miRNA expression is another hurdle. For instance, both inflammation, which is a well-known cancer hallmark, and hypoxia are frequent conditions in the tumor microenvironment, significantly perturbing miRNA expression and creating a foggy picture of candidate miRNAs dysregulated within the tumor [332,333,334,335].
Many issues linked with miRNA therapeutics prevent their journey from bench to bedside. Significant issues are associated with specific delivery and poor bioavailability at the target site. Due to their polyanionic nature, transport across the lipid bilayer becomes difficult. Also, vascular barricades, such as tight junctions between the cells, significantly prevent paravascular transport. Their polyanionic nature and conjugation with high molecular weight carriers, like antibodies, for target-specific delivery mean they cannot traverse lipid bilayers spontaneously and remain trapped in lysosomal compartments or are retro-gated back to the plasma membrane [336]. This decreases their bioavailability, and a small fraction remain available for miRISC to act on their target mRNAs in the cytosol. To promote RNA interference (RNAi) machinery activation, the so-called “endosomal escape” of payload (miRNA-conjugate or package) is crucial. Several endos-osmotic agents have been exploited to deliver siRNA and ASO conjugates effectively. Small fusogenic peptides are developed, which can disrupt the endosomal membrane and help release the payload. This concept came after profoundly understanding the mechanisms behind the lysosome escape of pathogens, such as bacteria, viruses, etc. For instance, an HA2 domain of hemagglutinin developed from the influenza virus can be conjugated to promote effective siRNA delivery [337]. Endosomatic polymers are also exploited to promote the “endosomal escape” of siRNA or ASO complexes. In an acidic environment, the hydrophilic polymer is transformed into a hydrophobic polymer causing endosomal membrane partitioning and destabilization [338]. Apart from poor bioavailability, miRNA therapeutics are associated with a high probability of off-target effects, which may compromise target specificity and therapeutic effect and sometimes cause cell death. The frequent off-target effect of therapeutic miRNA is a tendency tobehave as siRNA miRNA, since the seed sequence tolerates many mismatches and decreases the abundance of several non-target mRNAs while still retaining the on-target effect. Another off-target effect is the competitive behavior of therapeutic siRNA with endogenous miRNAs for landing on RNAi machinery; this disturbs the natural equilibrium between RISC and endogenous miRNAs and may lead to unpredictable off-target effects [339].

9. Conclusions

The study of miRNAs is still in its initial stages. Many issues prevent its transition from batch to bedside, i.e., from research batch to clinics, such as suitable delivery methods, and insufficient understanding of their off-target effects on the human body. There is a need for a better and deeper understanding of targets and biological pathways that affect miRNAs to avoid any possible off-target effects. A large-scale population-based study should be conducted to identify those miRNAs which possess diagnostic, therapeutic, and prognostic potential in lung cancer. Also, there are many such studies whose results contradict each other. People use different specimens, sample collection methods, sample storage conditions, and expression analysis techniques that have a tremendous effect on reproducibility and accuracy and thus the study’s final results. There is a need to develop a standard protocol.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sung, H.; Ferlay, J.; Siegel, R.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef] [PubMed]
  2. World Health Organization. Global Cancer Observatory-GLOBOCAN 2020. Available online: https://gco.iarc.fr (accessed on 16 April 2022).
  3. Eltayeb, K.; La Monica, S.; Tiseo, M.; Alfieri, R.; Fumarola, C. Reprogramming of Lipid Metabolism in Lung Cancer: An Overview with Focus on EGFR-Mutated Non-Small Cell Lung Cancer. Cells 2022, 11, 413. [Google Scholar] [CrossRef]
  4. Bray, F.; Ferlay, J.; Soerjomataram, I.; Siegel, R.L.; Torre, L.A.; Jemal, A. Global Cancer Statistics 2018: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2018, 68, 394–424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Lantuéjoul, S.; Salameire, D.; Salon, C.; Brambilla, E. Pulmonary Preneoplasia–Sequential Molecular Carcinogenetic Events. Histopathology 2009, 54, 43–54. [Google Scholar] [CrossRef] [PubMed]
  6. Murray, N.; Coy, P.; Pater, J.R.; Hodson, I.; Arnold, A.; Zee, B.C.; Payne, D.; Kostashuk, E.C.; Evans, W.K.; Dixon, P. Importance of Timing for Thoracic Irradiation in the Combined Modality Treatment of Limited-stage small-cell Lung Cancer. J. Clin. Oncol. 1993, 11, 336–344. [Google Scholar] [CrossRef]
  7. Maemondo, M.; Inoue, A.; Kobayashi, K.; Sugawara, S.; Oizumi, S.; Isobe, H.; Gemma, A.; Harada, M.; Yoshizawa, H.; Kinoshita, I.; et al. Gefitinib or Chemotherapy for Non-small-cell Lung Cancer with Mutated EGFR. N. Engl. J. Med. 2010, 362, 2380–2388. [Google Scholar] [CrossRef] [Green Version]
  8. Zhou, C.; Wu, Y.L.; Chen, G.; Feng, J.; Liu, X.Q.; Wang, C.; Zhang, S.; Wang, J.; Zhou, S.; Ren, S.; et al. Erlotinib Versus Chemotherapy as First-line Treatment for Patients with Advanced EGFR Mutation- Positive Non-small-cell Lung Cancer (optimal, ctong-0802): A Multicentre, Open-label, Randomised, Phase 3 Study. Lancet Oncol. 2011, 12, 735–742. [Google Scholar] [CrossRef]
  9. Reck, M.; Bondarenko, I.; Luft, A.; Serwatowski, P.; Barlesi, F.; Chacko, R.; Sebastian, M.; Lu, H.; Cuillerot, J.M.; Lynch, T.J. Ipilimumab in combination with paclitaxel and carboplatin as first-line therapy in extensive-disease-small-cell lung cancer: Results from a randomized double-blind multicentre phase 2 trial. Ann. Oncol. 2013, 24, 75–83. [Google Scholar] [CrossRef]
  10. Stahel, R.; Ginsberg, R.; Havemann, K.; Hirsch, F.R.; Ihde, D.C.; Jassem, J.; Karrer, K.; Maurer, L.H.; Osterlind, K.; Houtte, P.V. Staging and Prognostic Factors in Small Cell Lung Cancer: A Consensus Report. Lung Cancer 1989, 5, 119–126. [Google Scholar] [CrossRef]
  11. Schreiber, G.; McCrory, D.C. Performance Characteristics of Different Modalities for Diagnosis of Suspected Lung Cancer: Summary of Published Evidence. CHEST 2003, 123, 115Se28S. [Google Scholar] [CrossRef] [Green Version]
  12. Patz, E.F.; Pinsky, P.; Gatsonis, C.; Sicks, J.D.; Kramer, S.; Tammemägi, M.C.; Chiles, C.; Black, W.C.; Aberle, D.R. Overdiagnosis in Low-dose Computed Tomography Screening for Lung Cancer. JAMA Int. Med. 2014, 174, 269–274. [Google Scholar] [CrossRef] [PubMed]
  13. Higashi, K.; Ueda, Y.; Yagishita, M.; Arisaka, Y.; Sakurai, A.; Oguchi, M.; Seki, H.; Nambu, Y.; Tonami, H.; Yamamoto, I. FDG PET Measurement of the Proliferative Potential of Non-small Cell Lung Cancer. J. Nucl. Med. 2000, 41, 85–92. [Google Scholar] [PubMed]
  14. Kubota, R.; Yamada, S.; Kubota, K.; Ishiwata, K.; Tamahashi, N.; Ido, T. Intratumoral Distribution of Fluorine-18-fluorodeoxyglucose in vivo: High Accumulation in Macrophages and Granulation Tissues Studied by Micro Autoradiography. J. Nucl. Med. 1992, 33, 1972–1980. [Google Scholar]
  15. Osman, M.M.; Cohade, C.; Nakamoto, Y.; Wahl, R.L. Respiratory Motion Artifacts on PET Emission Images Obtained Using CT Attenuation Correction on PET-CT. Eur. J. Nucl. Med. Mol. Imaging 2003, 4, 603–606. [Google Scholar] [CrossRef] [PubMed]
  16. Goerres, G.W.; Burger, C.; Kamel, E.; Seifert, B.; Kaim, A.H.; Buck, A.; Buehler, T.C.; Von Schulthes, G.K. Respiration-Induced Attenuation Artifact at PET/CT: Technical Considerations. Radiology 2003, 3, 906–910. [Google Scholar] [CrossRef] [PubMed]
  17. Shroff, G.S.; Shroff, B.S.; Truong, M.T.; Carter, B.W.; Viswanathan, C. PET/CT Interpretative Pitfalls in Thoracic Malignancies. Semin. Ultrasound CT MRI 2018, 39, 282–288. [Google Scholar]
  18. Shimizu, T.; Takahiro, S.; Imai, N.; Hase, T.; Morise, M.; Hashimoto, N.; Sato, M.; Hasegawa, Y. Risk Factors for Pulmonary Infection after Diagnostic Bronchoscopy in Patients with Lung Cancer. Nagoya J. Med. Sci. 2020, 82, 69–75. [Google Scholar]
  19. Tang, S.; Li, S.; Liu, T.; He, Y.; Hu, H.; Zhu, Y.; Tang, S.; Zhou, H. MicroRNAs: Emerging Oncogenic and Tumor-suppressive Regulators, Biomarkers and Therapeutic Targets in Lung Cancer. Cancer Lett. 2021, 502, 71–83. [Google Scholar] [CrossRef]
  20. Lee, R.C.; Feinbaum, R.L.; Ambros, V. The C. elegans Heterochronic Gene lin-4 Encodes Small RNAs with Antisense Complementarity to lin-14. Cell 1993, 75, 843–854. [Google Scholar] [CrossRef]
  21. Takamizawa, J.; Konishi, H.; Yanagisawa, K.; Tomida, S.; Osada, H.; Endoh, H.; Harano, T.; Yatabe, Y.; Nagino, M.; Nimura, Y.; et al. Reduced Expression of the let-7 microRNAs in Human Lung Cancers in Association with Shortened Postoperative Survival. Cancer Res. 2004, 64, 3753–3756. [Google Scholar] [CrossRef] [Green Version]
  22. Ruvkun, G.; Wightman, B.; Ha, I. The 20 Years it Took to Recognize the Importance of Tiny RNAs. Cell 2004, 116, S93–S98. [Google Scholar] [CrossRef] [Green Version]
  23. Souza, C.P.; Cinegaglia, N.C.; Felix, T.F.; Evangelista, A.F.; Oliveira, R.A.; Hasimoto, E.N.; Cataneo, D.C.; Cataneo, A.J.M.; Scapulatempo, N.C.; Viana, C.R.; et al. De-regulated microRNAs Are Associated with Patient Survival and Predicted to Target Genes That Modulate Lung Cancer Signaling Pathways. Cancers 2020, 12, 2711. [Google Scholar] [CrossRef] [PubMed]
  24. Boeri, M.; Verri, C.; Conte, D.; Roz, L.; Modena, P.; Facchinetti, F.; Calabro, E.; Croce, C.M.; Pastorino, U.; Sozzi, G. MicroRNA Signatures in Tissues and Plasma Predict Development and Prognosis of Computed Tomography Detected Lung Cancer. Proc. Natl. Acad. Sci. USA 2011, 108, 3713–3718. [Google Scholar] [CrossRef] [Green Version]
  25. Mohammadi, A.; Mansoori, B.; Duijf, P.H.; Safarzadeh, E.; Tebbi, L.; Najafi, S.; Shokouhi, B.; Sorensen, G.L.; Holmskov, U.; Baradaran, B. Restoration of miR-330 Expression Suppresses Lung Cancer Cell Viability, Proliferation, and Migration. J. Cell. Physiol. 2021, 236, 273–283. [Google Scholar] [CrossRef]
  26. Ye, J.; Luo, W.; Luo, L.; Zhai, L.; Huang, P. MicroRNA 671 5p Inhibits Cell Proliferation, Migration and Invasion in Non small Cell Lung Cancer by Targeting MFAP3L. Mol. Med. Rep. 2022, 25, 30. [Google Scholar] [CrossRef] [PubMed]
  27. Yuan, Y.; Liao, H.; Pu, Q.; Ke, X.; Hu, X.; Ma, Y.; Luo, X.; Jiang, Q.; Gong, Y.; Wu, M.; et al. miR-410 Induces Both Epithelial–Mesenchymal Transition and Radioresistance Through Activation of the PI3K/mTOR Pathway in non-small Cell Lung Cancer. Signal Transduct. Target. Ther. 2020, 5, 85. [Google Scholar] [CrossRef]
  28. Pandey, M.; Mukhopadhyay, A.; Sharawat, S.K.; Kumar, S. Role of microRNAs in Regulating Cell Proliferation, Metastasis and Chemoresistance and their Applications as Cancer Biomarkers in Small Cell Lung Cancer. Biochim. Biophys. Acta Rev. Cancer 2021, 1876, 188552. [Google Scholar] [CrossRef]
  29. Pal, A.S.; Bains, M.; Agredo, A.; Kasinski, A.L. Identification of microRNAs that Promote Erlotinib Resistance in non-small Cell Lung Cancer. Biochem. Pharmacol. 2021, 189, 114154. [Google Scholar] [CrossRef]
  30. Chae, D.K.; Park, J.; Cho, M.; Ban, E.; Jang, M.; Yoo, Y.S.; Kim, E.E.; Baik, J.H.; Song, E.J. MiR-195 and miR-497 Suppress Tumorigenesis in Lung Cancer by Inhibiting SMURF2-Induced TGF-β Receptor I Ubiquitination. Mol. Oncol. 2019, 13, 2663–2678. [Google Scholar] [CrossRef] [Green Version]
  31. Huang, X.; Xiao, S.; Zhu, X.; Yu, Y.; Cao, M.; Zhang, X.; Li, S.; Zhu, W.; Wu, F.; Zheng, X.; et al. miR-196b-5p-mediated Downregulation of FAS Promotes NSCLC Progression by Activating IL6-STAT3 Signaling. Cell Death Dis. 2020, 11, 785. [Google Scholar] [CrossRef]
  32. Wang, H.; Deng, Q.; Lv, Z.; Ling, Y.; Hou, X.; Chen, Z.; Dinglin, X.; Ma, S.; Li, D.; Wu, Y.; et al. N6-methyladenosine Induced miR-143-3p Promotes the Brain Metastasis of Lung Cancer via Regulation of VASH1. Mol. Cancer 2019, 18, 181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Zou, P.; Zhu, M.; Lian, C.; Wang, J.; Chen, Z.; Zhang, X.; Yang, Y.; Chen, X.; Cui, X.; Liu, J.; et al. miR-192-5p Suppresses the Progression of Lung Cancer Bone Metastasis by Targeting TRIM44. Sci. Rep. 2019, 9, 19619. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Hanna, J.; Hossain, G.S.; Kocerha, J. The Potential for microRNA Therapeutics and Clinical Research. Front. Genet. 2019, 10, 478. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Zhang, J.; Hua, X.; Qi, N.; Han, G.; Yu, J.; Yu, Y.; Wei, X.; Li, H.; Chen, X.; Leng, C.; et al. MiR-27b Suppresses Epithelial-Mesenchymal Transition and Chemoresistance in Lung Cancer by Targeting Snail1. Life Sci. 2020, 254, 117238. [Google Scholar] [CrossRef] [PubMed]
  36. Sun, H.; Sun, Y. Lidocaine Inhibits Proliferation and Metastasis of Lung Cancer Cell via Regulation of miR-539/EGFR axis. Artif. Cells Nanomed. Biotechnol. 2019, 47, 2866–2874. [Google Scholar] [CrossRef] [Green Version]
  37. Jiang, Z.; Zhang, J.; Chen, F.; Sun, Y. MiR-148b Suppressed Non-small Cell Lung Cancer Progression via Inhibiting ALCAM through the NF-κB Signaling Pathway. Thorac. Cancer 2020, 11, 415–425. [Google Scholar] [CrossRef]
  38. Frixa, T.; Sacconi, A.; Cioce, M.; Roscilli, G.; Ferrara, F.F.; Aurisicchio, L.; Pulito, C.; Telera, S.; Carosi, M.; Muti, P.; et al. MicroRNA-128-3p-mediated Depletion of Drosha Promotes Lung Cancer Cell Migration. Carcinogenesis 2018, 39, 293–304. [Google Scholar] [CrossRef]
  39. Zafari, V.; Shanehbandi, D.; Bornehdeli, S.; Sadeghzadeh, M.; Zarredar, H.; Asadi, M.; Sharifi, A. MicroRNA Profiling in non-small Cell Lung Cancer and its Implications for the Disease Pathogenesis. Middle East J. Cancer 2021, 12, 79–85. [Google Scholar]
  40. Hashemi, S.; Yari, N.; Rahimi, J.F.; Mahdian, R.; Karimi, M.; Zeinali, S.; Rafiee, M.H.; Azizi, M. The Role of miRNA-377 as a Tumor Suppressor in Lung Cancer by Negative Regulation of Genes Belonging to ErbB Signaling Pathway. Mol. Biol. Rep. 2022, 49, 85–95. [Google Scholar] [CrossRef]
  41. Senedaa, A.L.; Lapac, R.M.; Felixa, T.F.; Minutentaga, I.W.; Camposa, C.F.; de Oliveirae, R.A.; Oliveiraf, C.C.; Hasimotoa, É.N.; Cataneoa, D.C.; Cataneoa, A.J.; et al. Typical Lung Carcinoids with Metastasis: Potential Role of MicroRNAs in the Regulation of Adaptive Immunity Associated with Disease: A Case Study. Cell Physiol. Biochem. 2021, 55, 1–12. [Google Scholar] [CrossRef]
  42. Deng, B.; Molina, J.; Aubry, M.C.; Sun, Z.; Wang, L.; Eckloff, B.W.; Vasmatzis, G.; You, M.; Wieben, E.D.; Jen, J.; et al. Clinical Biomarkers of Pulmonary Carcinoid Tumors in never Smokers via Profiling miRNA and Target mRNA. Cell Biosci. 2014, 4, 35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Rapa, I.; Votta, A.; Felice, B.; Righi, L.; Giorcelli, J.; Scarpa, A.; Speel, E.J.M.; Scagliotti, G.V.; Papotti, M.; Volante, M. Identification of MicroRNAs Differentially Expressed in Lung Carcinoid Subtypes and Progression. Neuroendocrinology 2015, 101, 246–255. [Google Scholar] [CrossRef] [PubMed]
  44. Davenport, M.L.; Echols, J.B.; Silva, A.D.; Anderson, J.C.; Owens, P.; Yates, C.; Wei, Q.; Harada, S.; Hurst, D.R.; Edmonds, M.D. miR-31 Displays Subtype Specificity in Lung Cancer. Cancer Res. 2021, 8, 1942–1953. [Google Scholar] [CrossRef]
  45. Cornett, A.L.; Lutz, C.S. Regulation of COX-2 Expression by miR-146a in Lung Cancer Cells. RNA 2014, 20, 1419–1430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Paterson, M.R.; Kriegel, A.J. MiR-146a/b: A Family with shared Seeds and Different Roots. Physiol. Genome 2017, 49, 243–252. [Google Scholar] [CrossRef] [PubMed]
  47. Yin, Z.; Cui, Z.; Ren, Y.; Xia, L.; Li, H.; Zhou, B. MiR-146a Polymorphism Correlates with Lung Cancer Risk in Chinese Nonsmoking Females. Oncotarget 2017, 8, 2275. [Google Scholar] [CrossRef]
  48. Li, J.; Yang, H.; Li, Y.; Liu, Y.; Chen, S.; Qi, C.; Zhang, Q.; Lan, T.; He, X.; Guan, X.Y.; et al. microRNA-146 Up-Regulation Predicts the Prognosis of Non-Small Cell Lung Cancer by miRNA in situ Hybridization. Exp. Mol. Pathol. 2014, 96, 195–199. [Google Scholar] [CrossRef]
  49. Mohamed, R.H.; Pasha, H.F.; Gad, D.M.; Toam, M.M. miR-146a and miR-196a-2 Genes Polymorphisms and its Circulating Levels in Lung Cancer Patients. J. Biochem. 2019, 166, 323–329. [Google Scholar] [CrossRef]
  50. Pavel, A.B.; Campbell, J.D.; Liu, G.; Elashoff, D.; Dubinett, S.; Smith, K.; Whitney, D.; Lenburg, M.E.; Spira, A. Alterations in Bronchial Airway miRNA Expression for Lung Cancer Detection. Cancer Prev. Res. 2017, 10, 651–659. [Google Scholar] [CrossRef] [Green Version]
  51. Wang, R.J.; Zheng, Y.H.; Wang, P.; Zhang, J.Z. Serum miR-125a-5p, miR-145 and miR-146a as Diagnostic Biomarkers in non-small Cell Lung Cancer. Int. J. Clin. Exp. Pathol. 2015, 8, 765. [Google Scholar]
  52. Wu, C.; Cao, Y.; He, Z.; He, J.; Hu, C.; Duan, H.; Jiang, J. Serum Levels of miR-19b and miR-146a as Prognostic Biomarkers for non-small Cell Lung Cancer. Tohoku J. Exp. Med. 2014, 232, 85–95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Lv, S.; Xue, J.; Wu, C.; Wang, L.; Wu, J.; Xu, S.; Liang, X.; Lou, J. Identification of a Panel of Serum microRNAs as Biomarkers for Early Detection of Lung Adenocarcinoma. J. Cancer 2017, 8, 48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Soliman, S.E.; Abdelaleem, A.H.; Alhanafy, A.M.; Ibrahem, R.A.; Elhaded, A.S.; Assar, M.F. Circulating miR-21-5p and miR-126-3p: Diagnostic, Prognostic Value, and Multivariate Analysis in Non-small-cell Lung Cancer. Mol. Biol. Rep. 2021, 48, 2543–2552. [Google Scholar] [CrossRef] [PubMed]
  55. Xu, S.; Shi, L. High Expression of miR-155 and miR-21 in the Recurrence or Metastasis of non-small Cell Lung Cancer. Oncol. Lett. 2019, 18, 758–763. [Google Scholar] [CrossRef] [Green Version]
  56. Zhao, W.; Zhao, J.J.; Zhang, L.; Xu, Q.F.; Zhao, Y.M.; Shi, X.Y.; Xu, A.G. Serum miR-21 level: A Potential Diagnostic and Prognostic Biomarker for non-small Cell Lung Cancer. Int. J. Clin. Exp. Med. 2015, 8, 14759. [Google Scholar]
  57. Capodanno, A.; Boldrini, L.; Proietti, A.; Alì, G.; Pelliccioni, S.; Niccoli, C.; D’Incecco, A.; Cappuzzo, F.; Chella, A.; Lucchi, M.; et al. Let-7g and miR-21 Expression in non-small Cell Lung Cancer: Correlation with Clinicopathological and Molecular Features. Int. J. Oncol. 2013, 43, 765–774. [Google Scholar] [CrossRef] [Green Version]
  58. Jiang, M.; Zhang, P.; Hu, G.; Xiao, Z.; Xu, F.; Zhong, T.; Huang, F.; Kuang, H.; Zhang, W. Relative Expressions of miR-205-5p, miR-205-3p, and miR-21 in Tissues and Serum of non-small Cell Lung Cancer Patients. Mol. Cell. Biochem. 2013, 383, 67–75. [Google Scholar] [CrossRef]
  59. Gao, W.; Yu, Y.; Cao, H.; Shen, H.; Li, X.; Pan, S.; Shu, Y. Deregulated Expression of miR-21, miR-143 and miR-181a in non Small Cell Lung Cancer is Related to Clinicopathologic Characteristics or Patient Prognosis. Biomed. Pharmacother. 2010, 64, 399–408. [Google Scholar] [CrossRef]
  60. Zhu, S.; Si, M.L.; Wu, H.; Mo, Y.Y. MicroRNA-21 Targets the Tumor Suppressor Gene Tropomyosin 1 (TPM1). J. Biol. Chem. 2007, 282, 14328–14336. [Google Scholar] [CrossRef] [Green Version]
  61. Qi, J.H.; Ebrahem, Q.; Moore, N.; Murphy, G.; Claesson-Welsh, L.; Bond, M.; Anand-Apte, B. A Novel Function for Tissue Inhibitor of Metalloproteinases-3 (TIMP3): Inhibition of Angiogenesis by Blockage of VEGF Binding to VEGF receptor-2. Nat. Med. 2003, 9, 407–415. [Google Scholar] [CrossRef]
  62. Clark, S.J. Action at a Distance: Epigenetic Silencing of Large Chromosomal Regions in Carcinogenesis. Hum. Mol. Genet. 2007, 16, 88–95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Hatley, M.E.; Patrick, D.M.; Garcia, M.R.; Richardson, J.A.; Bassel-Duby, R.; Van, R.E.; Olson, E.N. Modulation of K-Ras-dependent Lung Tumorigenesis by MicroRNA-21. Cancer Cell 2010, 18, 282–293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Gallardo, E.; Navarro, A.; Vinolas, N.; Marrades, R.M.; Diaz, T.; Gel, B.; Quera, A.; Bandres, E.; Garcia-Foncillas, J.; Ramirez, J.; et al. miR-34a as a Prognostic Marker of Relapse in Surgically Resected non-small-cell Lung Cancer. Carcinogenesis 2009, 30, 1903–1909. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Franchina, T.; Amodeo, V.; Bronte, G.; Savio, G.; Ricciardi, G.R.; Picciotto, M.; Russo, A.; Giordano, A.; Adamo, V. Circulating miR-22, miR-24 and miR-34a as Novel Predictive Biomarkers to Pemetrexed-based Chemotherapy in Advanced non-small Cell Lung Cancer. J. Cell. Physiol. 2014, 229, 97–99. [Google Scholar] [CrossRef] [Green Version]
  66. Kim, J.S.; Kim, E.J.; Lee, S.; Tan, X.; Liu, X.; Park, S.; Kang, K.; Yoon, J.S.; Ko, Y.H.; Kurie, J.M.; et al. MiR-34a and miR-34b/c Have Distinct Effects on the Suppression of Lung Adenocarcinomas. Exp. Mol. Med. 2019, 51, 1–10. [Google Scholar] [CrossRef] [Green Version]
  67. Wiggins, J.F.; Ruffino, L.; Kelnar, K.; Omotola, M.; Patrawala, L.; Brown, D.; Bader, A.G. Development of a Lung Cancer Therapeutic Based on the Tumor Suppressor microRNA-34. Cancer Res. 2010, 70, 5923–5930. [Google Scholar] [CrossRef] [Green Version]
  68. Stahlhut, C.; Slack, F.J. Combinatorial Action of microRNAs let 7 and miR-34 Effectively Synergizes with Erlotinib to Suppress non-small Cell Lung Cancer Cell Proliferation. Cell Cycle 2015, 14, 2171–2180. [Google Scholar] [CrossRef] [Green Version]
  69. Huang, D.; Ou, W.; Tong, H.; Peng, M.; Ou, Y.; Song, Z. Analysis of the Expression Levels and Clinical Value of miR 365 and miR 25 in Serum of Patients with non small Cell Lung Cancer. Oncol. Lett. 2020, 20, 191. [Google Scholar] [CrossRef]
  70. Liu, Y.; Zhang, G.; Li, H.; Han, L.; Fu, A.; Zhang, N.; Zheng, Y. Serum microRNA-365 in Combination with its Target gene TTF-1 as a Non-invasive Prognostic Marker for non-small Cell Lung Cancer. Biomed. Pharmacother. 2015, 75, 185–190. [Google Scholar] [CrossRef]
  71. Kang, S.M.; Lee, H.J.; Cho, J.Y. MicroRNA-365 Regulates NKX2-1, a Key Mediator of Lung Cancer. Cancer Lett. 2013, 335, 487–494. [Google Scholar] [CrossRef]
  72. Han, Q.; Cheng, P.; Yang, H.; Liang, H.; Lin, F. Altered Expression of microRNA-365 is Related to the Occurrence and Development of non-small-cell Lung Cancer by Inhibiting TRIM25 Expression. J. Cell Physiol. 2019, 234, 22321–22330. [Google Scholar] [CrossRef] [PubMed]
  73. Tian, F.; Wang, J.; Ouyang, T.; Lu, N.; Lu, J.; Shen, Y.; Bai, Y.; Xie, X.; Ge, Q. MiR-486-5p Serves as a Good Biomarker in non-small Cell Lung Cancer and Suppresses Cell Growth with the Involvement of a Target PIK3R1. Front. Genet. 2019, 10, 688. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Mohamed, M.A.; Mohamed, E.I.; El-Kaream, S.A.; Badawi, M.I.; Darwish, S.H. Underexpression of miR-486-5p but not Overexpression of miR-155 is Associated with Lung Cancer Stages. Microrna 2018, 7, 120–127. [Google Scholar] [CrossRef] [PubMed]
  75. Jin, X.; Pang, W.; Zhang, Q.; Huang, H. MicroRNA-486-5p Improves nonsmall-cell Lung Cancer Chemotherapy Sensitivity and Inhibits Epithelial–Mesenchymal Transition by Targeting Twinfilin Actin Binding Protein 1. Int. J. Med. Res. 2019, 47, 3745–3756. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Gao, Z.J.; Yuan, W.D.; Yuan, J.Q.; Yuan, K.; Wang, Y. miR-486-5p Functions as an Oncogene by Targeting PTEN in non-small Cell Lung Cancer. Pathol. Res. Pract. 2018, 214, 700–705. [Google Scholar] [CrossRef]
  77. Jin, X.; Chen, Y.; Chen, H.; Fei, S.; Chen, D.; Cai, X.; Liu, L.; Lin, B.; Su, H.; Zhao, L.; et al. Evaluation of Tumor-Derived Exosomal miRNA as Potential Diagnostic Biomarkers for Early-stage Non–small Cell lung Cancer using Next-generation Sequencing. Clin. Cancer Res. 2017, 17, 5311–5319. [Google Scholar] [CrossRef] [Green Version]
  78. Zhuang, Z.L.; Tian, F.M.; Sun, C.L. Downregulation of miR-361-5p Associates with Aggressive Clinicopathological Features and Unfavorable Prognosis in non-small Cell Lung Cancer. Eur. Rev. Med. Pharmacol. Sci. 2016, 20, 5132–5136. [Google Scholar]
  79. Roth, C.; Stückrath, I.; Pantel, K.; Izbicki, J.R.; Tachezy, M.; Schwarzenbach, H. Low Levels of Cell-free Circulating miR-361-3p and miR-625* as Blood-based Markers for Discriminating Malignant from Benign lung tumors. PLoS ONE. 2012, 7, e38248. [Google Scholar] [CrossRef]
  80. Hou, X.W.; Sun, X.; Yu, Y.; Zhao, H.M.; Yang, Z.J.; Wang, X.; Cao, X.C. miR-361-5p Suppresses Lung Cancer Cell Lines Progression by Targeting FOXM1. Neoplasma 2017, 64, 526–534. [Google Scholar] [CrossRef]
  81. Lee, H.Y.; Han, S.S.; Rhee, H.; Park, J.H.; Lee, J.S.; Oh, Y.M.; Choi, S.S.; Shin, S.H.; Kim, W.J. Differential Expression of microRNAs and their Target Genes in Non-small-cell Lung Cancer. Mol. Med. Rep. 2015, 11, 2034–2040. [Google Scholar] [CrossRef] [Green Version]
  82. Liu, J.; Jia, Y.; Jia, L.; Li, T.; Yang, L.; Zhang, G. MicroRNA 615-3p Inhibits the Tumor Growth and Metastasis of NSCLC via Inhibiting IGF2. Oncol. Res. 2019, 27, 269. [Google Scholar] [CrossRef] [PubMed]
  83. Pu, H.Y.; Xu, R.; Zhang, M.Y.; Yuan, L.J.; Hu, J.Y.; Huang, G.L.; Wang, H.Y. Identification of microRNA-615-3p as a Novel Tumor Suppressor in Non-small Cell Lung Cancer. Oncol. Lett. 2017, 13, 2403–2410. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Guo, L.; Wang, J.; Yang, P.; Lu, Q.; Zhang, T.; Yang, Y. Micro RNA-200 Promotes Lung Cancer Cell Growth Through FOG2-independent AKT Activation. IUBMB Life 2015, 67, 720–725. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Tejero, R.; Navarro, A.; Campayo, M.; Viñolas, N.; Marrades, R.M.; Cordeiro, A.; Ruíz-Martínez, M.; Santasusagna, S.; Molins, L.; Ramirez, J.; et al. miR-141 and miR-200c as Markers of Overall Survival in Early Stage non-small Cell Lung Cancer Adenocarcinoma. PLoS ONE 2014, 7, e101899. [Google Scholar] [CrossRef]
  86. Chen, G.; Umelo, I.A.; Lv, S.; Teugels, E.; Fostier, K.; Kronenberger, P.; Dewaele, A.; Sadones, J.; Geers, C.; De Grève, J. miR-146a Inhibits Cell Growth, Cell Migration and Induces Apoptosis in Non-small Cell lung Cancer Cells. PLoS ONE 2013, 8, e60317. [Google Scholar] [CrossRef] [Green Version]
  87. Gregory, P.A.; Bert, A.G.; Paterson, E.L.; Barry, S.C.; Tsykin, A.; Farshid, G.; Vadas, M.A.; Goodall, G.J. The miR-200 Family and miR-205 Regulate Epithelial to Mesenchymal Transition by Targeting ZEB1 and SIP1. Nat. Cell Biol. 2008, 10, 593–601. [Google Scholar] [CrossRef] [PubMed]
  88. Lee, J.S.; Ahn, Y.H.; Won, H.S.; Sun, D.S.; Kim, Y.H.; Ko, Y.H. Prognostic Role of the microRNA-200 Family in Various Carcinomas: A Systematic Review and Meta-analysis. Biomed. Res. Int. 2017, 15, 1928021. [Google Scholar] [CrossRef] [Green Version]
  89. Zhu, Y.; Li, T.; Chen, G.; Yan, G.; Zhang, X.; Wan, Y.; Li, Q.; Zhu, B.; Zhuo, W. Identification of a Serum microRNA Expression Signature for Detection of Lung Cancer, Involving miR-23b, miR-221, miR-148b and miR-423-3p. Lung Cancer 2017, 114, 6–11. [Google Scholar] [CrossRef]
  90. Heegaard, N.H.; Schetter, A.J.; Welsh, J.A.; Yoneda, M.; Bowman, E.D.; Harris, C.C. Circulating micro-RNA Expression Profiles in Early Stage Nonsmall Cell Lung Cancer. Int. J. Cancer Res. 2012, 130, 1378–1386. [Google Scholar] [CrossRef] [Green Version]
  91. Zhou, X.; Wen, W.; Shan, X.; Zhu, W.; Xu, J.; Guo, R.; Cheng, W.; Wang, F.; Qi, L.W.; Chen, Y.; et al. A six-microRNA Panel in Plasma was Identified as a Potential Biomarker for Lung Adenocarcinoma Diagnosis. Oncotarget 2017, 8, 6513–6525. [Google Scholar] [CrossRef] [Green Version]
  92. Garofalo, M.; Di Leva, G.; Romano, G.; Nuovo, G.; Suh, S.S.; Ngankeu, A.; Taccioli, C.; Pichiorri, F.; Alder, H.; Secchiero, P.; et al. miR-221&222 Regulate TRAIL Resistance and Enhance Tumorigenicity Through PTEN and TIMP3 Downregulation. Cancer Cell 2009, 16, 498–509. [Google Scholar] [PubMed] [Green Version]
  93. Zhang, C.; Zhang, J.; Zhang, A.; Wang, Y.; Han, L.; You, Y.; Pu, P.; Kang, C. PUMA was a Novel Target of miR-221/222 in Human Epithelial Cancers. Int. J. Oncol. 2010, 37, 1621–1626. [Google Scholar] [PubMed] [Green Version]
  94. Geng, Q.; Fan, T.; Zhang, B.; Wang, W.; Xu, Y.; Hu, H. Five microRNAs in Plasma as Novel Biomarkers for Screening of Early-Stage Non-small Cell Lung Cancer. Respir. Res. 2014, 15, 149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Esquela-Kerscher, A.; Trang, P.; Wiggins, J.F.; Patrawala, L.; Cheng, A.; Ford, L.; Weidhaas, J.B.; Brown, D.; Bader, A.G.; Slack, F.J. The let-7 microRNA Reduces Tumor Growth in Mouse Models of Lung Cancer. Cell Cycle 2008, 7, 759–764. [Google Scholar] [CrossRef] [Green Version]
  96. Kumar, M.S.; Erkeland, S.J.; Pester, R.E.; Chen, C.Y.; Ebert, M.S.; Sharp, P.A.; Jacks, T. Suppression of Non-small Cell Lung Tumor Development by the let-7 microRNA Family. Proc. Natl. Acad. Sci. USA 2008, 105, 3903–3908. [Google Scholar] [CrossRef] [Green Version]
  97. Wang, X.; Cao, L.E.; Wang, Y.; Wang, X.; Liu, N.; You, Y. Regulation of let-7 and its Target Oncogenes. Oncol. Lett. 2012, 3, 955–960. [Google Scholar] [CrossRef]
  98. Yin, J.; Zhao, J.; Hu, W.; Yang, G.; Yu, H.; Wang, R.; Wang, L.; Zhang, G.; Fu, W.; Dai, L.; et al. Disturbance of the let-7/LIN28 Double-negative Feedback Loop is Associated with Radio-and Chemo-resistance in Non-small Cell Lung Cancer. PLoS ONE 2017, 12, e0172787. [Google Scholar] [CrossRef]
  99. Jusufovic, E.; Rijavec, M.; Keser, D.; Korosec, P.; Sodja, E.; Iljazovic, E.; Radojevic, Z.; Kosnik, M. let-7b and miR-126 are Down-Regulated in Tumor Tissue and Correlate with Microvessel Density and Survival Outcomes in non–small–cell Lung Cancer. PLoS ONE 2012, 7, e45577. [Google Scholar] [CrossRef]
  100. Zhu, W.Y.; Luo, B.; An, J.Y.; He, J.Y.; Chen, D.D.; Xu, L.Y.; Huang, Y.Y.; Liu, X.G.; Le, H.B.; Zhang, Y.K. Differential Expression of miR-125a-5p and let-7e predicts the Progression and Prognosis of Non-small Cell Lung Cancer. Cancer Investig. 2014, 32, 394–401. [Google Scholar] [CrossRef]
  101. Xu, S.; Zheng, L.; Kang, L.; Xu, H.; Gao, L. microRNA-let-7e in Serum-derived Exosomes Inhibits the Metastasis of Non-small-cell Lung Cancer in a SUV39H2/LSD1/CDH1-Dependent Manner. Cancer Gene Ther. 2021, 3, 250–264. [Google Scholar] [CrossRef]
  102. Chen, Q.; Si, Q.; Xiao, S.; Xie, Q.; Lin, J.; Wang, C.; Chen, L.; Chen, Q.; Wang, L. Prognostic Significance of Serum miR-17-5p in Lung Cancer. Med. Oncol. 2013, 30, 353–365. [Google Scholar] [CrossRef] [PubMed]
  103. Chatterjee, A.; Chattopadhyay, D.; Chakrabarti, G. miR-17-5p Downregulation Contributes to Paclitaxel Resistance of Lung Cancer Cells through Altering Beclin1 Expression. PLoS ONE 2014, 9, e95716. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Serva, A.; Knapp, B.; Tsai, Y.T.; Claas, C.; Lisauskas, T.; Matula, P.; Harder, N.; Kaderali, L.; Rohr, K.; Erfle, H.; et al. miR-17-5p Regulates Endocytic Trafficking through Targeting TBC1D2/Armus. PLoS ONE 2012, 7, e52555. [Google Scholar] [CrossRef] [Green Version]
  105. Zhang, W.; Lin, J.; Wang, P.; Sun, J. miR-17-5p Down-regulation Contributes to Erlotinib Resistance in Non-small Cell Lung Cancer Cells. J. Drug Target. 2017, 25, 125–131. [Google Scholar] [CrossRef] [Green Version]
  106. Navarro, A.; Marrades, R.M.; Vinolas, N.; Quera, A.; Agusti, C.; Huerta, A.; Ramirez, J.; Torres, A.; Monzo, M. MicroRNAs Expressed During Lung Cancer Development are Expressed in Human Pseudoglandular Lung Embryogenesis. Oncology 2009, 76, 162–169. [Google Scholar] [CrossRef] [PubMed]
  107. Lin, Q.; Chen, T.; Lin, Q.; Lin, G.; Lin, J.; Chen, G.; Guo, L. Serum miR-19a Expression Correlates with Worse Prognosis of Patients with Non-small Cell Lung Cancer. J. Surg. Oncol. 2013, 107, 767–771. [Google Scholar] [CrossRef]
  108. Baumgartner, U.; Berger, F.; Gheinani, A.H.; Burgener, S.S.; Monastyrskaya, K.; Vassella, E. miR-19b Enhances Proliferation and Apoptosis Resistance via the EGFR Signaling Pathway by Targeting PP2A and BIM in Non-small Cell Lung Cancer. Mol. Cancer 2018, 17, 44–50. [Google Scholar] [CrossRef]
  109. Cao, X.; Lai, S.; Hu, F.; Li, G.; Wang, G.; Luo, X.; Fu, X.; Hu, J. miR-19a Contributes to Gefitinib Resistance and Epithelial Mesenchymal Transition in Non-small Cell Lung Cancer Cells by Targeting c-Met. Sci. Rep. 2017, 7, 2939. [Google Scholar] [CrossRef] [Green Version]
  110. Chen, X.; Wu, L.; Li, D.; Xu, Y.; Zhang, L.; Niu, K.; Kong, R.; Gu, J.; Xu, Z.; Chen, Z.; et al. Radiosensitizing Effects of miR-18a-5p on Lung Cancer Stem-like Cells via Downregulating both ATM and HIF-1α. Cancer Med. 2018, 8, 3834–3847. [Google Scholar] [CrossRef]
  111. Liang, C.; Zhang, X.; Wang, H.M.; Liu, X.M.; Zhang, X.J.; Zheng, B.; Qian, G.R.; Ma, Z.L. MicroRNA-18a-5p Functions as an Oncogene by Directly Targeting IRF2 in Lung Cancer. Cell Death. Dis. 2017, 8, e2764. [Google Scholar] [CrossRef] [Green Version]
  112. Xu, X.; Zhu, S.; Tao, Z.; Ye, S. High Circulating miR-18a, miR-20a, and miR-92a Expression Correlates with Poor Prognosis in Patients with Non-small Cell Lung Cancer. Cancer Med. 2018, 7, 21–31. [Google Scholar] [CrossRef] [PubMed]
  113. Zhang, X.; Li, Y.; Qi, P.; Ma, Z. Biology of MiR-17-92 Cluster and Its Progress in Lung Cancer. Int. J. Med. Sci. 2018, 15, 1443–1448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Liu, F.; Cai, Y.; Rong, X.; Chen, J.; Zheng, D.; Chen, L.; Zhang, J.; Luo, R.; Zhao, P.; Ruan, J. MiR-661 Promotes Tumor Invasion and Metastasis by Directly Inhibiting RB1 in Non Small Cell Lung Cancer. Mol. Cancer 2017, 16, 122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Zhou, G.H.; Yang, W.H.; Sun, B. Clinical Impact of Serum miR-661 in Diagnosis and Prognosis of Non-small Cell Lung Cancer. Eur. Rev. Med. Pharmacol. Sci. 2017, 21, 5696–5701. [Google Scholar]
  116. Wang, Y.; Li, Y.; Wu, B.; Shi, C.; Li, C. MicroRNA-661 Promotes Non-small Cell Lung Cancer Progression by Directly Targeting RUNX3. Mol. Med. Rep. 2017, 16, 2113–2120. [Google Scholar] [CrossRef] [Green Version]
  117. Ye, M.F.; Lin, D.; Li, W.J.; Xu, H.P.; Zhang, J. MiR-26a-5p Serves as an Oncogenic microRNA in Non-small Cell Lung Cancer by Targeting FAF1. Cancer Manag. Res. 2020, 12, 7131–7140. [Google Scholar] [CrossRef]
  118. Song, Q.; Liu, B.; Li, X.; Zhang, Q.; Cao, L.; Xu, M.; Meng, Z.; Wu, X.; Xu, K. MiR-26a-5p Potentiates Metastasis of Human Lung Cancer Cells by Regulating ITGβ8-JAK2/STAT3 Axis. Biochem. Biophys. Res. Commun. 2018, 50, 494–500. [Google Scholar] [CrossRef]
  119. Gasparini, P.; Cascione, L.; Landi, L.; Carasi, S.; Lovat, F.; Tibaldi, C.; Alì, G.; D’Incecco, A.; Minuti, G.; Chella, A.; et al. microRNA Classifiers are Powerful Diagnostic/Prognostic Tools in ALK-, EGFR-, and KRAS-driven Lung Cancers. Proc. Natl. Acad. Sci. USA 2015, 48, 14924–14929. [Google Scholar] [CrossRef] [Green Version]
  120. Ni, Y.; Yang, Y.; Ran, J.; Zhang, L.; Yao, M.; Liu, Z.; Zhang, L. miR-15a-5p Inhibits Metastasis and Lipid Metabolism by Suppressing Histone Acetylation in Lung Cancer. Free Radic. Biol. Med. 2020, 161, 150–162. [Google Scholar] [CrossRef]
  121. Ji, K.X.; Cui, F.; Qu, D.; Sun, R.Y.; Sun, P.; Chen, F.Y.; Wang, S.L.; Sun, H.S. MiR-378 Promotes the Cell Proliferation of Non-small Cell Lung Cancer by Inhibiting FOXG1. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 1011–1019. [Google Scholar]
  122. San, H.C.; Noor, S.M.; Nagoor, N.H. MiR-378 and MiR-1827 Regulate Tumor Invasion, Migration and Angiogenesis in Human Lung Adenocarcinoma by Targeting RBX1 and CRKL, Respectively. J. Cancer 2018, 9, 331–340. [Google Scholar]
  123. Chen, X.; Jiang, Y.; Huang, Z.; Li, D.; Chen, X.; Cao, M.; Meng, Q.; Pang, H.; Sun, L.; Zhao, Y.; et al. miRNA-378 Reverses Chemoresistance to Cisplatin in Lung Adenocarcinoma Cells by Targeting Secreted Clusterin. Sci. Rep. 2016, 19, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Li, Y.; Liang, M.; Zhang, Y.; Yuan, B.; Gao, W.; Shi, Z.; Bai, J. miR-93, miR-373, and miR-17-5p Negatively Regulate the Expression of TBP2 in Lung Cancer. Front. Oncol. 2020, 10, 526–535. [Google Scholar] [CrossRef] [PubMed]
  125. Shao, L.; Lu, X.; Zhou, Y.; Wang, Y.; Wang, X.; Zhuang, Z.; Gong, J. Altered miR-93-5p/miR-18a Expression in Serum for Diagnosing Non-small Cell Lung Cancer. Am. J. Transl. Res. 2021, 13, 5073–5075. [Google Scholar]
  126. Yang, W.; Bai, J.; Liu, D.; Wang, S.; Zhao, N.; Che, R.; Zhang, H. MiR-93-5p Up-Regulation is Involved in Non-small Cell Lung Cancer Cells Proliferation and Migration and Poor Prognosis. Gene 2018, 647, 13–20. [Google Scholar] [CrossRef]
  127. Du, L.; Zhao, Z.; Ma, X.; Hsiao, T.H.; Chen, Y.; Young, E.; Suraokar, M.; Wistuba, I.; Minna, J.D.; Pertsemlidis, A. miR-93-directed Downregulation of DAB2 Defines a Novel Oncogenic Pathway in Lung Cancer. Oncogene 2014, 34, 4307–4315. [Google Scholar] [CrossRef] [Green Version]
  128. Lin, C.W.; Chang, Y.L.; Chang, Y.C.; Lin, J.C.; Chen, C.C.; Pan, S.H.; Wu, C.T.; Chen, H.Y.; Yang, S.C.; Hong, T.M.; et al. MicroRNA-135b Promotes Lung Cancer Metastasis by Regulating Multiple Targets in the Hippo Pathway and LZTS1. Nat. Commun. 2013, 4, 1877. [Google Scholar] [CrossRef] [Green Version]
  129. Yang, Y.; Chen, K.; Zhou, Y.; Hu, Z.; Chen, S.; Huang, Y. Application of Serum microRNA-9-5p, 21–25p, and 223-3p Combined with Tumor Markers in the Diagnosis of Non-small-cell Lung Cancer in Yunnan in southwestern China. OncoTargets Ther. 2018, 11, 587–595. [Google Scholar] [CrossRef] [Green Version]
  130. Le, H.; Wang, X.; Zha, Y.; Wang, J.; Zhu, W.; Ye, Z.; Liu, X.; Ma, H.; Zhang, Y. Peripheral Lung Adenocarcinomas Harboring Epithelial Growth Factor Receptor Mutations with microRNA 135b Overexpression are more likely to Invade Visceral Pleura. Oncol. Lett. 2017, 14, 7931–7940. [Google Scholar] [CrossRef]
  131. Chen, T.; Xiao, Q.; Wang, X.; Wang, Z.; Hu, J.; Zhang, Z.; Gong, Z.; Chen, S. miR-16 Regulates Proliferation and Invasion of Lung Cancer Cells via the ERK/MAPK Signaling Pathway by Targeted Inhibition of MAPK kinase 1 (MEK1). Int. J. Med. Res. 2019, 10, 5194–5204. [Google Scholar] [CrossRef] [Green Version]
  132. Feng, Q.Q.; Dong, Z.Q.; Zhou, Y.; Zhang, H.; Long, C. miR-16-1-3p Targets TWIST1 to Inhibit Cell Proliferation and Invasion in NSCLC. Bratisl. Lek. Listy 2018, 119, 60–65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Reis, P.P.; Drigo, S.A.; Carvalho, R.F.; Lopez Lapa, R.M.; Felix, T.F.; Patel, D.; Cheng, D.; Pintilie, M.; Liu, G.; Tsao, M.S. Circulating miR-16-5p, miR-92a-3p, and miR-451a in plasma from lung cancer patients: Potential application in early detection and a regulatory role in tumorigenesis pathways. Cancers 2020, 12, 2071. [Google Scholar] [CrossRef] [PubMed]
  134. Ke, Y.; Zhao, W.; Xiong, J.; Cao, R. Downregulation of miR-16 promotes growth and motility by targeting HDGF in non-small cell lung cancer cells. FEBS Lett. 2013, 587, 3153–3157. [Google Scholar] [CrossRef] [Green Version]
  135. Navarro, A.; Diaz, T.; Gallardo, E.; Viñolas, N.; Marrades, R.M.; Gel, B.; Campayo, M.; Quera, A.; Bandres, E.; Garcia-Foncillas, J.; et al. Prognostic implications of miR-16 expression levels in resected non-small-cell lung cancer. J. Surg. Oncol. 2011, 103, 411–415. [Google Scholar] [CrossRef]
  136. Tung, Y.T.; Huang, P.W.; Chou, Y.C.; Lai, C.W.; Wang, H.P.; Ho, H.C.; Yen, C.C.; Tu, C.Y.; Tsai, T.C.; Yeh, D.C.; et al. Lung tumorigenesis induced by human vascular endothelial growth factor (hVEGF)-A165 overexpression in transgenic mice and amelioration of tumor formation by miR-16. Oncotarget 2015, 6, 10222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Ergun, S.; Güney, S.; Temiz, E.; Petrovic, N.; Gunes, S. Significance of miR-15a-5p and CNKSR3 as novel prognostic biomarkers in non-small cell lung cancer. Anticancer Agents Med. Chem. 2018, 18, 1695–1701. [Google Scholar] [CrossRef]
  138. Guo, S.; Li, M.; Li, J.; Lv, Y. Inhibition mechanism of lung cancer cell metastasis through targeted regulation of Smad3 by miR 15a. Oncol. Lett. 2020, 19, 1516–1522. [Google Scholar] [CrossRef]
  139. He, J. Knocking down MiR-15a expression promotes the occurrence and development and induces the EMT of NSCLC cells in vitro. Saudi J. Biol. Sci. 2017, 24, 1859. [Google Scholar]
  140. Yang, T.; Thakur, A.; Chen, T.; Yang, L.; Lei, G.; Liang, Y.; Zhang, S.; Ren, H.; Chen, M. MicroRNA-15a induces cell apoptosis and inhibits metastasis by targeting BCL2L2 in non-small cell lung cancer. Tumor Biol. 2015, 36, 4357–4365. [Google Scholar] [CrossRef]
  141. Kumar., S.; Sharawat, S.K.; Ali, A.; Gaur, V.; Malik, P.S.; Kumar, S.; Mohan, A.; Guleria, R. Identification of differentially expressed circulating serum microRNA for the diagnosis and prognosis of Indian non–small cell lung cancer patients. Curr. Probl. Cancer 2020, 44, 100540. [Google Scholar] [CrossRef]
  142. Liu, H.; Huang, J.; Peng, J.; Wu, X.; Zhang, Y.; Zhu, W.; Guo, L. Upregulation of the inwardly rectifying potassium channel Kir2.1 (KCNJ2) modulates multidrug resistance of small-cell lung cancer under the regulation of miR-7 and the Ras/MAPK pathway. Mol. Cancer 2015, 14, 59. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Lai, J.; Yang, H.; Zhu, Y.; Ruan, M.; Huang, Y.; Zhang, Q. MiR-7-5p-mediated downregulation of PARP1 impacts DNA homologous recombination repair and resistance to doxorubicin in small cell lung cancer. BMC Cancer 2019, 19, 602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Shang, A.Q.; Xie, Y.N.; Wang, J.; Sun, L.; Wei, J.; Lu, W.Y.; Lan, J.Y.; Wang, W.W.; Wang, L.; Wang, L.L. Predicative values of serum microRNA-22 and microRNA-126 levels for non-small cell lung cancer development and metastasis: A case-control study. Neoplasma 2017, 64, 453–459. [Google Scholar] [CrossRef]
  145. Yang, X.; Su, W.; Li, Y.; Zhou, Z.; Zhou, Y.; Shan, H.; Han, X.; Zhang, M.; Zhang, Q.; Bai, Y.; et al. MiR-22-3p suppresses cell growth via MET/STAT3 signaling in lung cancer. Am. J. Transl. Res. 2021, 13, 1221. [Google Scholar] [PubMed]
  146. Ling, B.; Wang, G.X.; Long, G.; Qiu, J.H.; Hu, Z.L. Tumor suppressor miR-22 suppresses lung cancer cell progression through post-transcriptional regulation of ErbB3. J. Cancer Res. Clin. Oncol. 2012, 138, 1355–1361. [Google Scholar] [CrossRef] [PubMed]
  147. Iorio, M.V.; Croce, C.M. Causes and Consequences of microRNA, Dysregulation. Cancer J. 2012, 18, 215–222. [Google Scholar] [CrossRef] [Green Version]
  148. Wang, I.X.; So, E.; Devlin, J.L.; Zhao, Y.; Wu, M.; Cheung, V.G. ADAR regulates RNA editing, transcript stability, and gene expression. Cell Rep. 2013, 5, 849–860. [Google Scholar] [CrossRef] [Green Version]
  149. Allegra, D.; Bilan, V.; Garding, A.; Döhner, H.; Stilgenbauer, S.; Kuchenbauer, F.; Mertens, D. Defective DROSHA processing contributes to downregulation of MiR-15/-16 in chronic lymphocytic leukemia. Leukemia 2014, 28, 98–107. [Google Scholar] [CrossRef]
  150. Su, X.; Chakravarti, D.; Cho, M.S.; Liu, L.; Gi, Y.J.; Lin, Y.L.; Leung, M.L.; El-Naggar, A.; Creighton, C.J.; Suraokar, M.B.; et al. TAp63 suppresses metastasis through coordinate regulation of Dicer and miRNAs. Nature 2010, 467, 986–990. [Google Scholar] [CrossRef] [Green Version]
  151. Karube, Y.; Tanaka, H.; Osada, H.; Tomida, S.; Tatematsu, Y.; Yanagisawa, K.; Yatabe, Y.; Takamizawa, J.; Miyoshi, S.; Mitsudomi, T.; et al. Reduced expression of Dicer associated with poor prognosis in lung cancer patients. Cancer Sci. 2005, 96, 111–115. [Google Scholar] [CrossRef]
  152. 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] [PubMed] [Green Version]
  153. Zhang, L.; Huang, J.; Yang, N.; Greshock, J.; Megraw, M.S.; Giannakakis, A.; Liang, S.; Naylor, T.L.; Barchetti, A.; Ward, M.R.; et al. microRNAs exhibit high frequency genomic alterations in human cancer. Proc. Natl. Acad. Sci. USA 2006, 103, 9136–9141. [Google Scholar] [CrossRef] [Green Version]
  154. Chin, L.J.; Ratner, E.; Leng, S.; Zhai, R.; Nallur, S.; Babar, I.; Muller, R.U.; Straka, E.; Su, L.; Burki, E.A.; et al. A SNP in a let-7 microRNA complementary site in the KRAS 3’ untranslated region increases non-small cell lung cancer risk. Cancer Res. 2008, 68, 8535–8540. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Shen, J.; Xia, W.; Khotskaya, Y.B.; Huo, L.; Nakanishi, K.; Lim, S.O.; Du, Y.; Wang, Y.; Chang, W.C.; Chen, C.H.; et al. EGFR modulates microRNA maturation in response to hypoxia through phosphorylation of AGO2. Nature 2013, 497, 383–387. [Google Scholar] [CrossRef] [PubMed]
  156. Davis-Dusenbery, B.N.; Hata, A. Mechanisms of control of microRNA biogenesis. J. Biochem. 2010, 148, 381–392. [Google Scholar]
  157. Du, J.; Johnson, L.M.; Jacobsen, S.E.; Patel, D.J. DNA methylation pathways and their crosstalk with histone methylation. Nat. Rev. Mol. Cell. Biol. 2015, 16, 519–532. [Google Scholar] [CrossRef] [Green Version]
  158. Liu, F.; Zhang, H.; Lu, S.; Wu, Z.; Zhou, L.; Cheng, Z.; Bai, Y.; Zhao, J.; Zhang, Q.; Mao, H. Quantitative assessment of gene promoter methylation in non small cell lung cancer using methylation sensitive high resolution melting. Oncol. Lett. 2018, 15, 7639–7648. [Google Scholar] [CrossRef] [Green Version]
  159. Wang, Z.; Chen, Z.; Gao, Y.; Li, N.; Li, B.; Tan, F.; Tan, X.; Lu, N.; Sun, Y.; Sun, L.; et al. DNA hypermethylation of microRNA-34b/c has prognostic value for stage nonsmall cell lung cancer. Cancer Biol. Ther. 2011, 11, 490–496. [Google Scholar] [CrossRef] [Green Version]
  160. Ceppi, P.; Mudduluru, G.; Kumarswamy, R.; Rapa, I.; Scagliotti, G.V.; Papotti, M.; Allgayer, H. Loss of miR-200c expression induces an aggressive, invasive and chemoresistant phenotype in non-small cell lung cancer. Mol. Cancer. Res. 2010, 8, 1207–1216. [Google Scholar] [CrossRef] [Green Version]
  161. Davalos, V.; Moutinho, C.; Villanueva, A.; Boque, R.; Silva, P.; Carneiro, F.; Esteller, M. Dynamic epigenetic regulation of the microRNA-200 family mediates epithelial and mesenchymal transitions in human tumorigenesis. Oncogene 2012, 31, 2062–2074. [Google Scholar] [CrossRef]
  162. Incoronato, M.; Urso, L.; Portela, A.; Laukkanen, M.O.; Soini, Y.; Quintavalle, C.; Keller, S.; Esteller, M.; Condorelli, G. Epigenetic regulation of miR-212 expression in lung cancer. PLoS ONE 2011, 6, e27722. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Yang, W.; Chendrimada, T.P.; Wang, Q.; Higuchi, M.; Seeburg, P.H.; Shiekhattar, R.; Nishikura, K. Modulation of microRNA processing and expression through RNA editing by ADAR deaminases. Nat. Struct. Mol. Biol. 2006, 13, 13–21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Nigita, G.; Distefano, R.; Veneziano, D.; Romano, G.; Rahman, M.; Wang, K.; Nana-Sinkam, P. Tissue and exosomal miRNA editing in non-small cell lung cancer. Sci. Rep. 2018, 8, 10222. [Google Scholar] [CrossRef] [PubMed]
  165. Yang, X.; Li, X.; Quan, X.; Li, H.; Hao, X.; Jiang, M.; Zhou, B. Association between two polymorphisms in the promoter region of miR-143/miR-145 and the susceptibility of lung cancer in northeast Chinese nonsmoking females. DNA Cell Biol. 2019, 38, 814–823. [Google Scholar] [CrossRef] [PubMed]
  166. Hayashita, Y.; Osada, H.; Tatematsu, Y.; Yamada, H.; Yanagisawa, K.; Tomida, S.; Yatabe, Y.; Kawahara, K.; Sekido, Y.; Takahashi, T. A polycistronic microRNA cluster, miR-17–92, is overexpressed in human lung cancers and enhances cell proliferation. Cancer Res. 2005, 65, 9628–9632. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Zhao, Y.L.; Zhang, J.X.; Yang, J.J.; Wei, Y.B.; Peng, J.F.; Fu, C.J.; Huang, M.H.; Wang, R.; Wang, P.Y.; Sun, G.B.; et al. MiR-205-5p promotes lung cancer progression and is valuable for the diagnosis of lung cancer. Thorac. Cancer 2022, 13, 832–843. [Google Scholar] [CrossRef]
  168. Lu, Y.; Zheng, W.; Rao, X.; Du, Y.; Xue, J. MicroRNA-9-5p Facilitates Lung Adenocarcinoma Cell Malignant Progression via Targeting STARD13. Biochem. Genet. 2022, 60, 1–6. [Google Scholar] [CrossRef]
  169. Qin, X.; Wang, X.Y.; Fei, J.W.; Li, F.H.; Han, J.; Wang, H.X. MiR-20a Promotes Lung Tumorigenesis by Targeting RUNX3 via TGF-β Signaling Pathway. J. Biol. Regul. Homeost. Agents 2020, 2, 34. [Google Scholar]
  170. Ren, P.; Gong, F.; Zhang, Y.; Jiang, J.; Zhang, H. MicroRNA-92a promotes growth, metastasis, and chemoresistance in non-small cell lung cancer cells by targeting PTEN, International Society of Oncology and BioMarkers (ISOBM). Tumor Biol. 2015, 37, 3215–3225. [Google Scholar] [CrossRef]
  171. Weidle, U.H.; Birzele, F.; Nopora, A. MicroRNAs as potential targets for therapeutic intervention with metastasis of non-small cell lung cancer. Cancer Genom. Proteom. 2019, 16, 99–119. [Google Scholar] [CrossRef] [Green Version]
  172. Kim, D.H.; Park, S.; Kim, H.; Choi, Y.J.; Kim, S.Y.; Sung, K.J.; Sung, Y.H.; Choi, C.M.; Yun, M.; Yi, Y.S.; et al. Tumor-derived exosomal miR-619-5p promotes tumor angiogenesis and metastasis through the inhibition of RCAN1. 4. Cancer Lett. 2020, 475, 2–13. [Google Scholar] [CrossRef] [PubMed]
  173. Chen, G.; Yu, L.; Dong, H.; Liu, Z.; Sun, Y. MiR-182 enhances radioresistance in non-small cell lung cancer cells by regulating FOXO3. Clin. Exp. Pharmacol. Physiol. 2019, 46, 137–143. [Google Scholar] [CrossRef] [PubMed]
  174. Fujita, S.; Ito, T.; Mizutani, T.; Minoguchi, S.; Yamamichi, N.; Sakurai, K.; Iba, H. miR-21 gene expression triggered by AP-1 was sustained through a double-negative feedback mechanism. J. Mol. Biol. 2008, 378, 492–504. [Google Scholar] [CrossRef] [PubMed]
  175. Cai, X.; Hagedorn, C.H.; Cullen, B.R. Human microRNAs were processed from capped, polyadenylated transcripts that can also function as mRNAs. RNA 2004, 10, 1957–1966. [Google Scholar] [CrossRef] [Green Version]
  176. Saab, A.A.R.; Raafat, R.H.; Alkady, M.S.; El Din, M.M.K. Clinical value of serum miR-21 as a potential biomarker in Non-Small Cell Lung Cancer (NSCLC). Med. Sci. 2021, 25, 1477–1485. [Google Scholar]
  177. Volinia, S.; Calin, G.A.; Liu, C.G.; Ambs, S.; Cimmino, A.; Petrocca, F.; Visone, R.; Iorio, M.; Roldo, C.; Ferracin, M.; et al. A microRNA expression signature of human solid tumors defines cancer gene targets. Proc. Natl. Acad. Sci. USA 2006, 103, 2257–2261. [Google Scholar] [CrossRef] [Green Version]
  178. Gabriely, G.; Wurdinger, T.; Kesari, S.; Esau, C.C.; Burchard, J.; Linsley, P.S.; Krichevsky, A.M. MicroRNA 21 promotes glioma invasion by targeting matrix metalloproteinase regulators. Mol. Cell. Biol. 2008, 28, 5369–5380. [Google Scholar] [CrossRef] [Green Version]
  179. Zhang, L.; Volinia, S.; Bonome, T.; Calin, G.A.; Greshock, J.; Yang, N.; Johnstone, C.N. Genomic and epigenetic alterations deregulate microRNA expression in human epithelial ovarian cancer. Proc. Natl. Acad. Sci. USA 2008, 105, 7004–7009. [Google Scholar] [CrossRef] [Green Version]
  180. Fei, J.; Lan, F.; Guo, M.; Li, Y.; Liu, Y. Inhibitory effects of anti-miRNA oligonucleotides (AMOs) on A549 cell growth. J. Drug Target. 2008, 16, 688–693. [Google Scholar] [CrossRef]
  181. Zhang, J.G.; Wang, J.J.; Zhao, F.; Liu, Q.; Jiang, K.; Yang, G.H. MicroRNA-21 (miR-21) represses tumor suppressor PTEN and promotes growth and invasion in non-small cell lung cancer (NSCLC). Clin. Chim. Acta 2010, 411, 846–852. [Google Scholar] [CrossRef]
  182. Seike, M.; Goto, A.; Okzhaano, T.; Bowman, E.D.; Schetter, A.J.; Horikawa, I.; Mathe, E.A.; Jen, J.; Yang, P.; Sugimura, H.; et al. MiR-21 was an EGFR-regulated antiapoptotic factor in lung cancer in never-smokers. Proc. Natl. Acad. Sci. USA 2009, 106, 12085–12090. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Shen, H.; Zhu, F.; Liu, J.; Xu, T.; Pei, D.; Wang, R.; Qian, Y.; Li, Q.; Wang, L.; Shi, Z.; et al. Alteration in Mir-21/PTEN expression modulates gefitinib resistance in non-small cell lung cancer. PLoS ONE 2014, 9, e103305. [Google Scholar] [CrossRef] [PubMed]
  184. Markou, A.; Tsaroucha, E.G.; Kaklamanis, L.; Fotinou, M.; Georgoulias, V.; Lianidou, E.S. Prognostic value of mature microRNA-21 and microRNA-205 overexpression in non- small cell lung cancer by quantitative real-time RT-PCR. Clin. Chem. 2008, 54, 54,1696–1704. [Google Scholar] [CrossRef] [PubMed]
  185. Wei, J.; Gao, W.; Zhu, C.J.; Liu, Y.Q.; Mei, Z.; Cheng, T.; Shu, Y.Q. Identification of plasma microRNA- 21 as a biomarker for early detection and chemosensitivity of non-small cell lung cancer. Chin. J. Cancer 2011, 30, 407–414. [Google Scholar] [CrossRef] [Green Version]
  186. Lin, L.; Tu, H.B.; Wu, L.; Liu, M.; Jiang, G.N. MicroRNA-21 regulates non-small cell lung cancer cell invasion and chemo-sensitivity through SMAD7. Cell Physiol. Biochem. 2016, 38, 2152–2162. [Google Scholar] [CrossRef] [Green Version]
  187. Zhang, Z.; Zhu, Z.; Watabe, K.; Zhang, X.; Bai, C.; Xu, M.; Wu, F.; Mo, Y.Y. Negative regulation of lncRNA GAS5 by miR-21. Cell Death Differ. 2013, 20, 1558–1568. [Google Scholar] [CrossRef] [Green Version]
  188. Cao, L.; Chen, J.; Ou, B.; Liu, C.; Zou, Y. GAS5 knockdown reduces the chemo-sensitivity of non-small cell lung cancer (NSCLC) cell to cisplatin (DDP) through regulating miR-21/ PTEN axis. Biomed. Pharmacother. 2017, 93, 570–579. [Google Scholar] [CrossRef]
  189. Tang, J.; Li, X.; Cheng, T.; Wu, J. miR-21-5p/SMAD7 axis promotes the progress of lung cancer. Thorac. Cancer 2021, 12, 2307–2313. [Google Scholar] [CrossRef]
  190. Zhou, X.; Liu, H.; Pang, Y.; Wang, M.; Liu, S. UTMD-mediated delivery of miR-21-5p inhibitor suppresses the development of lung cancer. Tissue Cell 2021, 23, 101719. [Google Scholar] [CrossRef]
  191. Grillari, J.; Hackl, M.; Grillari-Voglauer, R. miR-17-92 cluster: Ups and downs in cancer and aging. Biogerontology 2010, 11, 501–506. [Google Scholar] [CrossRef] [Green Version]
  192. Ota, A.; Tagawa, H.; Karnan, S.; Tsuzuki, S.; Karpas, A.; Kira, S.; Yoshida, Y.; Seto, M. Identification and characterization of a novel gene, C13orf25, as a target for 13q31-q32 amplification in malignant lymphoma. Cancer Res. 2004, 64, 3087–3095. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. He, L.; Thomson, J.M.; Hemann, M.T.; Hernando-Monge, E.; Mu, D.; Goodson, S.; Powers, S.; Cordon-Cardo, S.; Lowe, S.W.; Hannon, G.J.; et al. A microRNA polycistron as a potential human oncogene. Nature 2005, 435, 828–833. [Google Scholar] [CrossRef] [PubMed]
  194. Zhou, M.; Cai, J.; Tang, Y.; Zhao, Q. MiR-17-92 cluster is a novel regulatory gene of cardiac ischemic/reperfusion injury. Med. Hypotheses 2013, 81, 108–110. [Google Scholar] [CrossRef] [PubMed]
  195. Olive, V.; Bennett, M.J.; Walker, J.C.; Ma, C.; Jiang, I.; Cordon-Cardo, C.; Li, Q.J.; Lowe, S.W.; Hannon, G.J.; He, L. miR-19 is a key oncogenic component of mir-17-92. Genes Dev. 2009, 23, 2839–2849. [Google Scholar] [CrossRef] [Green Version]
  196. Zhang, Y.; Zhang, Y.; Yin, Y.; Li, S. Detection of circulating exosomal miR-17-5p serves as a novel non-invasive diagnostic marker for non-small cell lung cancer patients. Pathol. Res. Pract. 2019, 215, 152466. [Google Scholar] [CrossRef]
  197. Shen, Z.; Wu, X.; Wang, Z.; Li, B.; Zhu, X. Effect of miR-18a overexpression on the radiosensitivity of non-small cell lung cancer. Int. J. Clin. Exp. Pathol. 2015, 8, 643–648. [Google Scholar]
  198. Xiao, H.; Liu, Y.; Liang, P.; Wang, B.; Tan, H.; Zhang, Y.; Gao, J. TP53TG1 enhances cisplatin sensitivity of non-small cell lung cancer cells through regulating miR-18a/PTEN axis. Cell Biosci. 2018, 8, 23. [Google Scholar] [CrossRef]
  199. Li, J.; Yang, S.; Yan, W.; Yang, J.; Qin, Y.J.; Lin, X.L.; Xie, R.Y.; Wang, S.C.; Jin, W.; Gao, F.; et al. MicroRNA-19 triggers epithelial–mesenchymal transition of lung cancer cells accompanied by growth inhibition. Lab. Investig. 2015, 95, 1056–1070. [Google Scholar] [CrossRef] [Green Version]
  200. Woods, K.; Thomson, J.M.; Hammond, S.M. Direct regulation of an oncogenic micro-RNA cluster by E2F transcription factors. J. Biol. Chem. 2007, 282, 2130–2134. [Google Scholar] [CrossRef] [Green Version]
  201. Sylvestre, Y.; De Guire, V.; Querido, E.; Mukhopadhyay, U.K.; Bourdeau, V.; Major, F.; Ferbeyre, G.; Chartrand, P. An E2F/miR-20a autoregulatory feedback loop. J. Biol. Chem. 2007, 282, 2135–2143. [Google Scholar] [CrossRef] [Green Version]
  202. Novotny, G.W.; Sonne, S.B.; Nielsen, J.E.; Jonstrup, S.P.; Hansen, M.A.; Skakkebaek, N.E.; Meyts, E.; Kjems, J.; Leffers, H. Translational repression of E2F1 mRNA in carcinoma in situ and normal testis correlates with expression of the miR-17-92 cluster. Cell Death Differ. 2007, 14, 879–882. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Jiang, M.; Li, X.; Quan, X.; Li, X.; Zhou, B. MiR-92a Family: A Novel Diagnostic Biomarker and Potential Therapeutic Target in Human Cancers. Front. Mol. Biosci. 2019, 6, 98. [Google Scholar] [CrossRef] [PubMed]
  204. Ranade, A.R.; Cherba, D.; Sridhar, S.; Richardson, P.; Webb, C.; Paripati, A.; Bowles, B.; Weiss, G.J. MicroRNA 92a-2*: A biomarker predictive for chemoresistance and prognostic for survival in patients with small cell lung cancer. J. Thorac. Oncol. 2010, 5, 1273–1278. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Yang, C.; Jia, X.; Zhou, J.; Sun, Q.; Ma, Z. The MiR-17-92 gene cluster is a blood-based marker for cancer detection in non-small-cell lung cancer. Am. J. Med. Sci. 2020, 360, 248–260. [Google Scholar] [CrossRef]
  206. Galardi, S.; Mercatelli, N.; Giorda, E.; Massalini, S.; Frajese, G.V.; Ciafrè, S.A.; Farace, M.G. miR-221 and miR-222 expression affects the proliferation potential of human prostate carcinoma cell lines by targeting p27Kip1. J. Biol. Chem. 2007, 282, 23716–23724. [Google Scholar] [CrossRef] [Green Version]
  207. Mercatelli, N.; Coppola, V.; Bonci, D.; Miele, F.; Costantini, A.; Guadagnoli, M.; Bonanno, E.; Muto, G.; Frajese, G.V.; De Maria, R.; et al. The inhibition of the highly expressed miR-221 and miR-222 impairs the growth of prostate carcinoma xenografts in mice. PLoS ONE 2008, 3, e4029. [Google Scholar] [CrossRef]
  208. Visone, R.; Russo, L.; Pallante, P.; De Martino, I.; Ferraro, A.; Leone, V.; Borbone, E.; Petrocca, F.; Alder, H.; Croce, C.M.; et al. MicroRNAs (miR)-221 and miR-222, both overexpressed in human thyroid papillary carcinomas, regulate p27Kip1 protein levels and cell cycle. Endocr. Relat. Cancer 2007, 14, 791–798. [Google Scholar] [CrossRef]
  209. Zheng, P.L.; Zhang, N.; Liang, W.C.; Du, X.J.; Gong, X.W. miR-221 promotes proliferation of lung cancer A549 cells by down-regu-lating PTEN. Chin. J. Pathophysiol. 2017, 33, 2208–2211. [Google Scholar]
  210. Acunzo, M.; Visone, R.; Romano, G.; Veronese, A.; Lovat, F.; Palmieri, D.; Bottoni, A.; Garofalo, M.; Gasparini, P.; Condorelli, G.; et al. miR-130a targets MET and induces TRAIL-sensitivity in NSCLC by downregulating miR-221 and 222. Oncogene 2012, 31, 634–642. [Google Scholar] [CrossRef] [Green Version]
  211. Wu, S.; Chang, T.; Liu, Y.; Shih, J. MicroRNA in Lung Cancer Metastasis. Cancers 2019, 11, 265. [Google Scholar] [CrossRef] [Green Version]
  212. Sun, Y.J.; Li, J.; Chen, C.H. Effects of miR-221 on the apoptosis of non-small cell lung cancer cells by lncRNA HOTAIR. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 4226–4233. [Google Scholar] [PubMed]
  213. Ventura, L.; Gnetti, L.; Rossi, M.; Tiseo, M.; Giordano, G.; Corradi, M.; Silva, M.; Milanese, G.; Minari, R.; Leonetti, A.; et al. The role of miRNA-221 and miRNA-126 in patients with benign metastasizing leiomyoma of the lung: An overview with new interesting scenarios. Mol. Biol. Rep. 2021, 48, 3485–3494. [Google Scholar] [CrossRef] [PubMed]
  214. Abdel Ghany, S.M.; Ali, E.; Ahmed, A.; Hozayen, W.G.; Mohamed-Hussein, A.A.; Elnaggar, M.S.; Hetta, H.F. Circulating mirna-30a and mirna-221 AS novel biomarkers for the early detection of non-small-cell lung cancer. Middle East J. Cancer 2020, 11, 50–58. [Google Scholar]
  215. Wang, W.; Chen, D.; Chen, W.; Xin, Z.; Huang, Z.; Zhang, X.; Xi, K.; Wang, G.; Zhang, R.; Zhao, D.; et al. Early detection of Non-Small cell lung cancer by using a 12-microRNA panel and a nomogram for assistant diagnosis. Front. Oncol. 2020, 10, 855. [Google Scholar] [CrossRef]
  216. Mataki, H.; Seki, N.; Chiyomaru, T.; Enokida, H.; Goto, Y.; Kumamoto, T.; Machida, K.; Mizuno, K.; Nakagawa, M.; Inoue, H. Tumor-suppressive microRNA-206 as a dual inhibitor of MET and EGFR oncogenic signaling in lung squamous cell carcinoma. Int. J. Oncol. 2015, 46, 1039–1050. [Google Scholar] [CrossRef] [Green Version]
  217. Sun, C.; Liu, Z.; Li, S.; Yang, C.; Xue, R.; Xi, Y.; Wang, L.; Wang, S.; He, Q.; Huang, J.; et al. Down-regulation of c-Met and Bcl2 by microRNA-206, activates apoptosis, and inhibits tumor cell proliferation, migration and colony formation. Oncotarget 2015, 6, 625533–625574. [Google Scholar] [CrossRef] [Green Version]
  218. Chen, X.; Tong, Z.K.; Zhou, J.Y.; Yao, Y.K.; Zhang, S.M.; Zhou, J.Y. MicroRNA-206 inhibits the viability and migration of human lung adenocarcinoma cells partly by targeting MET. Oncol. Lett. 2016, 12, 1171–1177. [Google Scholar] [CrossRef] [Green Version]
  219. Zhang, J.X.; Yang, W.; Wu, J.Z.; Zhou, C.; Liu, S.; Shi, H.B.; Zhou, W.Z. MicroRNA-32-5p inhibits epithelial-mesenchymal transition and metastasis in lung adenocarcinoma by targeting SMAD family 3. J. Cancer 2021, 12, 2258. [Google Scholar] [CrossRef]
  220. Jasim, S.A.; Omear, H.A.; Al-Marzoqi, A.H. The role of miRNA-571 and miRNA-559 in lung cancer by affecting the expression of genes associated with the ErbB signaling pathway. Gene Rep. 2022, 26, 101436. [Google Scholar] [CrossRef]
  221. Gibbons, D.L.; Lin, W.; Creighton, C.J.; Rizvi, Z.H.; Gregory, P.A.; Goodall, G.J.; Thilaganathan, N.; Du, L.; Zhang, Y.; Pertsemlidis, A.; et al. Contextual extracellular cues promote tumor cell EMT and metastasis by regulating miR-200 family expression. Genes Dev. 2009, 23, 2140–2151. [Google Scholar] [CrossRef] [Green Version]
  222. Yang, N.; Liang, Y.; Zhu, T.; Long, Y.; Chen, Z.; Zhang, X.; Jiang, L. Epigenetic silencing of microRNA 199a 5p promotes the proliferation of non small cell lung cancer cells by increasing AKAP1 expression. Oncol. Lett. 2021, 21, 434. [Google Scholar] [CrossRef] [PubMed]
  223. Zhu, B.; Finch-Edmondson, M.; Lee, Y.; Wan, Y.; Sudol, M.; DasGupta, R. miR-582-5p is a tumor suppressor microRNA targeting the Hippo-YAP/TAZ signaling pathway in non-small cell lung cancer. Cancers 2021, 13, 756. [Google Scholar] [CrossRef] [PubMed]
  224. Liu, J.; Liu, S.; Deng, X.; Rao, J.; Huang, K.; Xu, G.; Wang, X. MicroRNA-582-5p suppresses non-small cell lung cancer cells growth and invasion via downregulating NOTCH1. PLoS ONE 2019, 14, e0217652. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Khandelwal, A.; Sharma, U.; Barwal, T.S.; Seam, R.K.; Gupta, M.; Rana, M.K.; Vasquez, K.M.; Jain, A. Circulating miR-320a Acts as a Tumor Suppressor and Prognostic Factor in Non-small Cell Lung Cancer. Front. Oncol. 2021, 11, 727. [Google Scholar] [CrossRef]
  226. Lee, S.B.; Park, Y.S.; Sung, J.S.; Lee, J.W.; Kim, B.; Kim, Y.H. Tumor suppressor miR-584-5p inhibits migration and invasion in smoking related non-small cell lung cancer cells by targeting YKT6. Cancers 2021, 13, 1159. [Google Scholar] [CrossRef]
  227. Luo, J.; Jin, Y.; Li, M.; Dong, L. Tumor suppressor miR 613 induces cisplatin sensitivity in non small cell lung cancer cells by targeting GJA1. Mol. Med. Rep. 2021, 23, 385. [Google Scholar]
  228. Reinhart, B.J.; Slack, F.J.; Basson, M.; Pasquinelli, A.E.; Bettinger, J. The 21 nucleotide let 7 RNA regulates developmental timing in Caenorhabditis elegans. Nature 2000, 403, 901–906. [Google Scholar] [CrossRef]
  229. Yamada, H.; Yanagisawa, K.; Tokumaru, S.; Taguchi, A.; Nimura, Y.; Osada, H.; Nagino, M.; Takahashi, T. Detailed characterization of a homozygously deleted region corresponding to a candidate tumor suppressor locus at 21q11-21 in human lung cancer. Genes Chromosomes Cancer 2008, 47, 810–818. [Google Scholar] [CrossRef]
  230. Boyerinas, B.; Park, S.M.; Hau, A.; Murmann, A.E.; Peter, M.E. The role of let-7 in cell differentiation and cancer. Endocr. Relat. Cancer 2010, 17, F19–F36. [Google Scholar] [CrossRef]
  231. Slack, F.J.; Joanne, B.W. MicroRNAs as a potential magic bullet in cancer. Future Oncol. 2006, 2, 73–82. [Google Scholar] [CrossRef]
  232. Trang, P.; Wiggins, J.F.; Daige, C.L.; Cho, C.; Omotola, M.; Brown, D.; Weidhaas, J.B.; Bader, A.G.; Slack, F.J. Systemic delivery of tumor suppressor microRNA mimics using a neutral lipid emulsion inhibits lung tumors in mice. Mol. Ther. 2011, 19, 1116–1122. [Google Scholar] [CrossRef] [PubMed]
  233. Xia, Y.; Zhu, Y.; Zhou, X.; Chen, Y. Low expression of let-7 predicts poor prognosis in patients with multiple cancers: A meta-analysis. Tumor Biol. 2014, 35, 5143–5148. [Google Scholar] [CrossRef] [PubMed]
  234. Ahmad, A.; Maitah, M.Y.; Ginnebaugh, K.R.; Li, Y.; Bao, B.; Gadgeel, S.M.; Sarkar, F.H. Inhibition of Hedgehog signalling sensitizes NSCLC cells to standard therapies through modulation of EMT-regulating miRNAs. J. Hematol. Oncol. 2013, 6, 77. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Yin, J.; Hu, W.; Pan, L.; Fu, W.; Dai, L.; Jiang, Z.; Zhang, F.; Zhao, J. let-7 and miR-17 promote self-renewal and drive gefitinib resistance in non-small cell lung cancer. Oncol. Rep. 2019, 42, 495–508. [Google Scholar] [CrossRef]
  236. Viswanathan, S.R.; Daley, G.Q. Lin28: A microRNA regulator with a macro role. Cell 2010, 140, 445–449. [Google Scholar] [CrossRef] [Green Version]
  237. Zhang, W.T.; Zhang, G.X.; Gao, S.S. The Potential Diagnostic Accuracy of Let-7 Family for Cancer: A Meta-Analysis. Technol. Cancer Res. Treat. 2021, 20, 15330338211033061. [Google Scholar] [CrossRef]
  238. Liu, J.K.; Liu, H.F.; Ding, Y.; Gao, G.D. Predictive value of microRNA let-7a expression for efficacy and prognosis of radiotherapy in patients with lung cancer brain metastasis: A case–control study. Medicine 2018, 97, e12847. [Google Scholar] [CrossRef]
  239. Wang, F.; Quan, Q. The long non-coding RNA SNHG4/microRNA-let-7e/KDM3A/p21 pathway is involved in the development of non-small cell lung cancer. Mol. Ther.-Oncolytics 2021, 20, 634–645. [Google Scholar] [CrossRef]
  240. Feng, X.; Wang, Z.; Fillmore, R.; Xi, Y. MiR-200, a new star miRNA in human cancer. Cancer Lett. 2014, 344, 166–173. [Google Scholar] [CrossRef] [Green Version]
  241. Wellner, U.; Schubert, J.; Burk, U.C.; Schmalhofer, O.; Zhu, F.; Sonntag, A.; Waldvogel, B.; Vannier, C.; Darling, D.; Zur Hausen, A.; et al. The EMT- activator ZEB1 promotes tumorigenicity by repressing stemness-inhibiting microRNAs. Nat. Cell Biol. 2009, 11, 1487–1495. [Google Scholar] [CrossRef]
  242. Hill, L.; Browne, G.; Tulchinsky, E. ZEB/miR-200 feedback loop: At the crossroads of signal transduction in cancer. Int. J. Cancer 2013, 132, 745–754. [Google Scholar] [CrossRef] [PubMed]
  243. Roybal, J.D.; Zang, Y.; Ahn, Y.H.; Yang, Y.; Gibbons, D.L.; Baird, B.N.; Alvarez, C.; Thilaganathan, N.; Liu, D.D.; Saintigny, P.; et al. miR-200 inhibits lung adenocarcinoma cell invasion and metastasis by targeting Flt1/VEGFR1. Mol. Cancer Res. 2011, 9, 25–35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Hess, A.K.; Muer, A.; Mairinger, F.D.; Weichert, W.; Stenzinger, A.; Hummel, M.; Budach, V.; Tinhofer, I. MiR-200b and miR-155 as predictive biomarkers for the efficacy of chemoradiation in locally advanced head and neck squamous cell carcinoma. Eur. J. Cancer 2017, 77, 3–12. [Google Scholar] [CrossRef] [PubMed]
  245. Cortez, M.A.; Valdecanas, D.; Zhang, X.; Zhan, Y.; Bhardwaj, V.; Calin, G.A.; Komaki, R.; Giri, D.K.; Quini, C.C.; Wolfe, T.; et al. Therapeutic delivery of miR-200c enhances radiosensitivity in lung cancer. Mol. Ther. 2014, 22, 1494–1503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Nishijima, N.; Seike, M.; Soeno, C.; Chiba, M.; Miyanaga, A.; Noro, R.; Sugano, T.; Matsumoto, M.; Kubota, K.; Gemma, A. miR-200/ZEB axis regulates sensitivity to nintedanib in non-small cell lung cancer cells. Int. J. Oncol. 2016, 48, 937–944. [Google Scholar] [CrossRef] [Green Version]
  247. Zhang, N.; Liu, Y.; Wang, Y.; Zhao, M.; Tu, L.; Luo, F. Decitabine reverses TGF-β1-induced epithelial–mesenchymal transition in non-small-cell lung cancer by regulating miR-200/ZEB axis. Drug Des. Dev. Ther. 2017, 11, 969. [Google Scholar] [CrossRef] [Green Version]
  248. Kim, E.J.; Kim, J.S.; Lee, S.; Lee, H.; Yoon, J.S.; Hong, J.H.; Chun, S.H.; Sun, D.S.; Won, H.S.; Hong, S.A.; et al. QKI, a miR-200 target gene, suppresses epithelial-to-mesenchymal transition and tumor growth. Int. J. Cancer 2019, 145, 1585–1595. [Google Scholar] [CrossRef]
  249. Hydbring, P.; De Petris, L.; Zhang, Y.; Brandén, E.; Koyi, H.; Novak, M.; Kanter, L.; Hååg, P.; Hurley, J.; Tadigotla, V.; et al. Exosomal RNA-profiling of pleural effusions identifies adenocarcinoma patients through elevated miR-200 and LCN2 expression. Lung Cancer 2018, 124, 45–52. [Google Scholar] [CrossRef]
  250. Xue, B.; Chuang, C.H.; Prosser, H.M.; Fuziwara, C.S.; Chan, C.; Sahasrabudhe, N.; Kühn, M.; Wu, Y.; Chen, J.; Biton, A. miR-200 deficiency promotes lung cancer metastasis by activating Notch signaling in cancer-associated fibroblasts. Genes Dev. 2021, 35, 1109–1122. [Google Scholar] [CrossRef]
  251. Koutsoulidou, A.; Mastroyiannopoulos, N.P.; Furling, D.; Uney, J.B.; Phylactou, L.A. Expression of miR-1, miR-133a, miR-133b and miR-206 increases during development of human skeletal muscle. BMC Dev. Biol. 2011, 11, 34. [Google Scholar] [CrossRef] [Green Version]
  252. Xue, D.; Yang, Y.; Liu, Y.; Wang, P.; Dai, Y.; Liu, Q.; Chen, L.; Shen, J.; Ju, H.; Li, Y.; et al. MicroRNA- 206 attenuates the growth and angiogenesis in non-small cell lung cancer cells by blocking the 14-3- 3zeta/STAT3/HIF-1alpha/VEGF signaling. Oncotarget 2016, 7, 79805–79813. [Google Scholar] [CrossRef] [PubMed]
  253. Shen, H.; Yu, X.; Yang, F.; Zhang, Z.; Shen, J.; Sun, J.; Choksi, S.; Jitkaew, S.; Shu, Y. Reprogramming of Normal Fibroblasts into Cancer-Associated Fibroblasts by miRNAs-Mediated CCL2/VEGFA Signaling. PLoS Genet. 2016, 12, e1006244. [Google Scholar] [CrossRef] [PubMed]
  254. Chen, Q.Y.; Jiao, D.M.; Wang, J.; Hu, H.; Tang, X.; Chen, J.; Mou, H.; Lu, W. miR-206 regulates cisplatin resistance and EMT in human lung adenocarcinoma cells partly by targeting MET. Oncotarget 2016, 7, 24510–24526. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Singh, A.; Happel, C.; Manna, S.K.; Acquaah-Mensah, G.; Carrerero, J.; Kumar, S.; Nasipuri, P.; Krausz, K.W.; Wakabayashi, N.; Dewi, R.; et al. Transcription factor NRF2 regulates miR-1 and miR-206 to drive tumorigenesis. J. Clin. Investig. 2013, 123, 2921–2934. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Jia, K.G.; Feng, G.; Tong, Y.S.; Tao, G.Z.; Xu, L. miR-206 regulates non-small-cell lung cancer cell aerobic glycolysis by targeting hexokinase 2. J. Biochem. 2020, 167, 365–370. [Google Scholar] [CrossRef]
  257. Zhao, L.; Zhang, X.; Shi, Y.; Teng, T. LncRNA SNHG14 contributes to the progression of NSCLC through miR-206/G6PD pathway. Thorac. Cancer 2020, 11, 1202–1210. [Google Scholar] [CrossRef]
  258. Liao, M.; Peng, L. MiR-206 may suppress non-small lung cancer metastasis by targeting CORO1C. Cell. Mol. Biol. Lett. 2020, 25, 22. [Google Scholar] [CrossRef]
  259. Li, W.; Liu, Y.; Li, Z.J.; Shi, Y.; Deng, J.; Bai, J.; Ma, L.; Zeng, X.X.; Feng, S.S.; Ren, J.L.; et al. Unravelling the role of LncRNA WT1-AS/mIR-206/NAMPT axis as prognostic biomarkers in lung adenocarcinoma. Biomolecules 2021, 11, 203. [Google Scholar] [CrossRef]
  260. Jiao, D.; Chen, J.; Li, Y.; Tang, X.; Wang, J.; Xu, W.; Song, J.; Li, Y.; Tao, H.; Chen, Q. miR-1-3p and miR-206 sensitizes HGF-induced gefitinib-resistant human lung cancer cells through inhibition of c-Met signalling and EMT. J. Cell. Mol. Med. 2018, 22, 3526–3536. [Google Scholar] [CrossRef]
  261. Wang, J.; Bao, Z.; Qiao, Y. miR-206 regulates EVI1 gene expression and cell biological behavior in stem cells of small cell lung cancer. Chin. J. Tissue Eng. Res. 2022, 26, 1027. [Google Scholar]
  262. Taganov, K.D.; Boldin, M.P.; Chang, K.J.; Baltimore, D. NF-kappaB-dependent induction of microRNA miR-146, an inhibitor targeted to signaling proteins of innate immune responses. Proc. Natl. Acad. Sci. USA. 2006, 103, 12481–12486. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  263. Li, Y.; VandenBoom, T.G.; Wang, Z.; Kong, D.; Ali, S.; Philip, P.A.; Sarkar, F.H. miR-146a suppresses invasion of pancreatic cancer cells. Cancer Res. 2010, 70, 1486–1495. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Iacona, J.R.; Monteleone, N.J.; Lutz, C.S. miR-146a suppresses 5-lipoxygenase activating protein (FLAP) expression and Leukotriene B4 production in lung cancer cells. Oncotarget 2018, 9, 26751–26769. [Google Scholar] [CrossRef] [PubMed]
  265. Boldin, M.P.; Taganov, K.D.; Rao, D.S.; Yang, L.; Zhao, J.L.; Kalwani, M.; Garcia-Flores, Y.; Luong, M.; Devrekanli, A.; Xu, J.; et al. miR-146a is a significant brake on autoimmunity, myeloproliferation, and cancer in mice. J. Exp. Med. 2011, 208, 1189–1201. [Google Scholar] [CrossRef]
  266. Xu, T.; Zhu, Y.; Wei, Q.K.; Yuan, Y.; Zhou, F.; Ge, Y.Y.; Yang, J.R.; Su, H.; Zhuang, S.M. A functional polymorphism in the miR-146a gene is associated with the risk for hepatocellular carcinoma. Carcinogenesis 2008, 29, 2126–2131. [Google Scholar] [CrossRef]
  267. Xu, B.; Feng, N.H.; Li, P.C.; Tao, J.; Wu, D.; Zhang, Z.D.; Tong, N.; Wang, J.F.; Song, N.H.; Zhang, W.; et al. A functional polymorphism in Pre-miR-146a gene is associated with prostate cancer risk and mature miR-146a expression in vivo. Prostate 2010, 70, 467–472. [Google Scholar] [CrossRef]
  268. Pang, L.; Lu, J.; Huang, J.; Xu, C.; Li, H.; Yuan, G.; Cheng, X.; Chen, J. Upregulation of miR 146a increases cisplatin sensitivity of the non small cell lung cancer A549 cell line by targeting JNK 2. Oncol. Lett. 2017, 14, 7745–7752. [Google Scholar] [CrossRef] [Green Version]
  269. Shi, L.; Xu, Z.; Wu, G.; Chen, X.; Huang, Y.; Wang, Y.; Jiang, W.; Ke, B. Up-regulation of miR-146a increases the sensitivity of non-small cell lung cancer to DDP by downregulating cyclin. BMC Cancer 2017, 17, 1–4. [Google Scholar] [CrossRef] [Green Version]
  270. Sun, X.; Cui, S.; Fu, X.; Liu, C.; Wang, Z.; Liu, Y. MicroRNA-146-5p promotes proliferation, migration and invasion in lung cancer cells by targeting claudin-12. Cancer Biomark. 2019, 25, 89–99. [Google Scholar] [CrossRef]
  271. Vashchenko, V.I.; Romashova, J.E.; Shabanov, P.D. Pathophysiology of miR-146a in lung cancer. Prospects of rising of efficiency of targeted therapy. Rev. Clin. Pharmacol. Drug Ther. 2021, 19, 359–381. [Google Scholar] [CrossRef]
  272. Pereira, D.M.; Rodrigues, P.M.; Borralho, P.M.; Rodrigues, C.M. Delivering the promise of miRNA cancer therapeutics. Drug Discov. Today 2013, 18, 282–289. [Google Scholar] [CrossRef] [PubMed]
  273. Ozcan, G.; Ozpolat, B.; Coleman, R.L.; Sood, A.K.; Lopez-Berestein, G. Preclinical and clinical development of siRNA-based therapeutics. Adv. Drug Deliv. Rev. 2015, 87, 108–119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  274. Bouchie, A. First microRNA mimic enters clinic. Nat. Biotechnol. 2013, 31, 577. [Google Scholar] [CrossRef] [PubMed]
  275. Talekar, M.; Trivedi, M.; Shah, P.; Ouyang, Q.; Oka, A.; Gandham, S.; Amiji, M.M. Combination wt-p53 and MicroRNA-125b transfection in a genetically engineered lung cancer model using dual CD44/EGFR-targeting nanoparticles. Mol. Ther. 2016, 24, 759–769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  276. Jiang, Q.; Yuan, Y.; Gong, Y.; Luo, X.; Su, X.; Hu, X.; Zhu, W. Therapeutic delivery of microRNA-143 by cationic lipoplexes for non-small cell lung cancer treatment in vivo. J. Cancer Res. Clin. Oncol. 2019, 145, 2951–2967. [Google Scholar] [CrossRef]
  277. Jeong, K.; Yu, Y.J.; You, J.Y.; Rhee, W.J.; Kim, J.A. Exosome-mediated microRNA-497 delivery for anti-cancer therapy in a microfluidic 3D lung cancer model. Lab Chip 2020, 20, 548–557. [Google Scholar] [CrossRef]
  278. Watashi, K.; Yeung, M.L.; Starost, M.F.; Hosmane, R.S.; Jeang, K.T. Identification of small molecules that suppress microRNA function and reverse tumorigenesis. J. Biol. Chem. 2010, 285, 24707–24716. [Google Scholar] [CrossRef] [Green Version]
  279. Guo, S.; Zhang, L.; Zhang, Y.; Wu, Z.; He, D.; Li, X.; Wang, Z. Long non-coding RNA TUG1 enhances chemosensitivity in non-small cell lung cancer by impairing microRNA-221-dependent PTEN inhibition. Aging 2019, 11, 7553. [Google Scholar] [CrossRef]
  280. Esau, C.; Davis, S.; Murray, S.F.; Yu, X.X.; Pandey, S.K.; Pear, M.; Watts, L.; Booten, S.L.; Graham, M.; McKay, R.; et al. miR-122 regulation of lipid metabolism revealed by in vivo antisense targeting. Cell Metabol. 2006, 3, 87–98. [Google Scholar] [CrossRef] [Green Version]
  281. Vester, B.; Wengel, J. LNA (locked nucleic acid): High-affinity targeting of complementary RNA and DNA. Biochemistry 2004, 43, 3233–3241. [Google Scholar] [CrossRef]
  282. Duan, F.G.; Wang, M.F.; Cao, Y.B.; Li, D.; Li, R.Z.; Fan, X.X.; Khan, I.; Lai, H.L.; Zhang, Y.Z.; Hsiao, W.W.; et al. MicroRNA-421 confers paclitaxel resistance by binding to the KEAP1 3′ UTR and predicts poor survival in non-small cell lung cancer. Cell Death Dis. 2019, 10, 1–4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  283. He, X.; Li, C.; Wu, X.; Yang, G. Docetaxel inhibits the proliferation of non-small-cell lung cancer cells via upregulation of microRNA-7 expression. Int. J. Clin. Exp. Pathol. 2015, 8, 9072. [Google Scholar] [PubMed]
  284. Jin, H.; Qiao, F.; Wang, Y.; Xu, Y.; Shang, Y. Curcumin inhibits cell proliferation and induces apoptosis of human non-small cell lung cancer cells through the upregulation of miR-192-5p and suppression of PI3K/Akt signaling pathway. Oncol. Rep. 2015, 34, 2782–2789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  285. Lelli, D.; Sahebkar, A.; Johnston, T.P.; Pedone, C. Curcumin use in pulmonary diseases: State of the art and future perspectives. Pharmacol. Res. 2017, 115, 133–148. [Google Scholar] [CrossRef]
  286. Wang, N.; Feng, T.; Liu, X.; Liu, Q. Curcumin inhibits migration and invasion of non-small cell lung cancer cells through up-regulation of miR-206 and suppression of PI3K/AKT/mTOR signaling pathway. Acta Pharm. 2020, 70, 399–409. [Google Scholar] [CrossRef] [Green Version]
  287. Cortez, M.A.; Bueso-Ramos, C.; Ferdin, J.; Lopez-Berestein, G.; Sood, A.K.; Calin, M.A. MicroRNAs in body fluids—the mix of hormones and biomarkers. Nat. Rev. Clin. Oncol. 2011, 8, 467–477. [Google Scholar] [CrossRef] [Green Version]
  288. Mitchell, P.S.; Parkin, R.K.; Kroh, E.M.; Fritz, B.R.; Wyman, S.K.; Pogosova-Agadjanyan, E.L.; Peterson, A.; Noteboom, J.; O’Briant, K.C.; Allen, A.; et al. Circulating microRNAs as stable blood-based markers for cancer detection. Proc. Natl. Acad. Sci. USA 2008, 105, 10513–10518. [Google Scholar] [CrossRef] [Green Version]
  289. Ferracin, M.; Veronese, A.; Negrini, M. Micromarkers: miRNAs in cancer diagnosis and prognosis. Expert Rev. Mol. Diagn. 2010, 10, 297–308. [Google Scholar] [CrossRef]
  290. Yu, S.L.; Chen, H.Y.; Chang, G.C.; Chen, C.Y.; Chen, H.W.; Singh, S.; Cheng, C.L.; Yu, C.J.; Lee, Y.C.; Chen, H.S.; et al. MicroRNA signature predicts survival and relapse in lung cancer. Cancer Cell 2008, 13, 48–57. [Google Scholar] [CrossRef] [Green Version]
  291. Dvinge, H.; Git, A.; Gräf, S.; Salmon-Divon, M.; Curtis, C.; Sottoriva, A.; Zhao, Y.; Hirst, M.; Armisen, J.; Miska, E.A.; et al. The shaping and functional consequences of the microRNA landscape in breast cancer. Nature 2013, 497, 378–382. [Google Scholar] [CrossRef]
  292. Yanaihara, N.; Caplen, N.; Bowman, E.; Seike, M.; Kumamoto, K.; Yi, M.; Stephens, R.M.; Okamoto, A.; Yokota, J.; Tanaka, T.; et al. Unique microRNA molecular profiles in lung cancer diagnosis and prognosis. Cancer Cell 2006, 9, 89–198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  293. Foss, K.M.; Sima, C.; Ugolini, D.; Neri, M.; Allen, K.E.; Weiss, G.J. miR-1254 and miR-574-5p: Serum based microRNA biomarkers for early-stage non-small cell lung cancer. J. Thorac. Oncol. 2011, 6, 482–488. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  294. Zheng, D.; Haddadin, H.; Wang, Y.; Gu, L.Q.; Perry, M.C.; Freter, C.E.; Wang, M.X. Plasma microRNAs as novel biomarkers for early detection of lung cancer. Int. J. Clin. Exp. Pathol. 2011, 4, 575–586. [Google Scholar] [PubMed]
  295. Hou, J.; Meng, F.; Chan, L.W.; Cho, W.; Wong, S.C. Circulating plasma MicroRNAs as diagnostic markers for NSCLC. Front. Gentics 2016, 7, 193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  296. Rosenfeld, N.; Aharonov, R.; Meiri, E.; Rosenwald, S.; Spector, Y.; Zepeniuk, M.; Benjamin, H.; Shabes, N.; Tabak, S.; Levy, A.; et al. MicroRNAs accurately identify cancer tissue origin. Nat. Biotechnol. 2008, 26, 462–469. [Google Scholar] [CrossRef] [PubMed]
  297. Peng, W.; Wang, J.; Shan, B.; Peng, Z.; Dong, Y.; Shi, W.; He, D.; Cheng, Y.; Zhao, W.; Zhang, C.; et al. Diagnostic and prognostic potential of circulating long non-coding RNAs in non-small cell lung cancer. Cell Physiol. Biochem. 2018, 49, 816–827. [Google Scholar] [CrossRef]
  298. Xie, Y.; Todd, N.W.; Liu, Z.; Zhan, M.; Fang, H.; Peng, H.; Alattar, M.; Deepak, J.; Stass, S.A.; Jiang, F. Altered miRNA expression in sputum for diagnosis of non-small cell lung cancer. Lung Cancer 2010, 67, 170–176. [Google Scholar] [CrossRef] [Green Version]
  299. Bianchi, F.; Nicassio, F.; Marzi, M.; Belloni, E.; Dall’olio, V.; Bernard, L.; Pelosi, G.; Maisonneuve, P.; Veronesi, G.; Di Fiore, P.P. A serum circulating miRNA diagnostic test to identify asymptomatic high-risk individuals with early-stage lung cancer. EMBO Mol. Med. 2011, 3, 495–503. [Google Scholar] [CrossRef]
  300. Liao, J.; Shen, J.; Leng, Q.; Qin, M.; Zhan, M.; Jiang, F. MicroRNA-based biomarkers for diagnosis of non-small cell lung cancer (NSCLC). Thorac. Cancer 2020, 11, 762–768. [Google Scholar] [CrossRef] [Green Version]
  301. Lebanony, D.; Benjamin, H.; Gilad, S.; Ezagouri, M.; Dov, A.; Ashkenazi, K.; Gefen, N.; Izraeli, S.; Rechavi, G.; Pass, H.; et al. Diagnostic assay based on hsa-miR-205 expression distinguishes squamous from nonsquamous non-small-cell lung carcinoma. J. Clin. Oncol. 2009, 27, 2030–2037. [Google Scholar] [CrossRef]
  302. Landi, M.T.; Zhao, Y.; Rotunno, M.; Koshiol, J.; Liu, H.; Bergen, A.W.; Rubagotti, M.; Goldstein, A.M.; Linnoila, I.; Marincola, F.M.; et al. MicroRNA expression differentiates histology and predicts survival of lung cancer. Clin. Cancer. Res. 2010, 16, 430–441. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  303. Qi, J.; Mu, D. MicroRNAs and lung cancers: From pathogenesis to clinical implications. Front. Med. 2012, 6, 134–155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Barshack, I.; Lithwick-Yanai, G.; Afek, A.; Rosenblatt, K.; Tabibian-Keissar, H.; Zepeniuk, M.; Cohen, L.; Dan, H.; Zion, O.; Strenov, Y.; et al. MicroRNA expression differentiates between primary lung tumors and metastases to the lung. Pathol. Res. Pract. 2010, 206, 578–584. [Google Scholar] [CrossRef] [PubMed]
  305. Kryczka, J.; Migdalska-Sęk, M.; Kordiak, J.; Kiszałkiewicz, J.M.; Pastuszak-Lewandoska, D.; Antczak, A.; Brzeziańska-Lasota, E. Serum Extracellular Vesicle-Derived miRNAs in Patients with Non-Small Cell Lung Cancer—Search for Non-Invasive Diagnostic Biomarkers. Diagnostics 2021, 11, 425. [Google Scholar] [CrossRef]
  306. Jiang, H.G.; Dai, C.H.; Xu, Y.P.; Jiang, Q.; Xia, X.B.; Shu, Y.; Li, J. Four plasma miRNAs act as biomarkers for diagnosis and prognosis of non small cell lung cancer. Oncol. Lett. 2021, 22, 792. [Google Scholar] [CrossRef]
  307. MacDonagh, L.; Gallagher, M.F.; Ffrench, B.; Gasch, C.; Gray, S.G.; Reidy, M.; Nicholson, S.; Leonard, N.; Ryan, R.; Young, V.; et al. MicroRNA expression profiling and biomarker validation in treatment-naïve and drug resistant non-small cell lung cancer. Transl. Lung Cancer Res. 2021, 10, 1773. [Google Scholar] [CrossRef]
  308. Kong, D.; Wang, K.; Zhang, Q.N.; Bing, Z.T. Systematic analysis reveals key microRNAs as diagnostic and prognostic factors in progressive stages of lung cancer. arXiv 2022, arXiv:2201.05408. [Google Scholar]
  309. Visan, K.S.; Lobb, R.J.; Wen, S.W.; Bedo, J.; Lima, L.G.; Krumeich, S.; Palma, C.; Ferguson, K.; Green, B.; Niland, C.; et al. Blood-Derived Extracellular Vesicle-Associated miR-3182 Detects Non-Small Cell Lung Cancer Patients. Cancers 2022, 14, 257. [Google Scholar] [CrossRef]
  310. Montani, F.; Marzi, M.J.; Dezi, F.; Dama, E.; Carletti, R.M.; Bonizzi, G.; Bertolotti, R.; Bellomi, M.; Rampinelli, C.; Maisonneuve, P.; et al. miR-Test: A blood test for lung cancer early detection. J. Natl. Cancer. Inst. 2015, 107, djv063. [Google Scholar] [CrossRef] [Green Version]
  311. Johnstone, R.M.; Adam, M.; Hammond, J.R.; Orr, L.; Turbide, C. Vesicle formation during reticulocyte maturation. Association of plasma membrane activities with released vesicles (exosomes). J. Biol. Chem. 1987, 262, 9412–9420. [Google Scholar] [CrossRef]
  312. Chen, J.; Chopp, M. Exosome therapy for stroke. Stroke 2018, 49, 1083–1090. [Google Scholar] [CrossRef] [PubMed]
  313. Liu, H.; Li, B. The functional role of exosome in hepatocellular carcinoma. J. Cancer Res. Clin. Oncol. 2018, 144, 2085–2095. [Google Scholar] [CrossRef] [PubMed]
  314. Prattichizzo, F.; Micolucci, L.; Cricca, M.; De Carolis, S.; Mensà, E.; Ceriello, A.; Procopio, A.D.; Bonafè, M.; Olivieri, F. Exosome-based immunomodulation during aging: A nano-perspective on inflamm-aging. Mech. Ageing Dev. 2017, 168, 44–53. [Google Scholar] [CrossRef]
  315. Sruthi, T.V.; Edatt, L.; Raji, G.R.; Kunhiraman, H.; Shankar, S.S.; Shankar, V.; Ramachandran, V.; Poyyakkara, A.; Kumar, S.V. Horizontal transfer of miR-23a from hypoxic tumor cell colonies can induce angiogenesis. J. Cell. Physiol. 2018, 233, 3498–3514. [Google Scholar] [CrossRef]
  316. Fu, X.; Liu, M.; Qu, S.; Ma, J.; Zhang, Y.; Shi, T.; Wen, H.; Yang, Y.; Wang, S.; Wang, J.; et al. Exosomal microRNA-32-5p induces multidrug resistance in hepatocellular carcinoma via the PI3K/Akt pathway. J. Exp. Clin. Cancer Res. 2018, 37, 52. [Google Scholar] [CrossRef] [Green Version]
  317. Głuszko, A.; Szczepański, M.J.; Ludwig, N. Exosomes in cancer: Circulating immune-related biomarkers. Biomed. Res. Int. 2019, 2019, 1628029. [Google Scholar] [CrossRef] [PubMed]
  318. Kok, V.C.; Yu, C.C. Cancer-derived exosomes: Their role in cancer biology and biomarker development. Int. J. Nanomed. 2020, 15, 8019–8036. [Google Scholar] [CrossRef] [PubMed]
  319. Cazzoli, R.; Buttitta, F.; Di Nicola, M.; Malatesta, S.; Marchetti, A.; Rom, W.N.; Pass, H.I. microRNAs derived from circulating exosomes as noninvasive biomarkers for screening and diagnosing lung cancer. J. Thorac. Oncol. 2013, 8, 1156–1162. [Google Scholar] [CrossRef] [Green Version]
  320. Munagala, R.; Aqil, F.; Gupta, R.C. Exosomal miRNAs as biomarkers of recurrent lung cancer. Tumor Biol. 2016, 37, 10703–10714. [Google Scholar] [CrossRef]
  321. Lamichhane, S.R.; Thachil, T.; De Ieso, P.; Gee, H.; Moss, S.A.; Milic, N. Prognostic role of MicroRNAs in human non-small-cell lung Cancer: A systematic review and Meta-analysis. Dis. Markers 2018, 2018, 8309015. [Google Scholar] [CrossRef] [Green Version]
  322. Xiao, W.; Zhong, Y.; Wu, L.; Yang, D.; Ye, S.; Zhang, M. Prognostic value of microRNAs in lung cancer: A systematic review and meta analysis. Mol. Clin. Oncol. 2019, 10, 67–77. [Google Scholar] [CrossRef] [Green Version]
  323. Zhang, Y.; Roth, J.A.; Yu, H.; Ye, Y.; Xie, K.; Zhao, H.; Chang, D.W.; Huang, M.; Li, H.; Qu, J.; et al. A 5-microRNA signature identified from serum microRNA profiling predicts survival in patients with advanced stage non-small cell lung cancer. Carcinogenesis 2019, 40, 643–650. [Google Scholar] [CrossRef] [PubMed]
  324. Yan, H.; Xin, S.; Ma, J.; Wang, H.; Zhang, H.; Liu, J. A three microRNA-based prognostic signature for small cell lung cancer overall survival. J. Cell. Biochem. 2019, 120, 8723–8730. [Google Scholar] [CrossRef]
  325. Nakhla, S.; Kamel, M.K.; Moussa, S.; Ali, I.M.; Elnaggar, M.; Talima, S.A.; Fitoury, S.; Ibrahim, L.; Morsi, M.I. The diagnostic and predictive values of miRNA-21, miRNA-126, miRNA-513a, and miRNA-98 in patients with non-small cell lung cancer (NSCLC). Int. J. Cancer. Biomed. Res. 2021, 5, 109–120. [Google Scholar] [CrossRef]
  326. Zhu, S.J.; Wang, X.; Hu, S.L.; Fang, Y.; Guan, B.X.; Li, J.; Li, G.; Xu, J.Y. Clinical Significance and Biological Function of miR-1274a in Non-small Cell Lung Cancer. Mol. Biotechnol. 2022, 64, 9–16. [Google Scholar] [CrossRef] [PubMed]
  327. Ma, W.; Ma, C.N.; Li, X.D.; Zhang, Y.J. Examining the effect of gene reduction in miR-95 and enhanced radiosensitivity in non-small cell lung cancer. Cancer Gene Ther. 2016, 23, 66–71. [Google Scholar] [CrossRef] [PubMed]
  328. Sun, Y.; Hawkins, P.G.; Bi, N.; Dess, R.T.; Tewari, M.; Hearn, J.W.; Hayman, J.A.; Kalemkerian, G.P.; Lawrence, T.S.; Ten Haken, R.K.; et al. Serum microRNA signature predicts response to high-dose radiation therapy in locally advanced non-small cell lung cancer. Int. J. Radiat. Oncol. Biol. Phy. 2018, 100, 107–114. [Google Scholar] [CrossRef] [PubMed]
  329. Pritchard., C.C.; Cheng, H.H.; Tewari, M. MicroRNA profiling: Approaches and considerations. Nat. Rev. Genet. 2012, 13, 358–369. [Google Scholar] [CrossRef]
  330. Condrat, C.E.; Thompson, D.C.; Barbu, M.G.; Bugnar, O.L.; Boboc, A.; Cretoiu, D.; Suciu, N.; Cretoiu, S.M.; Voinea, S.C. miRNAs as Biomarkers in Disease: Latest Findings Regarding Their Role in Diagnosis and Prognosis. Cells 2020, 9, 276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  331. Wani, J.A.; Majid, S.; Khan, A.; Arafah, A.; Ahmad, A.; Jan, B.L.; Shah, N.N.; Kazi, M.; Rehman, M.U. Clinico-Pathological Importance of miR-146a in Lung Cancer. Diagnostics 2021, 2, 274. [Google Scholar] [CrossRef]
  332. Coussens, L.M.; Werb, Z. Inflammation and cancer. Nature 2002, 420, 860–867. [Google Scholar] [CrossRef] [PubMed]
  333. KomarTili, E.; Michaille, J.J.; Croce, C.M. MicroRNAs play a central role in molecular dysfunctions linking inflammation with cancer. Immunol. Rev. 2013, 253, 167–184. [Google Scholar]
  334. Vaupel, P.; Mayer, A. Hypoxia in cancer: Significance and impact on clinical outcome. Cancer Metastasis Rev. 2007, 26, 225–239. [Google Scholar] [CrossRef]
  335. Rupaimoole, R.; Wu, S.Y.; Pradeep, S.; Ivan, C.; Pecot, C.V.; Gharpure, K.M.; Nagaraja, A.S.; Armaiz-Pena, G.N.; McGuire, M.; Zand, B.; et al. Hypoxia-mediated downregulation of miRNA biogenesis promotes tumour progression. Nat. Commun. 2014, 5, 5202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  336. Ova, Y.; Malik, A.B. Regulation of endothelial permeability via paracellular and transcellular transport pathways. Annu. Rev. Physiol. 2010, 72, 463–493. [Google Scholar]
  337. Varkouhi, A.K.; Scholte, M.; Storm, G.; Haisma, H.J. Endosomal escape pathways for delivery of biologicals. J. Control Release 2011, 151, 220–228. [Google Scholar] [CrossRef]
  338. Endoh, T.; Ohtsuki, T. Cellular siRNA delivery using cell-penetrating peptides modified for endosomal escape. Adv. Drug Deliv. Rev. 2009, 61, 704–709. [Google Scholar] [CrossRef]
  339. Lam, J.K.; Chow, M.Y.; Zhang, Y.; Leung, S.W. siRNA versus miRNA as therapeutics for gene silencing. Mol. Ther. Nucleic Acids 2015, 4, e252. [Google Scholar] [CrossRef] [Green Version]
Figure 1. This illustrates different stages of development in miRNA-based drug design for lung cancer.
Figure 1. This illustrates different stages of development in miRNA-based drug design for lung cancer.
Diagnostics 12 01610 g001
Figure 2. This figure illustrates tumor suppressor miRNAs (targeting oncogenes), tumor enhancer miRNAs (targeting tumor suppressors), and SNP that affect miRNA function (SNPs that render tumor suppressor miRNAs functionless) in the clinical significance of miRNAs in lung cancer. MiRNAs can be used in differential diagnosis of lung cancer subtypes, predicting chemoresistance, radio resistance, survival rate, patient response to treatment, and targeted medicines given to targeted persons.
Figure 2. This figure illustrates tumor suppressor miRNAs (targeting oncogenes), tumor enhancer miRNAs (targeting tumor suppressors), and SNP that affect miRNA function (SNPs that render tumor suppressor miRNAs functionless) in the clinical significance of miRNAs in lung cancer. MiRNAs can be used in differential diagnosis of lung cancer subtypes, predicting chemoresistance, radio resistance, survival rate, patient response to treatment, and targeted medicines given to targeted persons.
Diagnostics 12 01610 g002
Table 1. This table lists some of the miRNAs that have been frequently dysregulated in human lung cancer subjects and were further analyzed for their diagnostic and prognostic potential. The direct targets of these miRNAs are identified experimentally in NSCLC cell lines and animal models. (↑) indicates upregulation while as (↓) indicates downregulation. The number of arrows indicates number of studies in the respective human specimen.
Table 1. This table lists some of the miRNAs that have been frequently dysregulated in human lung cancer subjects and were further analyzed for their diagnostic and prognostic potential. The direct targets of these miRNAs are identified experimentally in NSCLC cell lines and animal models. (↑) indicates upregulation while as (↓) indicates downregulation. The number of arrows indicates number of studies in the respective human specimen.
miRNA Relative Expression Level in Lung CancerClinical AssociationExperimental ModelsExperimentally Validated TargetsEffect on Lung CarcinogenesisReferences
miR-146aSerum (↓↓), serum (↑↑), tissue (↓)DualXenograft mouse models, cell linesCOX-2, CCJN, FLAP, IRAK1, TRAF6Suppression[45,46,47,48,49,50,51,52]
miR-21Serum (↑↑↑↑) tissue (↑↑↑)chemoresistance and poor prognosisXenograft mouse models, cell lines, RECK, NFIB, TIMP3, TPM1, STAT3, Spry1, Spry2, Btg2, and Pdcd4Enhancement[53,54,55,56,57,58,59,60,61,62,63]
miR-34a/b/cTissue (↓), whole blood (↑)poor prognosis and relapseSyngenic mouse model, transgenic mouse model, murine cell lines Cdh2, Kras, Fn1 SNAIL, Suppression[64,65,66,67,68]
miR-365Serum (↓↓)poor prognosisKnock out, malignant cell linesCDC25, NKX2-1, TRIM25Suppression[69,70,71,72]
miR-486-5pSerum (↓), tissue (↓↓), endobronchial mucosa (↓), poor overall survival and chemoresistanceXenograft mouse model, Knock out, cell linesARHGAP5 (RhoA GTPase), mTOR, PtenDual[73,74,75,76]
miR-361tissue (↓), serum (↓poor prognosis and clinical outcomecell lines, xenograft mouse modelSH2B1, FOXM1Suppression[77,78,79,80]
miR-615-3pTissue (↓), tissue (↑↑), Differential diagnosiscell lines, xenograft mouse modelIGF2Suppression[81,82,83]
miR-200 familyTissue (↑, ↑)Dual Xenograft mouse model, Knock out, cell linesZEB1, ZEB2, VEGF, VEGFR1 PRDX2, GAPB/Nrf2, and SESN1, Suppression[84,85,86,87,88]
miR-221Tissue (↑↑), serum (↑↑), serum (↓), plasma (↑↑)Diagnosis, poor prognosis, and relapseXenograft mouse model, Knock out, cell linesP27kip1, TIMP3, PUMA, PTEN, MDM2, Enhancement[89,90,91,92,93]
Let-7a/btissue(↓↓), FFPE tissue (↓↓)Poor survival rate and clinical outcomeTransgenic mouse, Knock out, malignant cell linesKRAS, c-MYC, CDK6, HOXA9, TGFBR1, BCL-XL, MAP4K3Suppression[21,94,95,96,97,98]
Let-7eTissue (↓), FFPE tissue (↓↓↓)Poor survival rate and clinical outcomeTransgenic mouse, malignant cell linesSUV39H2Suppression[21,99,100,101]
miR-17-5pTissue (↓), tissue (↑), serum (↑), plasma (↓)DualTransgenic mouseBECN1, TBC1D2Enhancement[102,103,104,105]
miR-19a/bSerum (↑↑), tissue (↑)Poor prognosiscell lines, xenograft mouse modelc-MET, PP2A, BIM, E-cadherin, ZO-1, α-catenin, TNF-αEnhancement[90,106,107,108,109]
miR-18aPlasma (↑↑), Poor prognosis and radio resistance cell linesIRF2, ATM, HIF1-αEnhancement[102,110,111,112]
miR-661Tissue (↑↑), Serum (↑)Differential diagnosis and poor prognosis cell linesSOX7, RB1, RUNX3Enhancement[113,114,115,116]
miR-26a-5pFFPE tissue (↑)Differential diagnosisCell lines,Integrin-β8, FAF1Enhancement[117,118]
miR-128-3pTissue (↑), tissue (↓)Differential diagnosis, Cell lines, xenograft mouseSMURF2, cpp1, AXIN1, W1F1, SRFP2, DROSHA, DICEREnhancement[76,119]
miR-378Tissue (↑↑) Diagnosis and chemoresistanceCell lines, xenograft mouseRBX1, FOXG1, RBX1, clustinEnhancement[120,121,122,123]
miR-93Tissue (↑↑↑)
Serum (↑)
Diagnosis, Poor overall survivalCell lines, xenograft mouse LKB1, TBP2, DAB2Enhancement[124,125,126,127]
miR-135bSerum (not significant), tissue (↑)Diagnosis, EGFR mutations, invasion of visceral pleuraCell lines, xenograft mouseLZTS1, LATS1, MOB-1A, Dbf2, βTrCPEnhancement[128,129,130]
miR-16Plasma (↑↑)
Tissue (↓)
Lung cancer subtype diagnosis, poor prognosisCell lines, xenograft mouse, transgenic miceTWIST1, MEK1, HDGF, VEGF, Suppression[131,132,133,134,135,136]
miR-15aSerum (↓), Tissue (↓↓↓)Diagnosis, poor clinical outcomeCell lines, xenograft mouseACSS2, PDL1, FGFR1, DDX3X, SLC1A5 Smad3, FXR1, BCL2L2Suppression[137,138,139,140]
miR-7Tissue (↓)Shorter survival and chemoresistanceSCLC cell linesKIR2.1, ABCC1, PARP1Suppression[141,142,143]
miR-22Serum (↑), tissue (↓↓Differential diagnosis, cell lines, Murine xenograft mouse MET-STAT3, ErbB3Suppression[144,145,146]
Table 2. This table lists some miRNAs, their experimentally validated targets, and their clinical significance in lung cancer. MiRNAs upregulating (↑) in lung carcinoma (tumor enhancer miRNAs) regulate tumor suppressor genes.
Table 2. This table lists some miRNAs, their experimentally validated targets, and their clinical significance in lung cancer. MiRNAs upregulating (↑) in lung carcinoma (tumor enhancer miRNAs) regulate tumor suppressor genes.
miRNAs Expression (Lung Carcinoma)Effect on Lung CarcinomaType of miRNAExperimentally Validated TargetsClinical SignificanceReferences
miR-21EnhancementTumor enhancerRECK, NFIB, TIMP3, TPM1, STAT3, Spry1, Spry2, Btg2, and Pdcd4Promotes cell proliferation, metastasis and discourages apoptosis[60,61,62,63]
miR-205-5pEnhancementTumor enhancerTP53INP1Promotes proliferation and metastasis of lung cancer cells[167]
miR-9-5pEnhancementTumor enhancerSTARD13Promotes the progression of lung adenocarcinoma cell malignancy[168]
miR-221EnhancementTumor enhancerP27kip1, TIMP3, PUMA, PTEN,Promotes TRAIL resistance[92,93]
miR-17-92EnhancementTumor enhancerPTEN, RB1, P53, IRF2 SPRY4Promotes proliferation and metastasis and linked with short survival[101,102,112]
miR-95EnhancementTumor enhancerCaspase-3, Caspase-9, Bcl-2,Sensitises tumor tissue to radiotherapy, enhances apoptosis, and decreases proliferation[169,170]
miR-19aEnhancementTumor enhancerc-MET, PP2A, BIM, E-cadherin, ZO-1, and α-cateninPromotes gefitinib-resistance in NSCLC cells and is associated with poor prognosis in NSCLC patients[107,108,109]
miR-18aEnhancementTumor enhancerIRF2Associated with shorter survival and poor therapeutic response [111,112,113]
miR-150EnhancementTumor enhancerFOXO4Associated with metastatic malignant lung cells and tissues[171]
miR-619-5pEnhancementTumor enhancerRCAN1.4Promotes tumor angiogenesis and metastasis[172]
miR-135bEnhancementTumor enhancerLZTS1, LATS1, MOB-1A, Dbf2, βTrCPThe combined expression of LZTS1, TAZ, and miR-135b predict the prognosis of NSCLC patients.[171]
miRNA-182EnhancementTumor enhancerFOXO3Promotes tumor proliferation, chemo- and radioresistance [173]
Table 3. This table lists some miRNAs, their experimentally validated targets, and their clinical significance in lung cancer. MiRNAs downregulating (↓) in lung carcinoma (tumor suppressor miRNAs) regulate oncogenes.
Table 3. This table lists some miRNAs, their experimentally validated targets, and their clinical significance in lung cancer. MiRNAs downregulating (↓) in lung carcinoma (tumor suppressor miRNAs) regulate oncogenes.
miRNAs Expression (Lung Carcinoma)Effect on Lung CarcinomaType of miRNAExperimentally Validated TargetsClinical SignificanceReferences
miR-146SuppressionTumor suppressorCOX-2, CCJN, FLAP, IRAK1, TRAF6Discourages inflammation, associated with better overall survival, better response to chemotherapy (EGFR-TKI)[45,46,47]
miR-206SuppressionTumor suppressorc-MET, EGFR, Bcl2, VEGFA, VEGFDiscourages proliferation, tumour angiogenesis and promotes apoptosis[216,217,218]
miR-34aSuppressionTumor suppressorp21 WAF1/CIP1, MDM2, Adjunctive treatment of NSCLC patients with erlotinib along with miR-34a and Let7b sensitizes its action[67,68]
miR-32-5pSuppressionTumor suppressorSMAD family 3Inhibits EMT and metastasis in lung adenocarcinoma[219]
miR-377SuppressionTumor suppressorErbBReduces proliferation and induces apoptosis[52]
miR-205-5pEnhancementTumor enhancerTP53INP1Promotes proliferation and metastasis of lung cancer cells[167]
miR-571SuppressionTumor suppressorEGFR, MAPK1, PAK2Inhibits proliferation and induces apoptosis in lung cancer cells[220]
miR-486-5pSuppressionTumor suppressorARHGAP5 (RhoA GTPase)Inversely associated with lymph node metastasis [171]
miR-200SuppressionTumor suppressorZEB1, ZEB2, VEGF, VEGFR1 PRDX2, GAPB/Nrf2, and SESN1Suppresses angiogenesis, EMT and promotes radiosensitivity[87,88,221]
Let-7SuppressionTumor suppressorKRAS, c-MYC, CDK6, HOXA9, TGFBR1, BCL-XL, MAP4K3Associated with poor postoperative survival, chemoresistance or radio-resistance[21,97,98]
miR-199a-5pSuppressionTumor suppressorAKAP1Inhibits NSCLC proliferation and tumorigenecity [222]
miR-582SuppressionTumor suppressorHippo-YAP/TAZIncreases YAP/TAZ phosphorylation with a simultaneous reduction in cellular proliferation and promotion of apoptosis[223]
miR-582-5pSuppressionTumor suppressorNOTCH1Suppresses tumor growth and invasion[224]
miR-320aSuppressionTumor suppressorAKT3Lower levels correlated with poor prognosis and rate of survival[225]
miR-584SuppressorTumor suppressorYKT6Suppresses migration and invasion in NSCLC[226]
miR-613SuppressorTumor suppressorGJA1Inhibits lung cancer cell proliferation, migration, and formation of a colony[227]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wani, J.A.; Majid, S.; Imtiyaz, Z.; Rehman, M.U.; Alsaffar, R.M.; Shah, N.N.; Alshehri, S.; Ghoneim, M.M.; Imam, S.S. MiRNAs in Lung Cancer: Diagnostic, Prognostic, and Therapeutic Potential. Diagnostics 2022, 12, 1610. https://doi.org/10.3390/diagnostics12071610

AMA Style

Wani JA, Majid S, Imtiyaz Z, Rehman MU, Alsaffar RM, Shah NN, Alshehri S, Ghoneim MM, Imam SS. MiRNAs in Lung Cancer: Diagnostic, Prognostic, and Therapeutic Potential. Diagnostics. 2022; 12(7):1610. https://doi.org/10.3390/diagnostics12071610

Chicago/Turabian Style

Wani, Javaid Ahmad, Sabhiya Majid, Zuha Imtiyaz, Muneeb U. Rehman, Rana M. Alsaffar, Naveed Nazir Shah, Sultan Alshehri, Mohammed M. Ghoneim, and Syed Sarim Imam. 2022. "MiRNAs in Lung Cancer: Diagnostic, Prognostic, and Therapeutic Potential" Diagnostics 12, no. 7: 1610. https://doi.org/10.3390/diagnostics12071610

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