Next Article in Journal
Analysis of Prevalence and Clinical Features of Aortic Stenosis in Patients with and without Bicuspid Aortic Valve Using Machine Learning Methods
Previous Article in Journal
Exome Sequencing Reveals Novel Germline Variants in Breast Cancer Patients in the Southernmost Region of Thailand
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

miRNA-Based Technologies in Cancer Therapy

by
Maria Pagoni
1,*,
Claudia Cava
2,
Diamantis C. Sideris
3,
Margaritis Avgeris
4,
Vassilios Zoumpourlis
5,
Ioannis Michalopoulos
6 and
Nikolaos Drakoulis
1,*
1
Research Group of Clinical Pharmacology and Pharmacogenomics, Faculty of Pharmacy, School of Health Sciences, National and Kapodistrian University of Athens, 15701 Athens, Greece
2
Department of Science, Technology and Society, University School for Advanced Studies IUSS Pavia, 27100 Pavia, Italy
3
Department of Biochemistry and Molecular Biology, Faculty of Biology, National and Kapodistrian University of Athens, 15701 Athens, Greece
4
Laboratory of Clinical Biochemistry—Molecular Diagnostics, Second Department of Pediatrics, School of Medicine, “P. & A. Kyriakou” Children’s Hospital, National and Kapodistrian University of Athens, 11527 Athens, Greece
5
Biomedical Applications Unit, Institute of Chemical Biology, National Hellenic Research Foundation (NHRF), 11635 Athens, Greece
6
Centre of Systems Biology, Biomedical Research Foundation, Academy of Athens, 11527 Athens, Greece
*
Authors to whom correspondence should be addressed.
J. Pers. Med. 2023, 13(11), 1586; https://doi.org/10.3390/jpm13111586
Submission received: 27 September 2023 / Revised: 2 November 2023 / Accepted: 4 November 2023 / Published: 9 November 2023
(This article belongs to the Section Personalized Therapy and Drug Delivery)

Abstract

:
The discovery of therapeutic miRNAs is one of the most exciting challenges for pharmaceutical companies. Since the first miRNA was discovered in 1993, our knowledge of miRNA biology has grown considerably. Many studies have demonstrated that miRNA expression is dysregulated in many diseases, making them appealing tools for novel therapeutic approaches. This review aims to discuss miRNA biogenesis and function, as well as highlight strategies for delivering miRNA agents, presenting viral, non-viral, and exosomic delivery as therapeutic approaches for different cancer types. We also consider the therapeutic role of microRNA-mediated drug repurposing in cancer therapy.
Keywords:
miRNA; cancer

1. Introduction

The molecular taxonomy of cancer is recently exploiting new directions due to the discovery of a new class of RNA molecules [1,2,3]. Indeed, in 1993, microRNAs (miRNAs) were discovered by Lee et al. [4] in the nematode Caenorhabditis elegans as a large class of small non-coding evolutionarily conserved RNAs that were different from other small RNAs, due to the formation of a hairpin fold-back structure, derived from a precursor transcript, and the size of their mature sequence of approximately 22 nucleotides [5]. It is known that among the first identified miRNAs by Lee et. al. were lin-4 and let-7, which have important roles in controlling developmental timing and regulating mRNA translation [4,6,7]. Inactivation of lin-4 or let-7 causes a higher rate of cellular division in some epithelial cells than that of normal differentiation. The number of miRNA candidates has constantly increased. Indeed, the most complete public database for miRNA sequences and annotations, miRBase [8], contained, in 2002, just 218 miRNAs. Currently, 1917 miRNA precursors have been deposited into the last version of miRBase (v22.1), based on analyses of RNA deep-sequencing data [9,10,11]. By influencing protein translation, miRNAs have emerged as powerful regulators of key pathways that involve functions such as cell cycle control [12], apoptosis [13,14], hemopoiesis [15,16], adipocyte differentiation [17,18], and insulin secretion [19,20,21]. They are also involved in the regulation of human disease pathways, such as cancer [19,22], neurological disorders [23,24,25], viral infections [26,27], and metabolic diseases [19,28,29,30].
miRNAs repress over 30% of genes [31], representing the most powerful post-transcription regulators of gene expression, through direct binding to 3′-Untranslated Regions (UTRs) of target mRNA molecules. miRNA genes are located in intergenic regions (intergenic miRNAs) or within gene introns (intronic miRNAs) and are expressed mostly as independent transcriptional units (under the control of their promoters) either as single genes or forming gene clusters [32].
Although an increasing number of cancer biomarkers have been found through gene expression profiles, their reproducibility and overlap are poor. For example, various genetic tests have been developed and are currently available for breast cancer diagnosis and prognosis, such as Oncotype DX and MammaPrint [33]. However, patients with a similar expression profile of those biomarkers could present with different clinical outcomes. In addition, the invasive nature of the diagnostic tests limits their application. The most known and clinically used non-invasive biomarker is prostate-specific antigen (PSA) for prostate cancer screening and monitoring [34]. However, even PSA presents diagnostic limitations, as it is not prostate cancer-specific and elevated serum levels are also present in cases of benign prostate lesions (e.g., prostatitis, infections, or prostate hyperplasia) [35]. New affordable tools are therefore needed to support diagnosis and prognosis and to predict the most appropriate treatment for patients with cancer.
The physical properties of miRNAs, such as their high stability in body fluids and their resistance to various storage conditions (high or low pH levels and freeze–thaw cycles), support their suitability as modern molecular markers. Nonetheless, the biological mystery of their protection from RNase digestion is unknown. A potential explanation for the mechanisms attributing such protection and invulnerability could be their small length, their binding to protein complexes, as well as the fact that miRNAs are embedded in cell-secreted nanovesicles such as exosomes [36,37,38]. It is shown that within the cancerous microenvironment, tumor-derived vesicles behave as transporters of genetic information, exporting their protein content to tumor-surrounding cells [39,40]. The incorporation of determined microRNA species in nanoparticles is a molecular mechanism for cancer cells to affect the homeostasis of the surrounding microenvironment, like in the case of the exosome-mediated transfer of miR-105 that alters tight junctions of the vascular endothelial monolayer, promoting cancer metastasis and progression [41].
In addition to new genetic tests for early cancer detection, there is great interest in the development of new therapeutic agents. Indeed, although the use of new treatments has significantly reduced cancer mortality, resistance to anticancer drugs can cause treatment failure. Therefore, miRNAs are considered attractive targets for new therapeutics.
Despite significant enhancements in treatment, cancer is one of the main causes of death worldwide. Common problems related to the inefficacy of drugs are drug toxicity and resistance [42].
Common cytotoxic chemotherapy drugs are used to stop cancer cell proliferation. However, these drugs also act on non-cancer cells altering the functions of healthy tissues. Therapeutic alternatives are targeted therapies, such as erlotinib and rituximab, which are less toxic and better tolerated. Small-molecule targeted drugs such as tyrosine kinases, immunotherapy, antibody–drug conjugates, and proteolysis-targeting chimeras are the main new techniques in cancer therapy [42].
Epigenetic alterations, such as DNA methylation, histone modifications, and miRNA, influence the response to several anticancer treatments. In addition, treatment responses vary among different patients. Such dissimilarities can be due to genetic variability [43,44].
For example, the benefits of 5-fluorouracil, one of the most common anticancer drugs in gastrointestinal tract disease, are influenced by the expression of its first catabolite, namely, the enzyme dihydropyrimidine dehydrogenase (DPD). Some variants in the DPD gene influence DPD activity, increasing the life of the drug and leading to toxicity phenomena [45].
Pharmacogenetic tests are used to monitor the drug dosage, for example, based on the analysis of DPD allelic variants in the treatment with fluoropyrimidines and UGT1A1 for the use of irinotecan [46].
DNA hypomethylation can act on the sensitivity to Top I inhibitors by different processes [47]. Irinotecan, a topoisomerase I inhibitor, has been used to treat various malignancies for many years. SN38, the active metabolite of Irinotecan, is metabolized by the enzyme UDP glucuronosyltransferase 1A1 (UGT1A1). Variants of UGT1A1 are correlated with a reduction in its activity that leads to severe adverse drug responses [48]. In addition to genetic polymorphisms, recent studies have demonstrated a relationship between several miRNAs and UGT1As with a predictive role in drug efficacy. Among them, miR-548d-5p and miR-200a/-183 down-regulate UGT1A1 and UGT1A9, respectively, influencing the metabolism of drugs and thereby predicting their therapeutic effect [48]. Also, the up-regulation of several efflux transporters is associated with hypomethylation in some cancers, causing drug resistance [49]. ABCG2, a component of the ABC family of transporters, is implicated in the drug resistance of several drugs against cancer, such as adriamycin and platinum [50]. A variant of ABCG2 is associated with a better prognosis for anthracycline-based chemotherapy [51].
In this review, we discuss the recent findings of miRNAs in cancer therapy and how they can be used as potential tools in translational and clinical research.

2. Transcription and Processing of miRNAs

2.1. miRNA Biogenesis

Most miRNAs are transcribed by RNA polymerase II [52,53,54], which participates in the completion of the canonical pathway of miRNA biogenesis, generating primary transcripts (pri-miRNAs) that contain a stem–loop structure [55]. The initiation step (cropping) is regulated by the DROSHA-DGCR8 (Drosha Ribonuclease III—DGCR8 Microprocessor Complex Subunit enzymatic complex), producing ~65 nt pre-miRNAs. These have a short stem and 2-nt 3′ overhangs that are used by XPO5 (Exportin 5) [56,57]. After their transport to the cytoplasm, the endonuclease DICER1 (Dicer 1, Ribonuclease III) undertakes the catalysis of dicing and produces miRNA duplexes. These are 20-25 nt molecules consisting of a guide strand (referred to as miRNA) and a passenger strand (miRNA*) [58,59]. In humans, the recruitment of DICER1, TARBP2 (TARBP2 Subunit of RISC Loading Complex), or PRKRA (Protein Activator of Interferon Induced Protein Kinase EIF2AK2), and Argonaute components AGO1 (Argonaute RISC Component 1), AGO2 (Argonaute RISC Catalytic Component 2), AGO3 (Argonaute RISC Catalytic Component 3), and AGO4 (Argonaute RISC Component 4) serves to further process pre-miRNAs and assembly the miRISC (miRNA-induced silencing complex). One of the two strands of the RNA duplex binds to an AGO protein, whereas the remaining strand is degraded [60,61]. Canonical intronic miRNAs compared to their adjacent introns are generally spliced at a lower speed. pre-miRNAs move into the miRNA pathway, while other transcripts go through pre-mRNA splicing, producing mature mRNA for protein synthesis [38,62].
Moreover, miRNAs may also be produced via non-canonical pathways, through independent DROSHA and/or DICER1 synergy, but with splicesome-dependent mechanisms (miRtrons), or they can be transcribed by RNA polymerase III, especially when related to the regulation of metabolic activities, such as cell cycle growth and differentiation [63,64].
In addition to all the above mechanisms, the complexity of regulatory events also incorporates modifications by RNA editing, RNA methylation, uridylation and adenylation, AGO loading, and RNA decay [65,66].

2.2. miRNA Function

2.2.1. Intracellular miRNAs

miRNAs target mRNA molecules through base pair complementarity, leading to gene silencing. The grade of complementarity between miRNA and mRNA decides either mRNA degradation or protein synthesis repression [5,67]. In the case of extensive and near-perfect complementarity, which is mostly observed in plants, gene silencing is performed by the Ago-mediated cleavage of mRNAs between miRNA nucleotides 10 and 11 [68,69]. In metazoans, miRNA targeting involves perfect Watson and Crick base-pairing only between the 5′-end bases 2 to 7/8 of the miRNA, which is called “seed region”, and the 3′-UTR of the target mRNAs [13,70], although the binding 5′-UTR has also been demonstrated [71]. In this regard, miRISC induces, initially, the translational repression of the target mRNAs, and thereafter mRNA deadenylation and decay via mRNA decapping and 5′–3′ exonucleolytic degradation [72,73]. As each miRNA can regulate the expression of many mRNAs, each miRNA can simultaneously control different functional pathways. For example, let-7 can regulate several oncogenes, including KRAS, NRAS, HRAS [74], MYC, and CASP3, and various genes involved in epithelial-to-mesenchymal transition [75].

2.2.2. Extracellular miRNAs

Besides the well-studied intracellular RNAs, pioneering work in 2007 and 2008 [73,76] demonstrated the presence of extracellular miRNAs in cell-derived exosomes and plasma of pregnant women. Additionally, the presence of serum miRNAs in prostate cancer [77] and the clinical value of exosome-derived miRNAs (exomiRs) in ovarian carcinoma [78] were documented. Extracellular miRNAs have been detected either in extracellular vesicles, including exosomes and apoptotic bodies, or in forming complexes with proteins and lipoproteins [79].
In general, extracellular miRNAs have been detected in blood, urine, saliva, amniotic fluid, peritoneal fluid, breast milk, cerebrospinal fluid, and seminal fluid by either passive secretion following cell death or apoptosis, or active secretion through endosome-derived extracellular vesicles [79]. Recent studies have demonstrated the different quantity and profile of exomiRs in most human malignancies studied so far, highlighting their biomarker capabilities [80,81].

3. miRNA Dysregulation in Cancer

miRNA-based biomarker validation for malignancies divides miRNA research into three main sectors: (1) diagnosis, (2) prognosis, and (3) therapy [82,83,84,85].
Dysregulation of miRNA expression seems to have a fundamental role at the onset, progression, and dissemination of many diseases, including human malignancies. Numerous miRNAs map to regions of the human genome known to be deleted or amplified in cancer [86]. In addition, data emerging from large-scale miRNA research support the fact that almost 50% of miRNAs are found to be frequently located in cancer-associated regions or fragile sites of the genome [86]. In addition, while deletions and transcriptional modifications usually affect a single or a small group of miRNAs, deficiencies in their processing mechanisms, frequently lead to extensive changes associated with cancer [87,88,89]. Deletions or mutations in genes that encode for miRNA tumor suppressors may lead to the loss of one specific miRNA or of an miRNA cluster, with subsequent stabilization of targeted oncogenes [90,91]. For example, genetic aberrations of cancer-associated genes belonging to the miRNA processing protein complex, such as TARBP2 [92], DICER1 [93], and XPO5 [94], are of crucial importance in the cellular transformation pathways and contribute to an overall miRNA dysregulation in cancer [1,95].
Among the results of recent studies regarding the contribution of processing miRNA machinery to tumorigenesis is the role of DGCR8 and DROSHA in the development of Wilms tumors [96,97], while down-regulation and, controversially, up-regulation of both enzymes have been reported previously in many tumor types [89,98,99,100].
There are also several recognized disorders established as DICER1-related disorders, such as Sertoli–Leydig cell tumors, embryonal rhabdomyosarcoma, and multinodular goiter, which confirm that these mutations also promote tumorigenesis [101].
miRNA involvement in carcinogenesis through aberrant DROSHA and DICER1 expression was highlighted [102]. This was further confirmed by the preliminary results of the study on DICER1 differential expression, observed by immunohistochemistry in melanoma cells and in benign melanocytes derived from patients who were affected by cutaneous malignant melanomas (CMM), benign melanocytic nevi (BMN), and dysplastic melanocytic nevi [103].
Clinical studies related to miRNA biogenesis and processing machinery, concerning miRNA expression profiles in epithelial skin cancer in comparison to the premalignant state, showed that, except for TARBP2, there is a dysregulation of the miRNA pathway factors, in particular those of microprocessor complex and RISC [104].

4. miRNA as a Therapeutic Agent in Cancer

Therapeutic agents are categorized mainly into synthetic small molecules, monoclonal antibodies, or large proteins. Traditional drugs may be insufficient to hit intended therapeutic targets because of the inaccessibility of active sites in the target’s three-dimensional structure. RNA-based therapies may offer an excellent chance to potentially reach any therapeutic-relevant target [105] (see Table 1). Additionally, techniques based on nucleic acid technologies are less laborious towards synthesis procedures. Nevertheless, RNA-based therapies have specificity issues that carry the risk of off-target effects [106,107].
In this context, the therapeutic use of miRNA therapy is receiving attention in clinical trials of almost all human diseases Depending on the expression pattern in pathological conditions there are two main streams of miRNA therapies that include the development of synthetic molecules with an effect on protein expression [108]: miRNA mimics [109,110] and miRNA antagonists [111,112]. The restoration of miRNA functions and the inhibition of overexpressed miRNAs are fundamental for the development of miRNA-based cancer therapeutics [113,114]. This is because, the regulation of specific miRNA alterations through miRNA mimics or antagomirs normalizes the gene regulatory network and the signaling pathways, reversing the phenotypes in cancer cells [113,114,115].
Even though miRNA therapeutic potential is still at the preclinical and early clinical stage, researchers are encouraged to invest in miRNAs, as many miRNAs satisfy the stringent criteria that could make them successful therapeutic agents due to their function as oncogenes or tumor suppressors [116,117,118,119].

4.1. miRNA Inhibitors

The inhibition of oncogenic miRNAs has been achieved by antisense oligonucleotides (anti-microRNA oligonucleotides—AMOs), also called antagomirs, miRNA sponges, and miRNA-Masking Antisense Oligonucleotides [120,121,122,123]. Table 1 and Figure 1 show the main miRNA therapeutic strategies.

4.1.1. Anti-miRNA Oligonucleotides (AMOs) or Antagomirs

The most popular approach to inhibit miRNA function is the synthesis of antisense oligonucleotides with complementary sequences to endogenous miRNAs. The chemical structure that characterizes them helps to trap the endogenous miRNA that cannot be further processed by the RISC complex or where the endogenous miRNA is being degraded. During this process, the endogenous miRNA of interest serves as a biomarker optimizing the pharmacokinetic and pharmacodynamic properties of the antagonist [124,125,126,127].
First-generation pre-clinical compounds, the antisense phosphorothiolated oligodeoxynucleotides, were characterized by a low affinity for their target, negative immunostimulatory effects, and they also had very short half-lives, due to their immediate excretion by renal clearance [128,129].
These properties have been improved in the next generation of drug agents with the design of antisense and anti-miR oligonucleotides supported by better pharmacokinetic properties [130]. Because of their capacity to bind cognate sequences, miRNA’s action on target mRNAs can be inhibited by steric antisense oligonucleotides (ASOs) that have high affinity and specificity for some miRNAs [131]. ASOs are single-stranded antisense oligonucleotides, in particular chemically modified DNA-like molecules, 17 to 22 nt in length, which have been designed to have complementarity to a specific mRNA, inhibiting its translation. They have been used for more than thirty years in clinical trials phases II and III [124,132], but they have also been used with success to screen gene functions in high-throughput cellular assays. Their clinical and preclinical application in cancer is for the design of therapeutic solutions together with ribozymes, DNAzymes, small interfering RNAs (siRNAs), short hairpin RNAs (shRNAs), anti-miRNA agents such as ASOs-anti-miRNAs, and locked nucleic acids (LNA)-anti-miRNAs or antagomirs. Although their inhibitory action on the upregulated oncogenic miRNAs has an efficient and long-lasting effect [133], due to their weak pharmacodynamic and pharmacokinetic properties, the delivery of these agents is not ideal [134,135].
Experimental evidence on ASO applications for miR-21 and miR-221 showed increased expression levels of PTEN, RECK, and CDKN1B, but also reduced proliferation and increased apoptosis of pancreatic tumor cells [136,137]. The latest preliminary studies in pancreatic cancer on small side population (SP) cells with stem cell-like properties suggest that inhibition of miR-21 and miR-221 with combinatory ASOs technology against miR-21 and miR-221 inhibited primary tumor growth and metastasis compared to a single antagomir treatment [138]. ASOs can sensitize pancreatic ductal adenocarcinoma cells (PDAC) to gemcitabine causing synergistic anticancer outcomes [137].
A class of antisense oligonucleotides has been subjected to biochemical modifications, and they have been transformed into ssRNA analogs complementary only to specific miRNAs. They are also known under the term anti-miRNA ASOs (AMOs) [139,140]. In this mechanism of action, the modified synthetic anti-miRNA oligonucleotides (AMOs) inhibit specific individual miRNAs through competitive inhibition of base-pairing, as they block the interaction between miRNA and its target mRNAs.
miR-16, miR-21, miR-214, and miR-181a have been identified as potential drug targets for lung cancer therapy by design, synthesis, and benefit-specific AMOs on A549 lung carcinoma cells. Following transfection, some properties, such as cell viability, apoptosis, and miRNA expression, were tested under variable conditions of dose and time. It was observed that AMO-miR-21, AMO-miR-16, and AMO-miR-181a arrested cell growth by inducing apoptosis and S-phase suppression, suggesting that these miRNAs could be potential targets for cancer therapy and AMOs could be a functional technique for miRNA inhibition [141].

Chemical Modifications of AMOs

In the past, several chemical modifications of AMOs have been designed to advance their efficiency and stability such as the addition of 2′-O-methyl and 2′-O-methoxyethyl formations to the 5′ end of the molecule [142].
2′-O-methyl-group (OMe)
The 2′-O-methyl-group (OMe) is one of the most frequent chemical modifications to oligonucleotides. The methyl group has the function to contribute to nuclease resistance improvement and the binding affinity to RNA. Completely modified OMe-oligonucleotides have found application to prevent aberrant exon splicing in cells [143], whereas the hybrid backbone OMe/DNA that is formed in antisense oligonucleotides is under investigation in clinical applications [144]. Synthetic antisense oligonucleotides bearing 2′-O-Methyl (2′OMe) ribose modification and antagomirs were shown to decrease miRISC activity and to inhibit specifically miRNAs that have acquired a gain of function in human diseases including cancer [124,132].
Synthetic AMOs that bring the 2-O-methyl modification showed effective inhibition of target miRNAs in cell and xerograph models with the disadvantage of operating in high doses in the given study [145]. 2-O-methyl AMOs have been applied as inhibitors in glioblastoma and breast cancer in human cell lines and xenografts by targeting miR-21 [146]. Today, 2′OMe-modified AMOs represent the most used molecules to dissect miRNA roles, not only due to their high affinity for targeting miRNAs but also for the lack of high levels of toxicity and economic cost of synthesis [147].
2′-O-Methoxyethyl
2′-O-Methoxyethyl (MOE)-modified oligonucleotides show evidence of higher binding affinity and specificity to RNA compared to OMe-analogs. They have been used efficiently as chemically modified oligonucleotides with the property to re-address mRNA splicing. They also inhibit protein translation. MOE-ASOs are increasingly considered important modified oligonucleotides in clinical trials [148]. The combination of antisense oligonucleotides (ASOs) with 2′-O-(2-methoxyethyl) (2′-MOE) is a platform of RNA-based therapeutics with the property to hybridize to their target RNA via Watson–Crick base pairing, preventing the expression of the “disease-related” protein product. Recently, there has been high interest expressed in the number of 2′-MOE ASOs progressing to phase I, II, and III clinical trials for therapies in many diseases, including rheumatoid arthritis [149], cancer [150], and hypercholesterolemia [151].
Locked nucleic acid (LNA)
The chemistry-locked nucleic acid (LNA)-modified oligonucleotides consist of a 2′-O-modified RNA molecule where the 2′-O-oxygen links to the 4′-position through a methylene linker, making a bicyclic compound locked into a C3′-endo (RNA) sugar [152]. The LNA chemical modification provides thermodynamic stability to the duplex conformation with known RNA molecules. The use of LNA oligonucleotides as northern probes is important for the detection of miRNAs as mixed backbone oligonucleotides [153]. Potentially, they could be a new class of therapy. Among the types of nucleoside modifications applied, the addition of chemical groups to the 2′-hydroxyl group has been quite successful [154]. In addition, oligonucleotide derivatives could be applicable as curative AMOs [155].
Preclinical studies focusing on the understanding of the mechanisms describing osteosarcoma-initiating cells and their potential clinical significance showed that deregulation of miR-133a in a highly malignant CD133 cellular population affects cell invasiveness and identifies a lethal tumor phenotype. The inhibition of this miRNA by LNA reduced cell invasion, whereas administration of LNA in addition to chemotherapy suppressed lung cancer metastasis and increased survival outcomes in osteosarcoma-bearing mice. Clinically, overexpression of both CD133 and miR-133a has been associated with poor prognosis, whereas overexpression of four CD133 targets correlates to a good prognosis. Overall, silencing LNA miR-133a in combination with chemotherapy was proposed as an anticancer strategy, developed at the preclinical stage for targeting multiple regulatory pathways associated with metastasis of osteosarcoma cells [156]. In addition, studies investigating the effectiveness of such inhibitory approaches documented that although 2′ modifications were shown to improve affinity to target RNAs, their anti-miRNA activity was not fully correlated with affinity, suggesting that other variables may also be important for effective miRNA [157]. Studies analyzing the tissue toxicity of LNA in mice and monkeys focused on the effect induced by LNA-miR-221 inhibitors in vital organs. They confirmed the low toxicity of LNA and suggested its use for clinical purposes [158].

4.1.2. miRNA Sponges in Cancer Therapy

The miRNA “sponge” technology was introduced in 2007 with the purpose of ensuring miRNA loss of function in cell models and transgenic organisms. miRNA sponges are usually plasmid-encoding copies that contain binding sites complementary to the seed region of the target miRNA [111,159] and are products of transgenes within cells.
Following cell transfection, these plasmids transcribe high levels of sponge RNAs that bind to the seed region, which enables them to block a family of miRNAs that contain the same seed sequence. As competitive inhibitors, miRNA sponges exhibit similar inhibition efficiency with short nucleotide fragments [111,159]. In addition, it has been shown that miRNA sponges that are based on retroviral vectors are able to knock down miRNAs but also block an entire miRNA seed family using one vector [160].
The sponge binding sites act on the miRNA seeding region, and, in this way, they block a whole family of related miRNAs. The fine-tuning of miRNA sponge concentration compared to miRNA target concentration is crucial for the efficacy of miRNA inhibition. The highest expression of miRNA sponge is achieved when the affinity and avidity of binding sites are high, but also when the promoter used in the cell model of interest is strong, i.e., a cytomegalovirus promoter. Sponge inhibitors are used in long-term miRNA loss-of-function research and in vivo assays, for example, bone marrow reconstitution and cancer xenografts. The stability of miRNA sponge activity by expressing the transgene from chromosomal integrations has been tested by different groups [161,162,163]. The importance of miRNA sponges in cancer therapy is to mimic the down-regulation of specific miRNAs that are deregulated.
Before investigating the biological effects of miR-21 on A549 non-small cell lung cancer (NSCLC) cells, the expression of miR-21 in serum samples of patients affected by NSCLC was quantified. More precisely, they used miRNA sponge technology and transfection of A549 cells, suggesting that miR-21 could be an independent molecular biomarker for NSCLC, but also that modulating miR-21 or PDCD4 expression may provide a potentially novel therapeutic approach for NSCLC [164].

Competing Endogenous RNAs (ceRNAs)

Competing endogenous RNAs (ceRNAs) act as natural endogenous miRNA sponges, regulating the bioavailability and function of miRNAs. They include transcribed pseudogenes, long non-coding RNAs (lncRNAs), and circular RNAs (circRNAs). Their synergic action forming molecular networks aims to the regulation of protein expression [165]. Their advantage compared to antisense oligonucleotides, which target a single miRNA, is that they can have numerous different binding sites, coordinating simultaneous inhibition of a big subset of a miRNA cluster, or of distinct miRNAs that act on the same target. The development of quantitative methods for the determination of the absolute expression levels of miRNA and ceRNA molecules allows the possibility of estimating the efficiency of ceRNA crosstalk in many biological models. In addition, the use of mathematical models contributes to the calculation of fluctuations in ceRNA and miRNA levels, which respond to variations in additional parameters, such as structure and topology [166].

4.1.3. miRNA-Masking

The development of miRNA-mask technology that includes miRNA-masking antisense oligonucleotides uses single-stranded 2′-O-methyl-modified antisense oligonucleotides, which are complementary to the miRNA binding sites in the 3′-UTR of the target mRNA [161]. Due to the masking effect of these miRNAs to the target mRNA, their action is gene-specific. Recently, miRNA therapeutics technology used constructs of miRNA-masking oligonucleotide drugs (ONDs) that attach to mRNA. The binding of an miRNA-mask to its target site is assisted by the use of O-methyl groups and also LNAs, to improve the masking efficiency [167]. The high efficiency of miRNA masking is based on target gene selection. miRNA–gene interactions that are crucial for tumorigenesis, such as miRNA-203 and LASP1, and miRNA-29-b-1 and SPIN1, which are involved in proliferation, angiogenesis, migration, and apoptosis, have been investigated [168].

4.2. miRNA Mimics

miRNA replacement therapy is gaining traction worldwide using synthetic miRNAs or mimics, which, when incorporated, can restore the normal physiological activity of the organism. The miRNA replacement strategy is classified as either a viral or non-viral delivery therapy [169]. The overexpression of miRNAs can also be achieved by the synthesis of miRNA mimics, which are designed for protein-coding gene silencing [110,170]. The biochemistry of miRNA replacement technology includes an active strand of the mimic, which contains a sequence that is normally expressed to the cell, whereas the passenger strand is chemically modified to achieve the interaction of the active strand with the protein complex and guarantee the functionality of the miRNA.
miRNA mimics can also be described as double-stranded-like RNAs, which are composed frequently of siRNA-like oligoribonucleotide duplexes. [171,172]. They have a sequence motif on their 5′ end that is partially complementary to the target sequence in the 3′-UTR of the target mRNA. They mimic the functionality of mature endogenous miRNAs. The use of mimics is important for miRNA functionality assessment because it provides a tool for gain-of-function studies. The restoration of miRNAs that show a loss of function through miRNA mimics is fundamentally used to explore the therapeutic potentiality of tumor suppressors [173,174].
Technically, miRNA replacement therapy acts downstream of the miRISC complex and requires the enzymatic functions of cellular miRISC to be catalytically functional. miRNA mimics target multiple transcripts, and it seems that regulate the same set of genes as the endogenous miRNAs.
It was observed that the overexpression of miR-21 associated with keratinization of tumors in cases of oral squamous cell carcinomas was significantly correlated with the poor prognosis of patients. Transfection of miR-7- and miR-21-mimics reduced the expression of RECK, which is a tumor suppressor gene, through direct miRNA-mediated regulation. This study provided important information related to patient survival, with the aim of contributing to improved therapeutics for oral cancer [175].

5. Druggable miRNA Systems

The safety and efficiency of miRNA-based therapeutics, especially those focused on the delivery of miRNA mimics or antagomirs targeting human tissues, is still a challenge. The limitations of miRNA delivery mechanisms concern their susceptibility to degradation by nucleases upon addition into biological systems [176,177], the rapid clearance from blood [178], the off-target toxicity, the unwanted immune responses [179], and their poor binding affinity for complementary sequences [180]. In addition, since they target multiple pathways via imperfect matching in the 3′ UTR region, they might cause involuntary gene silencing [181].
Overall, there are several strategies to overcome these challenges. For example, the nuclease degradation of naked miRNAs is prevented with oligonucleotide chemical modifications [182], while the miRNA hydrophilic characteristics or their poor cellular uptake that is caused by charge repulsion between miRNAs and the cellular membranes can be overcome with several delivery vehicles [177].
The lack of a foolproof miRNA delivery strategy is the fundamental barrier to the clinical application of miRNA therapies. For this reason, we draw attention to non-viral synthetic and viral [119] delivery techniques [170,171] that are developed for local and systemic delivery [171].

5.1. Local Delivery

Local delivery (intratumor) is the most advantageous option for amenable malignancies [183]. It is specifically applicable in the experimentation of solid tumors, including primary and well-localized tumors, but it is not applicable in hematological malignancies such as leukemia and not suitable for metastasizing tumor cells that are present in late-stage malignancies as they are not exposed to the RNA drugs in circulation [181].

5.2. Systemic Delivery

The treatment of advanced metastatic tumors must be accomplished through systemic delivery. Nonetheless, there are many different strategies that have been developed in recent years to overcome the challenges faced by systemic delivery.
For this reason, as previously highlighted, the improvements aiming to achieve oligonucleotide stability while decreasing innate immunity can be indicatively represented by several chemical modifications on the 2′-OH ribose with a fluoro, amino, or methyl group that, although they can be easily degraded in serum, provides 1000-fold resistance to degradation in plasma compared to the unmodified RNA counterparts. Stability is also improved by modified anti-miRNAs with LNAs [177,184,185].

6. Categories of Delivery Vehicles for miRNA-Based Therapy

MiRNAs as therapeutic modalities employ vectors for gene delivery purposes that may be distinguished into two categories: viral carriers that incorporate genetic material and non-viral carriers that consist of cationic molecular carriers, in particular lipids and polymers, which interact electrostatically with nucleic acids for the gene delivery to cells [186]. Distant intercellular communication involved in RNA shuttling concerning miRNA delivery for cancer therapy via exosomes is also discussed [187].

6.1. Viral Vectors

Synthetic viral vectors have become valuable tools for gene therapy due to their transduction effectiveness and their permanent gene expression in many cell types [188]. Viral vectors may deliver miRNAs at different stages of biosynthesis (i.e., pri-miRNAs and pre-miRNAs) [189]. Driven by a viral promotor, pri-miRNAs and pre-miRNAs molecules, following cloning within the plasmids, may be transcribed and processed to mature miRNAs, enabling them to target the mRNAs [189,190].
The most frequently used viruses for therapeutic gene delivery that have been employed with miRNAs in various cancer models are Adenoviruses (AdVs), lentiviruses (LVs), and adeno-associated viruses (AAVs) [184,188,191,192]. The limitations regarding the adenoviruses and adeno-associated viruses concern the immunogenicity and the temporary miRNA expression, while lentiviruses may introduce a risk related to the safety of genomic integration procedure [184,193,194].
Oncolytic AdVs (OAdVs) have been successfully used in the delivery of miRNA mimics and anti-miRNAs. Specifically, the inhibition of tumor growth in xenograft models of triple-negative breast cancer (TNBC) was performed via OAdV-mediated delivery of anti-miRNAs in the form of long ncRNAs (lncRNAs) by simultaneous suppression of onco-miRNA levels [189,195]. In addition, Bhere et al. demonstrated the therapeutic efficacy of simultaneous up-regulation of miRNA-7 and down-regulation of miRNA-21 via AAV-mediated delivery of anti-miRNA-21 and miRNA-7 in mice bearing malignant brain tumors [189,196].

6.2. Non-Viral miRNA Delivery

Virus-mediated miRNA-based therapeutic delivery strategies, despite being very effective, are clinically insufficient due to several biosafety issues, including viral immunogenicity [189]. Non-viral delivery systems are indicated for the transport of endogenous miRNA or for miRNA-expressing vectors. They prevent nuclease degradation by use of organic, inorganic, or polymer-based carriers [119].
In the present review, we mention the chemical methodology of non-viral miRNA delivery, referring to lipid, polymer, inorganic, and extra-cellular vesicle carrier-based advances [191].

6.2.1. Lipid-Based Delivery Systems

The most employed non-viral vectors are organic-based carriers that use liposomes encapsulating nucleic acids. Lipid-based drug delivery exerted efficient therapeutic potential in preclinical and clinical cancer studies [197]. Cationic, anionic, and neutral liposomes are utilized extensively due to their significant affinity with the cellular membrane as they are amphipathic [198]. Liposomes are the main unit for all lipid nanoparticles (LNPs) undergoing chemical modifications, such as hyaluronic acid (HA) and polyethylene glycol (PEG), to enhance tumor-targeted capabilities and stability [198,199]. The optimization of lipid-based nanoparticles has generated ionizable liposomes with the ability to change the charge status depending on pH variation, making them clinically translatable [189]. On the other hand, even though a liposome formulation (MRX34, a liposomal miR-34a mimic) was used to treat liver cancer, unfortunately, the adverse effects were severe, and the trial was stopped following the death of four patients [200,201]. In addition, in another study, it was found that the delivery of an miR-199b-5p mimic by the use of ionizable liposomes could impair cancer stem cell markers in several cancer cell lines [189,191,202].

6.2.2. Polymeric Nanoparticles

Among the nanocarrier delivery systems for miRNA therapeutics in cancers are polymer-based delivery systems that have found success as efficient vectors for transferring nucleic acids due to their great stability, flexibility, and ease of functional group substitutions [189]. They are distinguished into natural and synthetic groups [188]. Natural polymers studied for gene therapy involve chitosan, collagen, gelatin, and their modified derivatives [203,204]. Chitosan and other natural polymers are characterized by muco adhesive capability, low toxicity, and pH-sensitive drug release properties, making them suitable candidates as drug carriers [188,205,206]. Cellular membranes may also be penetrated by other natural polymers, such as cell-penetrating peptides (CPP), allowing them to transfer a wide variety of active conjugates [207]. The CPP-conjugated nanoparticles have greater stability, improved cellular uptake, and lower toxicity, but they may also cause endosomal entrapment and particle aggregation [208].
Synthetic polymers that were successful in the delivery of specific miRNA-based cancer therapies comprise polyethyleneimines (PEIs), dendrimers, Polyamidoamines (PAMAM), and poly lactic-co-glycolic acid (PLGA) [188,189].
Polyethylenimines (PEI) are used in gene delivery. They are positively charged molecules that can assemble into nanoscale complexes with small RNAs, shielding them from degradation and released intracellularly [209]. Although frequently utilized, PEI alone are less favorable compared to other delivery vectors due to the excess positive charge and the limited degradability that is caused by the binding of serum proteins [190]. The high cytotoxicity associated with the high molecular weight of the compound has restricted their application in gene delivery. Further improvements at the preclinical stage showed satisfactory results on biosafety and anti-tumor effects, using fluorine-modified polyethyleneimine (PEI) 1.8 kDa for microRNA-942-5p-sponges non-coding RNA delivery.
Dendrimers have emerged as a significant sector in healthcare due to their ideal properties of having very strong drug delivery potential, making them good carriers. Dendrimer-based nanoparticles have a strategic role in the targeted administration of miRNAs in cancer therapy because they can transport large numbers of TNA (Therapeutic Nucleic Acids) into both cells and target tissues [210,211]. The main advantage that distinguishes dendrimers is their monodispersity because they are all identical and have a well-defined size after synthesis compared to other polymeric compounds that are produced by uncontrolled polymerization [212,213]. In addition, dendrimers can bind many molecules to their periphery because they have many functional terminal groups and establish lipoplexes and polyplexes that are called dendriplexes [214]. Furthermore, they are characterized by good stability, but they also have disadvantages that involve the uncontrolled release of drugs and toxicity because they have been reported to display cytotoxic and hemolytic activity at elevated concentrations, which excludes their intravenous use in anticancer therapy [215,216]. A reference example of their use is the advanced design of PCSTD-Gd/DOX/miR 21i polyplexes, which improved the combination of chemo-gene treatment guided by MR imaging and was studied on an orthotopic breast cancer model in vivo [217].
Polyamidoamine (PAMAM) is a positively charged and biodegradable synthetic polymer that, although it allows complexation with nucleic acids, is easily modifiable, has high buffering capacities, and has the disadvantage of inducing hepatic toxicity [189,213,218]. (PAMAM) dendrimers are the most investigated family of dendrimers [216]. As the highly positive charged surface of PAMAM can interact with negatively charged microRNAs, it can enable the anti-tumor effect of microRNA mimic molecules when loaded into PAMAM dendrimers. An example is the application of let-7b-loaded PAMAM-HA NPs in combination with chloroquine, which decreased the expression of oncogenic and anti-apoptotic genes and increased apoptotic gene expression [219].
Poly Lactic-Co-Glycolic Acid (PLGA) is one of the most thrivingly synthesized biodegradable polymers. It is an FDA-approved hydrophobic delivery vector that in synergy with lipids and polymers, of natural and synthetic origin, has been efficient in mediating miRNA delivery for the treatment of cancer [188]. The importance of this polymer is based on biological and technical features such as biocompatibility, well-defined formulation, easy processing, and the controllable release of drugs into vital organs, although disadvantages are also present and are related to high production costs [188,220]. An example of PLGA application is the delivery of synthetic miRNAs using clinically compatible PLGA-PEG nanoparticles before chemotherapy in glioblastoma (GBM) models. The performance of this study showed therapeutic efficiency that was genetically associated with signaling pathways in cancer [221].

6.3. Inorganic Based Delivery

Inorganic materials are frequently employed in nanotechnologies and have been created as vectors to carry miRNA. Examples of these materials are graphene oxide, mesoporous silicon, gold nanoparticles (AuNPs), and Fe3O4-mediated NPs. Chemically modified AuNPs are easily able to bind functional groups like thiol and amino groups to their surface, and these AuNPs have been used as miRNA carriers. The size, shape, and porosity of the particles may be controlled, and they can be made to be nontoxic, biocompatible, and non immunogenic [189,222].

6.4. Exosome-Mediated Delivery of miRNAs

Exosomes are cell-derived nanovesicles with a diameter range of 40–100 nm. They are mediators of cellular communication in health and disease conditions as they carry nucleic acids, proteins, lipids, and metabolites. Because they originate from the plasma membrane, they are safe, biocompatible, and induce immune tolerance in vivo. They can overcome the blood–brain barrier (BBB). Major challenges with this approach include the inability to produce significant quantities of highly purified exosomes, as well as their rapid clearance from the bloodstream and their accumulation in vital organs [188,223]. Solutions to these drawbacks are required before advancement into clinical practice.
Many studies used exosome nanovesicles for miRNA delivery derived from mesenchymal stem cells (MSCs). These MSC-derived nanovesicles carrying a synthetic miRNA-143 decreased the migration of osteosarcoma cells, although compared to lipofection, the efficiency was lower [224]. Another example is that for the diagnosis and miRNA therapy of pancreatic ductal adenocarcinoma (PDAC), a successful test was performed whereby the exosomal hsa_circ_0012634 restrained the PDAC progression via the miR-147b/HIPK2 pathway, highlighting its use as a candidate biomarker.

7. miRNA-Based Therapies in Clinical Practice

The contribution of miRNAs in the diagnostic area is high and some panels, notably for thyroid cancer, are covered by major insurance companies [225,226]. On the other hand, a multitude of clinical trials are currently underway to test new miRNA treatments, although their use in the therapeutic market is less advanced. For example, MiRagen Therapeutics is testing an oligonucleotide inhibitor of miR-155, MRG-106 (known by the name Cobomarsen), which is an LNA antagomiR for the treatment of cutaneous T-cell lymphoma [227,228,229,230]. ENGeneIC developed Mesomir, a miRNA mimic of tumor suppressor miR-16 for thoracic cancer patients [231,232,233]. Asbestos Diseases Research Foundation formulated nanoparticles for the delivery of TargomiR, a miR-16 mimic used against NSCL [234,235]. Synlogic (merged with MiRNA Therapeutics) created a liposomal–miRNA mimic formulation to target miR-34a in vitro and in vivo that was previously entered in clinical trials [201,236] under the license of Marina Biotech Inc. using Smarticles, a liposome technology [237,238]. Other biotechnological companies, like OPKO Health, Inc., Miami, FL, USA, Alnylam Pharmaceuticals, Cambridge, MA, USA, Interna Technologies, Utrecht, The Netherlands and Mello Biotech, Santa Fe Springs, CA, USA are in the screening phase [235]. Additionally, Rosetta Genomics, Princeton, NJ, USA [239], RXi Pharmaceuticals, Marlborough, MA, USA [240], Phio Pharmaceuticals, Marlborough, MA, USA [241], Asuragen, Inc., Austin, TX, USA [242], Sirnaomics, Inc., Germantown, TN, USA [243], Bristol-Myers Squibb, New York, NY, USA [244], and Dicerna Pharmaceuticals, Lexington, KY, USA [245] have developed miRNA-based biomarkers. A snapshot of the progress in the biopharmaceutical pipeline that could increase the number of miRNA therapeutics is shown in Table 2.

8. Drug Resistance

miRNAs target and regulate mRNAs, including several chemoresistance-related genes. They influence drug resistance mainly due to the regulation of cell survival and apoptosis signaling pathways [246] but also because they mediate the regulation of drug targets and the DNA repair system, as well as affect drug transport and metabolism-related enzymes [247]. Although the use of miRNAs for cancer chemotherapy has not yet been fully investigated in clinical practice, it has been experimentally demonstrated that miRNA-targeted therapy can be significant when combined with conventional chemo-radiotherapy to sensitize tumor cells [248]. For example, miRNA-modulatory strategies to circumvent tumor drug resistance have been reported, by using an antisense strategy for the inhibition of miR-21 and miR-200b while enhancing the response of cholangiocarcinomas to chemotherapy [249]. Additionally, it has been shown that delivery of functional anti-miR-9 through stem cell-produced exosomes to GBM cell lines (glioblastoma multiforme) circumvented resistance to temozolomide [250]. Transfection with anti-miR-92b regulated cisplatin resistance by targeting the PTEN gene in an A549 non-small cell lung cancer cell line [248,251].

Druggable miRNA Metabolic Pathways in Oncology

A dysregulated metabolism characterizes cancerous cells; thus, metabolic reprogramming is a hallmark of oncogenesis [252]. miRNAs positively and negatively regulate several metabolic genes [253] and contribute to modified levels of glycolysis [252,254,255,256], glucose uptake and transport [257,258], lactate production [259,260,261,262,263,264], tricarboxylic acid cycle, glutaminolysis [265,266], altered insulin production, and dyslipidemia, as well amino acid and nucleotide biogenesis [267,268,269,270,271]. Novel miRNA strategies specialized in metabolic plasticity in human cancers [272,273] are based on the pharmacological targeting of metabolic pathways aiming to drastically decrease cancer proliferation and progression [272,274]. miRNAs are prospective pharmacological targets through the drug-induced action of metabolic drugs that affect the impairment of signaling cascades and the regulation of the production of cellular energy [252].

9. Drug Repurposing Based on Drug–miRNA Associations in Cancer Therapy

Drug repurposing for oncological purposes is the identification of new applications of existing medications that already satisfy clinical and safety criteria [275]. New therapeutic indications for known drugs, with an oncological and non-oncological primary purpose, for example, metabolic-based drugs, offer new treatment options to oncologic patients [276]. Interestingly, there are growing findings on the antineoplastic effects and improved responses to these metabolic-based medications modulated to the induction of tumor suppressor miRNAs and the suppression of oncogenic miRNAs [275,277].

9.1. Metformin

Metformin, a drug for treating type 2 diabetes, despite targeting glucose metabolism [278,279,280,281,282,283], is characterized by an antineoplastic activity that can be partially interpreted through the variation in let-7, miR-26, and miR-200 in tumor types such as oral, colorectal, pancreatic, renal, and breast cancer.
The spread of reports supporting the potential efficacy of DCA, a PDK pyruvate dehydrogenase kinase (PDK) inhibitor, in cancer therapy, derives mainly from increased glycolysis and decreased mitochondrial oxidation, regardless of oxygen availability, due to the Warburg effect [58]. The combined effects of DCA and let-7a induce apoptosis, reduce reactive oxygen species generation and autophagy, and stimulate mitochondrial biogenesis in triple-negative MDA-MB-231 breast cancer cells [284].

9.2. Statins

Statins are cholesterol-lowering drugs that inhibit 3-Hydroxy-3-Methylglutaryl-CoA Reductase (HMGCR) during cholesterol biosynthesis [285]. They are traditionally used for the treatment of hypercholesterolemia, atherosclerosis, and obesity. Some statins, such as simvastatin, fluvastatin, and lovastatin, have been discovered to exert through schemes of monotherapy or combination therapy that help overcome anticancer drug resistance [286,287,288,289,290]. For example, simvastatin reduced NF-κB and LIN28B expression and subsequently increased let-7 levels, which, in summary, significantly inhibited cell viability and clonal proliferation [291].

9.3. Aspirin

Aspirin, a non-steroidal anti-inflammatory drug (NSAIDs), has shown metabolic and anti-tumor properties [292,293,294], with strong evidence in colorectal, lung, and ovarian cancer [294,295,296]. Aspirin improves lung cancer by targeting the miR-98/WNT1 axis, and its combination with radiotherapy improves the survival of pancreatic cancer patients through miR-194-5p [297,298,299].

9.4. Methotrexate

Methotrexate belongs to a group of chemotherapy drugs called antimetabolites. They stop cells from synthesizing and repairing DNA. miR-192 favors the chemosensitization of MG-63 cells to MTX, making it a candidate agent to overcome MTX resistance in OS, osteosarcoma cancer cells [300].

10. Discussion

MiRNA therapeutics in cancer is a potential strategy that goes beyond the drawbacks of conventional pharmaceutical medications. They provide significant advantages compared to small-molecule drug approaches that target single proteins because they are endogenous natural molecules, so their targeting potential can be more beneficial than many other alternative synthetic biomolecules. In addition, due to the capacity of these molecules to target several genes of a multitarget regulatory network, their action might lead to a broader drug response. Unlike small-molecule and protein-based drugs, miRNAs manage with specificity, the inhibition of all targets, including those that are non-druggable. Their capacity to target several genes of a multitarget regulatory network may lead to a broader drug response (Table 3).
Finally, the chemical modifications of the oligonucleotides improve their pharmacodynamic and pharmacokinetic features, attributing them to drug properties, but they have also significantly improved their stability and protection against nucleases [184,301]. In contrast with small molecules and protein-based drugs, they allow the rapid identification and optimization of potent compounds. Similarly, for small molecules, their production is easy, and the products are stable [302,303].
Further, several oligonucleotide carriers have been developed to enhance stability and improve tissue penetration through in vivo viral and non-viral delivery miRNA methods that have been previously reviewed (Table 4) [105,106,169,184,252,301,304,305]. Although miRNA-based therapeutics are promising, they also have limitations and safety disadvantages associated with the risk of immunogenicity. Low RNA stability, uncertain tumor-specific delivery, and local retention of miRNAs affect the target-specific shipment of miRNAs [248,306].
Genetically modified viral vectors have long been used for gene therapy and also designed to deliver transgenes encoding miRNA mimics or antagonists [111]. In this review, we identified distinct characteristics and limitations of major non-viral and viral-based vectors used for miRNA delivery, including retroviral, lentiviral, adenoviral, and adeno-associated virus (AAV) reviewed previously [191].
Members of the lentivirus genus of the Retroviridae family have been tailored to develop Lentiviral vectors (LVs). LVs can actively translocate across an intact nuclear membrane, targeting both quiescent and non-quiescent cells [307]. In contrast to RVs that can only access the host chromosome once the nuclear membrane is disintegrated during mitosis, lentiviral vectors have reduced the risk of insertional mutagenesis and oncogenesis associated with RVs due to their integration within actively transcribing units as they can integrate their reverse transcribed DNA into the host genome, leading to insertional mutagenesis and the activation of oncogenic pathways [112]. LVs have been more successful in the delivery of therapeutic miRNA mimics or antagonists in cancer studies [191].
In addition, non-integrating adenoviruses and AAVs have been also used as alternative miRNA carriers because they keep their genomes in episomal forms [112]. AAVs are gene delivery systems owing to their non-pathogenic nature, broad target tissue spectrum, and sustainable presence in the biological system of action [191].
Although viral vectors are used in delivery strategies, hurdles such as toxicity, immunogenicity, and manufacturing complexity shifted the miRNA therapeutic research toward non-viral carriers. To overcome and address these issues, non-viral delivery systems have been developed, without being vulnerable to nuclease degradation, improving the effective transfer of miRNA or miRNA-expressing vectors inside the cell. The comparison of the chemical methods for non-viral miRNA delivery, including lipid, polymer, inorganic, and exosome delivery is illustrated in Table 4.
The pleiotropic action of miRNA inhibition and restoration therapy through a variety of delivery strategies suggests that these molecules hold great promise for cancer treatment and may become a future medicine. Small RNA-based therapies are currently at the core of innovation and a crucial field for obtaining patent rights. The most evident drawback for the biopharmaceutical companies is that, to date, although miRNA-targeting drugs have progressed to clinical trials, none of them have been entered into the clinicaltrials.gov database for phase III [119]. To overcome this limitation, it is necessary for the science to proceed by filling research gaps on several pending issues and operative problems. Thus, it is crucial to accurately define the distinct miRNA profile in different biological samples across a broader spectrum of cancers. Another limitation is the experimental cost that many studies require, and the replicability of these experiments has frequently been disappointingly low. Therefore, it is important to have a good estimate of the appropriate sample size for a high-throughput miRNA screening.
Overall, considering the challenging characteristics that face miRNA-based drug design in translational research regarding the degradation by nucleases upon addition into biological models [179], the low cell membrane permeability [111], the mechanisms of endosomal entrapment [111], the weak binding affinity for complementary sequences [177], the poor delivery to target organs [308], the undesirable toxicity, and the activation of the innate immune response [309], we understand that stringent criteria must be met before bringing miRNAs from bench to bedside for cancer therapy [177]. In addition, all the steps of the workflow development of miRNA therapeutics, including the stages that begin from the proof-of-concept research to the preclinical phase and evolve towards clinical trial design and drug validation, should help the successful monitoring process for FDA approval and treatment scale-up of the candidate anticancer miRNA-based drug, so that it may be marketed [112,191].

11. Conclusions

In this review, the role of miRNAs in anticancer therapy was highlighted. The antineoplastic effect of evolving target miRNAs in existing drugs that already satisfy safety criteria was also examined. Attention was given to miRNA inhibition and replacement strategies, comparing the viral and non-viral-based delivery platforms, considering the challenges faced in clinical translation. Biopharmaceutical companies should progress towards advanced delivery systems for the use of commercial therapeutic miRNAs, satisfying the criteria of absorption, distribution, metabolism, and elimination properties for their localized and systemic delivery to minimize the off-target effects and enable their safe use.

Author Contributions

Conceptualization, M.P.; software, C.C. and I.M.; investigation, M.P.; writing—original draft preparation, M.P.; writing—review and editing, M.P., C.C., D.C.S., M.A., V.Z., I.M. and N.D.; visualization, M.P. and C.C.; supervision, I.M. and N.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lu, J.; Getz, G.; Miska, E.A.; Alvarez-Saavedra, E.; Lamb, J.; Peck, D.; Sweet-Cordero, A.; Ebert, B.L.; Mak, R.H.; Ferrando, A.A.; et al. MicroRNA expression profiles classify human cancers. Nature 2005, 435, 834–838. [Google Scholar] [CrossRef] [PubMed]
  2. Chung, C.H.; Bernard, P.S.; Perou, C.M. Molecular portraits and the family tree of cancer. Nat. Genet. 2002, 32, 533–540. [Google Scholar] [CrossRef]
  3. Ramaswamy, M.; Wasan, K.M. Differences in the method by which plasma is separated from whole blood influences amphotericin B plasma recovery and distribution following amphotericin B lipid complex incubation within whole blood. Drug Dev. Ind. Pharm. 2001, 27, 871–875. [Google Scholar] [CrossRef]
  4. 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]
  5. Bartel, D.P. MicroRNAs: Genomics, biogenesis, mechanism, and function. Cell 2004, 116, 281–297. [Google Scholar] [CrossRef]
  6. Ambros, V.; Horvitz, H.R. Heterochronic mutants of the nematode Caenorhabditis elegans. Science 1984, 226, 409–416. [Google Scholar] [CrossRef]
  7. Reinhart, B.J.; Slack, F.J.; Basson, M.; Pasquinelli, A.E.; Bettinger, J.C.; Rougvie, A.E.; Horvitz, H.R.; Ruvkun, G. The 21-nucleotide let-7 RNA regulates developmental timing in Caenorhabditis elegans. Nature 2000, 403, 901–906. [Google Scholar] [CrossRef] [PubMed]
  8. Kozomara, A.; Birgaoanu, M.; Griffiths-Jones, S. miRBase: From microRNA sequences to function. Nucleic Acids Res. 2019, 47, D155–D162. [Google Scholar] [CrossRef]
  9. Kozomara, A.; Griffiths-Jones, S. miRBase: Annotating high confidence microRNAs using deep sequencing data. Nucleic Acids Res. 2014, 42, D68–D73. [Google Scholar] [CrossRef] [PubMed]
  10. Alles, J.; Fehlmann, T.; Fischer, U.; Backes, C.; Galata, V.; Minet, M.; Hart, M.; Abu-Halima, M.; Grasser, F.A.; Lenhof, H.P.; et al. An estimate of the total number of true human miRNAs. Nucleic Acids Res. 2019, 47, 3353–3364. [Google Scholar] [CrossRef]
  11. Fromm, B.; Zhong, X.; Tarbier, M.; Friedlander, M.R.; Hackenberg, M. The limits of human microRNA annotation have been met. RNA 2022, 28, 781–785. [Google Scholar] [CrossRef] [PubMed]
  12. da Silveira, W.A.; Renaud, L.; Hazard, E.S.; Hardiman, G. miRNA and lncRNA Expression Networks Modulate Cell Cycle and DNA Repair Inhibition in Senescent Prostate Cells. Genes 2022, 13, 208. [Google Scholar] [CrossRef] [PubMed]
  13. Brennecke, J.; Stark, A.; Russell, R.B.; Cohen, S.M. Principles of microRNA-target recognition. PLoS Biol. 2005, 3, e85. [Google Scholar] [CrossRef] [PubMed]
  14. Fu, X.; He, Y.; Song, J.; Wang, L.; Guo, P.; Cao, J. MiRNA-181b-5p Modulates Cell Proliferation, Cell Cycle, and Apoptosis by Targeting SSX2IP in Acute Lymphoblastic Leukemia. Turk. J. Haematol. 2022, 39, 160–169. [Google Scholar] [CrossRef]
  15. Chen, C.Z.; Li, L.; Lodish, H.F.; Bartel, D.P. MicroRNAs modulate hematopoietic lineage differentiation. Science 2004, 303, 83–86. [Google Scholar] [CrossRef]
  16. Jeong, H.C.; Shukla, S.; Fok, W.C.; Huynh, T.N.; Batista, L.F.Z.; Parker, R. USB1 is a miRNA deadenylase that regulates hematopoietic development. Science 2023, 379, 901–907. [Google Scholar] [CrossRef]
  17. Esau, C.; Kang, X.; Peralta, E.; Hanson, E.; Marcusson, E.G.; Ravichandran, L.V.; Sun, Y.; Koo, S.; Perera, R.J.; Jain, R.; et al. MicroRNA-143 regulates adipocyte differentiation. J. Biol. Chem. 2004, 279, 52361–52365. [Google Scholar] [CrossRef]
  18. Tan, X.; Zhu, T.; Zhang, L.; Fu, L.; Hu, Y.; Li, H.; Li, C.; Zhang, J.; Liang, B.; Liu, J. miR-669a-5p promotes adipogenic differentiation and induces browning in preadipocytes. Adipocyte 2022, 11, 120–132. [Google Scholar] [CrossRef]
  19. Poy, M.N.; Eliasson, L.; Krutzfeldt, J.; Kuwajima, S.; Ma, X.; Macdonald, P.E.; Pfeffer, S.; Tuschl, T.; Rajewsky, N.; Rorsman, P.; et al. A pancreatic islet-specific microRNA regulates insulin secretion. Nature 2004, 432, 226–230. [Google Scholar] [CrossRef]
  20. Aghaei, M.; Khodadadian, A.; Elham, K.N.; Nazari, M.; Babakhanzadeh, E. Major miRNA Involved in Insulin Secretion and Production in Beta-Cells. Int. J. Gen. Med. 2020, 13, 89–97. [Google Scholar] [CrossRef]
  21. Qian, B.; Yang, Y.; Tang, N.; Wang, J.; Sun, P.; Yang, N.; Chen, F.; Wu, T.; Sun, T.; Li, Y.; et al. M1 macrophage-derived exosomes impair beta cell insulin secretion via miR-212-5p by targeting SIRT2 and inhibiting Akt/GSK-3beta/beta-catenin pathway in mice. Diabetologia 2021, 64, 2037–2051. [Google Scholar] [CrossRef]
  22. Gregory, R.I.; Shiekhattar, R. MicroRNA biogenesis and cancer. Cancer Res. 2005, 65, 3509–3512. [Google Scholar] [CrossRef] [PubMed]
  23. Dostie, J.; Mourelatos, Z.; Yang, M.; Sharma, A.; Dreyfuss, G. Numerous microRNPs in neuronal cells containing novel microRNAs. RNA 2003, 9, 180–186. [Google Scholar] [CrossRef] [PubMed]
  24. Rastegar-Moghaddam, S.H.; Ebrahimzadeh-Bideskan, A.; Shahba, S.; Malvandi, A.M.; Mohammadipour, A. Roles of the miR-155 in Neuroinflammation and Neurological Disorders: A Potent Biological and Therapeutic Target. Cell Mol. Neurobiol. 2023, 43, 455–467. [Google Scholar] [CrossRef]
  25. Kumar, M.; Li, G. Emerging Role of MicroRNA-30c in Neurological Disorders. Int. J. Mol. Sci. 2022, 24, 37. [Google Scholar] [CrossRef] [PubMed]
  26. Pfeffer, S.; Zavolan, M.; Grasser, F.A.; Chien, M.; Russo, J.J.; Ju, J.; John, B.; Enright, A.J.; Marks, D.; Sander, C.; et al. Identification of virus-encoded microRNAs. Science 2004, 304, 734–736. [Google Scholar] [CrossRef]
  27. Fernandez-Pato, A.; Virseda-Berdices, A.; Resino, S.; Ryan, P.; Martinez-Gonzalez, O.; Perez-Garcia, F.; Martin-Vicente, M.; Valle-Millares, D.; Brochado-Kith, O.; Blancas, R.; et al. Plasma miRNA profile at COVID-19 onset predicts severity status and mortality. Emerg. Microbes Infect. 2022, 11, 676–688. [Google Scholar] [CrossRef] [PubMed]
  28. Lee, W. MicroRNA, Insulin Resistance, and Metabolic Disorders. Int. J. Mol. Sci. 2022, 23, 16215. [Google Scholar] [CrossRef]
  29. Macvanin, M.; Obradovic, M.; Zafirovic, S.; Stanimirovic, J.; Isenovic, E.R. The Role of miRNAs in Metabolic Diseases. Curr. Med. Chem. 2023, 30, 1922–1944. [Google Scholar] [CrossRef]
  30. Ye, Z.; Wang, S.; Huang, X.; Chen, P.; Deng, L.; Li, S.; Lin, S.; Wang, Z.; Liu, B. Plasma Exosomal miRNAs Associated With Metabolism as Early Predictor of Gestational Diabetes Mellitus. Diabetes 2022, 71, 2272–2283. [Google Scholar] [CrossRef]
  31. Chen, Y.; Shen, Y.; Lin, P.; Tong, D.; Zhao, Y.; Allesina, S.; Shen, X.; Wu, C.I. Gene regulatory network stabilized by pervasive weak repressions: MicroRNA functions revealed by the May-Wigner theory. Natl. Sci. Rev. 2019, 6, 1176–1188. [Google Scholar] [CrossRef] [PubMed]
  32. Creemers, E.E.; Tijsen, A.J.; Pinto, Y.M. Circulating microRNAs: Novel biomarkers and extracellular communicators in cardiovascular disease? Circ. Res. 2012, 110, 483–495. [Google Scholar] [CrossRef] [PubMed]
  33. Xin, L.; Liu, Y.H.; Martin, T.A.; Jiang, W.G. The Era of Multigene Panels Comes? The Clinical Utility of Oncotype DX and MammaPrint. World J. Oncol. 2017, 8, 34–40. [Google Scholar] [CrossRef] [PubMed]
  34. Avgeris, M.; Mavridis, K.; Scorilas, A. Kallikrein-related peptidases in prostate, breast, and ovarian cancers: From pathobiology to clinical relevance. Biol. Chem. 2012, 393, 301–317. [Google Scholar] [CrossRef]
  35. Bertoli, G.; Cava, C.; Castiglioni, I. MicroRNAs as Biomarkers for Diagnosis, Prognosis and Theranostics in Prostate Cancer. Int. J. Mol. Sci. 2016, 17, 421. [Google Scholar] [CrossRef]
  36. Cortez, M.A.; Bueso-Ramos, C.; Ferdin, J.; Lopez-Berestein, G.; Sood, A.K.; Calin, G.A. MicroRNAs in body fluids--the mix of hormones and biomarkers. Nat. Rev. Clin. Oncol. 2011, 8, 467–477. [Google Scholar] [CrossRef] [PubMed]
  37. Arroyo, J.D.; Chevillet, J.R.; Kroh, E.M.; Ruf, I.K.; Pritchard, C.C.; Gibson, D.F.; Mitchell, P.S.; Bennett, C.F.; Pogosova-Agadjanyan, E.L.; Stirewalt, D.L.; et al. Argonaute2 complexes carry a population of circulating microRNAs independent of vesicles in human plasma. Proc. Natl. Acad. Sci. USA 2011, 108, 5003–5008. [Google Scholar] [CrossRef]
  38. Cruz-Burgos, M.; Cortes-Ramirez, S.A.; Losada-Garcia, A.; Morales-Pacheco, M.; Martinez-Martinez, E.; Morales-Montor, J.G.; Servin-Haddad, A.; Izquierdo-Luna, J.S.; Rodriguez-Martinez, G.; Ramos-Godinez, M.D.P.; et al. Unraveling the Role of EV-Derived miR-150-5p in Prostate Cancer Metastasis and Its Association with High-Grade Gleason Scores: Implications for Diagnosis. Cancers 2023, 15, 4148. [Google Scholar] [CrossRef]
  39. Melo, S.A.; Sugimoto, H.; O’Connell, J.T.; Kato, N.; Villanueva, A.; Vidal, A.; Qiu, L.; Vitkin, E.; Perelman, L.T.; Melo, C.A.; et al. Cancer exosomes perform cell-independent microRNA biogenesis and promote tumorigenesis. Cancer Cell 2014, 26, 707–721. [Google Scholar] [CrossRef]
  40. Tan, S.; Xia, L.; Yi, P.; Han, Y.; Tang, L.; Pan, Q.; Tian, Y.; Rao, S.; Oyang, L.; Liang, J.; et al. Exosomal miRNAs in tumor microenvironment. J. Exp. Clin. Cancer Res. CR 2020, 39, 67. [Google Scholar] [CrossRef]
  41. Zhou, J.; Nagarkatti, P.; Zhong, Y.; Ginsberg, J.P.; Singh, N.P.; Zhang, J.; Nagarkatti, M. Dysregulation in microRNA expression is associated with alterations in immune functions in combat veterans with post-traumatic stress disorder. PLoS ONE 2014, 9, e94075. [Google Scholar] [CrossRef] [PubMed]
  42. Kannampuzha, S.; Murali, R.; Gopalakrishnan, A.V.; Mukherjee, A.G.; Wanjari, U.R.; Namachivayam, A.; George, A.; Dey, A.; Vellingiri, B. Novel biomolecules in targeted cancer therapy: A new approach towards precision medicine. Med. Oncol. 2023, 40, 323. [Google Scholar] [CrossRef] [PubMed]
  43. Katoh, M.; Igarashi, M.; Fukuda, H.; Nakagama, H.; Katoh, M. Cancer genetics and genomics of human FOX family genes. Cancer Lett. 2013, 328, 198–206. [Google Scholar] [CrossRef]
  44. Antonino, M.; Nicolo, M.; Jerome Renee, L.; Federico, M.; Chiara, V.; Stefano, S.; Maria, S.; Salvatore, C.; Antonio, B.; Calvo-Henriquez, C.; et al. Single-nucleotide polymorphism in chronic rhinosinusitis: A systematic review. Clin. Otolaryngol. 2022, 47, 14–23. [Google Scholar] [CrossRef]
  45. Franczyk, B.; Rysz, J.; Gluba-Brzozka, A. Pharmacogenetics of Drugs Used in the Treatment of Cancers. Genes 2022, 13, 311. [Google Scholar] [CrossRef] [PubMed]
  46. Imyanitov, E.N.; Iyevleva, A.G. Molecular tests for prediction of tumor sensitivity to cytotoxic drugs. Cancer Lett. 2022, 526, 41–52. [Google Scholar] [CrossRef]
  47. Madkour, M.M.; Ramadan, W.S.; Saleh, E.; El-Awady, R. Epigenetic modulations in cancer: Predictive biomarkers and potential targets for overcoming the resistance to topoisomerase I inhibitors. Ann. Med. 2023, 55, 2203946. [Google Scholar] [CrossRef]
  48. Meng, C.L.; Zhao, W.; Zhong, D.N. Epigenetics and microRNAs in UGT1As. Hum. Genom. 2021, 15, 30. [Google Scholar] [CrossRef]
  49. Zappe, K.; Cichna-Markl, M. Aberrant DNA Methylation of ABC Transporters in Cancer. Cells 2020, 9, 2281. [Google Scholar] [CrossRef]
  50. Aboul-Soud, M.A.M.; Alzahrani, A.J.; Mahmoud, A. Decoding variants in drug-metabolizing enzymes and transporters in solid tumor patients by whole-exome sequencing. Saudi J. Biol. Sci. 2021, 28, 628–634. [Google Scholar] [CrossRef]
  51. Hu, X.; Qin, W.; Li, S.; He, M.; Wang, Y.; Guan, S.; Zhao, H.; Yao, W.; Wei, M.; Liu, M.; et al. Polymorphisms in DNA repair pathway genes and ABCG2 gene in advanced colorectal cancer: Correlation with tumor characteristics and clinical outcome in oxaliplatin-based chemotherapy. Cancer Manag. Res. 2019, 11, 285–297. [Google Scholar] [CrossRef]
  52. Saini, H.K.; Griffiths-Jones, S.; Enright, A.J. Genomic analysis of human microRNA transcripts. Proc. Natl. Acad. Sci. USA 2007, 104, 17719–17724. [Google Scholar] [CrossRef]
  53. Lee, Y.; Kim, M.; Han, J.; Yeom, K.H.; Lee, S.; Baek, S.H.; Kim, V.N. MicroRNA genes are transcribed by RNA polymerase II. EMBO J. 2004, 23, 4051–4060. [Google Scholar] [CrossRef]
  54. Bartel, D.P.; Chen, C.Z. Micromanagers of gene expression: The potentially widespread influence of metazoan microRNAs. Nat. Rev. Genet. 2004, 5, 396–400. [Google Scholar] [CrossRef]
  55. Li, Z.; Rana, T.M. Therapeutic targeting of microRNAs: Current status and future challenges. Nat. Rev. Drug Discov. 2014, 13, 622–638. [Google Scholar] [CrossRef]
  56. Yi, R.; Qin, Y.; Macara, I.G.; Cullen, B.R. Exportin-5 mediates the nuclear export of pre-microRNAs and short hairpin RNAs. Genes. Dev. 2003, 17, 3011–3016. [Google Scholar] [CrossRef] [PubMed]
  57. Zeng, Y.; Cullen, B.R. Structural requirements for pre-microRNA binding and nuclear export by Exportin 5. Nucleic Acids Res. 2004, 32, 4776–4785. [Google Scholar] [CrossRef] [PubMed]
  58. Chendrimada, T.P.; Gregory, R.I.; Kumaraswamy, E.; Norman, J.; Cooch, N.; Nishikura, K.; Shiekhattar, R. TRBP recruits the Dicer complex to Ago2 for microRNA processing and gene silencing. Nature 2005, 436, 740–744. [Google Scholar] [CrossRef] [PubMed]
  59. Feng, Y.; Zhang, X.; Graves, P.; Zeng, Y. A comprehensive analysis of precursor microRNA cleavage by human Dicer. Rna 2012, 18, 2083–2092. [Google Scholar] [CrossRef]
  60. Nakanishi, K. Anatomy of four human Argonaute proteins. Nucleic Acids Res. 2022, 50, 6618–6638. [Google Scholar] [CrossRef]
  61. Medley, J.C.; Panzade, G.; Zinovyeva, A.Y. microRNA strand selection: Unwinding the rules. Wiley Interdiscip. Rev. RNA 2021, 12, e1627. [Google Scholar] [CrossRef] [PubMed]
  62. Kim, V.N.; Han, J.; Siomi, M.C. Biogenesis of small RNAs in animals. Nat. Rev. Mol. Cell Biol. 2009, 10, 126–139. [Google Scholar] [CrossRef] [PubMed]
  63. Goodfellow, S.J.; White, R.J. Regulation of RNA polymerase III transcription during mammalian cell growth. Cell Cycle 2007, 6, 2323–2326. [Google Scholar] [CrossRef]
  64. Borchert, G.M.; Lanier, W.; Davidson, B.L. RNA polymerase III transcribes human microRNAs. Nat. Struct. Mol. Biol. 2006, 13, 1097–1101. [Google Scholar] [CrossRef]
  65. Czech, B.; Hannon, G.J. Small RNA sorting: Matchmaking for Argonautes. Nat. Rev. Genet. 2011, 12, 19–31. [Google Scholar] [CrossRef] [PubMed]
  66. Bartel, D.P. MicroRNAs: Target recognition and regulatory functions. Cell 2009, 136, 215–233. [Google Scholar] [CrossRef]
  67. Perron, M.P.; Provost, P. Protein interactions and complexes in human microRNA biogenesis and function. Front. Biosci. A J. Virtual Libr. 2008, 13, 2537–2547. [Google Scholar] [CrossRef]
  68. Eamens, A.L.; Wang, M.B. Alternate approaches to repress endogenous microRNA activity in Arabidopsis thaliana. Plant Signal. Behav. 2011, 6, 349–359. [Google Scholar] [CrossRef]
  69. Zhu, H.; Hu, F.; Wang, R.; Zhou, X.; Sze, S.H.; Liou, L.W.; Barefoot, A.; Dickman, M.; Zhang, X. Arabidopsis Argonaute10 specifically sequesters miR166/165 to regulate shoot apical meristem development. Cell 2011, 145, 242–256. [Google Scholar] [CrossRef]
  70. Doxakis, E. Principles of miRNA-target regulation in metazoan models. Int. J. Mol. Sci. 2013, 14, 16280–16302. [Google Scholar] [CrossRef]
  71. Orom, U.A.; Nielsen, F.C.; Lund, A.H. MicroRNA-10a binds the 5’UTR of ribosomal protein mRNAs and enhances their translation. Mol. Cell 2008, 30, 460–471. [Google Scholar] [CrossRef]
  72. Bartel, D.P. Metazoan MicroRNAs. Cell 2018, 173, 20–51. [Google Scholar] [CrossRef] [PubMed]
  73. Valadi, H.; Ekstrom, K.; Bossios, A.; Sjostrand, M.; Lee, J.J.; Lotvall, J.O. Exosome-mediated transfer of mRNAs and microRNAs is a novel mechanism of genetic exchange between cells. Nat. Cell Biol. 2007, 9, 654–659. [Google Scholar] [CrossRef] [PubMed]
  74. Chirshev, E.; Oberg, K.C.; Ioffe, Y.J.; Unternaehrer, J.J. Let-7 as biomarker, prognostic indicator, and therapy for precision medicine in cancer. Clin. Transl. Med. 2019, 8, 24. [Google Scholar] [CrossRef] [PubMed]
  75. Ameres, S.L.; Zamore, P.D. Diversifying microRNA sequence and function. Nat. Rev. Mol. Cell Biol. 2013, 14, 475–488. [Google Scholar] [CrossRef] [PubMed]
  76. Chim, S.S.; Shing, T.K.; Hung, E.C.; Leung, T.Y.; Lau, T.K.; Chiu, R.W.; Lo, Y.M. Detection and characterization of placental microRNAs in maternal plasma. Clin. Chem. 2008, 54, 482–490. [Google Scholar] [CrossRef] [PubMed]
  77. 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]
  78. Taylor, D.D.; Gercel-Taylor, C. MicroRNA signatures of tumor-derived exosomes as diagnostic biomarkers of ovarian cancer. Gynecol. Oncol. 2008, 110, 13–21. [Google Scholar] [CrossRef]
  79. O’Brien, J.; Hayder, H.; Zayed, Y.; Peng, C. Overview of MicroRNA Biogenesis, Mechanisms of Actions, and Circulation. Front. Endocrinol. 2018, 9, 402. [Google Scholar] [CrossRef]
  80. Avgeris, M.; Panoutsopoulou, K.; Papadimitriou, M.A.; Scorilas, A. Circulating exosomal miRNAs: Clinical significance in human cancers. Expert. Rev. Mol. Diagn. 2019, 19, 979–995. [Google Scholar] [CrossRef]
  81. Alotaibi, F. Exosomal microRNAs in cancer: Potential biomarkers and immunotherapeutic targets for immune checkpoint molecules. Front. Genet. 2023, 14, 1052731. [Google Scholar] [CrossRef] [PubMed]
  82. Calin, G.A.; Croce, C.M. MicroRNA signatures in human cancers. Nat. Rev. Cancer 2006, 6, 857–866. [Google Scholar] [CrossRef] [PubMed]
  83. Esquela-Kerscher, A.; Slack, F.J. Oncomirs—microRNAs with a role in cancer. Nat. Rev. Cancer 2006, 6, 259–269. [Google Scholar] [CrossRef]
  84. Mattick, J.S. The hidden genetic program of complex organisms. Sci. Am. 2004, 291, 60–67. [Google Scholar] [CrossRef]
  85. Muthamilselvan, S.; Ramasami Sundhar Baabu, P.; Palaniappan, A. Microfluidics for Profiling miRNA Biomarker Panels in AI-Assisted Cancer Diagnosis and Prognosis. Technol. Cancer Res. Treat. 2023, 22, 15330338231185284. [Google Scholar] [CrossRef] [PubMed]
  86. 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]
  87. Kim, M.S.; Oh, J.E.; Kim, Y.R.; Park, S.W.; Kang, M.R.; Kim, S.S.; Ahn, C.H.; Yoo, N.J.; Lee, S.H. Somatic mutations and losses of expression of microRNA regulation-related genes AGO2 and TNRC6A in gastric and colorectal cancers. J. Pathol. 2010, 221, 139–146. [Google Scholar] [CrossRef]
  88. 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]
  89. Kumar, M.S.; Lu, J.; Mercer, K.L.; Golub, T.R.; Jacks, T. Impaired microRNA processing enhances cellular transformation and tumorigenesis. Nat. Genet. 2007, 39, 673–677. [Google Scholar] [CrossRef]
  90. McManus, M.T. MicroRNAs and cancer. Semin. Cancer Biol. 2003, 13, 253–258. [Google Scholar] [CrossRef]
  91. Gong, H.; Liu, C.M.; Liu, D.P.; Liang, C.C. The role of small RNAs in human diseases: Potential troublemaker and therapeutic tools. Med. Res. Rev. 2005, 25, 361–381. [Google Scholar] [CrossRef] [PubMed]
  92. Melo, S.A.; Ropero, S.; Moutinho, C.; Aaltonen, L.A.; Yamamoto, H.; Calin, G.A.; Rossi, S.; Fernandez, A.F.; Carneiro, F.; Oliveira, C.; et al. A TARBP2 mutation in human cancer impairs microRNA processing and DICER1 function. Nat. Genet. 2009, 41, 365–370. [Google Scholar] [CrossRef]
  93. Hill, D.A.; Ivanovich, J.; Priest, J.R.; Gurnett, C.A.; Dehner, L.P.; Desruisseau, D.; Jarzembowski, J.A.; Wikenheiser-Brokamp, K.A.; Suarez, B.K.; Whelan, A.J.; et al. DICER1 mutations in familial pleuropulmonary blastoma. Science 2009, 325, 965. [Google Scholar] [CrossRef]
  94. Melo, S.A.; Moutinho, C.; Ropero, S.; Calin, G.A.; Rossi, S.; Spizzo, R.; Fernandez, A.F.; Davalos, V.; Villanueva, A.; Montoya, G.; et al. A genetic defect in exportin-5 traps precursor microRNAs in the nucleus of cancer cells. Cancer Cell 2010, 18, 303–315. [Google Scholar] [CrossRef]
  95. 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]
  96. Wegert, J.; Ishaque, N.; Vardapour, R.; Georg, C.; Gu, Z.; Bieg, M.; Ziegler, B.; Bausenwein, S.; Nourkami, N.; Ludwig, N.; et al. Mutations in the SIX1/2 pathway and the DROSHA/DGCR8 miRNA microprocessor complex underlie high-risk blastemal type Wilms tumors. Cancer Cell 2015, 27, 298–311. [Google Scholar] [CrossRef] [PubMed]
  97. Walz, A.L.; Ooms, A.; Gadd, S.; Gerhard, D.S.; Smith, M.A.; Guidry Auvil, J.M.; Meerzaman, D.; Chen, Q.R.; Hsu, C.H.; Yan, C.; et al. Recurrent DGCR8, DROSHA, and SIX homeodomain mutations in favorable histology Wilms tumors. Cancer Cell 2015, 27, 286–297. [Google Scholar] [CrossRef]
  98. Sugito, N.; Ishiguro, H.; Kuwabara, Y.; Kimura, M.; Mitsui, A.; Kurehara, H.; Ando, T.; Mori, R.; Takashima, N.; Ogawa, R.; et al. RNASEN regulates cell proliferation and affects survival in esophageal cancer patients. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2006, 12, 7322–7328. [Google Scholar] [CrossRef]
  99. Diaz-Garcia, C.V.; Agudo-Lopez, A.; Perez, C.; Lopez-Martin, J.A.; Rodriguez-Peralto, J.L.; de Castro, J.; Cortijo, A.; Martinez-Villanueva, M.; Iglesias, L.; Garcia-Carbonero, R.; et al. DICER1, DROSHA and miRNAs in patients with non-small cell lung cancer: Implications for outcomes and histologic classification. Carcinogenesis 2013, 34, 1031–1038. [Google Scholar] [CrossRef]
  100. Lin, R.J.; Lin, Y.C.; Chen, J.; Kuo, H.H.; Chen, Y.Y.; Diccianni, M.B.; London, W.B.; Chang, C.H.; Yu, A.L. microRNA signature and expression of Dicer and Drosha can predict prognosis and delineate risk groups in neuroblastoma. Cancer Res. 2010, 70, 7841–7850. [Google Scholar] [CrossRef]
  101. Doros, L.; Schultz, K.A.; Stewart, D.R.; Bauer, A.J.; Williams, G.; Rossi, C.T.; Carr, A.; Yang, J.; Dehner, L.P.; Messinger, Y.; et al. DICER1-Related Disorders. In GeneReviews®; Pagon, R.A., Adam, M.P., Ardinger, H.H., Wallace, S.E., Amemiya, A., Bean, L.J.H., Bird, T.D., Fong, C.T., Mefford, H.C., Smith, R.J.H., et al., Eds.; University of Washington: Seattle, WA, USA, 1993. [Google Scholar]
  102. Sand, M.; Gambichler, T.; Skrygan, M.; Sand, D.; Scola, N.; Altmeyer, P.; Bechara, F.G. Expression levels of the microRNA processing enzymes Drosha and dicer in epithelial skin cancer. Cancer Investig. 2010, 28, 649–653. [Google Scholar] [CrossRef]
  103. Sand, M.; Gambichler, T.; Sand, D.; Altmeyer, P.; Stuecker, M.; Bechara, F.G. Immunohistochemical expression patterns of the microRNA-processing enzyme Dicer in cutaneous malignant melanomas, benign melanocytic nevi and dysplastic melanocytic nevi. Eur. J. Dermatol. 2011, 21, 18–21. [Google Scholar] [CrossRef] [PubMed]
  104. Sand, M.; Skrygan, M.; Georgas, D.; Arenz, C.; Gambichler, T.; Sand, D.; Altmeyer, P.; Bechara, F.G. Expression levels of the microRNA maturing microprocessor complex component DGCR8 and the RNA-induced silencing complex (RISC) components argonaute-1, argonaute-2, PACT, TARBP1, and TARBP2 in epithelial skin cancer. Mol. Carcinog. 2012, 51, 916–922. [Google Scholar] [CrossRef] [PubMed]
  105. Iacomino, G. miRNAs: The Road from Bench to Bedside. Genes 2023, 14, 314. [Google Scholar] [CrossRef] [PubMed]
  106. Zhu, Y.; Zhu, L.; Wang, X.; Jin, H. RNA-based therapeutics: An overview and prospectus. Cell Death Dis. 2022, 13, 644. [Google Scholar] [CrossRef]
  107. Jo, S.J.; Chae, S.U.; Lee, C.B.; Bae, S.K. Clinical Pharmacokinetics of Approved RNA Therapeutics. Int. J. Mol. Sci. 2023, 24, 746. [Google Scholar] [CrossRef]
  108. Das, N.; Tripathi, N.; Khurana, S. Micro RNA Mimics And Antagonists. Int. J. Sci. Technol. Res. 2015, 4, 176–180. [Google Scholar]
  109. Manikkath, J.; Jishnu, P.V.; Wich, P.R.; Manikkath, A.; Radhakrishnan, R. Nanoparticulate strategies for the delivery of miRNA mimics and inhibitors in anticancer therapy and its potential utility in oral submucous fibrosis. Nanomedicine 2022, 17, 181–195. [Google Scholar] [CrossRef]
  110. Kang, E.; Kortylewski, M. Lipid Nanoparticle-Mediated Delivery of miRNA Mimics to Myeloid Cells. Methods Mol. Biol. 2023, 2691, 337–350. [Google Scholar] [CrossRef]
  111. Fu, Y.; Chen, J.; Huang, Z. Recent progress in microRNA-based delivery systems for the treatment of human disease. ExRNA 2019, 1, 24. [Google Scholar] [CrossRef]
  112. Reda El Sayed, S.; Cristante, J.; Guyon, L.; Denis, J.; Chabre, O.; Cherradi, N. MicroRNA Therapeutics in Cancer: Current Advances and Challenges. Cancers 2021, 13, 2680. [Google Scholar] [CrossRef] [PubMed]
  113. He, B.; Zhao, Z.; Cai, Q.; Zhang, Y.; Zhang, P.; Shi, S.; Xie, H.; Peng, X.; Yin, W.; Tao, Y.; et al. miRNA-based biomarkers, therapies, and resistance in Cancer. Int. J. Biol. Sci. 2020, 16, 2628–2647. [Google Scholar] [CrossRef]
  114. Fu, Z.; Wang, L.; Li, S.; Chen, F.; Au-Yeung, K.K.; Shi, C. MicroRNA as an Important Target for Anticancer Drug Development. Front. Pharmacol. 2021, 12, 736323. [Google Scholar] [CrossRef]
  115. Otoukesh, B.; Abbasi, M.; Gorgani, H.O.; Farahini, H.; Moghtadaei, M.; Boddouhi, B.; Kaghazian, P.; Hosseinzadeh, S.; Alaee, A. MicroRNAs signatures, bioinformatics analysis of miRNAs, miRNA mimics and antagonists, and miRNA therapeutics in osteosarcoma. Cancer Cell Int. 2020, 20, 254. [Google Scholar] [CrossRef] [PubMed]
  116. Check Hayden, E. Cancer complexity slows quest for cure. Nature 2008, 455, 148. [Google Scholar] [CrossRef]
  117. Parsons, D.W.; Jones, S.; Zhang, X.; Lin, J.C.; Leary, R.J.; Angenendt, P.; Mankoo, P.; Carter, H.; Siu, I.M.; Gallia, G.L.; et al. An integrated genomic analysis of human glioblastoma multiforme. Science 2008, 321, 1807–1812. [Google Scholar] [CrossRef] [PubMed]
  118. Medina, P.P.; Nolde, M.; Slack, F.J. OncomiR addiction in an in vivo model of microRNA-21-induced pre-B-cell lymphoma. Nature 2010, 467, 86–90. [Google Scholar] [CrossRef]
  119. Menon, A.; Abd-Aziz, N.; Khalid, K.; Poh, C.L.; Naidu, R. miRNA: A Promising Therapeutic Target in Cancer. Int. J. Mol. Sci. 2022, 23, 11502. [Google Scholar] [CrossRef]
  120. Garzon, R.; Marcucci, G.; Croce, C.M. Targeting microRNAs in cancer: Rationale, strategies and challenges. Nat. Rev. Drug Discov. 2010, 9, 775–789. [Google Scholar] [CrossRef]
  121. Cannataro, R.; Cione, E. miRNA as Drug: Antagomir and Beyond. Curr. Pharm. Des. 2023, 29, 462–465. [Google Scholar] [CrossRef]
  122. Ortega, M.M.; Bouamar, H. Guidelines on Designing MicroRNA Sponges: From Construction to Stable Cell Line. Methods Mol. Biol. 2023, 2595, 171–183. [Google Scholar] [CrossRef]
  123. Takegawa-Araki, T.; Kumagai, S.; Yasukawa, K.; Kuroda, M.; Sasaki, T.; Obika, S. Structure-Activity Relationships of Anti-microRNA Oligonucleotides Containing Cationic Guanidine-Modified Nucleic Acids. J. Med. Chem. 2022, 65, 2139–2148. [Google Scholar] [CrossRef]
  124. Hutvagner, G.; Simard, M.J.; Mello, C.C.; Zamore, P.D. Sequence-specific inhibition of small RNA function. PLoS Biol. 2004, 2, E98. [Google Scholar] [CrossRef] [PubMed]
  125. Orom, U.A.; Kauppinen, S.; Lund, A.H. LNA-modified oligonucleotides mediate specific inhibition of microRNA function. Gene 2006, 372, 137–141. [Google Scholar] [CrossRef] [PubMed]
  126. Zhang, J.; Sharma, R.; Ryu, K.; Shen, P.; Salaita, K.; Jo, H. Conditional Antisense Oligonucleotides Triggered by miRNA. ACS Chem. Biol. 2021, 16, 2255–2267. [Google Scholar] [CrossRef] [PubMed]
  127. Quemener, A.M.; Bachelot, L.; Forestier, A.; Donnou-Fournet, E.; Gilot, D.; Galibert, M.D. The powerful world of antisense oligonucleotides: From bench to bedside. Wiley Interdiscip. Rev. RNA 2020, 11, e1594. [Google Scholar] [CrossRef] [PubMed]
  128. Seth, P.P.; Siwkowski, A.; Allerson, C.R.; Vasquez, G.; Lee, S.; Prakash, T.P.; Wancewicz, E.V.; Witchell, D.; Swayze, E.E. Short antisense oligonucleotides with novel 2′-4′ conformationaly restricted nucleoside analogues show improved potency without increased toxicity in animals. J. Med. Chem. 2009, 52, 10–13. [Google Scholar] [CrossRef]
  129. Shadid, M.; Badawi, M.; Abulrob, A. Antisense oligonucleotides: Absorption, distribution, metabolism, and excretion. Expert. Opin. Drug Metab. Toxicol. 2021, 17, 1281–1292. [Google Scholar] [CrossRef]
  130. Monia, B.P.; Lesnik, E.A.; Gonzalez, C.; Lima, W.F.; McGee, D.; Guinosso, C.J.; Kawasaki, A.M.; Cook, P.D.; Freier, S.M. Evaluation of 2′-modified oligonucleotides containing 2′-deoxy gaps as antisense inhibitors of gene expression. J. Biol. Chem. 1993, 268, 14514–14522. [Google Scholar] [CrossRef] [PubMed]
  131. Lennox, K.A.; Behlke, M.A. Chemical modification and design of anti-miRNA oligonucleotides. Gene Ther. 2011, 18, 1111–1120. [Google Scholar] [CrossRef]
  132. Meister, G.; Landthaler, M.; Dorsett, Y.; Tuschl, T. Sequence-specific inhibition of microRNA- and siRNA-induced RNA silencing. Rna 2004, 10, 544–550. [Google Scholar] [CrossRef]
  133. Esau, C.C. Inhibition of microRNA with antisense oligonucleotides. Methods 2008, 44, 55–60. [Google Scholar] [CrossRef] [PubMed]
  134. Spizzo, R.; Rushworth, D.; Guerrero, M.; Calin, G.A. RNA inhibition, microRNAs, and new therapeutic agents for cancer treatment. Clin. Lymphoma Myeloma 2009, 9 (Suppl. S3), S313–S318. [Google Scholar] [CrossRef]
  135. Zhang, S.; Chen, L.; Jung, E.J.; Calin, G.A. Targeting microRNAs with small molecules: From dream to reality. Clin. Pharmacol. Ther. 2010, 87, 754–758. [Google Scholar] [CrossRef]
  136. Moriyama, T.; Ohuchida, K.; Mizumoto, K.; Yu, J.; Sato, N.; Nabae, T.; Takahata, S.; Toma, H.; Nagai, E.; Tanaka, M. MicroRNA-21 modulates biological functions of pancreatic cancer cells including their proliferation, invasion, and chemoresistance. Mol. Cancer Ther. 2009, 8, 1067–1074. [Google Scholar] [CrossRef] [PubMed]
  137. Park, J.K.; Lee, E.J.; Esau, C.; Schmittgen, T.D. Antisense inhibition of microRNA-21 or -221 arrests cell cycle, induces apoptosis, and sensitizes the effects of gemcitabine in pancreatic adenocarcinoma. Pancreas 2009, 38, e190–e199. [Google Scholar] [CrossRef]
  138. Zhao, Y.; Zhao, L.; Ischenko, I.; Bao, Q.; Schwarz, B.; Niess, H.; Wang, Y.; Renner, A.; Mysliwietz, J.; Jauch, K.W.; et al. Antisense inhibition of microRNA-21 and microRNA-221 in tumor-initiating stem-like cells modulates tumorigenesis, metastasis, and chemotherapy resistance in pancreatic cancer. Target. Oncol. 2015, 10, 535–548. [Google Scholar] [CrossRef] [PubMed]
  139. Krutzfeldt, J.; Rajewsky, N.; Braich, R.; Rajeev, K.G.; Tuschl, T.; Manoharan, M.; Stoffel, M. Silencing of microRNAs in vivo with ‘antagomirs’. Nature 2005, 438, 685–689. [Google Scholar] [CrossRef] [PubMed]
  140. Lima, J.F.; Cerqueira, L.; Figueiredo, C.; Oliveira, C.; Azevedo, N.F. Anti-miRNA oligonucleotides: A comprehensive guide for design. RNA Biol. 2018, 15, 338–352. [Google Scholar] [CrossRef]
  141. 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]
  142. Baker, B.F.; Lot, S.S.; Condon, T.P.; Cheng-Flournoy, S.; Lesnik, E.A.; Sasmor, H.M.; Bennett, C.F. 2′-O-(2-Methoxy)ethyl-modified anti-intercellular adhesion molecule 1 (ICAM-1) oligonucleotides selectively increase the ICAM-1 mRNA level and inhibit formation of the ICAM-1 translation initiation complex in human umbilical vein endothelial cells. J. Biol. Chem. 1997, 272, 11994–12000. [Google Scholar] [CrossRef] [PubMed]
  143. Friedman, K.J.; Kole, J.; Cohn, J.A.; Knowles, M.R.; Silverman, L.M.; Kole, R. Correction of aberrant splicing of the cystic fibrosis transmembrane conductance regulator (CFTR) gene by antisense oligonucleotides. J. Biol. Chem. 1999, 274, 36193–36199. [Google Scholar] [CrossRef]
  144. Mani, S.; Goel, S.; Nesterova, M.; Martin, R.M.; Grindel, J.M.; Rothenberg, M.L.; Zhang, R.; Tortora, G.; Cho-Chung, Y.S. Clinical studies in patients with solid tumors using a second-generation antisense oligonucleotide (GEM 231) targeted against protein kinase A type I. Ann. N. Y. Acad. Sci. 2003, 1002, 252–262. [Google Scholar] [CrossRef] [PubMed]
  145. Trang, P.; Medina, P.P.; Wiggins, J.F.; Ruffino, L.; Kelnar, K.; Omotola, M.; Homer, R.; Brown, D.; Bader, A.G.; Weidhaas, J.B.; et al. Regression of murine lung tumors by the let-7 microRNA. Oncogene 2010, 29, 1580–1587. [Google Scholar] [CrossRef] [PubMed]
  146. Si, M.L.; Zhu, S.; Wu, H.; Lu, Z.; Wu, F.; Mo, Y.Y. miR-21-mediated tumor growth. Oncogene 2007, 26, 2799–2803. [Google Scholar] [CrossRef]
  147. Lennox, K.A.; Owczarzy, R.; Thomas, D.M.; Walder, J.A.; Behlke, M.A. Improved Performance of Anti-miRNA Oligonucleotides Using a Novel Non-Nucleotide Modifier. Mol. Ther.-Nucleic Acids 2013, 2, e117. [Google Scholar] [CrossRef]
  148. Gleave, M.E.; Monia, B.P. Antisense therapy for cancer. Nat. Rev. Cancer 2005, 5, 468–479. [Google Scholar] [CrossRef]
  149. Sewell, K.L.; Geary, R.S.; Baker, B.F.; Glover, J.M.; Mant, T.G.; Yu, R.Z.; Tami, J.A.; Dorr, F.A. Phase I trial of ISIS 104838, a 2′-methoxyethyl modified antisense oligonucleotide targeting tumor necrosis factor-alpha. J. Pharmacol. Exp. Ther. 2002, 303, 1334–1343. [Google Scholar] [CrossRef] [PubMed]
  150. Chi, K.N.; Siu, L.L.; Hirte, H.; Hotte, S.J.; Knox, J.; Kollmansberger, C.; Gleave, M.; Guns, E.; Powers, J.; Walsh, W.; et al. A phase I study of OGX-011, a 2′-methoxyethyl phosphorothioate antisense to clusterin, in combination with docetaxel in patients with advanced cancer. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2008, 14, 833–839. [Google Scholar] [CrossRef] [PubMed]
  151. Raal, F.J.; Santos, R.D.; Blom, D.J.; Marais, A.D.; Charng, M.J.; Cromwell, W.C.; Lachmann, R.H.; Gaudet, D.; Tan, J.L.; Chasan-Taber, S.; et al. Mipomersen, an apolipoprotein B synthesis inhibitor, for lowering of LDL cholesterol concentrations in patients with homozygous familial hypercholesterolaemia: A randomised, double-blind, placebo-controlled trial. Lancet 2010, 375, 998–1006. [Google Scholar] [CrossRef]
  152. Vester, B.; Wengel, J. LNA (locked nucleic acid): High-affinity targeting of complementary RNA and DNA. Biochemistry 2004, 43, 13233–13241. [Google Scholar] [CrossRef] [PubMed]
  153. Valoczi, A.; Hornyik, C.; Varga, N.; Burgyan, J.; Kauppinen, S.; Havelda, Z. Sensitive and specific detection of microRNAs by northern blot analysis using LNA-modified oligonucleotide probes. Nucleic Acids Res. 2004, 32, e175. [Google Scholar] [CrossRef] [PubMed]
  154. Henry, S.P.; Geary, R.S.; Yu, R.; Levin, A.A. Drug properties of second-generation antisense oligonucleotides: How do they measure up to their predecessors? Curr. Opin. Investig. Drugs 2001, 2, 1444–1449. [Google Scholar]
  155. Weiler, J.; Hunziker, J.; Hall, J. Anti-miRNA oligonucleotides (AMOs): Ammunition to target miRNAs implicated in human disease? Gene Ther. 2006, 13, 496–502. [Google Scholar] [CrossRef] [PubMed]
  156. Fujiwara, T.; Katsuda, T.; Hagiwara, K.; Kosaka, N.; Yoshioka, Y.; Takahashi, R.U.; Takeshita, F.; Kubota, D.; Kondo, T.; Ichikawa, H.; et al. Clinical relevance and therapeutic significance of microRNA-133a expression profiles and functions in malignant osteosarcoma-initiating cells. Stem Cells 2014, 32, 959–973. [Google Scholar] [CrossRef]
  157. Davis, S.; Lollo, B.; Freier, S.; Esau, C. Improved targeting of miRNA with antisense oligonucleotides. Nucleic Acids Res. 2006, 34, 2294–2304. [Google Scholar] [CrossRef]
  158. Gallo Cantafio, M.E.; Nielsen, B.S.; Mignogna, C.; Arbitrio, M.; Botta, C.; Frandsen, N.M.; Rolfo, C.; Tagliaferri, P.; Tassone, P.; Di Martino, M.T. Pharmacokinetics and Pharmacodynamics of a 13-mer LNA-inhibitor-miR-221 in Mice and Non-human Primates. Mol. Ther. Nucleic Acids 2016, 5, E326. [Google Scholar] [CrossRef]
  159. Jie, J.; Liu, D.; Wang, Y.; Wu, Q.; Wu, T.; Fang, R. Generation of MiRNA sponge constructs targeting multiple MiRNAs. J. Clin. Lab. Anal. 2022, 36, e24527. [Google Scholar] [CrossRef]
  160. Ebert, M.S.; Neilson, J.R.; Sharp, P.A. MicroRNA sponges: Competitive inhibitors of small RNAs in mammalian cells. Nat. Methods 2007, 4, 721–726. [Google Scholar] [CrossRef]
  161. Xiao, J.; Yang, B.; Lin, H.; Lu, Y.; Luo, X.; Wang, Z. Novel approaches for gene-specific interference via manipulating actions of microRNAs: Examination on the pacemaker channel genes HCN2 and HCN4. J. Cell. Physiol. 2007, 212, 285–292. [Google Scholar] [CrossRef]
  162. Papapetrou, E.P.; Korkola, J.E.; Sadelain, M. A genetic strategy for single and combinatorial analysis of miRNA function in mammalian hematopoietic stem cells. Stem Cells 2010, 28, 287–296. [Google Scholar] [CrossRef]
  163. Ma, L.; Young, J.; Prabhala, H.; Pan, E.; Mestdagh, P.; Muth, D.; Teruya-Feldstein, J.; Reinhardt, F.; Onder, T.T.; Valastyan, S.; et al. miR-9, a MYC/MYCN-activated microRNA, regulates E-cadherin and cancer metastasis. Nat. Cell Biol. 2010, 12, 247–256. [Google Scholar] [CrossRef]
  164. Yang, Y.; Meng, H.; Peng, Q.; Yang, X.; Gan, R.; Zhao, L.; Chen, Z.; Lu, J.; Meng, Q.H. Downregulation of microRNA-21 expression restrains non-small cell lung cancer cell proliferation and migration through upregulation of programmed cell death 4. Cancer Gene Ther. 2015, 22, 23–29. [Google Scholar] [CrossRef] [PubMed]
  165. Bak, R.O.; Mikkelsen, J.G. miRNA sponges: Soaking up miRNAs for regulation of gene expression. Wiley Interdiscip. Rev. RNA 2014, 5, 317–333. [Google Scholar] [CrossRef] [PubMed]
  166. Krell, J.; Frampton, A.E.; Jacob, J.; Castellano, L.; Stebbing, J. miRNAs in breast cancer: Ready for real time? Pharmacogenomics 2012, 13, 709–719. [Google Scholar] [CrossRef]
  167. Murakami, K.; Miyagishi, M. Tiny masking locked nucleic acids effectively bind to mRNA and inhibit binding of microRNAs in relation to thermodynamic stability. Biomed. Rep. 2014, 2, 509–512. [Google Scholar] [CrossRef]
  168. Colangelo, T.; Polcaro, G.; Ziccardi, P.; Muccillo, L.; Galgani, M.; Pucci, B.; Milone, M.R.; Budillon, A.; Santopaolo, M.; Mazzoccoli, G.; et al. The miR-27a-calreticulin axis affects drug-induced immunogenic cell death in human colorectal cancer cells. Cell Death Dis. 2016, 7, e2108. [Google Scholar] [CrossRef] [PubMed]
  169. Saiyed, A.N.; Vasavada, A.R.; Johar, S.R.K. Recent trends in miRNA therapeutics and the application of plant miRNA for prevention and treatment of human diseases. Futur. J. Pharm. Sci. 2022, 8, 24. [Google Scholar] [CrossRef]
  170. Yang, H.; Liu, Y.; Chen, L.; Zhao, J.; Guo, M.; Zhao, X.; Wen, Z.; He, Z.; Chen, C.; Xu, L. MiRNA-Based Therapies for Lung Cancer: Opportunities and Challenges? Biomolecules 2023, 13, 877. [Google Scholar] [CrossRef]
  171. Hossain, A.; Kuo, M.T.; Saunders, G.F. Mir-17-5p regulates breast cancer cell proliferation by inhibiting translation of AIB1 mRNA. Mol. Cell. Biol. 2006, 26, 8191–8201. [Google Scholar] [CrossRef]
  172. Hutvagner, G.; Zamore, P.D. RNAi: Nature abhors a double-strand. Curr. Opin. Genet. Dev. 2002, 12, 225–232. [Google Scholar] [CrossRef] [PubMed]
  173. Takeshita, F.; Patrawala, L.; Osaki, M.; Takahashi, R.U.; Yamamoto, Y.; Kosaka, N.; Kawamata, M.; Kelnar, K.; Bader, A.G.; Brown, D.; et al. Systemic delivery of synthetic microRNA-16 inhibits the growth of metastatic prostate tumors via downregulation of multiple cell-cycle genes. Mol. Ther. J. Am. Soc. Gene Ther. 2010, 18, 181–187. [Google Scholar] [CrossRef]
  174. 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] [PubMed]
  175. Jung, H.M.; Phillips, B.L.; Patel, R.S.; Cohen, D.M.; Jakymiw, A.; Kong, W.W.; Cheng, J.Q.; Chan, E.K. Keratinization-associated miR-7 and miR-21 regulate tumor suppressor reversion-inducing cysteine-rich protein with kazal motifs (RECK) in oral cancer. J. Biol. Chem. 2012, 287, 29261–29272. [Google Scholar] [CrossRef] [PubMed]
  176. Segal, M.; Biscans, A.; Gilles, M.E.; Anastasiadou, E.; De Luca, R.; Lim, J.; Khvorova, A.; Slack, F.J. Hydrophobically Modified let-7b miRNA Enhances Biodistribution to NSCLC and Downregulates HMGA2 In Vivo. Mol. Ther. Nucleic Acids 2020, 19, 267–277. [Google Scholar] [CrossRef]
  177. Segal, M.; Slack, F.J. Challenges identifying efficacious miRNA therapeutics for cancer. Expert. Opin. Drug Discov. 2020, 15, 987–992. [Google Scholar] [CrossRef]
  178. Li, X.; Corbett, A.L.; Taatizadeh, E.; Tasnim, N.; Little, J.P.; Garnis, C.; Daugaard, M.; Guns, E.; Hoorfar, M.; Li, I.T.S. Challenges and opportunities in exosome research-Perspectives from biology, engineering, and cancer therapy. APL Bioeng. 2019, 3, 011503. [Google Scholar] [CrossRef]
  179. Stepanov, G.; Zhuravlev, E.; Shender, V.; Nushtaeva, A.; Balakhonova, E.; Mozhaeva, E.; Kasakin, M.; Koval, V.; Lomzov, A.; Pavlyukov, M.; et al. Nucleotide Modifications Decrease Innate Immune Response Induced by Synthetic Analogs of snRNAs and snoRNAs. Genes 2018, 9, 531. [Google Scholar] [CrossRef]
  180. Denzler, R.; McGeary, S.E.; Title, A.C.; Agarwal, V.; Bartel, D.P.; Stoffel, M. Impact of MicroRNA Levels, Target-Site Complementarity, and Cooperativity on Competing Endogenous RNA-Regulated Gene Expression. Mol. Cell 2016, 64, 565–579. [Google Scholar] [CrossRef]
  181. Abd-Aziz, N.; Kamaruzman, N.I.; Poh, C.L. Development of MicroRNAs as Potential Therapeutics against Cancer. J. Oncol. 2020, 2020, 8029721. [Google Scholar] [CrossRef]
  182. Bost, J.P.; Barriga, H.; Holme, M.N.; Gallud, A.; Maugeri, M.; Gupta, D.; Lehto, T.; Valadi, H.; Esbjorner, E.K.; Stevens, M.M.; et al. Delivery of Oligonucleotide Therapeutics: Chemical Modifications, Lipid Nanoparticles, and Extracellular Vesicles. ACS Nano 2021, 15, 13993–14021. [Google Scholar] [CrossRef]
  183. Inoue, J.; Fujiwara, K.; Hamamoto, H.; Kobayashi, K.; Inazawa, J. Improving the Efficacy of EGFR Inhibitors by Topical Treatment of Cutaneous Squamous Cell Carcinoma with miR-634 Ointment. Mol. Ther. Oncolytics 2020, 19, 294–307. [Google Scholar] [CrossRef] [PubMed]
  184. Romano, G.; Acunzo, M.; Nana-Sinkam, P. microRNAs as Novel Therapeutics in Cancer. Cancers 2021, 13, 1526. [Google Scholar] [CrossRef]
  185. Lennox, K.A.; Behlke, M.A. A direct comparison of anti-microRNA oligonucleotide potency. Pharm. Res. 2010, 27, 1788–1799. [Google Scholar] [CrossRef] [PubMed]
  186. Moraes, F.C.; Pichon, C.; Letourneur, D.; Chaubet, F. miRNA Delivery by Nanosystems: State of the Art and Perspectives. Pharmaceutics 2021, 13, 1901. [Google Scholar] [CrossRef]
  187. Ma, S.C.; Zhang, J.Q.; Yan, T.H.; Miao, M.X.; Cao, Y.M.; Cao, Y.B.; Zhang, L.C.; Li, L. Novel strategies to reverse chemoresistance in colorectal cancer. Cancer Med. 2023, 12, 11073–11096. [Google Scholar] [CrossRef] [PubMed]
  188. Arghiani, N.; Shah, K. Modulating microRNAs in cancer: Next-generation therapies. Cancer Biol. Med. 2021, 19, 289–304. [Google Scholar] [CrossRef]
  189. Holjencin, C.; Jakymiw, A. MicroRNAs and Their Big Therapeutic Impacts: Delivery Strategies for Cancer Intervention. Cells 2022, 11, 2332. [Google Scholar] [CrossRef]
  190. Hosseinahli, N.; Aghapour, M.; Duijf, P.H.G.; Baradaran, B. Treating cancer with microRNA replacement therapy: A literature review. J. Cell. Physiol. 2018, 233, 5574–5588. [Google Scholar] [CrossRef]
  191. Dasgupta, I.; Chatterjee, A. Recent Advances in miRNA Delivery Systems. Methods Protoc. 2021, 4, 10. [Google Scholar] [CrossRef]
  192. Jayandharan, G. Gene and Cell Therapy: Biology and Applications; Springer: Berlin/Heidelberg, Germany, 2018. [Google Scholar]
  193. Kortylewski, M.; Nechaev, S. How to train your dragon: Targeted delivery of microRNA to cancer cells in vivo. Mol. Ther. J. Am. Soc. Gene Ther. 2014, 22, 1070–1071. [Google Scholar] [CrossRef] [PubMed]
  194. Bulcha, J.T.; Wang, Y.; Ma, H.; Tai, P.W.L.; Gao, G. Viral vector platforms within the gene therapy landscape. Signal Transduct. Target. Ther. 2021, 6, 53. [Google Scholar] [CrossRef] [PubMed]
  195. Ang, L.; Guo, L.; Wang, J.; Huang, J.; Lou, X.; Zhao, M. Oncolytic virotherapy armed with an engineered interfering lncRNA exhibits antitumor activity by blocking the epithelial mesenchymal transition in triple-negative breast cancer. Cancer Lett. 2020, 479, 42–53. [Google Scholar] [CrossRef] [PubMed]
  196. Bhere, D.; Arghiani, N.; Lechtich, E.R.; Yao, Y.; Alsaab, S.; Bei, F.; Matin, M.M.; Shah, K. Simultaneous downregulation of miR-21 and upregulation of miR-7 has anti-tumor efficacy. Sci. Rep. 2020, 10, 1779. [Google Scholar] [CrossRef]
  197. Kim, M.W.; Kwon, S.H.; Choi, J.H.; Lee, A. A Promising Biocompatible Platform: Lipid-Based and Bio-Inspired Smart Drug Delivery Systems for Cancer Therapy. Int. J. Mol. Sci. 2018, 19, 3859. [Google Scholar] [CrossRef]
  198. O’Neill, C.P.; Dwyer, R.M. Nanoparticle-Based Delivery of Tumor Suppressor microRNA for Cancer Therapy. Cells 2020, 9, 521. [Google Scholar] [CrossRef]
  199. Wang, H.; Liu, S.; Jia, L.; Chu, F.; Zhou, Y.; He, Z.; Guo, M.; Chen, C.; Xu, L. Nanostructured lipid carriers for MicroRNA delivery in tumor gene therapy. Cancer Cell Int. 2018, 18, 101. [Google Scholar] [CrossRef]
  200. Beg, M.S.; Brenner, A.J.; Sachdev, J.; Borad, M.; Kang, Y.K.; Stoudemire, J.; Smith, S.; Bader, A.G.; Kim, S.; Hong, D.S. Phase I study of MRX34, a liposomal miR-34a mimic, administered twice weekly in patients with advanced solid tumors. Investig. New Drugs 2017, 35, 180–188. [Google Scholar] [CrossRef]
  201. Hong, D.S.; Kang, Y.K.; Borad, M.; Sachdev, J.; Ejadi, S.; Lim, H.Y.; Brenner, A.J.; Park, K.; Lee, J.L.; Kim, T.Y.; et al. Phase 1 study of MRX34, a liposomal miR-34a mimic, in patients with advanced solid tumours. Br. J. Cancer 2020, 122, 1630–1637. [Google Scholar] [CrossRef]
  202. de Antonellis, P.; Liguori, L.; Falanga, A.; Carotenuto, M.; Ferrucci, V.; Andolfo, I.; Marinaro, F.; Scognamiglio, I.; Virgilio, A.; De Rosa, G.; et al. MicroRNA 199b-5p delivery through stable nucleic acid lipid particles (SNALPs) in tumorigenic cell lines. Naunyn Schmiedebergs Arch. Pharmacol. 2013, 386, 287–302. [Google Scholar] [CrossRef]
  203. van den Berg, A.I.S.; Yun, C.O.; Schiffelers, R.M.; Hennink, W.E. Polymeric delivery systems for nucleic acid therapeutics: Approaching the clinic. J. Control Release 2021, 331, 121–141. [Google Scholar] [CrossRef] [PubMed]
  204. Madkhali, O.; Mekhail, G.; Wettig, S.D. Modified gelatin nanoparticles for gene delivery. Int. J. Pharm. 2019, 554, 224–234. [Google Scholar] [CrossRef] [PubMed]
  205. Garg, U.; Chauhan, S.; Nagaich, U.; Jain, N. Current Advances in Chitosan Nanoparticles Based Drug Delivery and Targeting. Adv. Pharm. Bull. 2019, 9, 195–204. [Google Scholar] [CrossRef] [PubMed]
  206. Wei, P.; Cornel, E.J.; Du, J. Ultrasound-responsive polymer-based drug delivery systems. Drug Deliv. Transl. Res. 2021, 11, 1323–1339. [Google Scholar] [CrossRef]
  207. McClorey, G.; Banerjee, S. Cell-Penetrating Peptides to Enhance Delivery of Oligonucleotide-Based Therapeutics. Biomedicines 2018, 6, 51. [Google Scholar] [CrossRef]
  208. Silva, S.; Almeida, A.J.; Vale, N. Combination of Cell-Penetrating Peptides with Nanoparticles for Therapeutic Application: A Review. Biomolecules 2019, 9, 22. [Google Scholar] [CrossRef]
  209. Hobel, S.; Aigner, A. Polyethylenimines for siRNA and miRNA delivery in vivo. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnolgy 2013, 5, 484–501. [Google Scholar] [CrossRef]
  210. Lim, C.C.; Chia, L.Y.; Kumar, P.V. Dendrimer-based nanocomposites for the production of RNA delivery systems. OpenNano 2023, 13, 100173. [Google Scholar] [CrossRef]
  211. Berger, A.G.; Chou, J.J.; Hammond, P.T. Approaches to Modulate the Chronic Wound Environment Using Localized Nucleic Acid Delivery. Adv. Wound Care New Rochelle 2021, 10, 503–528. [Google Scholar] [CrossRef]
  212. Mintzer, M.A.; Grinstaff, M.W. Biomedical applications of dendrimers: A tutorial. Chem. Soc. Rev. 2011, 40, 173–190. [Google Scholar] [CrossRef]
  213. Dzmitruk, V.; Apartsin, E.; Ihnatsyeu-Kachan, A.; Abashkin, V.; Shcharbin, D.; Bryszewska, M. Dendrimers Show Promise for siRNA and microRNA Therapeutics. Pharmaceutics 2018, 10, 126. [Google Scholar] [CrossRef] [PubMed]
  214. Zhu, B.; Zhong, W.; Cao, X.; Pan, G.; Xu, M.; Zheng, J.; Chen, H.; Feng, X.; Luo, C.; Lu, C.; et al. Loss of miR-31-5p drives hematopoietic stem cell malignant transformation and restoration eliminates leukemia stem cells in mice. Sci. Transl. Med. 2022, 14, eabh2548. [Google Scholar] [CrossRef]
  215. Pedziwiatr-Werbicka, E.; Gorzkiewicz, M.; Horodecka, K.; Abashkin, V.; Klajnert-Maculewicz, B.; Pena-Gonzalez, C.E.; Sanchez-Nieves, J.; Gomez, R.; de la Mata, F.J.; Bryszewska, M. Silver Nanoparticles Surface-Modified with Carbosilane Dendrons as Carriers of Anticancer siRNA. Int. J. Mol. Sci. 2020, 21, 4647. [Google Scholar] [CrossRef] [PubMed]
  216. Zenze, M.; Daniels, A.; Singh, M. Dendrimers as Modifiers of Inorganic Nanoparticles for Therapeutic Delivery in Cancer. Pharmaceutics 2023, 15, 398. [Google Scholar] [CrossRef] [PubMed]
  217. Gong, J.; Song, C.; Li, G.; Guo, Y.; Wang, Z.; Guo, H.; Xia, J.; Tao, Y.; Shi, Q.; Shi, X.; et al. Ultrasound-enhanced theranostics of orthotopic breast cancer through a multifunctional core-shell tecto dendrimer-based nanomedicine platform. Biomater. Sci. 2023, 11, 4385–4396. [Google Scholar] [CrossRef]
  218. Marcinkowska, M.; Sobierajska, E.; Stanczyk, M.; Janaszewska, A.; Chworos, A.; Klajnert-Maculewicz, B. Conjugate of PAMAM Dendrimer, Doxorubicin and Monoclonal Antibody-Trastuzumab: The New Approach of a Well-Known Strategy. Polymers 2018, 10, 187. [Google Scholar] [CrossRef]
  219. Maghsoudnia, N.; Eftekhari, R.B.; Sohi, A.N.; Dorkoosh, F.A. Chloroquine Assisted Delivery of microRNA Mimic Let-7b to NSCLC Cell Line by PAMAM (G5)—HA Nano-Carrier. Curr. Drug Deliv. 2021, 18, 31–43. [Google Scholar] [CrossRef] [PubMed]
  220. Mir, M.; Ahmed, N.; Rehman, A.U. Recent applications of PLGA based nanostructures in drug delivery. Colloids Surf. B Biointerfaces 2017, 159, 217–231. [Google Scholar] [CrossRef] [PubMed]
  221. Sadeghipour, N.; Kumar, S.U.; Massoud, T.F.; Paulmurugan, R. A rationally identified panel of microRNAs targets multiple oncogenic pathways to enhance chemotherapeutic effects in glioblastoma models. Sci. Rep. 2022, 12, 12017. [Google Scholar] [CrossRef]
  222. Labatut, A.E.; Mattheolabakis, G. Non-viral based miR delivery and recent developments. Eur. J. Pharm. Biopharm. 2018, 128, 82–90. [Google Scholar] [CrossRef]
  223. Lu, M.; Xing, H.; Xun, Z.; Yang, T.; Ding, P.; Cai, C.; Wang, D.; Zhao, X. Exosome-based small RNA delivery: Progress and prospects. Asian J. Pharm. Sci. 2018, 13, 1–11. [Google Scholar] [CrossRef]
  224. Shimbo, K.; Miyaki, S.; Ishitobi, H.; Kato, Y.; Kubo, T.; Shimose, S.; Ochi, M. Exosome-formed synthetic microRNA-143 is transferred to osteosarcoma cells and inhibits their migration. Biochem. Biophys. Res. Commun. 2014, 445, 381–387. [Google Scholar] [CrossRef] [PubMed]
  225. Bonneau, E.; Neveu, B.; Kostantin, E.; Tsongalis, G.J.; De Guire, V. How close are miRNAs from clinical practice? A perspective on the diagnostic and therapeutic market. EJIFCC 2019, 30, 114–127. [Google Scholar]
  226. Chakraborty, C.; Sharma, A.R.; Sharma, G.; Lee, S.S. Therapeutic advances of miRNAs: A preclinical and clinical update. J. Adv. Res. 2021, 28, 127–138. [Google Scholar] [CrossRef]
  227. Seto, A.G.; Beatty, X.; Lynch, J.M.; Hermreck, M.; Tetzlaff, M.; Duvic, M.; Jackson, A.L. Cobomarsen, an oligonucleotide inhibitor of miR-155, co-ordinately regulates multiple survival pathways to reduce cellular proliferation and survival in cutaneous T-cell lymphoma. Br. J. Haematol. 2018, 183, 428–444. [Google Scholar] [CrossRef]
  228. Anastasiadou, E.; Seto, A.G.; Beatty, X.; Hermreck, M.; Gilles, M.E.; Stroopinsky, D.; Pinter-Brown, L.C.; Pestano, L.; Marchese, C.; Avigan, D.; et al. Cobomarsen, an Oligonucleotide Inhibitor of miR-155, Slows DLBCL Tumor Cell Growth In Vitro and In Vivo. Clin. Cancer Res. 2021, 27, 1139–1149. [Google Scholar] [CrossRef] [PubMed]
  229. Witten, L.; Slack, F.J. miR-155 as a novel clinical target for hematological malignancies. Carcinogenesis 2020, 41, 2–7. [Google Scholar] [CrossRef] [PubMed]
  230. Querfeld, C.; Pacheco, T.; Foss, F.M.; Halwani, A.S.; Porcu, P.; Seto, A.G.; Ruckman, J.; Landry, M.L.; Jackson, A.L.; Pestano, L.A.; et al. Preliminary Results of a Phase 1 Trial Evaluating MRG-106, a Synthetic microRNA Antagonist (LNA antimiR) of microRNA-155, in Patients with CTCL. Blood 2016, 128, 1829. [Google Scholar] [CrossRef]
  231. Reid, G.; Kao, S.C.; Pavlakis, N.; Brahmbhatt, H.; MacDiarmid, J.; Clarke, S.; Boyer, M.; van Zandwijk, N. Clinical development of TargomiRs, a miRNA mimic-based treatment for patients with recurrent thoracic cancer. Epigenomics 2016, 8, 1079–1085. [Google Scholar] [CrossRef]
  232. van Zandwijk, N.; Pavlakis, N.; Kao, S.C.; Linton, A.; Boyer, M.J.; Clarke, S.; Huynh, Y.; Chrzanowska, A.; Fulham, M.J.; Bailey, D.L.; et al. Safety and activity of microRNA-loaded minicells in patients with recurrent malignant pleural mesothelioma: A first-in-man, phase 1, open-label, dose-escalation study. Lancet. Oncol. 2017, 18, 1386–1396. [Google Scholar] [CrossRef]
  233. Reid, G.; Pel, M.E.; Kirschner, M.B.; Cheng, Y.Y.; Mugridge, N.; Weiss, J.; Williams, M.; Wright, C.; Edelman, J.J.; Vallely, M.P.; et al. Restoring expression of miR-16: A novel approach to therapy for malignant pleural mesothelioma. Ann. Oncol. Off. J. Eur. Soc. Med. Oncol. 2013, 24, 3128–3135. [Google Scholar] [CrossRef]
  234. Calin, G.A.; Cimmino, A.; Fabbri, M.; Ferracin, M.; Wojcik, S.E.; Shimizu, M.; Taccioli, C.; Zanesi, N.; Garzon, R.; Aqeilan, R.I.; et al. MiR-15a and miR-16-1 cluster functions in human leukemia. Proc. Natl. Acad. Sci. USA. 2008, 105, 5166–5171. [Google Scholar] [CrossRef] [PubMed]
  235. Kumar, S.; Ashraf, M.U.; Kumar, A.; Bae, Y.S. Therapeutic Potential of microRNA Against Th2-associated Immune Disorders. Curr. Top. Med. Chem. 2021, 21, 753–766. [Google Scholar] [CrossRef] [PubMed]
  236. Kasina, V.; Wahane, A.; Liu, C.H.; Yang, L.; Nieh, M.P.; Slack, F.J.; Bahal, R. Next-generation poly-L-histidine formulations for miRNA mimic delivery. Mol. Ther. Methods Clin. Dev. 2023, 29, 271–283. [Google Scholar] [CrossRef]
  237. Fariha, A.; Hami, I.; Tonmoy, M.I.Q.; Akter, S.; Al Reza, H.; Bahadur, N.M.; Rahaman, M.M.; Hossain, M.S. Cell cycle associated miRNAs as target and therapeutics in lung cancer treatment. Heliyon 2022, 8, e11081. [Google Scholar] [CrossRef]
  238. Bouchie, A. First microRNA mimic enters clinic. Nat. Biotechnol. 2013, 31, 577. [Google Scholar] [CrossRef]
  239. Lithwick-Yanai, G.; Dromi, N.; Shtabsky, A.; Morgenstern, S.; Strenov, Y.; Feinmesser, M.; Kravtsov, V.; Leon, M.E.; Hajduch, M.; Ali, S.Z.; et al. Multicentre validation of a microRNA-based assay for diagnosing indeterminate thyroid nodules utilising fine needle aspirate smears. J. Clin. Pathol. 2017, 70, 500–507. [Google Scholar] [CrossRef]
  240. Chakraborty, C.; Sharma, A.R.; Sharma, G.; Doss, C.G.P.; Lee, S.S. Therapeutic miRNA and siRNA: Moving from Bench to Clinic as Next Generation Medicine. Mol. Ther. Nucleic Acids 2017, 8, 132–143. [Google Scholar] [CrossRef]
  241. Kohler, B.; Dubovik, S.; Horterer, E.; Wilk, U.; Stockl, J.B.; Tekarslan-Sahin, H.; Ljepoja, B.; Paulitschke, P.; Frohlich, T.; Wagner, E.; et al. Combating Drug Resistance by Exploiting miRNA-200c-Controlled Phase II Detoxification. Cancers 2022, 14, 5554. [Google Scholar] [CrossRef] [PubMed]
  242. Calin, G.A. MicroRNAs and cancer: What we know and what we still have to learn. Genome Med. 2009, 1, 78. [Google Scholar] [CrossRef]
  243. Available online: https://www.prnewswire.com/news-releases/sirnaomics-sirna-therapeutic-candidate-stp705-granted-orphan-drug-designation-by-us-fda-for-treatment-of-hepatocellular-carcinoma-301002347.html (accessed on 24 June 2022).
  244. Kobeissi, I.; Eljilany, I.; Achkar, T.; LaFramboise, W.A.; Santana-Santos, L.; Tarhini, A.A. A Tumor and Immune-Related Micro-RNA Signature Predicts Relapse-Free Survival of Melanoma Patients Treated with Ipilimumab. Int. J. Mol. Sci. 2023, 24, 8167. [Google Scholar] [CrossRef] [PubMed]
  245. Liang, X.; Li, D.; Leng, S.; Zhu, X. RNA-based pharmacotherapy for tumors: From bench to clinic and back. Biomed. Pharmacother. 2020, 125, 109997. [Google Scholar] [CrossRef]
  246. Si, W.; Shen, J.; Zheng, H.; Fan, W. The role and mechanisms of action of microRNAs in cancer drug resistance. Clin. Epigenetics 2019, 11, 25. [Google Scholar] [CrossRef] [PubMed]
  247. Giovannetti, E.; Erozenci, A.; Smit, J.; Danesi, R.; Peters, G.J. Molecular mechanisms underlying the role of microRNAs (miRNAs) in anticancer drug resistance and implications for clinical practice. Crit. Rev. Oncol./Hematol. 2012, 81, 103–122. [Google Scholar] [CrossRef] [PubMed]
  248. To, K.K. MicroRNA: A prognostic biomarker and a possible druggable target for circumventing multidrug resistance in cancer chemotherapy. J. Biomed. Sci. 2013, 20, 99. [Google Scholar] [CrossRef]
  249. Meng, F.; Henson, R.; Lang, M.; Wehbe, H.; Maheshwari, S.; Mendell, J.T.; Jiang, J.; Schmittgen, T.D.; Patel, T. Involvement of human micro-RNA in growth and response to chemotherapy in human cholangiocarcinoma cell lines. Gastroenterology 2006, 130, 2113–2129. [Google Scholar] [CrossRef]
  250. Munoz, J.L.; Bliss, S.A.; Greco, S.J.; Ramkissoon, S.H.; Ligon, K.L.; Rameshwar, P. Delivery of Functional Anti-miR-9 by Mesenchymal Stem Cell-derived Exosomes to Glioblastoma Multiforme Cells Conferred Chemosensitivity. Mol. Ther. Nucleic Acids 2013, 2, e126. [Google Scholar] [CrossRef]
  251. Li, Y.; Li, L.; Guan, Y.; Liu, X.; Meng, Q.; Guo, Q. MiR-92b regulates the cell growth, cisplatin chemosensitivity of A549 non small cell lung cancer cell line and target PTEN. Biochem. Biophys. Res. Commun. 2013, 440, 604–610. [Google Scholar] [CrossRef]
  252. Pedroza-Torres, A.; Romero-Cordoba, S.L.; Justo-Garrido, M.; Salido-Guadarrama, I.; Rodriguez-Bautista, R.; Montano, S.; Muniz-Mendoza, R.; Arriaga-Canon, C.; Fragoso-Ontiveros, V.; Alvarez-Gomez, R.M.; et al. MicroRNAs in Tumor Cell Metabolism: Roles and Therapeutic Opportunities. Front. Oncol. 2019, 9, 1404. [Google Scholar] [CrossRef]
  253. Subramaniam, S.; Jeet, V.; Clements, J.A.; Gunter, J.H.; Batra, J. Emergence of MicroRNAs as Key Players in Cancer Cell Metabolism. Clin. Chem. 2019, 65, 1090–1101. [Google Scholar] [CrossRef]
  254. Jiang, S.; Zhang, L.F.; Zhang, H.W.; Hu, S.; Lu, M.H.; Liang, S.; Li, B.; Li, Y.; Li, D.; Wang, E.D.; et al. A novel miR-155/miR-143 cascade controls glycolysis by regulating hexokinase 2 in breast cancer cells. EMBO J. 2012, 31, 1985–1998. [Google Scholar] [CrossRef] [PubMed]
  255. Sikand, K.; Singh, J.; Ebron, J.S.; Shukla, G.C. Housekeeping gene selection advisory: Glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and beta-actin are targets of miR-644a. PLoS ONE 2012, 7, e47510. [Google Scholar] [CrossRef] [PubMed]
  256. Khanna, M.; Saini, S.; Shariff, M.; Ronsard, L.; Singh, J.K.; Kumar, H. Data highlighting miR-155 and GAPDH correlation. Data Brief. 2019, 24, 103945. [Google Scholar] [CrossRef]
  257. Qu, W.; Ding, S.M.; Cao, G.; Wang, S.J.; Zheng, X.H.; Li, G.H. miR-132 mediates a metabolic shift in prostate cancer cells by targeting Glut1. FEBS Open Bio 2016, 6, 735–741. [Google Scholar] [CrossRef] [PubMed]
  258. Liu, M.; Gao, J.; Huang, Q.; Jin, Y.; Wei, Z. Downregulating microRNA-144 mediates a metabolic shift in lung cancer cells by regulating GLUT1 expression. Oncol. Lett. 2016, 11, 3772–3776. [Google Scholar] [CrossRef]
  259. Kinoshita, T.; Nohata, N.; Yoshino, H.; Hanazawa, T.; Kikkawa, N.; Fujimura, L.; Chiyomaru, T.; Kawakami, K.; Enokida, H.; Nakagawa, M.; et al. Tumor suppressive microRNA-375 regulates lactate dehydrogenase B in maxillary sinus squamous cell carcinoma. Int. J. Oncol. 2012, 40, 185–193. [Google Scholar] [CrossRef]
  260. Xiao, X.; Huang, X.; Ye, F.; Chen, B.; Song, C.; Wen, J.; Zhang, Z.; Zheng, G.; Tang, H.; Xie, X. The miR-34a-LDHA axis regulates glucose metabolism and tumor growth in breast cancer. Sci. Rep. 2016, 6, 21735. [Google Scholar] [CrossRef]
  261. Chen, H.; Gao, S.; Cheng, C. MiR-323a-3p suppressed the glycolysis of osteosarcoma via targeting LDHA. Hum. Cell 2018, 31, 300–309. [Google Scholar] [CrossRef]
  262. He, Y.; Chen, X.; Yu, Y.; Li, J.; Hu, Q.; Xue, C.; Chen, J.; Shen, S.; Luo, Y.; Ren, F.; et al. LDHA is a direct target of miR-30d-5p and contributes to aggressive progression of gallbladder carcinoma. Mol. Carcinog. 2018, 57, 772–783. [Google Scholar] [CrossRef]
  263. Pullen, T.J.; da Silva Xavier, G.; Kelsey, G.; Rutter, G.A. miR-29a and miR-29b contribute to pancreatic beta-cell-specific silencing of monocarboxylate transporter 1 (Mct1). Mol. Cell. Biol. 2011, 31, 3182–3194. [Google Scholar] [CrossRef]
  264. Romero-Cordoba, S.L.; Rodriguez-Cuevas, S.; Bautista-Pina, V.; Maffuz-Aziz, A.; D’Ippolito, E.; Cosentino, G.; Baroni, S.; Iorio, M.V.; Hidalgo-Miranda, A. Loss of function of miR-342-3p results in MCT1 over-expression and contributes to oncogenic metabolic reprogramming in triple negative breast cancer. Sci. Rep. 2018, 8, 12252. [Google Scholar] [CrossRef]
  265. Dong, J.; Xiao, D.; Zhao, Z.; Ren, P.; Li, C.; Hu, Y.; Shi, J.; Su, H.; Wang, L.; Liu, H.; et al. Epigenetic silencing of microRNA-137 enhances ASCT2 expression and tumor glutamine metabolism. Oncogenesis 2017, 6, e356. [Google Scholar] [CrossRef] [PubMed]
  266. Anderton, B.; Camarda, R.; Balakrishnan, S.; Balakrishnan, A.; Kohnz, R.A.; Lim, L.; Evason, K.J.; Momcilovic, O.; Kruttwig, K.; Huang, Q.; et al. MYC-driven inhibition of the glutamate-cysteine ligase promotes glutathione depletion in liver cancer. EMBO Rep. 2017, 18, 569–585. [Google Scholar] [CrossRef] [PubMed]
  267. Chen, D.; Wang, H.; Chen, J.; Li, Z.; Li, S.; Hu, Z.; Huang, S.; Zhao, Y.; He, X. MicroRNA-129-5p Regulates Glycolysis and Cell Proliferation by Targeting the Glucose Transporter SLC2A3 in Gastric Cancer Cells. Front. Pharmacol. 2018, 9, 502. [Google Scholar] [CrossRef] [PubMed]
  268. Guo, W.; Qiu, Z.; Wang, Z.; Wang, Q.; Tan, N.; Chen, T.; Chen, Z.; Huang, S.; Gu, J.; Li, J.; et al. MiR-199a-5p is negatively associated with malignancies and regulates glycolysis and lactate production by targeting hexokinase 2 in liver cancer. Hepatology 2015, 62, 1132–1144. [Google Scholar] [CrossRef]
  269. Yang, Y.; Ishak Gabra, M.B.; Hanse, E.A.; Lowman, X.H.; Tran, T.Q.; Li, H.; Milman, N.; Liu, J.; Reid, M.A.; Locasale, J.W.; et al. MiR-135 suppresses glycolysis and promotes pancreatic cancer cell adaptation to metabolic stress by targeting phosphofructokinase-1. Nat. Commun. 2019, 10, 809. [Google Scholar] [CrossRef] [PubMed]
  270. Zhang, X.; Li, Z.; Xuan, Z.; Xu, P.; Wang, W.; Chen, Z.; Wang, S.; Sun, G.; Xu, J.; Xu, Z. Novel role of miR-133a-3p in repressing gastric cancer growth and metastasis via blocking autophagy-mediated glutaminolysis. J. Exp. Clin. Cancer Res. CR 2018, 37, 320. [Google Scholar] [CrossRef]
  271. 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 Metab. 2006, 3, 87–98. [Google Scholar] [CrossRef]
  272. Schiliro, C.; Firestein, B.L. Mechanisms of Metabolic Reprogramming in Cancer Cells Supporting Enhanced Growth and Proliferation. Cells 2021, 10, 1056. [Google Scholar] [CrossRef]
  273. Moldogazieva, N.T.; Mokhosoev, I.M.; Terentiev, A.A. Metabolic Heterogeneity of Cancer Cells: An Interplay between HIF-1, GLUTs, and AMPK. Cancers 2020, 12, 862. [Google Scholar] [CrossRef]
  274. Li, W.; Hao, J.; Zhang, L.; Cheng, Z.; Deng, X.; Shu, G. Astragalin Reduces Hexokinase 2 through Increasing miR-125b to Inhibit the Proliferation of Hepatocellular Carcinoma Cells in Vitro and in Vivo. J. Agric. Food Chem. 2017, 65, 5961–5972. [Google Scholar] [CrossRef] [PubMed]
  275. Nosengo, N. Can you teach old drugs new tricks? Nature 2016, 534, 314–316. [Google Scholar] [CrossRef]
  276. Pantziarka, P.; Bouche, G.; Meheus, L.; Sukhatme, V.; Sukhatme, V.P.; Vikas, P. The Repurposing Drugs in Oncology (ReDO) Project. Ecancermedicalscience 2014, 8, 442. [Google Scholar] [CrossRef] [PubMed]
  277. Ye, Q.; Raese, R.A.; Luo, D.; Feng, J.; Xin, W.; Dong, C.; Qian, Y.; Guo, N.L. MicroRNA-Based Discovery of Biomarkers, Therapeutic Targets, and Repositioning Drugs for Breast Cancer. Cells 2023, 12, 1917. [Google Scholar] [CrossRef] [PubMed]
  278. Evans, J.M.; Donnelly, L.A.; Emslie-Smith, A.M.; Alessi, D.R.; Morris, A.D. Metformin and reduced risk of cancer in diabetic patients. Bmj 2005, 330, 1304–1305. [Google Scholar] [CrossRef]
  279. Lee, J.H.; Kim, T.I.; Jeon, S.M.; Hong, S.P.; Cheon, J.H.; Kim, W.H. The effects of metformin on the survival of colorectal cancer patients with diabetes mellitus. Int. J. Cancer 2012, 131, 752–759. [Google Scholar] [CrossRef]
  280. Noto, H.; Goto, A.; Tsujimoto, T.; Noda, M. Cancer risk in diabetic patients treated with metformin: A systematic review and meta-analysis. PLoS ONE 2012, 7, e33411. [Google Scholar] [CrossRef]
  281. Gandini, S.; Puntoni, M.; Heckman-Stoddard, B.M.; Dunn, B.K.; Ford, L.; DeCensi, A.; Szabo, E. Metformin and cancer risk and mortality: A systematic review and meta-analysis taking into account biases and confounders. Cancer Prev. Res. 2014, 7, 867–885. [Google Scholar] [CrossRef]
  282. Shaw, R.J.; Kosmatka, M.; Bardeesy, N.; Hurley, R.L.; Witters, L.A.; DePinho, R.A.; Cantley, L.C. The tumor suppressor LKB1 kinase directly activates AMP-activated kinase and regulates apoptosis in response to energy stress. Proc. Natl. Acad. Sci. USA 2004, 101, 3329–3335. [Google Scholar] [CrossRef]
  283. Oliveras-Ferraros, C.; Cufi, S.; Vazquez-Martin, A.; Torres-Garcia, V.Z.; Del Barco, S.; Martin-Castillo, B.; Menendez, J.A. Micro(mi)RNA expression profile of breast cancer epithelial cells treated with the anti-diabetic drug metformin: Induction of the tumor suppressor miRNA let-7a and suppression of the TGFbeta-induced oncomiR miRNA-181a. Cell Cycle 2011, 10, 1144–1151. [Google Scholar] [CrossRef]
  284. Sharma, P.; Singh, S. Combinatorial Effect of DCA and Let-7a on Triple-Negative MDA-MB-231 Cells: A Metabolic Approach of Treatment. Integr. Cancer Ther. 2020, 19, 1534735420911437. [Google Scholar] [CrossRef]
  285. Istvan, E.S.; Deisenhofer, J. Structural mechanism for statin inhibition of HMG-CoA reductase. Science 2001, 292, 1160–1164. [Google Scholar] [CrossRef]
  286. Wiley, J.J.; Baxter, M.P. Tibial spine fractures in children. Clin. Orthop. Relat. Res. 1990, 255, 54–60. [Google Scholar] [CrossRef]
  287. Lin, J.J.; Ezer, N.; Sigel, K.; Mhango, G.; Wisnivesky, J.P. The effect of statins on survival in patients with stage IV lung cancer. Lung Cancer 2016, 99, 137–142. [Google Scholar] [CrossRef]
  288. Hung, M.S.; Chen, I.C.; Lee, C.P.; Huang, R.J.; Chen, P.C.; Tsai, Y.H.; Yang, Y.H. Statin improves survival in patients with EGFR-TKI lung cancer: A nationwide population-based study. PLoS ONE 2017, 12, e0171137. [Google Scholar] [CrossRef]
  289. Tilija Pun, N.; Jeong, C.H. Statin as a Potential Chemotherapeutic Agent: Current Updates as a Monotherapy, Combination Therapy, and Treatment for Anti-Cancer Drug Resistance. Pharmaceuticals 2021, 14, 470. [Google Scholar] [CrossRef] [PubMed]
  290. Beckwitt, C.H.; Brufsky, A.; Oltvai, Z.N.; Wells, A. Statin drugs to reduce breast cancer recurrence and mortality. Breast Cancer Res. BCR 2018, 20, 144. [Google Scholar] [CrossRef] [PubMed]
  291. Karlic, H.; Thaler, R.; Gerner, C.; Grunt, T.; Proestling, K.; Haider, F.; Varga, F. Inhibition of the mevalonate pathway affects epigenetic regulation in cancer cells. Cancer Genet. 2015, 208, 241–252. [Google Scholar] [CrossRef]
  292. Hundal, R.S.; Petersen, K.F.; Mayerson, A.B.; Randhawa, P.S.; Inzucchi, S.; Shoelson, S.E.; Shulman, G.I. Mechanism by which high-dose aspirin improves glucose metabolism in type 2 diabetes. J. Clin. Investig. 2002, 109, 1321–1326. [Google Scholar] [CrossRef]
  293. Paez Espinosa, E.V.; Murad, J.P.; Khasawneh, F.T. Aspirin: Pharmacology and clinical applications. Thrombosis 2012, 2012, 173124. [Google Scholar] [CrossRef]
  294. Barnard, M.E.; Poole, E.M.; Curhan, G.C.; Eliassen, A.H.; Rosner, B.A.; Terry, K.L.; Tworoger, S.S. Association of Analgesic Use With Risk of Ovarian Cancer in the Nurses’ Health Studies. JAMA Oncol. 2018, 4, 1675–1682. [Google Scholar] [CrossRef]
  295. Simon, T.G.; Ma, Y.; Ludvigsson, J.F.; Chong, D.Q.; Giovannucci, E.L.; Fuchs, C.S.; Meyerhardt, J.A.; Corey, K.E.; Chung, R.T.; Zhang, X.; et al. Association Between Aspirin Use and Risk of Hepatocellular Carcinoma. JAMA Oncol. 2018, 4, 1683–1690. [Google Scholar] [CrossRef]
  296. Jiang, M.J.; Chen, Y.Y.; Dai, J.J.; Gu, D.N.; Mei, Z.; Liu, F.R.; Huang, Q.; Tian, L. Dying tumor cell-derived exosomal miR-194-5p potentiates survival and repopulation of tumor repopulating cells upon radiotherapy in pancreatic cancer. Mol. Cancer 2020, 19, 68. [Google Scholar] [CrossRef]
  297. Gan, H.; Lin, L.; Hu, N.; Yang, Y.; Gao, Y.; Pei, Y.; Chen, K.; Sun, B. Aspirin ameliorates lung cancer by targeting the miR-98/WNT1 axis. Thorac. Cancer 2019, 10, 744–750. [Google Scholar] [CrossRef]
  298. Watari, J.; Ito, C.; Shimoda, T.; Tomita, T.; Oshima, T.; Fukui, H.; Das, K.M.; Miwa, H. DNA methylation silencing of microRNA gene methylator in the precancerous background mucosa with and without gastric cancer: Analysis of the effects of H. pylori eradication and long-term aspirin use. Sci. Rep. 2019, 9, 12559. [Google Scholar] [CrossRef]
  299. Yiannakopoulou, E. Targeting epigenetic mechanisms and microRNAs by aspirin and other non steroidal anti-inflammatory agents--implications for cancer treatment and chemoprevention. Cell. Oncol. 2014, 37, 167–178. [Google Scholar] [CrossRef] [PubMed]
  300. Bazavar, M.; Fazli, J.; Valizadeh, A.; Ma, B.; Mohammadi, E.; Asemi, Z.; Alemi, F.; Maleki, M.; Xing, S.; Yousefi, B. miR-192 enhances sensitivity of methotrexate drug to MG-63 osteosarcoma cancer cells. Pathol. Res. Pract. 2020, 216, 153176. [Google Scholar] [CrossRef] [PubMed]
  301. Long, J.; Danesh, F.R. Promises and challenges of miRNA therapeutics. Am. J. Physiol. Ren. Physiol. 2022, 323, F673–F674. [Google Scholar] [CrossRef] [PubMed]
  302. 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]
  303. Bumcrot, D.; Manoharan, M.; Koteliansky, V.; Sah, D.W. RNAi therapeutics: A potential new class of pharmaceutical drugs. Nat. Chem. Biol. 2006, 2, 711–719. [Google Scholar] [CrossRef]
  304. Chakraborty, C.; Wen, Z.H.; Agoramoorthy, G.; Lin, C.S. Therapeutic microRNA Delivery Strategies with Special Emphasis on Cancer Therapy and Tumorigenesis: Current Trends and Future Challenges. Curr. Drug Metab. 2016, 17, 469–477. [Google Scholar] [CrossRef] [PubMed]
  305. Saraniti, C.; Speciale, R.; Santangelo, M.; Massaro, N.; Maniaci, A.; Gallina, S.; Serra, A.; Cocuzza, S. Functional outcomes after supracricoid modified partial laryngectomy. J. Biol. Regul. Homeost. Agents 2019, 33, 1903–1907. [Google Scholar] [CrossRef]
  306. Orellana, E.A.; Abdelaal, A.M.; Rangasamy, L.; Tenneti, S.; Myoung, S.; Low, P.S.; Kasinski, A.L. Enhancing MicroRNA Activity through Increased Endosomal Release Mediated by Nigericin. Mol. Ther. Nucleic Acids 2019, 16, 505–518. [Google Scholar] [CrossRef] [PubMed]
  307. Milone, M.C.; O’Doherty, U. Clinical use of lentiviral vectors. Leukemia 2018, 32, 1529–1541. [Google Scholar] [CrossRef] [PubMed]
  308. Wang, J.; Tan, M.; Wang, Y.; Liu, X.; Lin, A. Advances in modification and delivery of nucleic acid drugs. Zhejiang Da Xue Xue Bao Yi Xue Ban. 2023, 52, 417–428. [Google Scholar] [CrossRef]
  309. Gareev, I.; de Jesus Encarnacion Ramirez, M.; Goncharov, E.; Ivliev, D.; Shumadalova, A.; Ilyasova, T.; Wang, C. MiRNAs and lncRNAs in the regulation of innate immune signaling. Noncoding RNA Res. 2023, 8, 534–541. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of miRNA therapeutic strategies: 1. anti-miRNA oligonucleotides (AMOs) bind with miRNA, preventing the binding of miRNA to mRNA; 2. miRNA sponges include multiple sites for miRNA binding, blocking the binding of miRNA to mRNA; 3. miRNA mask binds with the specific gene (target of miRNA).
Figure 1. Schematic representation of miRNA therapeutic strategies: 1. anti-miRNA oligonucleotides (AMOs) bind with miRNA, preventing the binding of miRNA to mRNA; 2. miRNA sponges include multiple sites for miRNA binding, blocking the binding of miRNA to mRNA; 3. miRNA mask binds with the specific gene (target of miRNA).
Jpm 13 01586 g001
Table 1. Methods available for the inhibition of miRNAs.
Table 1. Methods available for the inhibition of miRNAs.
Method of DeliveryCharacteristics
anti-miRNA ASOs (AMOs)complementary sequence to endogenous miRNA
Modified AMOs with 2′-O-methyl-group (OMe)Increase binding affinity and nuclease resistance
Modified AMOs with 2′ methoxyethyl (MOE)More stable and specific
Modified AMOs with locked nucleic acid (LNA)High binding affinity and low toxicity
miRNA spongesInhibit a whole family of associated miRNAs
miRNA maskingGene-specific
Table 2. Examples of Applied miRNA-based therapies in cancer.
Table 2. Examples of Applied miRNA-based therapies in cancer.
Biotech CompanyExperimental ProductTarget miRNAPathologic ConditionClinical Phase
MiRagen TherapeuticsMRG-106miR-155Lymphoma and LeukemiaI, II
ENGeneICMesomirmiR-16MesotheliomaIII
SynlogicPLGA-poly-L-His NPsmiR-34aAdvanced solid tumorsPre-Clinical
Asbestos Diseases ResearchFoundationTargomiRsmiR-16Malignant Pleural Mesothelioma, NS-CLCI
Alnylam PharmaceuticalsScreening
Interna TechnologiesScreening
Mello BiotechScreening
OpkoScreening
Table 3. Key features of the major classes of pharmaceutical drugs, small molecules, proteins, and antibodies, in comparison with miRNAs therapeutics.
Table 3. Key features of the major classes of pharmaceutical drugs, small molecules, proteins, and antibodies, in comparison with miRNAs therapeutics.
Small MoleculesProteins and AntibodiesmiRNA Therapeutics
Nature of actionActivation or Inhibition of targetActivation or Inhibition of targetsInhibition of targets
Site of target proteinsExtracellular and intracellular targetsExtracellular targetsAll targets including non-druggable targets
Selectivity and potencyVariable, depending on binding site and ligand specificity their affinity efficacyHighly selective and potentHighly selective and potent
Lead optimizationLead ID and optimization slowLead ID and optimization slowRapid lead ID and optimization
ManufactureEasy to synthesizeDifficult to produceEasy to synthesize
StabilityStableUnstableStable
DeliveryEasyDifficultDifficult
Safety/toxicityRisk of off-target effectsRisk of immunogenicityRisk of immunogenicity
Table 4. Comparative description of viral- and non-viral-based delivery of miRNAs in cancer.
Table 4. Comparative description of viral- and non-viral-based delivery of miRNAs in cancer.
Viral-Based Delivery VectorsAdvantagesDisadvantages
Retroviral vectorsStability of transgene expressionPropensity for carcinogenesis due to insertional mutagenesis
Unable to transduce not dividing cells.
Lentiviral vectors (LVs)Stability of transgene expression
Can transduce both dividing and nondividing cells
Lower risk of insertional mutagenesis and oncogenesis
Adenoviral (AdVs) and Adeno-associated vectors (AAVs)Low immunogenicity
High transduction efficiency in a variety of cells
Packaging capacity low
Expensive manufacturing
Lipid-based deliveryNon-Immunogenic Biocompatible
Easy production
Cytotoxicity
Polymeric deliverHigh packaging capacityCytotoxicity
Nonspecific delivery
InorganicNon-Immunogenic
Biocompatible
Easy production
Nontoxic
Control of physical features
Low efficacy
Exosome based deliveryBiocompatible
Non immunogenic
Tissue organ-specific delivery
Lack of strong experimental evidence/data
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Pagoni, M.; Cava, C.; Sideris, D.C.; Avgeris, M.; Zoumpourlis, V.; Michalopoulos, I.; Drakoulis, N. miRNA-Based Technologies in Cancer Therapy. J. Pers. Med. 2023, 13, 1586. https://doi.org/10.3390/jpm13111586

AMA Style

Pagoni M, Cava C, Sideris DC, Avgeris M, Zoumpourlis V, Michalopoulos I, Drakoulis N. miRNA-Based Technologies in Cancer Therapy. Journal of Personalized Medicine. 2023; 13(11):1586. https://doi.org/10.3390/jpm13111586

Chicago/Turabian Style

Pagoni, Maria, Claudia Cava, Diamantis C. Sideris, Margaritis Avgeris, Vassilios Zoumpourlis, Ioannis Michalopoulos, and Nikolaos Drakoulis. 2023. "miRNA-Based Technologies in Cancer Therapy" Journal of Personalized Medicine 13, no. 11: 1586. https://doi.org/10.3390/jpm13111586

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