Modulating microRNAs in cancer: next-generation therapies ========================================================= * Nahid Arghiani * Khalid Shah ## Abstract MicroRNAs (miRNAs) are a class of endogenously expressed non-coding regulators of the genome with an ability to mediate a variety of biological and pathological processes. There is growing evidence demonstrating frequent dysregulation of microRNAs in cancer cells, which is associated with tumor initiation, development, migration, invasion, resisting cell death, and drug resistance. Studies have shown that modulation of these small RNAs is a novel and promising therapeutic tool in the treatment of a variety of diseases, especially cancer, due to their broad influence on multiple cellular processes. However, suboptimal delivery of the appropriate miRNA to the cancer sites, quick degradation by nucleases in the blood circulation, and off target effects have limited their research and clinical applications. Therefore, there is a pressing need to improve the therapeutic efficacy of miRNA modulators, while at the same time reducing their toxicities. Several delivery vehicles for miRNA modulators have been shown to be effective *in vitro* and *in vivo*. In this review, we will discuss the role and importance of miRNAs in cancer and provide perspectives on currently available carriers for miRNA modulation. We will also summarize the challenges and prospects for the clinical translation of miRNA-based therapeutic strategies. * MiRNAs * dysregulation * delivery systems * cancer therapy * clinical translation ## Introduction Cancer is a major global health concern that imposes a significant financial burden. Despite advances in cancer treatment, new therapeutic approaches are needed. Targeting molecules that have key regulatory functions on up to 200 mRNAs and multiple pathways may eventually provide an efficient cancer cure. Increasing evidence shows that miRNAs, which are small untranslated molecules, can regulate multiple genes important in cancer formation and development. These small RNAs have emerged as highly conserved non-coding endogenous RNAs in 142 species, and have been implicated as key regulators of different biological processes, mostly through gene-silencing pathways1,2. MiRNAs can target gene expression *via* direct interaction with DNA, translational repression, or mRNA degradation. Thus, the biogenesis and expression of miRNAs are carefully controlled by multiple molecular mechanisms; a certain set of miRNAs are expressed in specific cell types during particular stages in development3. Furthermore, aborted expression of miRNAs is involved in initiation, development, and progression of different diseases including cancer4. The therapeutic potential of miRNA modulation is supported by its use in treating different cancer types in preclinical models. Recently, research groups have focused on optimizing strategies for miRNA delivery to tumor sites, as well as reducing their toxicities to normal cells in order to accelerate clinical translation. Here, we first review the role of miRNAs in cancer and shed a light on the current status and recent progress of therapeutic miRNA delivery systems. We also discuss the challenges and advances in miRNA-based therapeutics, leading to a number of clinical trials. ## Dysregulation of microRNAs in cancer Oncogenic miRNAs can be classified into 2 groups based on expression levels. Generally, miRNAs that are overexpressed in cancer cells that promote tumorigenesis by inhibiting tumor suppressors are known as “oncomiRs,” whereas down-regulated miRNAs that prevent tumor development by inhibiting the expressions of oncogenes are known as “tumor suppressor miRs.” Dysregulation can occur in a single oncogenic microRNA or global miRNA expression through various mechanisms. Upregulation of oncomiRs has been shown to result in a gain-of-function mutation, hypomethylation, gene amplification, or aberrant expression of oncomiRs5. Furthermore, evidence from multiple sources has shown that downregulation of tumor suppressor miRs is the consequence of their deletion or inactivation through mutation or heterochromatin formation. In such cases, epigenetic alterations such as aberrant DNA methylation and histone modifications play important roles5. Additionally, aborted expression of miRNAs can be associated with aberrant activity of transcription factors such as MYC, TP53, STAT3, SP1, and RREB16,7. For example, elevated expression of c-MYC induces increased levels of oncogenic miR-155 and miR-17–92 clusters *via* interaction with the E-box in the promoter region of these miRNAs. In addition, c-MYC induces cancer initiation through repression of tumor-suppressive microRNAs such as let-7, miR-23b and miR-29a8,9. Other studies have demonstrated that miRNAs such as miR-34 and miR-605 are regulated by P53, leading to cell cycle arrest and induction of apoptosis10,11. Disturbing the interactions of STAT3 and NF-kB with binding sites at the hsa-miR-21 promoter results in downregulation of this small RNA12. This study indicated that upregulation of miR-21 was regulated by STAT3 and NF-kB. Transcription factors like the zinc finger protein, RREB1, are activated by the RAS oncogene and bind to RAS-responsive element (RRE), which is located in the miR-143/145 cluster promoter. RREB1 activation leads to repression of miR-143/145 cluster expression in pancreatic cancers. In addition, expressions of KRAS and RREB1 are adversely controlled by miR-143/14513. The transcriptional repressor ZEB1 also inhibits the expression of miR-141 and miR-200c *via* binding to the ZEB-type E-box element located within target gene promoters. ZEB1 causes the epithelial-to-mesenchymal transition (EMT) in cancer cells14. Increasing evidence also suggests that global changes in miRNA expression reflect defective biogenesis systems. Copy number abnormalities or deregulated expression of components involved in the miRNA pathway such as Drosha, DGCR8, Dicer, and TRBP have been related to promotion of cancers15. Utilizing matched primary and metastatic tumor pairs and robust metastasis models, miRNAs have been shown to be the main regulatory components of molecular networks during metastasis. MetastamiRs refer to those regulatory miRNAs, which promote or inhibit invasion and metastasis of cancer cells through regulation of key steps in the metastatic program such as cancer stem cell properties, EMT, adhesion and proteolysis, and migration16. Stemness is responsible for initiating metastasis and cancer recurrence, and miRNAs are involved in the maintenance of the cancer stem cell *via* targeting main signaling pathways such as WNT/β-catenin, NOTCH, NF-kB, PI3K/AKT/mTOR, BMI-1, and STAT3. These pathways are targeted by many families such as miR-7, miR-10b, miR-21, miR-31, miR-140-5p, and miR-34a17. Additionally, studies have shown miR-200 and miR-205 increased the EMT by increasing the expression of ZEB1. In contrast, the forced expression of miR-1199-5p significantly decreases ZEB1 and hinders EMT18. Dysregulation of miR-101 and miR-9 has been shown to induce metastasis by activating Zeste Homolog 2 (EZH2) and the β-catenin pathway, as well as by targeting the EMT *via* targeting E-cadherin19. A growing body of evidence indicates that several miRNAs participate in cancer cell resistance by directly or indirectly regulating different signaling pathways20. MicroRNAs seem to reduce the effectiveness of treatments by regulating the cell cycle, DNA damage repair systems, intake of anticancer drugs and drug efflux, apoptosis resistance, drug inactivation, and suppression of intracellular prodrug activation. For example, miR-34a, miR-124, miR-106a, miR-188, miR-195 miR-449a, and miR-320c can affect cell cycle progression and the sensitivity of cancer cells by targeting CDK, while miR-192, miR-221, and miR-223 are involved in radiation and chemotherapy resistance *via* regulating the cell cycle by targeting p2721. Furthermore, the upregulation of miR-21, miR-27a, and miR-451 contributes to overexpression of P-glycoprotein, a major efflux pump in humans, whereas miR-138 and miR-298 increase the sensitivity to anticancer drug resistance of the MDR cell line in leukemia and breast cancer22. A large number of miRNAs, such as miR-9, miR-19a, miR-19b-3p, miR-21, miR-125b, miR-192, miR-221, and miR-222 regulate core apoptotic signaling pathways and play key roles in radiotherapy and drug resistance *via* targeting of PTEN, BCL-2, PARP, and PDCD4 in different cancer cells23,24. It is well-known that miR-21, miR-122–5p, miR-132–5p, miR-320, and miR-892a alter expressions of enzymes involved in intracellular activation of some chemotherapeutic agents and/or Cytochrome P450 (P450)-related enzymes, which are important for the clearance of toxic drugs25. Additionally, down-regulating miR-135a, miR-23a, and miR-203 can increase IL-17 and IL-8 and activate STAT3, which has a negative impact on the outcomes of radiotherapy20,24. Overall, changes in miRNA levels have potential value in clinical applications as effective diagnostic and prognostic markers or predictive markers of responses to treatments. In different studies, profiles of miRNA expressions have been used to identify cancer types and stages. In addition, correlations between miRNA signatures and efficacies of responses to different therapeutic regimens have great clinical value26. Understanding the molecular mechanisms involved in miRNA dysregulation could therefore be useful in selecting appropriate target molecules or treatment approaches. ## The role of microRNAs in the tumor microenvironment The environments surrounding tumors are recognized as key contributors to tumor growth, angiogenesis, metastasis, and therapeutic resistance27,28. Several studies have shown that molecular interactions among tumor cells and other cell types such as endothelial cells (ECs), fibroblasts, and innate and adaptive immune cells in the tumor microenvironments regulate tumor development29. MiRNAs can participate in intracellular communication as mediators of signaling by amplifying a signal or tuning its response (**Figure 1**). However, they are also released by cells and mediate crosstalk between tumor cells and their microenvironments30. Previous studies have demonstrated that miRNAs have important effects in transdifferentiation of normal fibroblasts to cancer-associated fibroblasts (CAFs). For example, dysregulation of miRNAs like miR-9, miR-21, miR-15, miR-16, miR-31, miR-155, miR-200s, miR-320, and miR-214 reprogram fibroblasts to CAFs31. In addition, tumor-secreted factors such as TGF-β are known to influence miRNA expression in CAFs. For example, miR-21 expression is upregulated in CAFs by cancer cells secreting TGF-β, which results in uncontrolled proliferation of cancer cells32. In addition, miRNAs can affect chemokines and growth factors derived from CAFs, to participate in cancer cell invasion and metastasis. In such cases, low expression of miR-214 in the tumor microenvironment increases chemokine (C-C motif) ligand 5 and fibroblast growth factor 9 (FGF9), which contributes to tumor growth and motility33. Furthermore, significant downregulations of miR-15 and miR-16 have been reported in CAFs, which promotes proliferation and migration in cancer cells through activation of the FGFR1 pathway34. ![Figure 1](http://www.cancerbiomed.org/https://www.cancerbiomed.org/content/cbm/19/3/289/F1.medium.gif) [Figure 1](http://www.cancerbiomed.org/content/19/3/289/F1) Figure 1 Dysregulation of miRNAs in tumor microenvironment. MiRNAs mediate crosstalk between tumor cells and their microenvironment. Red and black colors represent oncomiRs and tumor suppressor miRNAs, respectively. ARG1: Arginase 1; CAF: Cancer associated fibroblast; CC: Cancer cell; DC: Dendritic cells; EC: Endothelial cell; IDO: Indoleamine 2,3-dioxygenase; iNOS: inducible NO synthase; M: Macrophage; MDSC: Myeloid-derived suppressor cell; NFb: Normal fibroblast; NK: Natural killer cells; TL: T lymphocyte. MiRNAs also play key regulatory roles in pathways related to angiogenesis within the tumor microenvironment. Some of the most commonly known miRNAs, such as miR-93, miR-29, and miR-519c negatively control angiogenesis while miR-155 and miR-210 can positively influence this complex process35–37. During hypoxia, the overexpression of miR-130a has been shown to promote endothelial cell migration *via* targeting c-MYB38. Other studies have also revealed that miR-21 can induce angiogenesis *via* activation of the AKT and ERK1/2 signaling pathways, which elevate levels of VEGF and HIF-1α39. MiR-9 can stimulate EC migration and angiogenesis *via* aberrantly activating the JAK-STAT pathway40. Furthermore, miR-7 has emerged as an inhibitor of angiogenesis by targeting the EGFR/PI3K/AKT signaling pathways41. Moreover, miRNAs like miR-29b reduce expression and activity of ECM signaling genes such as MMP2 and MMP9. Thus, downregulation of this miRNA attenuates angiogenesis through the MAPK/ERK and PI3K/AKT signaling pathways, directly affecting cancer metastasis42. Diverse infiltrated immune cells are present in the tumor microenvironment. Previous studies have revealed that miRNAs also affected the immune cells recruited to cancer sites. For example, miR-155 and miR-21 are mostly upregulated in myeloid-derived suppressor cells (MDSCs), and promote immunosuppressive effects *via* activation of the STAT3 pathway in cancer cells43. These miRNAs can induce MDSC expansion, which contributes to tumor progression. Conversely, reduction of the anti-inflammatory miRNA, miR-199a, leads to activation of the NF-kB signaling pathway and tumor progression44,45. Previous studies have also shown that miRNAs such as miR-21, miR-29a, and miR-147 released from cancer cells can regulate the inflammatory response and result in cancer progression and invasion46–48. Nakano and co-workers reported that cancer cell-derived exosomal miR-92b inhibited natural killer (NK) cell-mediated cytotoxicity by deregulating CD6949. Furthermore, transfer of cancer cell-secreted miR-214 to regulatory T cells (Tregs) suppresses immune responses at the tumor site *via* promoting Treg expansion50. Taken together, given the pivotal roles of miRNAs in the tumor microenvironment, targeting these small molecules presents a very promising strategy for cancer treatment. ## Therapeutic strategies for microRNAs Several methods have been characterized to find appropriate target miRNAs for the most effective cancer treatments based on the cancer type. Among them, a combination of miRNA sequencing (miRNA-seq) and bioinformatics analysis has been shown to facilitate the identification of miRNAs. MicroRNA sequencing has also been used to quantitatively profile small RNAs and discover new miRNAs, as well as investigate their functions51–54. Recently, Wang and co-workers have shown that co-sequencing microRNAs and mRNAs in the same cell revealed the non-genetic heterogeneity of small RNAs, while at the same time improving our understanding of the contribution of different pathways in controlling miRNA expression55. Different strategies have been reported for regulating miRNA expression, 1 of which is restoration of suppressed miRNA levels using synthetic miRNA mimics or miR-expressing vectors. Other strategies developed to suppress oncomiRs include using anti-miR (AMOs), miRNA sponges, miRNA masking, and small molecule inhibitors (SMIRs) that can epigenetically influence miRNA expression56,57. AMOs are chemically modified synthetic antisense sequences, and despite their transient inhibition due to dilution during cell replication, they are still the most popular and are frequently used for their simple structures. The miRsponge consists of multiple tandem complementary binding sites designed to trap a specific miRNA, whereas miRNA masks are single-stranded antisense oligonucleotides designed to block the desired miRNA target site in specific mRNAs58,59. In contrast, SMIRs can block miRNA biogenesis at the pre-transcription, transcription, and post-transcription levels. They offer some advantages such as short production time, low cost, and long circulation times in the blood60. However, SMIRs may have off-target effects in healthy tissues58,59. Recent studies have also shown that CRISPR/Cas9, a feasible and simple strategy for creating gene knockout or knock-in human cells, could also efficiently generate miRNA knockout mutants61. All of these strategies are more efficient *in vitro* than *in vivo* due to insufficient *in vivo* delivery of these molecules. Therefore, it is necessary to identify better delivery systems and proper administration routes to overcome this limitation. The major miRNA administration routes are local and systemic administration. Compared with systemic administration, local administration has shown fewer side effects and toxicity. In addition, local delivery seems to be a promising approach to robustly protect small RNAs from endogenous RNase, and requires only a minimal dose of these molecules to obtain maximum therapeutic effects62–64. Although systemic administration is more effective in metastatic cancers, rapid clearance from the body is still an issue63. Additionally, negatively charged hydrophilic miRNAs have difficultly passing through the cell membrane and penetrating cells. Thus, for maximizing the killing potential of miRNA, carrier systems must protect small RNAs from degradation, avoid rapid clearance, and increase their circulation time in the blood stream. This would facilitate their specific uptake by cancer cells with less toxicity to normal cells. Several microRNA carrier systems that have recently been developed (**Figure 2**) are briefly discussed below. ![Figure 2](http://www.cancerbiomed.org/https://www.cancerbiomed.org/content/cbm/19/3/289/F2.medium.gif) [Figure 2](http://www.cancerbiomed.org/content/19/3/289/F2) Figure 2 Schematic illustration of different vehicles applied for miRNAs delivery to cancer cells. The 5 major groups of microRNA delivery systems are nano delivery systems, viral vectors, virus-like particles, exosomes and mesenchymal stem cells that have their own advantages and disadvantages. Choosing delivery systems based on the type and stage of cancer might accelerate translating miRNA-based therapeutics into the clinic. ### Nanoparticle-based microRNA delivery systems Although nanocarrier delivery systems are a relatively new area of research, they are rapidly establishing their own niche in the pharmaceutical industry due to their small sizes and low molecular weights. Nanoparticle-based microRNA delivery systems are divided into 4 major groups, each of which has its advantages and disadvantages, involving lipid-based carriers, polymer-based delivery systems, inorganic nanoparticles, and nucleic acid-based delivery systems (**Table 1**). Most of these nanoparticles can be functionalized by conjugating with antibodies. View this table: [Table 1](http://www.cancerbiomed.org/content/19/3/289/T1) Table 1 Comparison of various nanocarriers for miRNA-based therapy #### Lipid-based carriers Lipid nanocarriers are made of both solid and liquid lipids by homogenization-emulsification methods to encapsulate different drugs114. Three groups of lipids (cationic, neutral, and ionizable) are commonly used for delivery of miRNAs to cancer cells65. Recently, ionizable lipids have been used more often than cationic and neutral lipids for safe and efficient *in vivo* delivery. At acidic pH, negatively charged miRNAs are encapsulated in pH-sensitive ionizable lipids *via* electrostatic interactions. Because these particles have essentially no surface charge at near-physiological pH, they are less likely to form aggregates with plasma proteins and do not induce a robust immune response65. In the acidic tumor microenvironment, positively charged ionizable lipid particles can easily bind to negatively charged cell membranes, increasing cellular uptake. Thus, these smart lipid particles are suitable for systemic and local administration of drugs and can reduce undesirable side effects115–117. However, a higher loading capacity and modifications of the lipids are needed to translate these lipid carriers into the clinical setting. #### Polymer-based systems Polymer-based delivery systems are divided into natural and synthetic groups. Abundant natural polymers like chitosan possess mucoadhesive properties, have low toxicity, and have pH-tunable drug release characteristics, making them ideal candidates as drug carriers73,118. Other natural polymers, such as cell-penetrating peptides (CPP) can pass through cellular membranes to deliver a variety of biologically active conjugates87. However, despite the higher stability, good cellular uptake and lower toxicity of CPP-conjugated nanoparticles, they may increase particle aggregation and endosomal entrapment86. Synthetic polymers include dendrimers, poly(lactic-co-glycolic acid) (PLGA), and polyethyleneimines (PEIs). Although synthetic polymers are easily modifiable and have high buffering capacities, their toxicity and uncontrolled release of drug still need to be addressed44,76,77,79,119. #### Nucleic acid-based delivery systems DNA nanotechnology has shown great potential for preparing structures that serve as delivery systems. The advantages of using nucleic acids as delivery vehicles as compared to peptide-based delivery systems include lower induced immunity, higher solubility, greater thermodynamic stability, and lower toxicity. Aptamers and pRNAs are 2 classes of these carriers. Aptamers are a group of nucleic acid delivery vehicles often called chemical antibodies, which are capable of binding various molecules with high affinity and specificity. These molecules can serve as cancer therapeutics as well as diagnostic and delivery tools109. To increase drug-loading capacity, a polyvalent aptamer system composed of Mucin1 and AS1411 with pH-sensitive drug-release properties was developed120. Although aptamers have rapidly advanced the drug delivery field, some limitations that reduce their efficacy still need to be addressed, such as degradation in the presence of blood nucleases and nonspecific interactions with bloodstream proteins109. Previous studies have shown that packaging RNA (pRNA), a conserved noncoding RNA in the phi29 DNA packaging motor, has a strong tendency to form multimeric structures and therefore represents an ideal drug delivery system for several therapeutic agents121,122. Additional nuclease stability is conferred to this RNA by chemical modifications using 2′-fluoro on the ribose sugar of nucleotides111. Furthermore, to enhance the mechanical and thermal stability of these carriers, the pRNA contains an ultra-stable three-way junction (3WJ), which was developed for successful translation of small RNAs to cancer cells111,123. ### Viral-based microRNA delivery methods Synthetic viral vectors are emerging as promising tools for gene therapy due to their high transduction efficacy and long-lasting gene expression in different types of cells. The viruses most commonly utilized for delivery of therapeutic genes are lentiviruses (LVs), adenoviruses (Ads), and adeno-associated viruses (AAVs)63,124,125. HIV-based vectors have been developed to deliver therapeutic genes, although other LVs including SIV and FIV are also being explored87,126. LVs have also been used for microRNA replacement therapy63. However, insertional mutagenesis, genetic instability, activation of oncogenes, and histone modifications may occur after integration of the retroviral genome into the host cell genome, limiting their clinical use78,127–129. To overcome the risk of insertional mutagenesis in cytotoxic cancer therapies, integration-defective lentiviral vectors have been developed using mutations in the viral integrase or attachment sites. For example, overexpression of miR-145 using episomal lentiviral vector can stop proliferation and induce apoptosis in esophageal cancer cells130. Third-generation adenoviruses can express miRNAs in a stable manner without integrating into the host genome131. In 1 study, engineered oncolytic adenoviruses were used to deliver miRNA-148a to pancreatic cancer cells without risking viral-associated toxicity in normal cells132. However, a major drawback of these vectors is that they strongly induce immune responses. To overcome this problem, adenoviruses have been chemically or physically modified using polymers. In addition, to enhance their tumor tropism, a variety of ligands have been introduced into coated-Ads133,134. AAVs also have good safety and display low immunogenicity, and their wide tissue tropism makes them promising tools for gene delivery *in vivo*135,136. So far, there are more than 13 AAV serotypes identified, most of which have been tested in more than 200 gene delivery trial studies135,137. Though limited insert capacity is the most familiar disadvantage of AAVs138; they remain good candidates for delivery of small DNA such as miRNAs and siRNAs. Bhere et al. showed that co-delivery of miR-21 and miR-7 mediated by AVVs effectively induced apoptosis in mice bearing brain tumors56. Additionally, AAV-miRNAs can also be used in combination with other therapeutic methods. In a study characterizing the therapeutic potential of miR-7-AAVs and S-TRAIL-MSCs for glioblastomas, the delivery of miR-7 *via* AAV led to upregulation of death receptor5 (DR 5), also known as TRAIL. Co-delivery of miR-7-AAVs and S-TRIAL-MSCs was shown to result in strong tumor growth inhibition and prolonged mouse survival139. Overall, delivery of AAV-mediated miRNAs holds great promise in cancer therapy. In summary, viral vectors provide long-term stable expression of the desired gene. They could become an efficient means of clinical application of miRNAs if their stimulation of immune responses can be overcome. ### Virus-like particle (VLP) delivery systems VLPs, or noninfectious viral protein structures, were successfully developed for drug delivery in the early 1990s. VLPs can reassemble *in vitro*, after which they can be loaded with drugs or small molecules140–142. MS2 is a member of the Leviviridae family, with single strand-RNA that infects *Enterobacteriaceae*143,144. MS2 can be produced as pseudo-viral particles and possesses multiple unique features that makes it an interesting candidate for targeted gene delivery, such as quick production, self-assembly in the presence of nucleic acids, and easy capsid modification145. Several studies suggested that Tat modification of VLPs improved the cellular uptake of these particles. Delivery of miR-146a was increased 1−14-fold *in vitro* and 2-fold *in vivo* after Tat peptide grafting146. Likewise, VLPs derived from PP7 are nontoxic and protect RNAs against nucleolytic attack. Interestingly, they are more resistant to high temperatures than MS2. Sun et al. demonstrated that TAT peptide and pre-microRNA-23b co-loaded with PP7 VLP could be effectively delivered to hepatoma cells and significantly inhibited their migration147. Furthermore, it is well documented that the capsid protein (VP1) of the John Cunningham virus (JC), which belongs to the Polyomavirus family and causes central nervous disease, has self-assembling properties both *in vitro* and *in vivo*, and is useful in creating VLPs. The superiority of these VLPs over PP7 and MS2 VLPs for treating brain diseases is that VP1 mediates cellular entry through interaction with cell receptors in the central nervous system148,149. In brief, VLPs are thermoresistant and more stable than lipid-based particles, which protect miRNAs from rapid degradation by nucleases. The 2 most distinctive characteristics of these particles are their lower toxicities and infectivities due to their lack of a viral genome. The pH-dependent assembly-disassembly behavior of VLPs also makes them attractive candidates for drug delivery to target cells *via* exogenously produced virus structures. Although VLPs can be easily modified to specifically target cancer cells, overall use and successful clinical translation of these particles will depend on solving their immunogenicity problems150–152. ### Delivery of exosome-mediated microRNAs Exosomes are nanosized cell-derived vesicles with a diameter range of 40−100 nm. These small particles originate from the endocytic pathway and are present in most body fluids. Exosomes can be taken-up by recipient cells, whose biological functions are altered in a paracrine manner. Because exosomes originate from the plasma membrane, they are safe and biocompatible, and induce immune tolerance *in vivo*153–155. An important feature of exosomes is their ability to overcome otherwise impermeable biological barriers such as the blood-brain barrier (BBB). Most previous studies used exosomes derived from mesenchymal stem cells (MSCs) and HEK cells as vehicles for loading miRNAs. Shimbo et al. demonstrated that MSC-derived exosomes containing synthetic miR-143 reduced the migration of osteosarcoma cells. However, the efficacy of exosome-mediated delivery of miR-143 was lower than that of lipofection156. One strategy to improve the pharmacokinetic properties of these small vesicles is to introduce appropriate ligands to specifically target cancer cells. In 1 study, Ohno et al. engineered donor cells to express the transmembrane domain of the platelet-derived growth factor receptor fused to the GE11 peptide, which specifically bound EGFR-expressing breast cancer cells. Their results showed that let-7a-containing GE11-exosomes suppressed the growth of xenograft breast tumor in mice157. In another study, exosomes derived from dendritic cells were modified with aptamer AS1411 to facilitate targeting of breast tumor cells158. Another strategy for increasing extracellular vesicle (EV) therapeutic efficacy is to enhance drug loading into these nanocarriers without affecting their membrane integrity. However, the inability to produce large quantities of highly purified EVs, as well as their rapid clearance from the bloodstream and accumulation in lung, liver, and spleen are major obstacles associated with this method159–161. Solutions to the drawbacks associated with the application of EVs as miRNA delivery systems are therefore needed before clinical studies can be initiated. ### MSC-based microRNA delivery MSCs are 1 of the most popular types of stem cells used in clinical trials. Their immunomodulatory properties and suppression of inflammatory responses *in vivo* make them good candidates for stem cell-based therapy162–164. Furthermore, MSCs are easily engineered to improve their therapeutic potential or modify their adverse effects165,166. As small molecules can easily translocate between MSCs and various other cell types through gap junctions and exosomes, MSCs have been examined as carriers for therapeutic RNAs. Munoz et al. reported that delivery of anti-miR-9 to chemoresistant glioblastoma multiforme (GBM) cells sensitized them to temozolomide167. MiR-124 and miR-145 mimics were also delivered using MSCs to glioma cells and glioma stem cells *in vitro* and *in vivo*. The results of that study showed that delivery of miR-124 or miR-145 decreased both migration and self-renewal in GBMs168. Contradicting their results, our study indicated that co-culture of LV-anti-miR-21-Ad-MSCs with prostate, colon, and GBM cancer cells did not significantly inhibit tumor cell growth56. However, these contradictions may arise from the source from which the MSCs are derived or the enrichment of the desired miRNA in exosomes. Despite the many advantages of MSCs, the major drawback with using stem cells as biological carriers for cancer treatment is that tumor cells can produce factors that enhance proliferation and invasion of stem cells169. To overcome this shortcoming, it would be better to use a combination of stem cell-based delivery systems and other therapeutic approaches170. ## Clinical translation As miRNAs are easily detected in body fluids, they could be used as noninvasive markers for early diagnosis of cancers. The first commercial miR-based diagnostic kit, miRNA-7™, was designed for liver cancer detection, and well-defined panels of miRNAs improved the sensitivity of cancer detection170. For example, ThyraMIR offers miR-based thyroid cancer screening and stratification based on the expression profiles of miRNAs such as miR-29b-1-5p, miR-31-5p, miR-138-1-3p, miR-139-5p, miR-146b-5p, miR-155, miR-204-5p, miR-222-3p, miR-375, and miR-551b-3p. ThyraMIR has been used in combination with ThyGeNEXT® to distinguish between benign and malignant lesions171. Clinical trials utilizing microRNA for cancer therapy have shown promising results. In 2013, the first report of miR mimics entering clinical trials involved MRX34172. MiR-34 is lost or down-regulated in a wide range of cancer types and plays an important regulatory role in tumor suppression. MRX34 is a double-stranded RNA oligonucleotide that has been incorporated into amphoteric liposomes and examined in different cancer types. Amphoteric liposomes increase the circulation time of MRX34 in the blood stream and deliver higher amounts of it to target tissues. Immune-related adverse reactions associated with MRX34 in 2016 limited its application. However, in 2017, the first in-human phase I clinical study showed that MRX34 with dexamethasone premedication had acceptable safety and tolerability in patients with advanced solid tumors. In this study, MRX34 was injected twice weekly for 3 weeks and showed anti-tumor activity in a subset of patients173. However, we still need an efficient carrier for miR-34 that avoids immune activation during delivery to tumors174. Correspondingly, Miravirsen, a LNA-based antisense oligonucleotide to miR-122, is nontoxic even in humans and has successfully completed Phase II clinical trials. Miravirsen is being developed by Santaris Pharma for hepatitis C treatment60,175. An LNA-based anti-miR-155, MRG-106, has shown tolerability and pharmacokinetic benefits in patients with cutaneous T-cell lymphoma in a first-in-human phase I clinical trial176. Additionally, MesomiR 1 is a bacterially-derived small spherical minicell designed to deliver double-stranded 23-base-pair miR-16 mimics to pleural mesothelioma *in vivo*. MesomiR 1 has now passed the phase I clinical study due to its safety and tolerability, and entered a phase II clinical trial177,178. MRG-201, a LNA-miRNA mimic, which restores miR-29b in human cancer, is currently undergoing phase II clinical trials *via* intradermal injection171. Despite the small number of clinical trial studies on miRNAs, many small RNAs have shown beneficial effects in tumor treatment that warrant their progression to clinical trials. Some of them, like miR-21, miR-182, miR-15a, miR-103/107, miR-122, miR-195, and miR-208, are now in preclinical studies and await further clinical investigations. ## Challenges and perspectives Several lines of evidence demonstrate that miRNAs are emerging as promising noninvasive biomarkers and cancer therapies. MiR-based therapy is divided into 2 therapeutic strategies: modulation of miRNA biogenesis and miRNA replacement or antisense therapy179,180. A major advantage of miRNA replacement therapy is that a single small molecule is able to regulate several hundred targets in different oncogenic pathways181. Additionally, this approach can restore normal function in target cells with fewer side effects involving noncancerous tissues. Furthermore, the phenomenon of acquired resistance to therapy is a challenge in the field of cancer. Studies have shown that several miRNAs are involved in radiation-, immune-, and chemoresistance in various cancer types20. We envision that targeting miRNAs combined with other cancer therapy strategies could successfully increase treatment efficiency, overcome cancer resistance, and improve patient outcomes. However, the translation of promising preclinical findings to the clinical setting faces technical challenges related to tolerance issues and the lack of a proper delivery system. To overcome these limitations, successful carriers of miRNA molecules must be designed to enhance bioavailabilities and the endosomal escape of particles, and to evade the host’s immune response63,182. So far, diverse miRNA delivery vehicles have been developed that have their own advantages and disadvantages. Some of these systems are currently being used in phase I and II clinical trials for gene therapies. Furthermore, it is likely that designing and choosing delivery systems and routes of administration based on the type and stage of cancer will accelerate clinical translation of miRNA-replacement therapies. For example, oral delivery can be helpful for cancers related to the digestive tract183. In addition, use of mucoadhesive chitosan-based nanoparticles is effective for oral administration of drugs. Less invasive intranasal administration can overcome the BBB and be used for targeting CNS diseases184; cell-penetrating peptide-mediated delivery systems will be useful for this purpose. Additionally, local delivery of miRNAs *via* viral vectors or nanofibers seems more efficient and less toxic in treating early-stage cancers, whereas for metastatic cancer, more stable and less immunogenic delivery vehicles with high affinity and specificity for cancer sites elicit better responses63,185. Furthermore, the use of “cocktail” carriers or multiple simultaneous approaches may increase the transfer efficiency of miRNAs to target cells. More recently, a report by Wang et al. reported that a let-7a core dendrimer encapsulated in natural killer cell-derived exosomes significantly inhibited neuroblastoma growth without inducing immune response186. At the same time, to improve the therapeutic effectiveness of miRNAs, efforts should be made to optimize therapeutic doses and schedules based on personalized medicine. Moreover, due to intratumor heterogeneity, targeting multiple molecules or pathways simultaneously with the goal of synergic effects will likely prove more efficient. In this regard, manipulation of different miRNAs in the same cells would be a more promising strategy to fully eradicate certain tumors. ## Conflict of interest statement K.S. owns equity in and is a member of the Board of Directors of AMASA Therapeutics, a company developing stem cell-based therapies for cancer. K.S.’s interests were reviewed and are managed by Brigham and Women’s Hospital and Partners HealthCare in accordance with their conflict of interest policies. The other authors declare that they have no competing interests. ## Acknowledgments This work was supported by NIH grants R01-CA201148 (K.S.) and by DoD grant LC180495. * Received May 13, 2021. * Accepted July 27, 2021. * Copyright: © 2022, Cancer Biology & Medicine [https://creativecommons.org/licenses/by/4.0/](https://creativecommons.org/licenses/by/4.0/) This is an open access article distributed under the terms of the [Creative Commons Attribution License (CC BY) 4.0](https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution and reproduction in any medium, provided the original author and source are credited. ## References 1. 1. Anglicheau D, Muthukumar T, Suthanthiran M. MicroRNAs: small RNAs with big effects. Transplantation. 2010; 90: 105–12. [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=20574417&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000280098400001&link_type=ISI) 2. 2. Martinez VD, Cohn DE, Telkar N, Minatel BC, Pewarchuk ME, Marshall EA, et al. Profiling the small non-coding RNA transcriptome of the human placenta. Sci Data. 2021; 8: 166. 3. 3. Peng Y, Croce CM. The role of microRNAs in human cancer. Signal Transduct Target Ther. 2016; 1: 15004. 4. 4.1. 2. Walayat A, 3. Yang M, 4. Xiao D Sharad S, Kapur S. Therapeutic implication of miRNA in human disease. In: Walayat A, Yang M, Xiao D, editors. Antisense Therapy. London: IntechOpen; 2018. p. 82738. 5. 5. Hata A, Lieberman J. Dysregulation of microRNA biogenesis and gene silencing in cancer. Sci Signal. 2015; 8: re3. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6ODoic2lndHJhbnMiO3M6NToicmVzaWQiO3M6OToiOC8zNjgvcmUzIjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTkvMy8yODkuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 6. 6. Iorio MV, Croce CM. Causes and consequences of microRNA dysregulation. Cancer J. 2012; 18: 215–22. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1097/PPO.0b013e318250c001&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=22647357&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 7. 7. Zeinali T, Mansoori B, Mohammadi A, Baradaran B. Regulatory mechanisms of miR-145 expression and the importance of its function in cancer metastasis. Biomed Pharmacother. 2019; 109: 195–207. 8. 8. Wang B, Hsu SH, Wang X, Kutay H, Bid HK, Yu J, et al. Reciprocal regulation of microRNA-122 and c-Myc in hepatocellular cancer: role of E2F1 and transcription factor dimerization partner 2. Hepatology. 2014; 59: 555–66. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1002/hep.26712&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=24038073&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 9. 9. Misiewicz-Krzeminska I, Krzeminski P, Corchete LA, Quwainder D, Rojas EA, Herreero AB, et al. Factors regulating microRNA expression and function in multiple myeloma. Noncoding RNA. 2019; 5: 9. 10. 10. Xiao J, Lin H, Luo X, Luo X, Wang Z. MiR-605 joins p53 network to form a p53:miR-605:Mdm2 positive feedback loop in response to stress. EMBO J. 2011; 30: 5021. [FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiRlVMTCI7czoxMToiam91cm5hbENvZGUiO3M6NzoiZW1ib2pubCI7czo1OiJyZXNpZCI7czoxMDoiMzAvMjQvNTAyMSI7czo0OiJhdG9tIjtzOjE4OiIvY2JtLzE5LzMvMjg5LmF0b20iO31zOjg6ImZyYWdtZW50IjtzOjA6IiI7fQ==) 11. 11. Hermeking H. The miR-34 family in cancer and apoptosis. Cell Death Differ. 2010; 17: 193–9. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/cdd.2009.56&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=19461653&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000273518700003&link_type=ISI) 12. 12. Ma J, Gong W, Liu S, Li Q, Guo M, Wang J, et al. Ibrutinib targets microRNA-21 in multiple myeloma cells by inhibiting NF-κB and STAT3. Tumour Biol. 2018; 40: 1010428317731369. 13. 13. Kent OA, Chivukula RR, Mullendore M, Wentzel EA, Feldmann G, Lee KL, et al. Repression of the miR-143/145 cluster by oncogenic Ras initiates a tumor-promoting feed-forward pathway. Genes Dev. 2010; 24: 2754–9. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6ODoiZ2VuZXNkZXYiO3M6NToicmVzaWQiO3M6MTA6IjI0LzI0LzI3NTQiO3M6NDoiYXRvbSI7czoxODoiL2NibS8xOS8zLzI4OS5hdG9tIjt9czo4OiJmcmFnbWVudCI7czowOiIiO30=) 14. 14. Bracken CP, Gregory PA, Kolesnikoff N, Bert AG, Wang J, Shannon MF, et al. A double-negative feedback loop between ZEB1-SIP1 and the microRNA-200 family regulates epithelial-mesenchymal transition. Cancer Res. 2008; 68: 7846–54. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NjoiY2FucmVzIjtzOjU6InJlc2lkIjtzOjEwOiI2OC8xOS83ODQ2IjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTkvMy8yODkuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 15. 15. Jiang RL, Li Y. Regulation of miRNA pathway and roles of microRNAs in tumorigenesis and metastasis. Human Genet Embryol S. 2013; 2: 2161–436. 16. 16. Kim J, Yao F, Xiao Z, Sun Y, Ma L. MicroRNAs and metastasis: small RNAs play big roles. Cancer Metastasis Rev. 2018; 37: 5–15. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1007/s10555-017-9712-y&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=29234933&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 17. 17. Yoshida K, Yamamoto Y, Ochiya T. MiRNA signaling networks in cancer stem cells. Regen Ther. 2021; 17: 1–7. 18. 18. Diepenbruck M, Tiede S, Saxena M, Ivanek R, Kalathur RKR, Lüönd F, et al. MiR-1199-5p and Zeb1 function in a double-negative feedback loop potentially coordinating EMT and tumour metastasis. Nat Commun. 2017; 8: 1168. 19. 19. Zhu X, Pan H, Liu L. Long noncoding RNA network: Novel insight into hepatocellular carcinoma metastasis (Review). Int J Mol Med. 2021; 48: 1–12. 20. 20. Arghiani N, Matin MM. MiR-21: A key small molecule with great effects in combination cancer therapy. Nucleic Acid Ther. 2021; 31: 271–83. 21. 21. Cai Z, Zhang F, Chen W, Zhang J, Li H. MiRNAs: A promising target in the chemoresistance of bladder cancer. Onco Targets Ther. 2019; 12: 11805–16. 22. 22. An X, Sarmiento C, Tan T, Zhu H. Regulation of multidrug resistance by microRNAs in anti-cancer therapy. Acta Pharm Sin B. 2017; 7: 38–51. 23. 23. Jamialahmadi K, Zahedipour F, Karimi G. The role of microRNAs on doxorubicin drug resistance in breast cancer. J Pharm Pharmacol. 2021; 73: 997–1006. 24. 24. Tian Y, Tang L, Yi P, Pan Q, Han Y, Shi Y, et al. MiRNAs in radiotherapy resistance of nasopharyngeal carcinoma. J Cancer. 2020; 11: 3976–85. 25. 25. Li D, Tolleson WH, Yu D, Chen S, Guo L, Xiao W, et al. Regulation of cytochrome P450 expression by microRNAs and long noncoding RNAs: Epigenetic mechanisms in environmental toxicology and carcinogenesis. J Environ Sci Health C Environ Carcinog Ecotoxicol Rev. 2019; 37: 180–214. 26. 26. Lan H, Lu H, Wang X, Jin H. MicroRNAs as potential biomarkers in cancer: opportunities and challenges. Biomed Res Int. 2015; 2015: 125094. 27. 27. Arneth B. Tumor microenvironment. Medicina. 2019; 56: 15. 28. 28. Caner A, Asik E, Ozpolat B. SRC signaling in cancer and tumor microenvironment. Adv Exp Med Biol. 2021; 1270: 57–71. 29. 29. Antomarchi J, Ambrosetti D, Cohen C, Delotte J, Chevallier A, Karimdjee-Soilihi B, et al. Immunosuppressive tumor microenvironment status and histological grading of endometrial carcinoma. Cancer Microenviron. 2019; 12: 169–79. 30. 30. Avraham R, Yarden Y. Regulation of signalling by microRNAs. Biochem Soc Trans. 2012; 40: 26–30. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NzoicHBiaW9zdCI7czo1OiJyZXNpZCI7czo3OiI0MC8xLzI2IjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTkvMy8yODkuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 31. 31. Eichelmann AK, Matuszcak C, Hummel R, Haier J. Role of miRNAs in cell signaling of cancer associated fibroblasts. Int J Biochem Cell Biol. 2018; 101: 94–102. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.biocel.2018.05.015&link_type=DOI) 32. 32. Boguslawska J, Kryst P, Poletajew S, Piekielko-Witkowska A. TGF-β and microRNA Interplay in genitourinary cancers. Cells. 2019; 8: 1619. 33. 33. Wang R, Sun Y, Yu W, Yan Y, Qiao M, Jiang R, et al. Downregulation of miRNA-214 in cancer-associated fibroblasts contributes to migration and invasion of gastric cancer cells through targeting FGF9 and inducing EMT. J Exp Clin Cancer Res. 2019; 38: 20. 34. 34. Musumeci M, Coppola V, Addario A, Patrizii M, Maugeri-Sacca M, Memeo L, et al. Control of tumor and microenvironment cross-talk by miR-15a and miR-16 in prostate cancer. Oncogene. 2011; 30: 4231–42. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/onc.2011.140&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=21532615&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000296356300002&link_type=ISI) 35. 35. Loh HY, Norman BP, Lai KS, Rahman NM, Alitheen NB, Osman MA. The regulatory role of microRNAs in breast cancer. Int J Mol Sci. 2019; 20: 4940. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.3390/ijms20194940&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=31590453&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 36. 36. Lou W, Liu J, Gao Y, Zhong G, Chen D, Shen J, et al. MicroRNAs in cancer metastasis and angiogenesis. Oncotarget. 2017; 8: 115787–802. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.18632/oncotarget.23115&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=29383201&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 37. 37. Wang H, Li K, Mei Y, Huang X, Li Z, Yang Q, et al. Sp1 suppresses miR-3178 to promote the metastasis invasion cascade via upregulation of TRIOBP. Mol Ther Nucleic Acids. 2018; 12: 1–11. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.omtn.2018.04.008&link_type=DOI) 38. 38. Yang H, Zhang H, Ge S, Nng T, Bai M, Li J, et al. Exosome-derived miR-130a activates angiogenesis in gastric cancer by targeting C-MYB in vascular endothelial cells. Mol Ther. 2018; 26: 2466–75. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.ymthe.2018.07.023&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=30120059&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 39. 39. Liu LZ, Li C, Chen Q, Jing Y, Carpenter R, Jiang Y, et al. MiR-21 induced angiogenesis through AKT and ERK activation and HIF-1α expression. PLoS One. 2011; 6: e19139. 40. 40. Zhuang G, Wu X, Jiang Z, Kasman I, Yao J, Guan Y, et al. Tumour-secreted miR-9 promotes endothelial cell migration and angiogenesis by activating the JAK-STAT pathway. EMBO J. 2012; 31: 3513–23. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NzoiZW1ib2pubCI7czo1OiJyZXNpZCI7czoxMDoiMzEvMTcvMzUxMyI7czo0OiJhdG9tIjtzOjE4OiIvY2JtLzE5LzMvMjg5LmF0b20iO31zOjg6ImZyYWdtZW50IjtzOjA6IiI7fQ==) 41. 41. Goradel NH, Mohammadi N, Haghi-Aminjan H, Farhood B, Negshdari B, Sahebkar A. Regulation of tumor angiogenesis by microRNAs: State of the art. J Cell Physiol. 2019; 234: 1099–110. 42. 42. Chung HJ, Choi YE, Kim ES, Han YH, Park MJ, Bae IH. miR-29b attenuates tumorigenicity and stemness maintenance in human glioblastoma multiforme by directly targeting BCL2L2. Oncotarget. 2015; 6: 18429–44. 43. 43. Chen S, Zhang Y, Kuzel TM, Zhang B. Regulating tumor myeloid-derived suppressor cells by microRNAs. Cancer Cell Microenviron. 2015; 2: 1–8. 44. 44. 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. 45. 45. Sirotkin AV, Alexa R, Kišová G, Harrath AH, Alwasel S, Ovcharenko D, et al. MicroRNAs control transcription factor NF-kB (p65) expression in human ovarian cells. Funct Integr Genomics. 2015; 15: 271–5. 46. 46. An Y, Yang Q. MiR-21 modulates the polarization of macrophages and increases the effects of M2 macrophages on promoting the chemoresistance of ovarian cancer. Life Sci. 2020; 242: 117162. 47. 47. Afshar AS, Xu J, Goutsias J. Integrative identification of deregulated miRNA/TF-mediated gene regulatory loops and networks in prostate cancer. PLoS One. 2014; 9: e100806. 48. 48. Shoeibi S. Diagnostic and theranostic microRNAs in the pathogenesis of atherosclerosis. Acta Physiol. 2020; 228: e13353. 49. 49. Nakano T, Chen IH, Wang CC, Chen PJ, Tseng HP, Huang KT, et al. Circulating exosomal miR-92b: Its role for cancer immunoediting and clinical value for prediction of posttransplant hepatocellular carcinoma recurrence. Am J Transplant. 2019; 19: 3250–62. 50. 50. Kogure A, Kosaka N, Ochiya T. Cross-talk between cancer cells and their neighbors via miRNA in extracellular vesicles: an emerging player in cancer metastasis. J Biomed Sci. 2019; 26: 7. 51. 51. Dai X, Huang R, Hu S, Zhou Y, Sun X, Gui P, et al. A novel miR-0308-3p revealed by miRNA-seq of HBV-positive hepatocellular carcinoma suppresses cell proliferation and promotes G1/S arrest by targeting double CDK6/Cyclin D1 genes. Cell Biosci. 2020; 10: 24. 52. 52. Lee WH, Chen KP, Wang K, Huang HC, Juan HF. Characterizing the cancer-associated microbiome with small RNA sequencing data. Biochem Biophys Res Commun. 2020; 522: 776–82. 53. 53. Hanna J, Hossain GS, Kocerha J. The potential for microRNA therapeutics and clinical research. Front Genet. 2019; 10: 478. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.3389/fgene.2019.00478&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=31156715&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 54. 54. Kim H, Kim J, Kim K, Chang H, You K, Kim VN. Bias-minimized quantification of microRNA reveals widespread alternative processing and 3’ end modification. Nucleic Acids Res. 2019; 47: 2630–40. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1093/nar/gky1293&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=30605524&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 55. 55. Wang N, Zheng J, Chen Z, Liu Y, Dura B, Kwak M, et al. Single-cell microRNA-mRNA co-sequencing reveals non-genetic heterogeneity and mechanisms of microRNA regulation. Nat Commun. 2019; 10: 95. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/s41467-018-07981-6&link_type=DOI) 56. 56. Bhere D, Arghiani N, Lechtich ER, Yao Y, Alsaab S, Bei F, et al. Simultaneous downregulation of miR-21 and upregulation of miR-7 has anti-tumor efficacy. Sci Rep. 2020; 10: 1779. 57. 57. Corsten MF, Miranda R, Kasmieh R, Krichevsky AM, Weissleder R, Shah K. MicroRNA-21 knockdown disrupts glioma growth in vivo and displays synergistic cytotoxicity with neural precursor cell delivered S-TRAIL in human gliomas. Cancer Res. 2007; 67: 8994–9000. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NjoiY2FucmVzIjtzOjU6InJlc2lkIjtzOjEwOiI2Ny8xOS84OTk0IjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTkvMy8yODkuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 58. 58. Lima JF, Cerqueira L, Figueiredo C, Oliveira C, Azevedo NF. Anti-miRNA oligonucleotides: A comprehensive guide for design. RNA Biol. 2018; 15: 338–52. 59. 59.1. 2. Balacescu O, 3. Visan S, 4. Baldasici O, 5. Balacescu L Sharad S, Kapur S. MiRNA-based therapeutics in oncology, realities, and challenges. In: Balacescu O, Visan S, Baldasici O, Balacescu L, editors. Antisense Therapy. London: IntechOpen. 2018; 31. 60. 60. Zhong S, Jeong JH, Chen Z, Chen Z, Luo JL. Targeting tumor microenvironment by small-molecule inhibitors. Transl Oncol. 2020; 13: 57–69. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.tranon.2019.10.001&link_type=DOI) 61. 61. Aquino-Jarquin G. Emerging role of CRISPR/Cas9 technology for microRNAs editing in cancer research. Cancer Res. 2017; 77: 6812–7. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NjoiY2FucmVzIjtzOjU6InJlc2lkIjtzOjEwOiI3Ny8yNC82ODEyIjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTkvMy8yODkuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 62. 62. Nam L, Coll C, Erthal LCS, de la Torre C, Serrano D, Martínez-Máñez R, et al. Drug delivery nanosystems for the localized treatment of glioblastoma multiforme. Materials (Basel). 2018; 11: 779. 63. 63. Chen Y, Gao DY, Huang L. In vivo delivery of miRNAs for cancer therapy: challenges and strategies. Adv Drug Deliv Rev. 2015; 81: 128–41. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.addr.2014.05.009&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=24859533&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 64. 64.1. 2. Delcassian D, 3. Patel AK Ladame S, Cheng JY. Nanotechnology and drug delivery. In: Delcassian D, Patel AK, editors. Bioengineering innovative solutions for cancer. Cambridge: Academic Press; 2020. p. 197–219. 65. 65. Rietwyk S, Peer D. Next-generation lipids in RNA interference therapeutics. ACS Nano. 2017; 11: 7572–86. 66. 66. Chen Y, Zhu X, Zhang X, Liu B, Huang L. Nanoparticles modified with tumor-targeting scFv deliver siRNA and miRNA for cancer therapy. Mol Ther. 2010; 18: 1650–6. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/mt.2010.136&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=20606648&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000281502300011&link_type=ISI) 67. 67. Zhang JS, Liu F, Huang L. Implications of pharmacokinetic behavior of lipoplex for its inflammatory toxicity. Adv Drug Deliv Rev. 2005; 57: 689–98. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.addr.2004.12.004&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=15757755&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 68. 68. Trang P, Wiggins JF, Daige CL, Cho C, Omotola M, Brown D, et al. Systemic delivery of tumor suppressor microRNA mimics using a neutral lipid emulsion inhibits lung tumors in mice. Mol Ther. 2011; 19: 1116–22. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/mt.2011.48&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=21427705&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 69. 69.1. 2. Myoung SS, 3. Kasinski AL Peplow PV, Martinez B, Calin GA, Esquela-Kerscher A. Strategies for safe and targeted delivery of microRNA therapeutics. In: Myoung SS, Kasinski AL, editors. MicroRNAs in diseases and disorders. London: Royal Society of Chemistry; 2019. p. 386–415. 70. 70.1. 2. Malik SS, 3. Masood N, 4. Sherrard A, 5. Bishop PN Malik B. Small non-coding RNAs as a tool for personalized therapy in familial cancers. In: Malik SS, Masood N, Sherrard A, Bishop PN, editors. AGO-driven non-coding RNAs. Cambridge: Academic Press; 2019. p. 179–208. 71. 71. Daige CL, Wiggins JF, Priddy L, Nelligan-Davis T, Zhao J, Brown D. Systemic delivery of a miR34a mimic as a potential therapeutic for liver cancer. Mol Cancer Ther. 2014; 13: 2352–60. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6MTA6Im1vbGNhbnRoZXIiO3M6NToicmVzaWQiO3M6MTA6IjEzLzEwLzIzNTIiO3M6NDoiYXRvbSI7czoxODoiL2NibS8xOS8zLzI4OS5hdG9tIjt9czo4OiJmcmFnbWVudCI7czowOiIiO30=) 72. 72. Bai Z, Wei J, Yu C, Han X, Qin X, Zhanf C, et al. Non-viral nanocarriers for intracellular delivery of microRNA therapeutics. J Mater Chem B. 2019; 7: 1209–25. 73. 73. 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. 74. 74. Palmerston Mendes L, Pan J, Torchilin VP. Dendrimers as nanocarriers for nucleic acid and drug delivery in cancer therapy. Molecules. 2017; 22: 1404. 75. 75. Ren L, Huang C, Liu YH, Yu Y, Lin L, Wen LJ. The effect and mechanism of microRNA-21 on cis-dichlorodiamineplatinum resistance in lung cancer cell strain. Zhonghua Yi Xue Za Zhi. 2016; 96: 1454–8. 76. 76. Mir M, Ahmed N, Rehman AU. Recent applications of PLGA based nanostructures in drug delivery. Colloid Surface B Biointerfaces. 2017; 159: 217–31. 77. 77. Navarro SM, Morgan TW, Astete CE, Stout RW, Coulon D, Mottram P, et al. Biodistribution and toxicity of orally administered poly (lactic-co-glycolic) acid nanoparticles to F344 rats for 21 days. Nanomedicine. 2016; 11: 1653–69. 78. 78. Sharma S, Parmar A, Kori S, Sandhir R. PLGA-based nanoparticles: a new paradigm in biomedical applications. Trac-Trend Anal Chem. 2016; 80: 30–40. 79. 79. Fernandez-Piñeiro I, Badiola I, Sanchez A. Nanocarriers for microRNA delivery in cancer medicine. Biotechnol Adv. 2017; 35: 350–60. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.biotechadv.2017.03.002&link_type=DOI) 80. 80. Zhang T, Xue X, He D, Hsieh JT. A prostate cancer-targeted polyarginine-disulfide linked PEI nanocarrier for delivery of microRNA. Cancer Lett. 2015; 365: 156–65. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.canlet.2015.05.003&link_type=DOI) 81. 81. Debnath S, Karan S, Debnath M, Dash J, Chatterjee TK. Poly-L-lysine inhibits tumor angiogenesis and induces apoptosis in ehrlich ascites carcinoma and in sarcoma S-180 tumor. Asian Pac J Cancer Prev. 2017; 18: 2255–68. 82. 82. Shukla RS, Qin B, Cheng K. Peptides used in the delivery of small noncoding RNA. Mol Pharm. 2014; 11: 3395–408. 83. 83. Jin H, Yu Y, Chrisler WB, Xiong Y, Hu D, Lei C. Delivery of microRNA-10b with polylysine nanoparticles for inhibition of breast cancer cell wound healing. Breast Cancer. 2012; 6: 9–19. [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=22259248&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 84. 84. Li T, Meng Z, Zhu X, Gan H, Gu R, Wu Z, et al. New synthetic peptide with efficacy for heparin reversal and low toxicity and immunogenicity in comparison to protamine sulfate. Biochem Biophys Res Commun. 2015; 467: 497–502. 85. 85. Yoo H, Mok H. Evaluation of multimeric siRNA conjugates for efficient protamine-based delivery into breast cancer cells. Arch Pharm Res. 2015; 38: 129–36. [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=24687259&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 86. 86. Silva S, Almeida AJ, Vale N. Combination of cell-penetrating peptides with nanoparticles for therapeutic application: A review. Biomolecules. 2019; 9: 22. 87. 87. Mcclorey G, Banerjee S. Cell-penetrating peptides to enhance delivery of oligonucleotide-based therapeutics. Biomedicines. 2018; 6: 51. 88. 88. Oh B, Song H, Lee D, Oh J, Kim G, Ihm SH, et al. Anti-cancer effect of R3V6 peptide-mediated delivery of an anti-microRNA-21 antisense-oligodeoxynucleotide in a glioblastoma animal model. J Drug Target. 2017; 25: 132–9. 89. 89. Song H, Oh B, Choi M, Oh J, Lee M. Delivery of anti-microRNA-21 antisense-oligodeoxynucleotide using amphiphilic peptides for glioblastoma gene therapy. J Drug Target. 2015; 23: 360–70. 90. 90. Takei Y, Takigahira M, Mihara K, Tarumi Y, Yanagihara K. The metastasis-associated microRNA miR-516a-3p is a novel therapeutic target for inhibiting peritoneal dissemination of human scirrhous gastric cancer. Cancer Res. 2011; 71: 1442–53. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NjoiY2FucmVzIjtzOjU6InJlc2lkIjtzOjk6IjcxLzQvMTQ0MiI7czo0OiJhdG9tIjtzOjE4OiIvY2JtLzE5LzMvMjg5LmF0b20iO31zOjg6ImZyYWdtZW50IjtzOjA6IiI7fQ==) 91. 91. Takeshita F, Patrawala L, Osaki M, Takahashi RU, Yamamoto Y, Kosaka N, et al. Systemic delivery of synthetic microRNA-16 inhibits the growth of metastatic prostate tumors via downregulation of multiple cell-cycle genes. Mol Ther. 2010; 18: 181–7. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/mt.2009.207&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=19738602&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000274447200025&link_type=ISI) 92. 92. Tazawa H, Tsuchiya N, Izumiya M, Nakagama H. Tumor-suppressive miR-34a induces senescence-like growth arrest through modulation of the E2F pathway in human colon cancer cells. Proc Natl Acad Sci U S A. 2007; 104: 15472–7. [Abstract/FREE Full Text](http://www.cancerbiomed.org/lookup/ijlink/YTozOntzOjQ6InBhdGgiO3M6MTQ6Ii9sb29rdXAvaWpsaW5rIjtzOjU6InF1ZXJ5IjthOjQ6e3M6ODoibGlua1R5cGUiO3M6NDoiQUJTVCI7czoxMToiam91cm5hbENvZGUiO3M6NDoicG5hcyI7czo1OiJyZXNpZCI7czoxMjoiMTA0LzM5LzE1NDcyIjtzOjQ6ImF0b20iO3M6MTg6Ii9jYm0vMTkvMy8yODkuYXRvbSI7fXM6ODoiZnJhZ21lbnQiO3M6MDoiIjt9) 93. 93. Ben-Shushan D, Markovsky E, Gibori H, Triam G, Scomparin A, Satchi-Fainaro R. Overcoming obstacles in microRNA delivery towards improved cancer therapy. Drug Deliv Transl Res. 2014; 4: 38–49. 94. 94. Di Pietro P, Strano G, Zuccarello L, Satriano C. Gold and silver nanoparticles for applications in theranostics. Curr Top Med Chem. 2016; 16: 3069–102. 95. 95. Vallet-Regí M, Colilla M, Izquierdo-Barba I, Manzano M. Mesoporous silica nanoparticles for drug delivery: current insights. Molecules. 2017; 23: 47. 96. 96. El-Boubbou K. Magnetic iron oxide nanoparticles as drug carriers: clinical relevance. Nanomedicine. 2018; 13: 953–71. 97. 97. Zhao MX, Zhu BJ. The research and applications of quantum dots as nano-carriers for targeted drug delivery and cancer therapy. Nanoscale Res Lett. 2016; 11: 207. 98. 98. Li JM, Zhao MX, Su H, Wang YY, Tan CP, Ji LN, et al. Multifunctional quantum-dot-based siRNA delivery for HPV18 E6 gene silence and intracellular imaging. Biomaterials. 2011; 32: 7978–87. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.biomaterials.2011.07.011&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=21784514&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000295072600024&link_type=ISI) 99. 99. Choi AO, Brown SE, Szyf M, Maysinger D. Quantum dot-induced epigenetic and genotoxic changes in human breast cancer cells. J Mol Med (Berl). 2008; 86: 291–302. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1007/s00109-007-0274-2&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=17965848&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 100.100. Dobrovolskaia MA, Mcneil SE. Immunological properties of engineered nanomaterials. Nat Nanotechnol. 2007; 2: 469–78. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/nnano.2007.223&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=18654343&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000249062900010&link_type=ISI) 101.101. Valentini F, Mari E, Zicari A, Calcaterra A, Talano M, Scioli MG, et al. Metal free graphene oxide (GO) nanosheets and pristine-single wall carbon nanotubes (p-SWCNTs) biocompatibility investigation: A comparative study in different human cell lines. Int J Mol Sci. 2018; 19: 1316. 102.102. Yang HW, Huang CY, Lin CW, Liu HL, Huang CW, Liao SS, et al. Gadolinium-functionalized nanographene oxide for combined drug and microRNA delivery and magnetic resonance imaging. Biomaterials. 2014; 35: 6534–42. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.biomaterials.2014.04.057&link_type=DOI) 103.103. Dahlin RL, Kasper FK, Mikos AG. Polymeric nanofibers in tissue engineering. Tissue Eng Part B Rev. 2011; 17: 349–64. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1089/ten.teb.2011.0238&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=21699434&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000286661600009&link_type=ISI) 104.104.1. 2. Estanqueiro MVH, 3. Lobo JMS, 4. Amaral H Grumezescu AM. Delivering miRNA modulators for cancer treatment. In: Estanqueiro MVH, Lobo JMS, Amaral H. Drug targeting and stimuli sensitive drug delivery systems. London: William Andrew Publishing. 2018: 517–65. 105.105.1. 2. Agrahari V, 3. Agrahari V, 4. Meng J, 5. Mitra AK Mitra AK, Cholkar K, Mandal A. Electrospun Nanofibers in Drug Delivery: Fabrication, Advances, and Biomedical Applications. In: Agrahari V, Agrahari V, Meng J, Mitra AK, editors. Emerging nanotechnologies for diagnostics, drug delivery and medical devices. Amsterdam: Elsevier; 2017. p. 189–215. 106.106. Shi X, Zhou W, Ma D, Ma Q, Bridges D, Ma Y, et al. Ectrospinning of nanofibers and their applications for energy devices. J Nanomat. 2015; 16: 1–21. 107.107. Sasikanth K, Nama S, Suresh S, Brahmaiah B. Nanofibers-a new trend in nano drug delivery systems. Pharma Innovation. 2013; 2: 118. 108.108. Orellana EA, Tenneti S, Rangasamy L, Lyle LT, Low PS, Kasinski AL. FolamiRs: ligand-targeted, vehicle-free delivery of microRNAs for the treatment of cancer. Sci Transl Med. 2017; 9: 1–10. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1126/scitranslmed.aai7514&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=28931657&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 109.109. He F, Wen N, Xiao D, Yan J, Xiong H, Cai S, et al. Aptamer-based targeted drug deliverysystems: Current potential and challenges. Curr Med Chem. 2020; 27: 2189–219. 110.110. Esposito CL, Catuogno S, de Franciscis V. Aptamer-mediated selective delivery of short RNA therapeutics in cancer cells. J RNAi Gene Silencing. 2014; 10: 500–6. [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=25414727&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 111.111. Yin H, Xiong G, Guo S, Xu C, Xu R, Guo P, et al. Delivery of anti-miRNA for triple-negative breast cancer therapy using RNA nanoparticles targeting sem cell marker CD133. Mol Ther. 2019; 27: 1252–61. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.ymthe.2019.04.018&link_type=DOI) 112.112. Guo S, Huang F, Guo P. Construction of folate-conjugated pRNA of bacteriophage phi29 DNA packaging motor for delivery of chimeric siRNA to nasopharyngeal carcinoma cells. Gene Ther. 2006; 13: 814–20. [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=16482206&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000237272600002&link_type=ISI) 113.113. Ye X, Liu Z, Hemida MG, Yang D. Targeted delivery of mutant tolerant anti-coxsackievirus artificial microRNAs using folate conjugated bacteriophage Phi29 pRNA. PLoS One. 2011; 6: e21215. 114.114.1. 2. Kovacevic AB Shegokar R. Lipid nanocarriers for delivery of poorly soluble and poorly permeable drugs. In: Kovacevic AB, editor. Nanopharmaceuticals. Amsterdam: Elsevier; 2020. p. 151–74. 115.115. Khalil IA, Younis MA, Kimura S, Harashima H. Lipid nanoparticles for cell-specific in vivo targeted delivery of nucleic acids. Biol Pharm Bull. 2020; 43: 584–95. 116.116. Costa PM, Cardoso AL, Custódia C, Cunha P, de Almedia LP, de Lima MC. MiRNA-21 silencing mediated by tumor-targeted nanoparticles combined with sunitinib: A new multimodal gene therapy approach for glioblastoma. J Control Release. 2015; 207: 31–9. 117.117. Schlich M, Palomba R, Costabile G, Mizrahy S, Pannuzzo M, Peer D, et al. Cytosolic delivery of nucleic acids: the case of ionizable lipid nanoparticles. Bioeng Transl Med. 2021; 6: e10213. 118.118. Wei P, Cornel EJ, Du J. Ultrasound-responsive polymer-based drug delivery systems. Drug Deliv Transl Res. 2021; 11: 1323–39. 119.119. Dzmitruk V, Apartsin E, Ihnatsyeu-Kachan A, Abashkin V, Scharbin D, Bryszewska M. Dendrimers show promise for siRNA and microRNA therapeutics. Pharmaceutics. 2018; 10: 126. 120.120. Yazdian-Robati R, Ramezani M, Jalalian SH, Abnous K, Taghdisi SM. Targeted delivery of epirubicin to cancer cells by polyvalent aptamer system in vitro and in vivo. Pharm Res. 2016; 33: 2289–97. 121.121. Hao Y, Kieft JS. Three-way junction conformation dictates self-association of phage packaging RNAs. RNA Biol. 2016; 13: 635–45. 122.122. Hao C, Li X, Tian C, Jiang W, Wang G, Mao C. Construction of RNA nanocages by re-engineering the packaging RNA of Phi29 bacteriophage. Nat Commun. 2014; 5: 3890. 123.123. Binzel DW, Shu Y, Li H, Sun M, Zhang Q, Shu D, et al. Specific delivery of miRNA for high efficient inhibition of prostate cancer by RNA nanotechnology. Mol Ther. 2016; 24: 1267–77. 124.124.1. 2. Mahato M, 3. Jayandharan GR, 4. Vemula PK Jayandharan GR. Viral-and non-viral-based hybrid vectors for gene therapy. In: Mahato M, Jayandharan GR, Vemula PK, editors. Gene and cell therapy: biology and applications. New York: Springer; 2018. p. 111–30. 125.125. Dasgupta I, Chatterjee A. Recent advances in miRNA delivery systems. Methods Protoc. 2021; 4: 10. 126.126. Frank AM, Weidner T, Brynza J, Uckert W, Buchholz CJ, Hartmann J. CD8-Specific designed ankyrin repeat proteins improve selective gene delivery into human and primate T lymphocytes. Hum Gene Ther. 2020; 31: 679–91. 127.127. Stein S, Ott MG, Schultze-Strasser S, Jauch A, Burwinkel B, Kinner A, et al. Genomic instability and myelodysplasia with monosomy 7 consequent to EVI1 activation after gene therapy for chronic granulomatous disease. Nat Med. 2010; 16: 198–204. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/nm.2088&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=20098431&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000274297000041&link_type=ISI) 128.128. Hacein-Bey-Abina S, von Kalle C, Schmidt M, Le Deist F, Wulffraat N, Mclntyre E, et al. A serious adverse event after successful gene therapy for X-linked severe combined immunodeficiency. N Engl J Med. 2003; 348: 255–6. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1056/NEJM200301163480314&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=12529469&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000180390600013&link_type=ISI) 129.129. Cavazzana-Calvo M, Payen E, Negre O, Wang G, Hehir K, Fusil F, et al. Transfusion independence and HMGA2 activation after gene therapy of human β-thalassaemia. Nature. 2010; 467: 318–22. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/nature09328&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=20844535&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000281824900037&link_type=ISI) 130.130. Zhang JH, Du AL, Wang L, Wang XY, Gao JH, Wang TY. Episomal lentiviral vector-mediated miR-145 overexpression inhibits proliferation and induces apoptosis of human esophageal carcinomas cells. Recent Pat Anticancer Drug Discov. 2016; 11: 453–60. 131.131. Aravalli RN. Development of microRNA therapeutics for hepatocellular carcinoma. Diagnostics (Basel). 2013; 3: 170–91. 132.132. Bofill-De Ros X, Gironella M, Fillat C. MiR-148a- and miR-216a-regulated oncolytic adenoviruses targeting pancreatic tumors attenuate tissue damage without perturbation of miRNA activity. Mol Ther. 2014; 22: 1665–77. 133.133. Kim J, Kim PH, Kim SW, Yun CO. Enhancing the therapeutic efficacy of adenovirus in combination with biomaterials. Biomaterials. 2012; 33: 1838–50. 134.134.1. 2. Shams S, 3. Silva EA Fernandes TG, Diogo MM, Cabral JMS. Bioengineering strategies for gene delivery. In: Shams S, Silva EA, editors. Engineering strategies for regenerative medicine. Academic Press; 2020. p. 107–48. 135.135. Aponte-Ubillus JJ, Barajas D, Peltier J, Bardliving C, Shamlou P, Gold D. Molecular design for recombinant adeno-associated virus (rAAV) vector production. Appl Microbiol Biotechnol. 2018; 102: 1045–54. 136.136. Latronico MV, Condorelli G. Therapeutic use of microRNAs in myocardial diseases. Curr Heart Fail Rep. 2011; 8: 193–7. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1007/s11897-011-0068-2&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=21713604&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 137.137. Brunetti-Pierri N. AAV vector-based gene therapy, progress and current challenges. In Safety and Efficacy of Gene-Based Therapeutics for Inherited Disorders. New York: Springer; 2017. p. 77–112. 138.138. Vannucci L, Lai M, Chiuppesi F, Ceccherini-Nelli L, Pistello M. Viral vectors: a look back and ahead on gene transfer technology. New Microbiol. 2013; 36: 1–22. [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=23435812&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 139.139. Bhere D, Tamura K, Wakimoto H, Choi SH, Purow B, Debatisse J, et al. MicroRNA-7 upregulates death receptor 5 and primes resistant brain tumors to caspase-mediated apoptosis. Neuro Oncol. 2018; 20: 215–24. 140.140.1. 2. Saha G, 3. Saudagar P, 4. Dubey VK Sonawane SH, Bhanvase BA, Sivakumar M. Virus-like particles: nano-carriers in targeted therapeutics. In: Saha G, Saudagar P, Dubey VK, editors. Encapsulation of active molecules and their delivery system. Amsterdam: Elsevier; 2020. p. 197. 141.141. Yao Y, Jia T, Pan Y, Gou H, Li Y, Sun Y, et al. Using a novel microRNA delivery system to inhibit osteoclastogenesis. Int J Mol Sci. 2015; 16: 8337–50. 142.142. Pushko P, Pumpens P, Grens E. Development of virus-like particle technology from small highly symmetric to large complex virus-like particle structures. Intervirology. 2013; 56: 141–65. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1159/000346773&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=23594863&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 143.143. Callanan J, Stockdale SR, Shkoporov A, Draper LA, Ross RP, Hill C. RNA phage biology in a metagenomic era. Viruses. 2018; 10: 386. 144.144. Pinto AM, Silva MD, Pastrana LM, Bañobre-López M, Sillankorva S. The clinical path to deliver encapsulated phages and lysins. FEMS Microbiol Rev. Published online first: March 30, 2021. DOI: 10.1093/femsre/fuab019. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1093/femsre/fuab019&link_type=DOI) 145.145.1. 2. Dashti NH, 3. Sainsbury F Gerrard JA, Domigan LJ. Virus-derived nanoparticles. In: Dashti NH, Sainsbury F, editors. Protein nanotechnology. New York: Springer; 2020. p. 149–62. 146.146. Pan Y, Zhang Y, Jia T, Zhang K, Li J, Wang L. Development of a microRNA delivery system based on bacteriophage MS2 virus-like particles. FEBS J. 2012; 279: 1198–208. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1111/j.1742-4658.2012.08512.x&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=22309233&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 147.147. Sun Y, Sun Y, Zhao R. Establishment of microRNA delivery system by PP7 bacteriophage-like particles carrying cell-penetrating peptide. J Biosci Bioeng. 2017; 124: 242–9. 148.148. Hoffmann DB, Böker KO, Schneider S, Eckermann-Felkl E, Schuder A, Komrakova M, et al. In vivo siRNA delivery using JC virus-like particles decreases the expression of RANKL in rats. Mol Ther Nucleic Acids. 2016; 5: e298. 149.149. Chao CN, Yang YH, Wu MS, Chou MC, Fang CY, Lin MC, et al. Gene therapy for human glioblastoma using neurotropic JC virus-like particles as a gene delivery vector. Sci Rep. 2018; 8: 2213. 150.150. Chen W, Kang T, Yuan R, Shao C, Jing S. Immunogenicity and protective potency of Norovirus GII.17 virus-like particle-based vaccine. Biotechnol Lett. 2020; 42: 1211–8. 151.151. Armstrong JK, Hempel G, Koling S, Chan LS, Fisher T, Meiselman HJ, et al. Antibody against poly(ethylene glycol) adversely affects PEG-asparaginase therapy in acute lymphoblastic leukemia patients. Cancer. 2007; 110: 103–11. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1002/cncr.22739&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=17516438&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000247384200013&link_type=ISI) 152.152. Galaway FA, Stockley PG. MS2 viruslike particles: a robust, semisynthetic targeted drug delivery platform. Mol Pharm. 2013; 10: 59–68. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1021/mp3003368&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=23110441&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 153.153. Lin Y, Lu Y, Li X. Biological characteristics of exosomes and genetically engineered exosomes for the targeted delivery of therapeutic agents. J Drug Target. 2020; 28: 129–41. 154.154. Sun Z, Shi K, Yang S, Liu J, Zhou Q, Wang G, et al. Effect of exosomal miRNA on cancer biology and clinical applications. Mol Cancer. 2018; 17: 147. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1186/s12943-018-0897-7&link_type=DOI) 155.155. del Pozo-Acebo L, Hazas MLL, Tomé-Carneiro J, Gil-Cabrerizo P, San-Cristobal R, Busto R, et al. Bovine milk-derived exosomes as a drug delivery vehicle for miRNA-based therapy. Int J Mol Sci. 2021; 22: 1105. 156.156. Shimbo K, Miyaki S, Ishitobi H, Kato Y, Kubo T, Shimose S, et al. Exosome-formed synthetic microRNA-143 is transferred to osteosarcoma cells and inhibits their migration. Biochem Biophys Res Commun. 2014; 445: 381–7. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.bbrc.2014.02.007&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=24525123&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 157.157. Ohno S, Takanashi M, Sudo K, Ueda S, Ishikawa A, Matsuyama N, et al. Systemically injected exosomes targeted to EGFR deliver antitumor microRNA to breast cancer cells. Mol Ther. 2013; 21: 185–91. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/mt.2012.180&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=23032975&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 158.158. Wang Y, Chen X, Tian B, Liu J, Yang L, Zeng L, et al. Nucleolin-targeted extracellular vesicles as a versatile platform for biologics delivery to breast cancer. Theranostics. 2017; 7: 1360–72. 159.159. Lu M, Xing H, Xun Z, Yang T, Ding P, Cai C, et al. Exosome-based small RNA delivery: progress and prospects. Asian J Pharm Sci. 2018; 13: 1–11. 160.160. Morishita M, Takahashi Y, Nishikawa M, Sano K, Kato K, Yamashita T, et al. Quantitative analysis of tissue distribution of the B16BL6-derived exosomes using a streptavidin-lactadherin fusion protein and iodine-125-labeled biotin derivative after intravenous injection in mice. J Pharm Sci. 2015; 104: 705–13. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1002/jps.24251&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=http://www.n&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 161.161. Takahashi Y, Nishikawa M, Shinotsuka H, Matsui Y, Ohara S, Imai T, et al. Visualization and in vivo tracking of the exosomes of murine melanoma B16-BL6 cells in mice after intravenous injection. J Biotechnol. 2013; 165: 77–84. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.jbiotec.2013.03.013&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=23562828&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) [Web of Science](http://www.cancerbiomed.org/lookup/external-ref?access_num=000319244000001&link_type=ISI) 162.162. Pittenger MF, Discher DE, Péault BM, Phinney DG, Hare JM, Caplan AI. Mesenchymal stem cell perspective: cell biology to clinical progress. NPJ Regen Med. 2019; 4: 22. 163.163.1. 2. Korsgren O, 3. Scholz H 1. 2. Orlando G, 3. Piemonti L, 4. Ricordi C, 5. Stratta RJ , eds. Cellular therapies in preclinical and clinical islet transplantation: Mesenchymal stem cells. In: Korsgren O, Scholz H, editors. Transplantation, Bioengineering, and Regeneration of the Endocrine Pancreas. Cambridge: Academic Press; 2020. p. 821–31. 164.164. Su Y, Zhang T, Huang T, Gao J. Current advances and challenges of mesenchymal stem cells-based drug delivery system and their improvements. Int J Pharm. 2021; 600: 120477. 165.165. Fan XL, Zhang Y, Li X, Fu QL. Mechanisms underlying the protective effects of mesenchymal stem cell-based therapy. Cell Mol Life Sci. 2020; 77: 2771–94. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1007/s00018-020-03454-6&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=http://www.n&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 166.166. Shah K. Mesenchymal stem cells engineered for cancer therapy. Adv Drug Deliv Rev. 2012; 64: 739–48. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/j.addr.2011.06.010&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=21740940&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 167.167. Munoz JL, Bliss SA, Greco SJ, Ramkissoon SH, Ligon KL, 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. 168.168. Lee HK, Finniss S, Cazacu S, Bucris E, Ziv-Av A, Xiang C, et al. Mesenchymal stem cells deliver synthetic microRNA mimics to glioma cells and glioma stem cells and inhibit their cell migration and self-renewal. Oncotarget. 2013; 4: 346–61. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.18632/oncotarget.868&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=23548312&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 169.169. Zhang CL, Huang T, Wu BL, He WX, Liu D. Stem cells in cancer therapy: opportunities and challenges. Oncotarget. 2017; 8: 75756–66. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.18632/oncotarget.20798&link_type=DOI) 170.170. Chu DT, Nguyen TT, Tien NLB, Tran DK, Jeong JH, Anh PG, et al. Recent progress of stem cell therapy in cancer treatment: molecular mechanisms and potential applications. Cells. 2020; 9: 563. 171.171. Bonneau E, Neveu B, Kostantin E, Tsongalis GJ, de Guire V. How close are miRNAs from clinical practice? A perspective on the diagnostic and therapeutic market. EJIFCC. 2019; 30: 114–27. [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=31263388&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 172.172. Baumann V, Winkler J. miRNA-based therapies: strategies and delivery platforms for oligonucleotide and non-oligonucleotide agents. Future Med Chem. 2014; 6: 1967–84. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.4155/fmc.14.116&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=25495987&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 173.173. Beg MS, Brenner AJ, Sachdev J, Borad M, Kang YK, Stoudemire J, et al. Phase I study of MRX34, a liposomal miR-34a mimic, administered twice weekly in patients with advanced solid tumors. Invest New Drugs. 2017; 35: 180–8. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1007/s10637-016-0407-y&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=27917453&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 174.174. Hong DS, Kang YK, Borad M, Sachdev J, Ejadi S, Lim HY, et al. Phase I study of MRX34, a liposomal miR-34a mimic, in patients with advanced solid tumours. Br J Cancer. 2020; 122: 1630–7. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1038/s41416-020-0802-1&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=http://www.n&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 175.175. Titze-De-Almeida R, David C, Titze-De-Almeida SS. The race of 10 synthetic RNAi-based drugs to the pharmaceutical market. Pharm Res. 2017; 34: 1339–63. 176.176. Kreth S, Hübner M, Hinske LC. MicroRNAs as clinical biomarkers and therapeutic tools in perioperative medicine. Anesth Analg. 2018; 126: 670–81. 177.177. Reid G, Johnson TG, van Zandwijk N. Manipulating microRNAs for the treatment of malignant pleural mesothelioma: Past, present and future. Front Oncol. 2020; 10: 105. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.3389/fonc.2020.00105&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=32117755&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 178.178. van Zandwijk N, Pavlakis N, Kao SC, Linton A, Boyer MJ, Clarke S, et al. Safety and activity of microRNA-loaded minicells in patients with recurrent malignant pleural mesothelioma: a first-in-man, phase I, open-label, dose-escalation study. Lancet Oncol. 2017; 18: 1386–96. [CrossRef](http://www.cancerbiomed.org/lookup/external-ref?access_num=10.1016/S1470-2045(17)30621-6&link_type=DOI) [PubMed](http://www.cancerbiomed.org/lookup/external-ref?access_num=28870611&link_type=MED&atom=%2Fcbm%2F19%2F3%2F289.atom) 179.179. Bajan S, Hutvagner G. RNA-based therapeutics: From antisense oligonucleotides to miRNAs. Cells. 2020; 9: 137. 180.180. Shah MY, Ferrajoli A, Sood AK, Lopez-Berestein G, Calin GA. MicroRNA therapeutics in cancer - an emerging concept. EBioMedicine. 2016; 12: 34–42. 181.181.1. 2. Souckova K, 3. Ivkovic TC, 4. Slaby O Faintuch J, Faintuch S. Non-coding RNA therapy in cancer. In: Souckova K, Ivkovic TC, Slaby O, editors. Precision medicine for investigators, practitioners and providers. Cambridge: Academic Press; 2020. p. 211–20. 182.182. Segal M, Biscans A, Gilles ME, Anastasiadou E, de Luca R, Lim J, et al. Hydrophobically modified let-7b miRNA enhances biodistribution to NSCLC and downregulates HMGA2 in vivo. Mol Ther Nucleic Acids. 2020; 19: 267–77. 183.183. Hossian A, Mackenzie GG. Mattheolabakis G. MiRNAs in gastrointestinal diseases: can we effectively deliver RNA-based therapeutics orally? Nanomedicine (Lond). 2019; 14: 2873–89. 184.184. Tao H, Zhao J, Liu T, Cai Y, Zhou X, Xing H, et al. Intranasal delivery of miR-146a mimics delayed seizure onset in the lithium-pilocarpine mouse model. Mediators Inflamm. 2017; 2017: 6512620. 185.185. Topcu B, Gultekinoglu M, Timur SS, Eroglu I, Ulubayram K, Eroglu H. Current approaches and future prospects of nanofibers: a special focus on antimicrobial drug delivery. J Drug Target. 2021; 29: 563–75. 186.186. Wang G, Hu W, Chen H, Shou X, Ye T, Xu Y. Cocktail strategy based on NK cell-derived exosomes and their biomimetic nanoparticles for dual tumor therapy. Cancers (Basel). 2019; 11: 1560.