Delivery Platforms for miRNA-Based Cancer Therapeutics: History
Please note this is an old version of this entry, which may differ significantly from the current revision.

Restoration of microRNA (miRNA) expression or downregulation of aberrantly expressed miRNAs using miRNA mimics or anti-miRNA inhibitors (anti-miRs/antimiRs), respectively, continues to show therapeutic potential for the treatment of cancer. Although the manipulation of miRNA expression presents a promising therapeutic strategy for cancer treatment, it is predominantly reliant on nucleic acid-based molecules for their application, which introduces an array of hurdles, with respect to in vivo delivery. Because naked nucleic acids are quickly degraded and/or removed from the body, they require delivery vectors that can help overcome the many barriers presented upon their administration into the bloodstream.

  • microRNA
  • miRNA therapeutics
  • miRNA delivery
  • cancer
  • cell-penetrating peptides
  • CPPs

1. Introduction

Some of the most significant barriers to the widespread use of miRNA replacement/inhibition therapies for the treatment of cancer are consequences of systemic delivery and limitations in tumor-tissue specificity. Thus, because the therapeutic manipulation of miRNA expression is predominantly reliant on nucleic acid-based molecules, delivery vectors are needed to help them overcome not only the systemic barriers, such as rapid renal clearance, but also clinical barriers, such as biosafety [1]. The following sections highlight the current state-of-the-art viral and nonviral delivery systems that have been demonstrated to deliver miRNA-based cancer therapies in vitro and in vivo, with a focus on nonviral nanotechnology-based miRNA delivery systems. The graphical depictions of these varied delivery systems, including their generalized modes of cell entry, as well as their strengths and weaknesses, can be found in Figure 1 and Table 1 below.
Figure 1. Graphical depictions highlighting the varied miRNA delivery platforms developed for cancer therapeutics, as well as their mechanisms of cellular internalization. Schematic diagram illustrating the generalized modes of cell entry for both viral and nonviral miRNA delivery systems. To cross the plasma membrane of the targeted cell, many of the delivery systems utilize multiple different cellular entry routes, but in general, utilize either direct cellular entry mechanisms or endocytosis-based uptake pathways [2][3]. For example, viral vectors, such as adenovirus (AV), adeno-associated virus (AAV), and lentivirus (LV), can utilize either direct entry mechanisms or endocytosis-based uptake pathways in their delivery of miRNA mimic-encoded RNA/DNA cargo into cells. Other examples, in terms of nonviral delivery systems, can include direct cytoplasmic entry via lipid fusion of a lipid-based vector with the plasma membrane or direct cellular entry of a polymer (e.g., polyamidoamine (PAMAM))-based vector in the delivery of antagomiR cargo. Additionally depicted are examples of cellular internalizations via endocytosis of inorganic (e.g. gold (Au)), extracellular vesicle (EV), and peptide-based vectors in complex with miRNA mimic cargo. Select images within the figure were acquired from BioRender.com. OncomiR, oncogenic miRNA; TS miRNA, tumor suppressor miRNA.
Table 1. The highlighted advantages and disadvantages of both viral and nonviral delivery systems in their transport of miRNA-based therapeutics for the treatment of cancer.

2. Viral Delivery

Virus-mediated delivery of miRNAs has been shown to be highly efficacious, where viral vectors can be designed to deliver miRNAs at different stages of biogenesis (i.e., pri-miRNAs and pre-miRNAs). Driven by a viral promotor, pri/pre-miRNA cloned within a plasmid can be transcribed and further processed to the mature miRNA form, enabling it to subsequently act on the target mRNA [4]. Adenoviruses (AdVs), particularly oncolytic AdVs (OAdVs), have been successful in the delivery of antimiRs and miRNA mimics. In fact, OAdV-mediated delivery of antimiRs in the form of long noncoding RNAs (lncRNAs), which has the therapeutic design advantage of targeting multiple copies of the same miRNA or different miRNAs within a single lncRNA molecule, has been shown to inhibit tumor growth in xenograft murine models of triple-negative breast cancer (TNBC) by simultaneously suppressing oncomiR levels of miR-9-5p, miR-10b-5p, miR-21-5p, miR-23a-3p, miR-29a-3p, miR-155-5p, miR-222-3p, miR-301a-3p, and miR-373-3p [5] and to decrease sorafenib resistance in sorafenib-resistant hepatocellular carcinoma (HCC) by concomitantly targeting miR-21, miR-153, miR-216a, miR-217, miR-494, and miR-10a-5p [6]. Virus-mediated overexpression of miRNAs can also have its advantages by suppressing particular oncogenes, as it has been shown that overexpression of miR-143 can inhibit cell growth and induce apoptosis by targeting KRAS in human colorectal cancer, in vitro [7]. Moreover, virus-mediated overexpression of miR-199 and miR-34a has been found to lead to control of tumor growth by targeting mTOR, c-MET, HIF-1α, and CD44, as well as complete tumor regression by targeting Bcl-2 and SIRT1 in xenograft murine models of HCC [8][9], respectively. Similarly to AdV, adeno-associated virus (AAV)-mediated delivery systems have also had success in the treatment of HCC, as overexpression of miR-342-3p and miR-26a has each demonstrated anti-tumor effects in murine models [10][11] by targeting MCT1 and cyclins D2 and E2, respectively. Additionally, Bhere and colleagues recently showed that AAV-mediated delivery of anti-miR-21 and miR-7 resulted in decreased cell proliferation, migration, and invasion of human prostate and colon cancer cells, in vitro, and a significant reduction in tumor burden in glioblastoma murine models [12] through targeting of the PI3K/AKT and JAK/STAT3 signaling pathways (via anti-miR-21) and down-regulation of EGFR and p-AKT (via miR-7). Lentivirus-mediated delivery of miR-15a and miR-16 in murine models of chronic lymphoid leukemia [13] and a miRNA sponge targeting miR-494 in murine models of breast cancer [14] have also resulted in beneficial therapeutic effects; however, concerns of lentiviral integration into the host genome have limited their clinical application as a delivery vector.

3. Nonviral Delivery

Despite being highly efficacious, virus-mediated miRNA-based therapeutic delivery platforms lack clinical desirability due to a number of biosafety concerns, including viral immunogenicity. Although classically thought of as inefficient compared to viral vectors, recent advancements in nonviral delivery platforms, however, are paving the way for nucleic acid-based therapies that make their application that much more feasible in the clinic [1]. As such, the following sections highlight and discuss the pros/cons of the varied nonviral delivery technologies that have been developed for miRNA-based cancer therapeutics, including polymer nanoparticles, lipid-based nanoparticles, inorganic nanoparticles, extracellular vesicles, and an emerging technology - peptide carriers.

3.1. Polymer Nanoparticles

Polymeric delivery systems have found success as suitable vectors for delivery of nucleic acids due to their high stability and flexibility, and the facile ability to make substitutions and/or additions of functional groups [15]. In fact, one particular polymer, poly lactic-co-gycolic acid, has gained FDA approval as a delivery vector and is in phase II clinical trials for the delivery of therapeutic small interfering RNA (siRNA) molecules (NCT01676259) [16]. The ability to control the molecular weight, polymer composition, and architecture of polymers allows for the manipulation of size, morphology, charge, pKa, membrane interactions, and biodegradability [17]. Moreover, polymer-based vectors are composed of a variety of materials, including natural polymers, such as collagen, gelatin, and chitosan, synthetic polymers, and combinations of natural and synthetic polymers. The following subsections pertaining to polymer nanoparticles detail polymer-based vectors that have found success in the delivery of specific miRNA-based cancer therapies.

Polyethylenimines

Polyethylenimines (PEI) are second-generation cationic polymers that are frequently utilized for therapeutic gene delivery [17]. PEI is composed of many positively charged amino groups, allowing for complexation with anionic RNA molecules and shielding from degradation, as well as enabling the proton sponge effect, which promotes escape from endosomes after endocytosis [18]. Indeed, PEI has been shown to be an effective delivery vector of miRNA for treatment of various cancers, including miR-33a and miR-145 for the treatment of colon cancer [19] and miR-708-5p mimics for the treatment of metastatic non-small cell lung cancer [20]. Although successful, PEI alone is not as desirable as other delivery vectors, due to the excess positive charge and low degradability due to the binding of serum proteins [4]. As such, PEI in combination/conjugation with other lipids and polymers has been investigated to mitigate these undesirable effects and has found some success. For example, PEI-polyethylene glycol (PEG)-mediated delivery of miR-34a and miR-150 has been shown to be effective for treatment of HCC [21] and to effectively reduce the cell viability of chronic myeloid leukemia cells [22], respectively. Additionally, PEI-hyaluronic acid (HA)-PEG-mediated delivery of a plasmid encoding miR-125b has been demonstrated to inhibit tumor growth and induce apoptosis in a murine lung cancer model [23]. Likewise, polyurethane-PEI-mediated delivery of a plasmid encoding miR-145 has shown success for the treatment of glioblastoma [24], and PEI-antagomiR-126 complexes, loaded into liposomes, have been demonstrated to effectively target leukemic stem cells in vivo for the treatment of acute myeloid leukemia [25].

Polyamidoamine

Polyamidoamine (PAMAM) is a hyperbranched synthetic polymer that is positively charged and biodegradable. While the overall positive charge of the polymer allows for complexation with nucleic acids, it promotes hepatic accumulation and toxicity [26]. Consequently, PAMAM is frequently found in combination/conjugation with other lipids and polymers. For instance, PAMAM-PEG-mediated delivery of a miR-34a-expressing plasmid has been shown to have anti-tumor effects in non-small cell lung cancer [27], and nanographene oxide (NGO)-PEG-PAMAM-mediated delivery of anti-miR-21 has proven effective in reducing migration and invasion of non-small-cell lung cancer A549 cells, in vitro [28].

Chitosan

Chitosan is a biocompatible, natural polysaccharide. It is a deacetylated derivative of chitin, which is found in the exoskeleton of insects, crustaceans, and fungi, making it the second most abundant natural polymer [29]. Chitosan consists of repeating units of β-1,4 linked N-acetyl-D-glucosamine and D-glucosamine [30], and has been described as having a profound binding affinity for miRNAs [31]. Chitosan has been shown to be an effective delivery vector in the treatment of breast cancer, through the complexation of miR-200c [32][33], and in the treatment of prostate bone metastasis, through the complexation of miR-34a mimics [34].

Poly Lactic-Co-Gycolic Acid

As previously mentioned, poly lactic-co-gycolic acid (PLGA) is an FDA-approved polymeric delivery vector. These polymers are polyesters and are negatively charged, biodegradable, and biocompatible. PLGA is also hydrophobic, which is thought to impair its miRNA delivery efficacy [35]. As a result, PLGA in combination/conjugation with various lipids and polymers, both synthetic and natural, have been investigated, with some combinations proving effective in mediating miRNA delivery for the treatment of various cancers. In particular, PLGA-chitosan complexes containing miR-34a mimics have been shown to inhibit tumor growth of multiple myeloma xenografts and resulted in the greater survival of treated NOD-SCID tumor-bearing mice [36]. Additionally, PLGA-HA-PEI in complex with a miR-145-encoding plasmid was shown to reduce tumor growth in a murine xenograft model of colon cancer [37], and PLGA-PEG-anti-miR-21, PLGA-PEG-anti-miR-10b/21, and PLGA-PEG-miR-122 complexes were demonstrated to be effective in treating HCC [38], breast cancer [39][40], and colon cancer [41], respectively.

3.2. Lipid-Based Nanoparticles

The ease of use and versatility of lipid-based nanoparticles in the form of liposomes have made them the most widely used nanoparticle for the delivery of nucleic acid-based therapies, which includes miRNAs. Liposomes are spherical structures with a hydrophilic core that is separated from the external environment by a phospholipid bilayer. Liposomes can accommodate hydrophobic molecules within the bilayer, hydrophilic molecules within the liposome core, and amphiphilic molecules at the interphase between the bilayer and core [42]. Due to their phospholipid composition, liposomes can interact with cell membranes, which leads to efficient delivery of cargo.

Cationic Liposomes

The first generation of liposomes were cationic in nature, which allowed for electrostatic interactions with nucleic acid-based cargos, as well as with the negatively charged surfaces of cells [43][44]. While advantageous for drug loading and delivery, this positive-charge property, however, was found to limit the cell specificity of cationic liposomes, allowed interactions with serum proteins, and increased the susceptibility of uptake by RES [45][46]. Despite these challenges within the circulatory system, a significant tumor reduction in a xenograft tumor mouse model of colorectal cancer was still observed through cationic liposome-based nanoparticles loaded with miR-139-5p mimics, albeit with the liposomes also possessing other functionalized moieties [47]. Decreased levels of miR-143 and miR-145, which are associated with colorectal carcinoma, and the delivery of miR-143 and miR-145 mimics using cationic liposomes to restore their levels have also been shown to reduce cell proliferation in a number of colorectal cancer cell lines [48]. In addition to colorectal cancer, cationic liposome-mediated delivery of a miR-7-encoding plasmid and miR-29 mimics have been demonstrated to significantly reduce tumor sizes in xenograft tumor mouse models of lung cancer [49][50].

Neutral Liposomes

To reduce the charge-associated shortcomings of cationic liposomes, neutral liposomes were developed by the inclusion of helper lipids, such as 1,2-dioleoyl-sn-glycero-3-phosphoethanolamine (DOPE) [51], PEG [52], phosphatidylcholines (PCs) [53], and cholesterol [54]. These modifications have led to reduced RES uptake, which allows for increased half-life of the neutral liposomes within the bloodstream [45][46]. Taking advantage of the helper lipid phosphocholine, an intermediate of PC, Trang and colleagues found that neutral liposome-mediated delivery of miR-34a and let-7b mimics resulted in the significant reduction in tumor burden in a K-Ras-activated autochthonous mouse model of non-small cell lung cancer (NSCLC) [55].

Ionizable Liposomes

Further optimization of lipid-based nanoparticles has resulted in the generation of ionizable liposomes, which are cationic at low pH and neutral/anionic at neutral or higher pH levels. The ability to change charge states with respect to extracellular pH gives ionizable liposomes enhanced cell selectively characteristics that makes them more clinically translatable. In fact, an ionizable liposome-miRNA complex (comprising miR-34 mimics; MRX34) made it as far as phase I clinical trials for treatment of liver cancer and metastasis (NCT01829971) [56][57]. However, this trial was stopped due to severe immune-related adverse events, which resulted in the death of four patients [57]. Despite this unfortunate setback, another group has shown that ionizable liposomes could still hold promise as delivery vehicles for miRNA therapeutics, as they have found, at least at the preclinical level, that ionizable liposome-mediated delivery of miR-200c mimics could result in enhanced radiosensitivity in a xenograft mouse model of lung cancer [58]. Additionally, another study found that delivery of a miR-199b-5p mimic using an ionizable cationic liposome [59] could significantly impair Hes-1 (a downstream effector of the canonical Notch and noncanonical SHH pathways) and cancer stem cell markers in a number of different tumorigenic cell lines, including colon (HT-29, CaCo-2, and SW480), breast (MDA-MB231T and MCF-7), prostate (PC-3), glioblastoma (U-87), and medulloblastoma (Daoy, ONS-76, and UW-228) cancer cell lines [60].

3.3. Inorganic Nanoparticles

Inorganic nanoparticles are desirable as delivery vectors because they can be designed to be biocompatible, nonimmunogenic, and nontoxic, and the size, shape, and porosity of particles can be controlled [31][61]. Nevertheless, the use of inorganic materials for delivery of miRNAs still faces challenges, such as protection from degradation in vivo, as well as endosomal escape [1]. The following subsections describe examples of common inorganic vectors utilized for delivery of miRNAs, however, the examples discussed herein are not a complete representation of all developed technologies.

Calcium Phosphate

Calcium phosphate (CaP) nanoparticles, composed of hydroxyapatite [Ca5(PO4)3OH], the inorganic component of bone and teeth, are described as being the most successful inorganic vectors for miRNA therapeutics [31], particularly for the treatment of colon/colorectal cancers. CaP owes this success to its unique in vivo characteristics, including its biocompatibility and biodegradability properties. Moreover, CaP’s susceptibility to acidic conditions allows for endocytic escape, where once these CaP nanoparticles are endocytosed, the acidic environment of the endosome dissolves them, resulting in subsequent increases in ionic strength that lead to osmotic swelling and the release of cargo [62]. As previously mentioned, CaP nanoparticle-mediated delivery of miRNAs has been particularly successful in the treatment of colon/colorectal cancers, as delivery of miR-4711-5p [63], miR-4689 [64], and miR-29b [65] mimics were found to effectively inhibit tumor growth in xenograft colon/colorectal cancer mouse models.

Silica

Silica-based nanoparticle technologies are desirable due to their biocompatibility, large surface area, well-defined chemical properties, and ability to control characteristics, such as pore structure [66]. One type of silica-based nanoparticle that has found success with miRNA delivery is the mesoporous silica nanoparticle (MSNP). MSNPs have a/an: (i) tunable particle size, which is important for endocytosis; (ii) stable and rigid framework, making them more resistant to heat, pH, mechanical stress, and hydrolysis-mediated degradation; (iii) uniform and tunable pore size, allowing for controlled drug loading; (iv) high surface area (>900 m2/g) and large pore volume (>0.9 cm3/g), which allows for increased drug loading; (v) interior and exterior surface, permitting selective functionalization of either surface; (vi) unique “honeycomb-like” porous structure, which aids in decreased premature drug release or leaking [67]. Taking advantage of these MSNP properties, Bertucci and colleagues successfully induced apoptosis in temozolomide (TMZ)-resistant T98G glioblastoma cells, in vitro, by loading the MSNPs with the anti-cancer drug TMZ and decorating them on the surface with a polyarginine-peptide nucleic acid (R8-PNA) antimiR conjugate designed to target miR-221, a miRNA, whose downregulation was previously reported to sensitize glioma cells to TMZ [68][69].

Gold

Gold (Au)-based nanoparticles (AuNPs) are well suited for delivery of nucleic acids, particularly after the addition of various functional groups. AuNPs have multifunctional monolayers, allowing for the addition of multiple functional moieties, which can control cytotoxicity, biodistribution, and excretion [70][71][72][73][74][75]. AuNPs can also be easily scaled with low size dispersity [76]. Due to these characteristics, AuNPs have found success in delivery of miRNAs for the treatment of various cancers. In particular, miR-375 mimic-coated AuNPs were observed to reduce tumor volume in primary and xenograft tumor mouse models of HCC [77]. Additionally, AuNPs formulated with PEG were found to mediate the highly efficient cell uptake of miRNAs and could decrease cell proliferation upon delivery of a miR-31 mimic into neuroblastoma (NGP and SH-SY5Y) and ovarian (OVCAR8 and HEYA8) cancer cell lines [78]. Moreover, Gilam and colleagues showed that in combination with the chemotherapy drug, cisplatin, AuNPs functionalized with PEG and a tumor-homing peptide, embedded within a hydrogel, could mediate the efficient local, selective, and sustained release of co-complexed miR-96 or miR-182 mimics, leading to the reduction in primary tumor size and metastasis in a breast cancer mouse model [79].

3.4. Extracellular Vesicles

Extracellular vesicles (EVs) are cell-derived nanovesicles that transport DNA, RNA, proteins, and lipids for cellular communication and activation of signaling pathways [80]. While EVs transport mRNA and other RNA species, such as lncRNAs and circular RNAs, miRNAs are perhaps the most abundant cargo molecule in EVs, particularly in exosomes [81]. In fact, these exosome-associated miRNAs have significant roles in the post-transcriptional regulation of gene expression and participate in the mediation of inflammatory reactions, cell migration, proliferation, apoptosis, autophagy, and epithelial-mesenchymal transition [81]. It, therefore, stands to reason that exploiting EVs for therapeutic miRNA delivery has its advantages over other delivery vectors. As a natural biomolecular carrier possessing specific ligands, EVs can be selectively delivered to cell types bearing specific surface receptors [82]. Additionally, EV ligands, such as CD47, can aid in their protection from phagocytes [83]. Moreover, lipid bilayer-encapsulated miRNA cargo is protected from RNase-mediated degradation, as well as from other circulatory system obstacles. This lipid bilayer also allows for direct fusion of the EV with the target cell membrane, with the subsequent release of cargo directly into the cytoplasm of the target cell, thus evading potential endosomal entrapment of the cargo; however, it should be noted that EVs can also undergo receptor-mediated endocytosis [84][85][86]. Though a relatively new delivery platform for therapeutic nucleic acid-based drugs, EV-mediated delivery of miRNAs has already shown promising therapeutic responses in various cancers. For example, EV-mediated delivery of a chemically modified miR-143 mimic, a plasmid expressing miR-146b, and a miR-145 mimic has each been observed to have therapeutic effects in colon cancer [87], glioma [88], and lung cancer [89], respectively. Other therapeutic uses of EVs have also been reported in the delivery of a plasmid expressing miR-122 in HCC [90] and a chemically synthesized miR-199a-3p mimic in ovarian cancer [91]. Similar to nonviral vectors discussed thus far, exosomes also have the advantage of being modified to contain different functional surface moieties. One interesting example of this type of modification was reported by Ohno and colleagues, where they engineered exosomes to express an epidermal growth factor receptor (EGFR)-specific targeting peptide, GE11, on their surfaces, which were then subsequently used to target a TS miRNA to EGFR-expressing breast cancer cells [92]. In particular, when GE11-positive exosomes containing the TS miRNA, let-7a, were administered to EGFR-expressing breast cancer xenograft tumor-bearing mice, these GE11-positive let-7a-loaded exosomes were observed to not only target the tumors, but also impair their development [92].

3.5. Peptides

The use of peptides as a delivery vector of nucleic acid-based therapeutics was initially described over two decades ago [93][94][95], but has only recently begun to gain popularity. Peptides are favorable delivery vectors because of the diversity of their physiochemical properties and functions [96]. The controlled ability to modify their amino acid sequences and ease of synthesis allows for the production of peptides that can overcome many of the systemic circulation-associated barriers faced by nucleic acid-based therapies. As such, many different classes of peptide carriers exist, one of which, the cell-penetrating peptides (CPPs), are proving to be highly efficacious and clinically translatable for the treatment of various cancers, as suggested by their presence in phase I and II clinical trials [97][98][99][100][101], as either a therapeutic agent alone or as a delivery agent for macromolecular therapeutics [102]. CPPs are typically 4–40 amino acid residues in length [103], can penetrate the plasma membrane of a cell and facilitate the delivery of different cargos [104], and are considered by some to be the most promising nonviral delivery platform for improvement of intracellular trafficking of nucleic acid-based cargos [105], which have included DNA, RNA, siRNA [106], and more recently, even those associated with miRNA-based therapeutics. For instance, one particular peptide carrier, named FA-R9-FPcas3, comprising a folate receptor-targeting ligand, folic acid (FA), a nona-arginine CPP (R9), and a Caspase-3-sensitive imaging probe (FPcas3), was used to form a multi-functional peptide-miRNA nanocomplex consisting of the miR-34a mimic that was capable of suppressing tumor growth upon tail vein injection into living mice bearing subcutaneous HeLa tumors [107]. Moreover, because molecular imaging is such a powerful tool for visualization and quantification of pathological processes, such as cancer, Yang and colleagues recently demonstrated that a CPP, PepFect6, could also be used in complex with a radioactively-labeled antisense miRNA oligonucleotide (AMO) designed to target the oncomiR, miR-21, to successfully image miR-21 expression in A549 lung adenocarcinoma xenografts, thus demonstrating a promising method for the noninvasive imaging of miRNA expression levels in vivo [108]. Additionally, although not examples of cancer-related applications, peptides such as LMWP and PepFect6 have also been shown to successfully deliver miRNA mimics, including miR-29b, to stem cells to promote osteoblastic differentiation [109] and miR-146a (a known anti-inflammatory miRNA) to inhibit inflammatory responses in a murine model of irritant contact dermatitis [110], respectively. Lastly, as noted above, it has also been reported that CPPs can be effective carriers of therapeutic small noncoding RNAs, particularly siRNAs, which are similar to miRNAs in function in that they both can induce post-transcriptional gene silencing, but differ in that siRNAs typically inhibit the expression of a single mRNA target, whereas miRNAs normally regulate the expression of multiple mRNA targets [111]. More specifically, researchers have demonstrated that a CPP, named 599, could enhance the intracellular delivery and bioavailability of siRNAs in oral cancer cells in vitro, as well as induce oncogene silencing upon intratumoral administration, resulting in significant inhibition of tumor growth in an orthotopic oral cancer mouse model [112][113]. In subsequent work, it was demonstrated that co-complexation of 599 with a cancer cell-targeting peptide could synergistically mediate the effective targeting/delivery of siRNAs to xenograft oral cancer tumors upon systemic administration and significantly enhance silencing of the targeted oncogene [114]. More recently, in an effort to improve upon the efficacy of siRNA uptake and gene silencing mediated by 599, it was further found that a 599 peptide variant, RD3AD, which exhibited enhanced siRNA uptake and gene silencing in comparison to 599, also directed siRNAs to specific cell-surface protrusions, identified as filopodia [115]. Intriguingly, filopodia are highly dynamic, elongated, and thin cellular processes that have been reported to facilitate the highly efficient cell entry of viruses, bacteria, activated receptors, lipo/polyplexes, and exosomes by mediating their retrograde transport and/or “surfing” along the structures toward the cell body [116][117][118][119][120], where, at the filopodial base, endocytic hotspots potentially allow for easier cell entry [117]. Of particular relevance regarding exosomes, which are known transporters of miRNAs [121], is that they can utilize filopodia to “surf” toward endocytic hotspots at the filopodial base, internalize, and then traffic within endosomes to the ER [116], which coincidentally is the central nucleation site of siRNA and miRNA-mediated RNA silencing [122][123]. Hence, one can envision how the targeting of filopodia and the subsequent directed transport of RD3AD-siRNA/miRNA complexes from the filopodia to the ER would potentially allow for a very efficient trafficking route of exogenous siRNAs/miRNAs into the cellular RNA silencing machinery. In fact, recent preliminary findings have found that complexation of RD3AD with a synthetic fluorescently-labeled let-7b miRNA duplex could similarly localize the miRNA mimic to filopodia and direct its entry into cancer cells (Figure 2). The significance is that one can, thus, potentially exploit filopodia-directed cell-entry machineries and subcellular-trafficking routes via CPPs for the development of more effective miRNA therapeutics.
Figure 2. RD3AD peptide-mediated localization of complexed let-7b miRNAs to filopodia and delivery into cancer cells. Confocal fluorescence microscopy analysis of CAL 27 oral cancer cells 2 hours post-treatment with a synthetic Cy3-labeled let-7b miRNA duplex (Cy3-let-7b; red) in complex with the RD3AD peptide. Filopodia (green) were stained with the F-actin label Alex Fluor 488 phalloidin, and nuclei (blue) were counterstained with DAPI. The merged images are also presented. Scale bar: 15 µm.

This entry is adapted from the peer-reviewed paper 10.3390/cells11152332

References

  1. Chen, Y.; Gao, D.-Y.; Huang, L. In vivo delivery of miRNAs for cancer therapy: Challenges and strategies. Adv. Drug Deliv. Rev. 2014, 81, 128–141.
  2. Kootstra, N.A.; Verma, I.M. Gene Therapy with Viral Vectors. Annu. Rev. Pharmacol. Toxicol. 2003, 43, 413–439.
  3. Donahue, N.D.; Acar, H.; Wilhelm, S. Concepts of nanoparticle cellular uptake, intracellular trafficking, and kinetics in nanomedicine. Adv. Drug Deliv. Rev. 2019, 143, 68–96.
  4. 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.
  5. 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.
  6. Tang, S.; Tan, G.; Jiang, X.; Han, P.; Zhai, B.; Dong, X.; Qiao, H.; Jiang, H.; Sun, X. An artificial lncRNA targeting multiple miRNAs overcomes sorafenib resistance in hepatocellular carcinoma cells. Oncotarget 2016, 7, 73257–73269.
  7. Luo, Q.; Basnet, S.; Dai, Z.; Li, S.; Zhang, Z.; Ge, H. A novel E1B55kDa-deleted oncolytic adenovirus carrying microRNA-143 exerts specific antitumor efficacy on colorectal cancer cells. Am. J. Transl. Res. 2016, 8, 3822–3830.
  8. Callegari, E.; Elamin, B.K.; D’Abundo, L.; Falzoni, S.; Donvito, G.; Moshiri, F.; Milazzo, M.; Altavilla, G.; Giacomelli, L.; Fornari, F.; et al. Anti-Tumor Activity of a miR-199-dependent Oncolytic Adenovirus. PLoS ONE 2013, 8, e73964.
  9. Lou, W.; Chen, Q.; Ma, L.; Liu, J.; Yang, Z.; Shen, J.; Cui, Y.; Bian, X.-W.; Qian, C. Oncolytic adenovirus co-expressing miRNA-34a and IL-24 induces superior antitumor activity in experimental tumor model. Klin. Wochenschr. 2013, 91, 715–725.
  10. Komoll, R.-M.; Hu, Q.; Olarewaju, O.; von Döhlen, L.; Yuan, Q.; Xie, Y.; Tsay, H.-C.; Daon, J.; Qin, R.; Manns, M.P.; et al. MicroRNA-342-3p is a potent tumour suppressor in hepatocellular carcinoma. J. Hepatol. 2020, 74, 122–134.
  11. Kota, J.; Chivukula, R.R.; O’Donnell, K.A.; Wentzel, E.A.; Montgomery, C.L.; Hwang, H.-W.; Chang, T.-C.; Vivekanandan, P.; Torbenson, M.; Clark, K.R.; et al. Therapeutic microRNA Delivery Suppresses Tumorigenesis in a Murine Liver Cancer Model. Cell 2009, 137, 1005–1017.
  12. 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, 1–10.
  13. Kasar, S.; Salerno, E.; Yuan, Y.; Underbayev, C.; Vollenweider, D.; Laurindo, M.F.; Fernandes, H.; Bonci, D.; Addario, A.; Mazzella, F.; et al. Systemic in vivo lentiviral delivery of miR-15a/16 reduces malignancy in the NZB de novo mouse model of chronic lymphocytic leukemia. Genes Immun. 2011, 13, 109–119.
  14. Liu, Y.; Lai, L.; Chen, Q.; Song, Y.; Xu, S.; Ma, F.; Wang, X.; Wang, J.; Yu, H.; Cao, X.; et al. MicroRNA-494 Is Required for the Accumulation and Functions of Tumor-Expanded Myeloid-Derived Suppressor Cells via Targeting of PTEN. J. Immunol. 2012, 188, 5500–5510.
  15. Fernandez, C.A.; Rice, K.G. Engineered Nanoscaled Polyplex Gene Delivery Systems. Mol. Pharm. 2009, 6, 1277–1289.
  16. Yin, H.; Kanasty, R.L.; Eltoukhy, A.A.; Vegas, A.J.; Dorkin, J.R.; Anderson, D.G. Non-viral vectors for gene-based therapy. Nat. Rev. Genet. 2014, 15, 541–555.
  17. Berg, A.I.v.D.; 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.
  18. Höbel, S.; Aigner, A. Polyethylenimines for siRNA and miRNA delivery in vivo. WIREs Nanomed. Nanobiotechnology 2013, 5, 484–501.
  19. Ibrahim, A.F.; Weirauch, U.; Thomas, M.; Grünweller, A.; Hartmann, R.K.; Aigner, A. MicroRNA Replacement Therapy for miR-145 and miR-33a Is Efficacious in a Model of Colon Carcinoma. Cancer Res. 2011, 71, 5214–5224.
  20. Wu, X.; Liu, T.; Fang, O.; Dong, W.; Zhang, F.; Leach, L.; Hu, X.; Luo, Z. MicroRNA-708-5p acts as a therapeutic agent against metastatic lung cancer. Oncotarget 2015, 7, 2417–2432.
  21. Hu, Q.; Wang, K.; Sun, X.; Li, Y.; Fu, Q.; Liang, T.; Tang, G. A redox-sensitive, oligopeptide-guided, self-assembling, and efficiency-enhanced (ROSE) system for functional delivery of microRNA therapeutics for treatment of hepatocellular carcinoma. Biomaterials 2016, 104, 192–200.
  22. Biray Avcı, Ç.; Özcan, İ.; Balcı, T.; Özer, Ö.; Gündüz, C. Design of polyethylene glycol-polyethylenimine nanocomplexes as non-viral carriers: Mir-150 delivery to chronic myeloid leukemia cells. Cell Biol. Int. 2013, 37, 1205–1214.
  23. Talekar, M.; Trivedi, M.; Shah, P.; Ouyang, Q.; Oka, A.; Gandham, S.; Amiji, M.M. Combination wt-p53 and MicroRNA-125b Transfection in a Genetically Engineered Lung Cancer Model Using Dual CD44/EGFR-targeting Nanoparticles. Mol. Ther. 2016, 24, 759–769.
  24. Yang, Y.-P.; Chien, Y.; Chiou, G.-Y.; Cherng, J.-Y.; Wang, M.-L.; Lo, W.-L.; Chang, Y.-L.; Huang, P.-I.; Chen, Y.-W.; Shih, Y.-H.; et al. Inhibition of cancer stem cell-like properties and reduced chemoradioresistance of glioblastoma using microRNA145 with cationic polyurethane-short branch PEI. Biomaterials 2012, 33, 1462–1476.
  25. Dorrance, A.M.; Neviani, P.; Ferenchak, G.J.; Huang, X.; Nicolet, D.; Maharry, K.S.; Ozer, H.G.; Hoellarbauer, P.; Khalife, J.; Hill, E.B.; et al. Targeting leukemia stem cells in vivo with antagomiR-126 nanoparticles in acute myeloid leukemia. Leukemia 2015, 29, 2143–2153.
  26. Duncan, R.; Izzo, L. Dendrimer biocompatibility and toxicity. Adv. Drug Deliv. Rev. 2005, 57, 2215–2237.
  27. Wang, H.; Zhao, X.; Guo, C.; Ren, D.; Zhao, Y.; Xiao, W.; Jiao, W. Aptamer-Dendrimer Bioconjugates for Targeted Delivery of miR-34a Expressing Plasmid and Antitumor Effects in Non-Small Cell Lung Cancer Cells. PLoS ONE 2015, 10, e0139136.
  28. Wang, F.; Zhang, B.; Zhou, L.; Shi, Y.; Li, Z.; Xia, Y.; Tian, J. Imaging Dendrimer-Grafted Graphene Oxide Mediated Anti-miR-21 Delivery With an Activatable Luciferase Reporter. ACS Appl. Mater. Interfaces 2016, 8, 9014–9021.
  29. Ravi Kumar, M.N.V. A review of chitin and chitosan applications. React. Funct. Polym. 2000, 46, 1–27.
  30. Martirosyan, A.; Olesen, M.J.; Howard, K.A. Chitosan-Based Nanoparticles for Mucosal Delivery of RNAi Therapeutics. Adv. Genet. 2014, 88, 325–352.
  31. Forterre, A.; Komuro, H.; Aminova, S.; Harada, M. A Comprehensive Review of Cancer MicroRNA Therapeutic Delivery Strategies. Cancers 2020, 12, 1852.
  32. Kaban, K.; Salva, E.; Akbuga, J. The effects of chitosan/miR-200c nanoplexes on different stages of cancers in breast cancer cell lines. Eur. J. Pharm. Sci. 2016, 95, 103–110.
  33. Kaban, K.; Salva, E.; Akbuga, J. In Vitro Dose Studies on Chitosan Nanoplexes for microRNA Delivery in Breast Cancer Cells. Nucleic Acid Ther. 2017, 27, 45–55.
  34. Gaur, S.; Wen, Y.; Song, J.H.; Parikh, N.U.; Mangala, L.S.; Blessing, A.M.; Ivan, C.; Wu, S.Y.; Varkaris, A.; Shi, Y.; et al. Chitosan nanoparticle-mediated delivery of miRNA-34a decreases prostate tumor growth in the bone and its expression induces non-canonical autophagy. Oncotarget 2015, 6, 29161–29177.
  35. Cheng, C.J.; Saltzman, W.M. Polymer Nanoparticle-Mediated Delivery of MicroRNA Inhibition and Alternative Splicing. Mol. Pharm. 2012, 9, 1481–1488.
  36. Cosco, D.; Cilurzo, F.; Maiuolo, J.; Federico, C.; Di Martino, M.T.; Cristiano, M.C.; Tassone, P.; Fresta, M.; Paolino, D. Delivery of miR-34a by chitosan/PLGA nanoplexes for the anticancer treatment of multiple myeloma. Sci. Rep. 2015, 5, 17579.
  37. Liang, G.; Zhu, Y.; Jing, A.; Wang, J.; Hu, F.; Feng, W.; Xiao, Z.; Chen, B. Cationic microRNA-delivering nanocarriers for efficient treatment of colon carcinoma in xenograft model. Gene Ther. 2016, 23, 829–838.
  38. Devulapally, R.; Foygel, K.; Sekar, T.V.; Willmann, J.K.; Paulmurugan, R. Gemcitabine and Antisense-microRNA Co-encapsulated PLGA–PEG Polymer Nanoparticles for Hepatocellular Carcinoma Therapy. ACS Appl. Mater. Interfaces 2016, 8, 33412–33422.
  39. Devulapally, R.; Sekar, N.M.; Sekar, T.V.; Foygel, K.; Massoud, T.F.; Willmann, J.K.; Paulmurugan, R. Polymer Nanoparticles Mediated Codelivery of AntimiR-10b and AntimiR-21 for Achieving Triple Negative Breast Cancer Therapy. ACS Nano 2015, 9, 2290–2302.
  40. Devulapally, R.; Sekar, T.V.; Paulmurugan, R. Formulation of Anti-miR-21 and 4-Hydroxytamoxifen Co-loaded Biodegradable Polymer Nanoparticles and Their Antiproliferative Effect on Breast Cancer Cells. Mol. Pharm. 2015, 12, 2080–2092.
  41. Wang, T.-Y.; Choe, J.W.; Pu, K.; Devulapally, R.; Bachawal, S.; Machtaler, S.; Chowdhury, S.M.; Luong, R.; Tian, L.; Khuri-Yakub, B.; et al. Ultrasound-guided delivery of microRNA loaded nanoparticles into cancer. J. Control. Release 2015, 203, 99–108.
  42. Ewert, K.K.; Scodeller, P.; Simón-Gracia, L.; Steffes, V.M.; Wonder, E.A.; Teesalu, T.; Safinya, C.R. Cationic Liposomes as Vectors for Nucleic Acid and Hydrophobic Drug Therapeutics. Pharmaceutics 2021, 13, 1365.
  43. Malone, R.W.; Felgner, P.L.; Verma, I.M. Cationic liposome-mediated RNA transfection. Proc. Natl. Acad. Sci. USA 1989, 86, 6077–6081.
  44. A Mislick, K.; Baldeschwieler, J.D. Evidence for the role of proteoglycans in cation-mediated gene transfer. Proc. Natl. Acad. Sci. USA 1996, 93, 12349–12354.
  45. Cheng, X.; Lee, R.J. The role of helper lipids in lipid nanoparticles (LNPs) designed for oligonucleotide delivery. Adv. Drug Deliv. Rev. 2016, 99, 129–137.
  46. Sun, X.; Yan, X.; Jacobson, O.; Sun, W.; Wang, Z.; Tong, X.; Xia, Y.; Ling, D.; Chen, X. Improved Tumor Uptake by Optimizing Liposome Based RES Blockade Strategy. Theranostics 2017, 7, 319–328.
  47. Zhao, Y.; Xu, J.; Le, V.M.; Gong, Q.; Li, S.; Gao, F.; Ni, L.; Liu, J.; Liang, X. EpCAM Aptamer-Functionalized Cationic Liposome-Based Nanoparticles Loaded with miR-139-5p for Targeted Therapy in Colorectal Cancer. Mol. Pharm. 2019, 16, 4696–4710.
  48. Su, J.; Liang, H.; Yao, W.; Wang, N.; Zhang, S.; Yan, X.; Feng, H.; Pang, W.; Wang, Y.; Wang, X.; et al. MiR-143 and MiR-145 Regulate IGF1R to Suppress Cell Proliferation in Colorectal Cancer. PLoS ONE 2014, 9, e114420.
  49. Rai, K.; Takigawa, N.; Ito, S.; Kashihara, H.; Ichihara, E.; Yasuda, T.; Shimizu, K.; Tanimoto, M.; Kiura, K. Liposomal Delivery of MicroRNA-7–Expressing Plasmid Overcomes Epidermal Growth Factor Receptor Tyrosine Kinase Inhibitor-Resistance in Lung Cancer Cells. Mol. Cancer Ther. 2011, 10, 1720–1727.
  50. Wu, Y.; Crawford, M.; Mao, Y.; Lee, R.J.; Davis, I.C.; Elton, T.S.; Lee, L.J.; Nana-Sinkam, S.P. Therapeutic Delivery of MicroRNA-29b by Cationic Lipoplexes for Lung Cancer. Mol. Ther.-Nucleic Acids 2013, 2, e84.
  51. Hattori, Y.; Suzuki, S.; Kawakami, S.; Yamashita, F.; Hashida, M. The role of dioleoylphosphatidylethanolamine (DOPE) in targeted gene delivery with mannosylated cationic liposomes via intravenous route. J. Control. Release 2005, 108, 484–495.
  52. Jokerst, J.V.; Lobovkina, T.; Zare, R.N.; Gambhir, S.S. Nanoparticle PEGylation for imaging and therapy. Nanomedicine 2011, 6, 715–728.
  53. Thewalt, J.L.; Bloom, M. Phosphatidylcholine: Cholesterol phase diagrams. Biophys. J. 1992, 63, 1176–1181.
  54. Semple, S.C.; Chonn, A.; Cullis, P.R. Influence of Cholesterol on the Association of Plasma Proteins with Liposomes. Biochemistry 1996, 35, 2521–2525.
  55. Trang, P.; Wiggins, J.F.; Daige, C.L.; Cho, C.; Omotola, M.; Brown, D.; Weidhaas, J.B.; Bader, A.G.; Slack, F.J. Systemic Delivery of Tumor Suppressor microRNA Mimics Using a Neutral Lipid Emulsion Inhibits Lung Tumors in Mice. Mol. Ther. 2011, 19, 1116–1122.
  56. 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 2016, 35, 180–188.
  57. 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.
  58. Cortez, M.A.; Valdecanas, D.; Zhang, X.; Zhan, Y.; Bhardwaj, V.; Calin, G.A.; Komaki, R.; Giri, D.K.; Quini, C.C.; Wolfe, T.; et al. Therapeutic Delivery of miR-200c Enhances Radiosensitivity in Lung Cancer. Mol. Ther. J. Am. Soc. Gene Ther. 2014, 22, 1494–1503.
  59. Dasgupta, I.; Chatterjee, A. Recent Advances in miRNA Delivery Systems. Methods Protoc. 2021, 4, 10.
  60. 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. fur Exp. Pathol. und Pharmakol. 2013, 386, 287–302.
  61. Labatut, A.E.; Mattheolabakis, G. Non-viral based miR delivery and recent developments. Eur. J. Pharm. Biopharm. 2018, 128, 82–90.
  62. Sokolova, V.; Epple, M. Inorganic Nanoparticles as Carriers of Nucleic Acids into Cells. Angew. Chem. Int. Ed. 2008, 47, 1382–1395.
  63. Morimoto, Y.; Mizushima, T.; Wu, X.; Okuzaki, D.; Yokoyama, Y.; Inoue, A.; Hata, T.; Hirose, H.; Qian, Y.; Wang, J.; et al. miR-4711-5p regulates cancer stemness and cell cycle progression via KLF5, MDM2 and TFDP1 in colon cancer cells. Br. J. Cancer 2020, 122, 1037–1049.
  64. Hiraki, M.; Nishimura, J.; Takahashi, H.; Wu, X.; Takahashi, Y.; Miyo, M.; Nishida, N.; Uemura, M.; Hata, T.; Takemasa, I.; et al. Concurrent Targeting of KRAS and AKT by MiR-4689 Is a Novel Treatment Against Mutant KRAS Colorectal Cancer. Mol. Ther.-Nucleic Acids 2015, 4, e231.
  65. Inoue, A.; Mizushima, T.; Wu, X.; Okuzaki, D.; Kambara, N.; Ishikawa, S.; Wang, J.; Qian, Y.; Hirose, H.; Yokoyama, Y.; et al. A miR-29b Byproduct Sequence Exhibits Potent Tumor-Suppressive Activities via Inhibition of NF-κB Signaling in KRAS-Mutant Colon Cancer Cells. Mol. Cancer Ther. 2018, 17, 977–987.
  66. Mamaeva, V.; Sahlgren, C.; Lindén, M. Mesoporous silica nanoparticles in medicine—Recent advances. Adv. Drug Deliv. Rev. 2013, 65, 689–702.
  67. Slowing, I.I.; Vivero-Escoto, J.L.; Wu, C.-W.; Lin, V.S.-Y. Mesoporous silica nanoparticles as controlled release drug delivery and gene transfection carriers. Adv. Drug Deliv. Rev. 2008, 60, 1278–1288.
  68. Bertucci, A.; Prasetyanto, E.A.; Septiadi, D.; Manicardi, A.; Brognara, E.; Gambari, R.; Corradini, R.; De Cola, L. Combined Delivery of Temozolomide and Anti-miR221 PNA Using Mesoporous Silica Nanoparticles Induces Apoptosis in Resistant Glioma Cells. Small 2015, 11, 5687–5695.
  69. Jiang, C.; Chen, L.; Zhang, J.; Han, L.; Zhang, A.; Zhang, C.; Zheng, Y.; Jiang, T.; Pu, P.; Kang, C. Downregulation of miR-221/222 sensitizes glioma cells to temozolomide by regulating apoptosis independently of p53 status. Oncol. Rep. 2011, 27, 854–860.
  70. Bowman, M.-C.; Ballard, T.E.; Ackerson, C.J.; Feldheim, D.L.; Margolis, D.M.; Melander, C. Inhibition of HIV Fusion with Multivalent Gold Nanoparticles. J. Am. Chem. Soc. 2008, 130, 6896–6897.
  71. Ryan, J.A.; Overton, K.W.; Speight, M.E.; Oldenburg, C.N.; Loo, L.; Robarge, W.; Franzen, S.; Feldheim, D.L. Cellular Uptake of Gold Nanoparticles Passivated with BSA−SV40 Large T Antigen Conjugates. Anal. Chem. 2007, 79, 9150–9159.
  72. Kim, S.T.; Saha, K.; Kim, C.; Rotello, V.M. The Role of Surface Functionality in Determining Nanoparticle Cytotoxicity. Accounts Chem. Res. 2013, 46, 681–691.
  73. Chompoosor, A.; Saha, K.; Ghosh, P.S.; Macarthy, D.J.; Miranda, O.R.; Zhu, Z.-J.; Arcaro, K.F.; Rotello, V.M. The Role of Surface Functionality on Acute Cytotoxicity, ROS Generation and DNA Damage by Cationic Gold Nanoparticles. Small 2010, 6, 2246–2249.
  74. De Jong, W.H.; Hagens, W.I.; Krystek, P.; Burger, M.C.; Sips, A.J.; Geertsma, R.E. Particle size-dependent organ distribution of gold nanoparticles after intravenous administration. Biomaterials 2008, 29, 1912–1919.
  75. Zhu, M.; Perrett, S.; Nie, G. Understanding the Particokinetics of Engineered Nanomaterials for Safe and Effective Therapeutic Applications. Small 2012, 9, 1619–1634.
  76. Daniel, M.-C.; Astruc, D. Gold Nanoparticles: Assembly, Supramolecular Chemistry, Quantum-Size-Related Properties, and Applications toward Biology, Catalysis, and Nanotechnology. Chem. Rev. 2003, 104, 293–346.
  77. Xue, H.-Y.; Liu, Y.; Liao, J.-Z.; Lin, J.-S.; Li, B.; Yuan, W.-G.; Lee, R.J.; Li, L.; Xu, C.-R.; He, X.-X. Gold nanoparticles delivered miR-375 for treatment of hepatocellular carcinoma. Oncotarget 2016, 7, 86675–86686.
  78. Ghosh, R.; Singh, L.C.; Shohet, J.M.; Gunaratne, P.H. A gold nanoparticle platform for the delivery of functional microRNAs into cancer cells. Biomaterials 2012, 34, 807–816.
  79. Gilam, A.; Conde, J.; Weissglas-Volkov, D.; Oliva, N.; Friedman, E.; Artzi, N.; Shomron, N. Local microRNA delivery targets Palladin and prevents metastatic breast cancer. Nat. Commun. 2016, 7, 12868.
  80. Tkach, M.; Théry, C. Communication by Extracellular Vesicles: Where We Are and Where We Need to Go. Cell 2016, 164, 1226–1232.
  81. Zheng, D.; Huo, M.; Li, B.; Wang, W.; Piao, H.; Wang, Y.; Zhu, Z.; Li, D.; Wang, T.; Liu, K. The Role of Exosomes and Exosomal MicroRNA in Cardiovascular Disease. Front. Cell Dev. Biol. 2021, 8.
  82. Valadi, H.; Ekström, K.; Bossios, A.; Sjöstrand, M.; Lee, J.J.; Lötvall, 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.
  83. De Jong, O.G.; Kooijmans, S.A.A.; Murphy, D.E.; Jiang, L.; Evers, M.J.W.; Sluijter, J.P.G.; Vader, P.; Schiffelers, R.M. Drug Delivery with Extracellular Vesicles: From Imagination to Innovation. Acc. Chem. Res. 2019, 52, 1761–1770.
  84. Théry, C.; Ostrowski, M.; Segura, E. Membrane vesicles as conveyors of immune responses. Nat. Rev. Immunol. 2009, 9, 581–593.
  85. Cocucci, E.; Racchetti, G.; Meldolesi, J. Shedding microvesicles: Artefacts no more. Trends Cell Biol. 2009, 19, 43–51.
  86. Escrevente, C.; Keller, S.; Altevogt, P.; Costa, J. Interaction and uptake of exosomes by ovarian cancer cells. BMC Cancer 2011, 11, 108.
  87. Akao, Y.; Iio, A.; Itoh, T.; Noguchi, S.; Itoh, Y.; Ohtsuki, Y.; Naoe, T. Microvesicle-mediated RNA Molecule Delivery System Using Monocytes/Macrophages. Mol. Ther. 2011, 19, 395–399.
  88. Katakowski, M.; Buller, B.; Zheng, X.; Lu, Y.; Rogers, T.; Osobamiro, O.; Shu, W.; Jiang, F.; Chopp, M. Exosomes from marrow stromal cells expressing miR-146b inhibit glioma growth. Cancer Lett. 2013, 335, 201–204.
  89. Vázquez-Ríos, A.J.; Molina-Crespo, Á.; Bouzo, B.L.; López-López, R.; Moreno-Bueno, G.; De La Fuente, M. Exosome-mimetic nanoplatforms for targeted cancer drug delivery. J. Nanobiotechnol. 2019, 17, 85.
  90. Lou, G.; Song, X.; Yang, F.; Wu, S.; Wang, J.; Chen, Z.; Liu, Y. Exosomes derived from miR-122-modified adipose tissue-derived MSCs increase chemosensitivity of hepatocellular carcinoma. J. Hematol. Oncol. 2015, 8, 122.
  91. Kobayashi, M.; Sawada, K.; Miyamoto, M.; Shimizu, A.; Yamamoto, M.; Kinose, Y.; Nakamura, K.; Kawano, M.; Kodama, M.; Hashimoto, K.; et al. Exploring the potential of engineered exosomes as delivery systems for tumor-suppressor microRNA replacement therapy in ovarian cancer. Biochem. Biophys. Res. Commun. 2020, 527, 153–161.
  92. Ohno, S.-I.; Takanashi, M.; Sudo, K.; Ueda, S.; Ishikawa, A.; Matsuyama, N.; Fujita, K.; Mizutani, T.; Ohgi, T.; Ochiya, T.; et al. Systemically Injected Exosomes Targeted to EGFR Deliver Antitumor MicroRNA to Breast Cancer Cells. Mol. Ther. 2013, 21, 185–191.
  93. Langel, Ü. Cell-Penetrating Peptides; Springer: New York, NY, USA, 2015; Volume 1324.
  94. Ramsey, J.D.; Flynn, N.H. Cell-penetrating peptides transport therapeutics into cells. Pharmacol. Ther. 2015, 154, 78–86.
  95. Kauffman, W.B.; Fuselier, T.; He, J.; Wimley, W.C. Mechanism Matters: A Taxonomy of Cell Penetrating Peptides. Trends Biochem. Sci. 2015, 40, 749–764.
  96. Cummings, J.C.; Zhang, H.; Jakymiw, A. Peptide carriers to the rescue: Overcoming the barriers to siRNA delivery for cancer treatment. Transl. Res. 2019, 214, 92–104.
  97. Meric-Bernstam, F.; Saleh, M.N.; Infante, J.R.; Goel, S.; Falchook, G.S.; Shapiro, G.; Chung, K.Y.; Conry, R.M.; Hong, D.S.; Wang, J.S.-Z.; et al. Phase I trial of a novel stapled peptide ALRN-6924 disrupting MDMX- and MDM2-mediated inhibition of WT p53 in patients with solid tumors and lymphomas. J. Clin. Oncol. 2017, 35, 2505.
  98. Søgaard, C.K.; Blindheim, A.; Røst, L.M.; Petrović, V.; Nepal, A.; Bachke, S.; Liabakk, N.-B.; Gederaas, O.A.; Viset, T.; Arum, C.-J.; et al. “Two hits - one stone”; increased efficacy of cisplatin-based therapies by targeting PCNA’s role in both DNA repair and cellular signaling. Oncotarget 2018, 9, 32448–32465.
  99. Arrouss, I.; Nemati, F.; Roncal, F.; Wislez, M.; Dorgham, K.; Vallerand, D.; Rabbe, N.; Karboul, N.; Carlotti, F.; Bravo, J.; et al. Specific Targeting of Caspase-9/PP2A Interaction as Potential New Anti-Cancer Therapy. PLoS ONE 2013, 8, e60816.
  100. Lulla, R.R.; Goldman, S.; Yamada, T.; Beattie, C.W.; Bressler, L.; Pacini, M.; Pollack, I.F.; Fisher, P.G.; Packer, R.J.; Dunkel, I.J.; et al. Phase I trial of p28 (NSC745104), a non-HDM2-mediated peptide inhibitor of p53 ubiquitination in pediatric patients with recurrent or progressive central nervous system tumors: A Pediatric Brain Tumor Consortium Study. Neuro-Oncology 2016, 18, 1319–1325.
  101. Bennett, G.; Lutz, R.; Park, P.; Harrison, H.; Lee, K. Abstract 1167: Development of BT1718, a novel Bicycle Drug Conjugate for the treatment of lung cancer. Cancer Res. 2017, 77, 1167.
  102. Kurrikoff, K.; Vunk, B.; Langel, Ü. Status update in the use of cell-penetrating peptides for the delivery of macromolecular therapeutics. Expert Opin. Biol. Ther. 2020, 21, 361–370.
  103. Srimanee, A.; Arvanitidou, M.; Kim, K.; Hällbrink, M.; Langel, Ü. Cell-penetrating peptides for siRNA delivery to glioblastomas. Peptides 2018, 104, 62–69.
  104. Lehto, T.; Ezzat, K.; Wood, M.J.A.; EL Andaloussi, S. Peptides for nucleic acid delivery. Adv. Drug Deliv. Rev. 2016, 106, 172–182.
  105. Roberts, T.C.; Langer, R.; Wood, M.J.A. Advances in oligonucleotide drug delivery. Nat. Rev. Drug Discov. 2020, 19, 673–694.
  106. Guidotti, G.; Brambilla, L.; Rossi, D. Cell-Penetrating Peptides: From Basic Research to Clinics. Trends Pharmacol. Sci. 2017, 38, 406–424.
  107. Xiao, X.; Wang, X.; Wang, Y.; Yu, T.; Huang, L.; Chen, L.; Li, J.; Zhang, C.; Zhang, Y. Multi-Functional Peptide-MicroRNA Nanocomplex for Targeted MicroRNA Delivery and Function Imaging. Chem.–A Eur. J. 2018, 24, 2277–2285.
  108. Yang, G.; Zhao, Y.; Gong, A.; Miao, W.; Yan, L.; Nie, P.; Wang, Z. Improved Cellular Delivery of Antisense Oligonucleotide for miRNA-21 Imaging In Vivo Using Cell-Penetrating Peptide-Based Nanoprobes. Mol. Pharm. 2021, 18, 787–795.
  109. Suh, J.S.; Lee, J.Y.; Choi, Y.S.; Chong, P.C.; Park, Y.J. Peptide-mediated intracellular delivery of miRNA-29b for osteogenic stem cell differentiation. Biomaterials 2013, 34, 4347–4359.
  110. Urgard, E.; Lorents, A.; Klaas, M.; Padari, K.; Viil, J.; Runnel, T.; Langel, K.; Kingo, K.; Tkaczyk, E.; Langel, Ü.; et al. Pre-administration of PepFect6-microRNA-146a nanocomplexes inhibits inflammatory responses in keratinocytes and in a mouse model of irritant contact dermatitis. J. Control. Release 2016, 235, 195–204.
  111. Lam, J.K.W.; Chow, M.Y.T.; Zhang, Y.; Leung, S.W.S. siRNA Versus miRNA as Therapeutics for Gene Silencing. Mol. Ther.-Nucleic Acids 2015, 4, e252.
  112. Alexander-Bryant, A.A.; Dumitriu, A.; Attaway, C.C.; Yu, H.; Jakymiw, A. Fusogenic-oligoarginine peptide-mediated silencing of the CIP2A oncogene suppresses oral cancer tumor growth in vivo. J. Control. Release 2015, 218, 72–81.
  113. Cantini, L.; Attaway, C.C.; Butler, B.; Andino, L.M.; Sokolosky, M.L.; Jakymiw, A. Fusogenic-Oligoarginine Peptide-Mediated Delivery of siRNAs Targeting the CIP2A Oncogene into Oral Cancer Cells. PLoS ONE 2013, 8, e73348.
  114. Alexander-Bryant, A.A.; Zhang, H.; Attaway, C.C.; Pugh, W.; Eggart, L.; Sansevere, R.M.; Andino, L.M.; Dinh, L.; Cantini, L.P.; Jakymiw, A. Dual peptide-mediated targeted delivery of bioactive siRNAs to oral cancer cells in vivo. Oral Oncol. 2017, 72, 123–131.
  115. Holjencin, C.E.; Feinberg, C.R.; Hedrick, T.; Halsey, G.; Williams, R.D.; Patel, P.V.; Biles, E.; Cummings, J.C.; Wagner, C.; Vyavahare, N.; et al. Advancing peptide siRNA-carrier designs through L/D-amino acid stereochemical modifications to enhance gene silencing. Mol. Ther.-Nucleic Acids 2021, 24, 462–476.
  116. Heusermann, W.; Hean, J.; Trojer, D.; Steib, E.; Von Bueren, S.; Graff-Meyer, A.; Genoud, C.; Martin, K.; Pizzato, N.; Voshol, J.; et al. Exosomes surf on filopodia to enter cells at endocytic hot spots, traffic within endosomes, and are targeted to the ER. J. Cell Biol. 2016, 213, 173–184.
  117. Lehmann, M.J.; Sherer, N.M.; Marks, C.B.; Pypaert, M.; Mothes, W. Actin- and myosin-driven movement of viruses along filopodia precedes their entry into cells. J. Cell Biol. 2005, 170, 317–325.
  118. Lidke, D.S.; Lidke, K.A.; Rieger, B.; Jovin, T.M.; Arndt-Jovin, D.J. Reaching out for signals. J. Cell Biol. 2005, 170, 619–626.
  119. Rehman, Z.U.; Sjollema, K.A.; Kuipers, J.; Hoekstra, D.; Zuhorn, I.S. Nonviral Gene Delivery Vectors Use Syndecan-Dependent Transport Mechanisms in Filopodia To Reach the Cell Surface. ACS Nano 2012, 6, 7521–7532.
  120. Romero, S.; Grompone, G.; Carayol, N.; Mounier, J.; Guadagnini, S.; Prevost, M.-C.; Sansonetti, P.J.; Van Nhieu, G.T. ATP-Mediated Erk1/2 Activation Stimulates Bacterial Capture by Filopodia, which Precedes Shigella Invasion of Epithelial Cells. Cell Host Microbe 2011, 9, 508–519.
  121. Rodriguez-Dorantes, M.; Romero-Cordoba, S.L.; Peralta-Zaragoza, O.; Salido-Guadarrama, I.; Hidalgo-Miranda, A. MicroRNAs transported by exosomes in body fluids as mediators of intercellular communication in cancer. OncoTargets Ther. 2014, 7, 1327–1338.
  122. Stalder, L.; Heusermann, W.; Sokol, L.; Trojer, D.; Wirz, J.; Hean, J.; Fritzsche, A.; Aeschimann, F.; Pfanzagl, V.; Basselet, P.; et al. The rough endoplasmatic reticulum is a central nucleation site of siRNA-mediated RNA silencing. EMBO J. 2013, 32, 1115–1127.
  123. Li, S.; Liu, L.; Zhuang, X.; Yu, Y.; Liu, X.; Cui, X.; Ji, L.; Pan, Z.; Cao, X.; Mo, B.; et al. MicroRNAs Inhibit the Translation of Target mRNAs on the Endoplasmic Reticulum in Arabidopsis. Cell 2013, 153, 562–574.
More
This entry is offline, you can click here to edit this entry!
Video Production Service