Aptamer-Assisted Delivery of Nucleotides with Tumor-Suppressing Properties for Targeted Cancer Therapies

authors:

avatar Pouya Safarzadeh Kozani ORCID 1 , 2 , avatar Pooria Safarzadeh Kozani ORCID 3 , avatar Fatemeh Rahbarizadeh ORCID 3 , 4 , *

Department of Medical Biotechnology, Faculty of Paramedicine, Guilan University of Medical Sciences, Rasht, Iran
Student Research Committee, Medical Biotechnology Research Center, School of Nursing, Midwifery, and Paramedicine, Guilan University of Medical Sciences, Rasht, Iran
Department of Medical Biotechnology, Faculty of Medical Sciences, Tarbiat Modares University, Tehran, Iran
Research and Development Center of Biotechnology, Tarbiat Modares University, Tehran, Iran

how to cite: Safarzadeh Kozani P , Safarzadeh Kozani P, Rahbarizadeh F. Aptamer-Assisted Delivery of Nucleotides with Tumor-Suppressing Properties for Targeted Cancer Therapies. J Adv Immunopharmacol. 2021;1(4):e114909. https://doi.org/10.5812/tms.114909.

Abstract

Targeted cancer therapy is developing rapidly according to the fact that it has been demonstrated that this type of therapy can reduce various side effects and adverse events of the commonly available cancer treatment approaches such as chemotherapy and radiotherapy. This selective type of cancer therapy can mediate encouraging outcomes where the frontline cancer treatment methods have failed to do so. Aptamer-assisted delivery of various types of cargoes or the utilization of aptamer for the redirection of delivery vehicles is among various fields of targeted cancer therapy that have gained significant attention lately. Aptamers are single-stranded oligonucleotides or peptide molecules that harbor significant levels of specificity and affinity toward various types of targets such as cell surface antigens, ions, toxins, chemicals, etc. They have shown encouraging results in several types of targeted cancer therapy for the redirection of a variety of cargoes. In this review, we shed the light on the application of aptamers for the delivery of nucleotides such as MicroRNAs (miRNAs), short or small interfering RNAs (siRNAs), and short hairpin RNA or small hairpin RNAs (shRNAs) that harbor tumor suppression properties in various kinds of malignancies.

1. Context

Novel approaches in the field of cancer treatment have revolutionized the way cancer patients are treated nowadays (1-3). The advent of targeted cancer treatment modalities, such as monoclonal antibodies or chimeric antigen receptor (CAR) T cells, has prompted the idea that targeted cancer therapies can ameliorate the side effects or enhance the therapeutic benefit of conventional cancer treatments. One of such targeted cancer therapies can be based on aptamer (1-3). Aptamers are short single-stranded oligonucleotides (either DNA- or RNA-based) or peptide molecules that have the ability to bind to a specific target molecule (4, 5). These oligonucleotides harbor significant binding affinity toward various targets, which can be of a wide range from cell surface antigens to soluble ligands (4, 5). Aptamers exhibit high affinity and specificity, similar to those of antibodies, because of their unique folding properties, which enable them to fold into tertiary structures (4, 5). However, aptamers suffer from several limitations, including their susceptibility to degradation in biological media or the high rate of renal clearance of naked aptamers, which is due to their small size.

The utilization of aptamers in various fields of research is mainly due to their multiple favorable properties that can be efficiently exploited for the redirection of various delivery platforms towards the tumor cells of interest with a great level of specificity (4, 5). Aptamers came to the spotlight of attention when in 2004, Macugen® (also known as Pegaptanib) became first aptamer approved by the US Food and Drug Administration (FDA) as an anti-angiogenic agent for the treatment of age-related macular degeneration (AMD) (6, 7). In comparison with antibodies, aptamers have a shorter generation time, exhibit more capability for modifications, harbor significant thermal stability, and their production is more cost-effective (4, 5).

To this date, aptamers have been utilized in many fields of investigation. The high affinity and specificity of aptamers allow for their application for clinical diagnostic purposes. They are also used in environmental protection and food safety fields. Aptamers are also used in the detection of pathogen microorganisms such as various types of viruses, bacteria, and parasites (8-23). Cancer recognition is another field in which aptamers are utilized for the detection of cancer-associated biomarkers such as mucin 1 (MUC1) and human epidermal growth factor receptor 2 (HER2) (24). For recognition purposes, aptamers are also used for the detection of the surface biomarkers of stem cells such as EpCAM, CD133, CD117, and CD44 (25). In addition to the abovementioned applications of aptamers, they are also used for monitoring environmental contaminations, including chemicals and toxins, for the production of biosensors capable of detecting various types of disease-related biomarkers, and they are also used as therapeutic agents (26, 27). Aptamers can also be conjugated to different types of molecules, such as cytotoxic drugs or nucleotides with tumor-suppressing properties (28). Moreover, they can be exploited for the redirection of cargo-loaded delivery vehicles such as liposomes, nanoparticles, and micelles (28). In this review, we discuss examples of aptamers used for the delivery of nucleotides that exhibit tumor-suppressing properties. We also discuss how these delivery platforms can be uniquely beneficial in the field of targeted cancer therapy.

2. Selective Delivery of Oligonucleotides Using Aptamer-Armed Platforms

2.1. Delivery of miRNAs

MicroRNAs, also known as miRNAs or miRs, are a class of 20 - 22 nucleotide-long non-coding small RNAs considered important regulators of various vital cellular functions such as proliferation, differentiation, and apoptosis (29-35). In addition, miRNAs exert such effects through the mechanism of complementary base-paring with their target mRNAs in a perfect or imperfect matching pattern leading to the subsequent degradation of the target; thus directly causing a transcriptional down-regulation or translational repression of the relative genes, which could include various tumor suppressor genes (29-35). Cancer therapeutic strategies leading to the loss of function of various cancer-specific miRNAs through their binding to miRNA-associated gene silencing complexes via fully-complementary base-pairing with synthetic oligonucleotides called “antagomir” or “antimiR” can lead to significantly reduced gene expression profiles of oncogenes as well as a considerably diminished cell viability of tumor cells (36-38).

Specific delivery of particular antimiR oligonucleotides to cancer cells through the targeting of nucleolin (the foremost nucleolar protein in growing eukaryotic cells, which is overexpressed in various types of malignancies) in cells that overexpress this cell surface protein can be utilized as a potent strategy to disrupt the miRNA-mediated oncogenic circuits in these cells (39, 40). Zhang and colleagues have reported the fabrication of a traceable and dual-targeted drug delivery system based on DNA-hybrid-capped mesoporous silica-coated quantum dots (MSQDs) in which the release of the loaded drug (doxorubicin) is controlled by miRNA (miR-21) (41). MiR-21 is one of the oncogenic miRNAs, which is overexpressed in various human cancers. Therefore, the suppression of its expression through delivering antisense oligonucleotides such as antimiR-21 can lead to the activation of caspase-dependent apoptosis and subsequent eradication of the tumor cells in a specific way. Moreover, the antimiR-21 strand can be coupled with a DNA aptamer, which leads to the formation of a DNA hybrid that can specifically recognize antigens overexpressed on the surfaces of the target tumor cells alongside having an exclusive response to miR-21 (which is proceeded through a complementary base-pairing mechanism) (41). In the study by Zhang and colleagues, the mentioned multifunctional MSQDs were loaded with doxorubicin, and they were capped with the DNA hybrid (synthesized by coupling antimiR-21 at the 3' end of the AS1411 aptamer, a nucleolin-targeting aptamer) by forming 12 base pairs between parts of anti-miR-21 and the anchor-DNA on the nanoparticles resulting in the formation of a DNA gate for the prevention of doxorubicin leakage. These nanocarriers enter the tumor cells upon the recognition of nucleolin by AS1411. Since miR-21 is overexpressed in the cytoplasm of the tumor cells, they play the role of an exclusive key to unlock the doxorubicin gate and meditate its release from the delivery vehicle complex by competing with anchor-DNA for full hybridization with anti-miR-21. Additionally, further enhanced efficacy of the chemotherapy is achieved by the complementary base pairing of anti-miR-21 with miR-21 resulting in the suppression of miR-21 expression (41). In a nutshell, this platform might elevate the therapeutic efficacy while diminishing unwanted adverse effects (41).

Aptamer-redirection of miRNAs has been investigated in various types of cancers, including non-small cell lung cancer (NSCLC), glioblastoma, prostate cancer, breast cancer, and gastric cancer (42-44). Esposito et al. have investigated specific aptamers for the receptor tyrosine kinase Axl conjugated to the let-7g miRNA (42). They have demonstrated that these constructs selectively target Axl-expressing tumor cells and effectively suppress tumor growth in xenograft models of lung adenocarcinoma (42). Moreover, Russo et al. have investigated the Axl-specific aptamer-redirected reintroduction of miR-34c-3p to NSCLC cells (44). NSCLC cells exhibit a decreased level of miR-34c-3p as compared with normal lung cells (44). Therefore, the authors of this study hypothesized that this reintroduction might decrease the proliferation of NSCLC tumor cells in vitro (44). They demonstrated that this method can suppress tumor cell growth in an efficient manner (44). Additionally, researchers have also investigated the Axl-aptamer-assisted delivery of miR-212 to NSCLC cells (43). TNF-related apoptosis-inducing ligand (TRAIL) is a well-recognized tumor suppressor pathway downregulated in many types of malignancies such as NSCLC (43). Recovering the activity of TRAIL can be achieved through reintroduction or overexpression of miR-212, which can lead to a targeted tumor cell apoptosis-mediated elimination (43). Iaboni et al. demonstrated that Axl-aptamer-assisted delivery of miR-212 to NSCLC cells could lead to selective tumor cell elimination (43).

2.2. Delivery of siRNAs

Short or small interfering RNAs (siRNAs) are a class of 20 - 25 base pair long synthetic double-stranded non-coding RNA molecules, which similar to endogenous microRNA, can operate within the RNA interference (RNAi) pathway to mediate highly efficient and specific post-transcriptional expression silencing of genes that are traditionally considered undruggable (45). Owing to their high therapeutic potential, siRNA-based approaches are being considered for various types of disease treatments, including several cancer types and viral infections (46-49). The application of siRNA-based therapeutics is still hindered by several drawbacks such as the instability of unmodified siRNAs in the bloodstream, their immunogenicity, and their weak cell-membrane crossing capabilities (45, 50). These limitations have encouraged researchers to develop safe siRNA delivery methods for redirecting them to their specific action sites without off-target toxicities or adverse effects (45, 50).

Targeted therapeutic agents consisting of aptamer-siRNA chimeras are currently being appraised for the treatment of several cancer types such as prostate cancer [by targeting prostate-specific membrane antigen (PSMA) and integrin alpha-V beta-3 (αVβ3)], B cell non-Hodgkin lymphoma [by targeting B-cell-activating factor-receptor (BAFF-R)], and breast cancer (by targeting HER2) (51-56). Aptamer-siRNA chimeras have also been investigated in other fields such as drug hypersensitivity (57) and HIV-1 treatment (58-60).

In one study, an AS1411 aptamer-redirected nanoliposome-based delivery system has been utilized for the co-delivery of the chemotherapeutic drug paclitaxel (PTX) and Polo-like kinase 1-targeted siRNA (PLK1-targeted siRNA) to breast cancer cells (61). PLK1 is a highly conserved serine/threonine protein kinase with important regulatory mitotic effects whose high expression levels have been significantly associated with abnormal tumor cell proliferation, metastasis, angiogenesis, and tumor prognosis in various types of cancers, including breast cancer (62, 63). Therefore, PLK1 can be considered a promising primary target candidate for cancer treatment, including PLK1-targeting RNAi-based gene therapy (61, 64-66). The simultaneous co-delivery of PTX and siRNA proposed by the mentioned study could result in a synergistic incremental pattern of apoptotic cancer cells and diminished angiogenesis (61). Therefore, this method may exhibit various advantages over methods separately delivering PTX and siRNA (61). It could also demonstrate a valuable potential for suppressing the growth of breast cancer in preclinical models (61).

In another example, Zhou et al. have utilized anti-BAFF-R aptamers for the redirection and delivery of nanoparticles loaded with the STAT3 siRNAs (67). They have demonstrated that the BAFF-R aptamers can specifically redirect the nanoparticles toward various B cell lines (67). This action is followed by the internalization of the nanoparticles, which eventually leads to the disruption of STAT3 mRNAs (67).

Moreover, PSMA is a very popular target antigen targeted in investigations studying aptamer-assisted redirection platforms. In this regard, Wullner et al. conjugated siRNAs specific for eukaryotic elongation factor 2 mRNA (eEF2K) to PSMA-targeting aptamers (68). Inhibiting EEF2 can mediate protein synthesis blockade leading to apoptosis in the PMSA-expressing prostate cancer cells (68). Moreover, other researchers have generated aptamer-siRNA chimeras made of two anti-PSMA aptamers in between which two siRNAs, one specific for EGFR and the other one specific for survivin, are located (69). The authors have reported that these chimeras can suppress EGFR and survivin expression at the same time and mediate apoptosis both in vitro and in vivo in an efficient manner (69).

2.3. Delivery of DNAzymes

Deoxyribozyme (also known as DNA enzyme, DNAzyme, Dz, or catalytic DNA), is another example of nucleotides with therapeutic properties. They are synthetic single-stranded DNA molecules capable of mediating chemical or catalytic reactions on particular nucleic acid targets similar to those of other biological protein-based enzymes or ribozymes (70-72). DNAzymes have been at the center of attention mainly due to their outstanding advantages, including their affordability, stability properties, and easy biosynthesis process (70-72).

DNAzymes has been proven to be capable of cleaving β-catenin and survivin mRNA and BCR-ABL transcripts, which further proves their potent role in growth inhibition of tumor cells alongside justifying the numerous attempts made for the development of efficient DNAzyme delivery platforms such as Poly(lactic-co-glycolic acid) (PLGA) microspheres, transferrin modified PEGylated polyplexes, poly-L-Lysine (PLL) microspheres, nanoparticulate systems, and dendrimers (73-77). These delivery platforms can be redirected towards tumor cells of interest using aptamers targeting tumor cell surface antigens. Such platforms can selectively deliver these delivery vehicles without targeting normal cells.

2.4. Delivery of shRNAs

Short hairpin RNAs or small hairpin RNAs (shRNAs), also known as hairpin vectors, are artificial RNA molecules biosynthesized exogenously or transcribed from RNA polymerase III promoters in vivo (78). These molecules are capable of inducing stable and heritable gene silencing effects with high specificity via RNAi pathway, thus allowing for the generation of continuous gene-modified cell lines or transgenic animals (78). After the generation of the shRNA transcripts, they are processed and loaded into RNA-induced silencing complex (RISC) in the cytoplasm undergoing further cytoplasmic RNAi processing (79). As the story is with plasmids, shRNAs encounter difficulties passing cellular membranes and migrating to the nucleus; therefore, their efficient delivery into target cells requires specific carriers such as nanocarriers or dendrimers capable of overcoming such obstacles (80-82).

One study has developed a novel targeted delivery platform for specific delivery of shRNA plasmids through the targeting of nucleolin ligand on target cancer cells (83). This targeted shRNA delivery system is composed of alkyl-modified polyamidoamine (PAMAM) dendrimers with 10-bromodecanoic acid (10C) and 10C-PEG to improve the efficiency of transfection, shRNA plasmid for specific knockdown of Bcl-xL protein, and the AS1411 aptamer for targeted delivery towards nucleolin over-expressing cancer cells (83). Dendrimers are star-shaped structures with numerous branches whose dimensions do not exceed nanometer scales. The fate of living cells is determined by the balance between the pro-apoptotic members of the Bcl-2 family, such as BAX, BAK, and BOK, which act by protecting the outer mitochondrial membrane and inhibiting the release of cytochrome c and the anti-apoptotic members, including Bcl-2, Bcl-xL, and MCL1. Selective silencing of Bcl-xL can be exploited as a strategy for apoptosis induction in cancer cells since the high level of Bcl-xL expression has been reported in numerous solid tumors such as bladder and gastric cancer (83-86). Without causing considerable cytotoxicity, the abovementioned targeted shRNA delivery system could efficiently downregulate the expression of Bcl-xL up to 25% and induce strongly selective late apoptosis in 14% of target cancer cells while exhibiting improved transfection efficiency in comparison to non-targeted vectors (83). In a nutshell, this strategic delivery system demonstrates that efficient and targeted apoptosis induction in various cancer cells through the knockdown of Bcl-xL expression using shRNAs can be achieved through aptamer-assisted redirection of delivery vehicles such as PAMAM dendrimers (83).

Moreover, Kim et al. have investigated the co-delivery of shRNAs specific for Bcl-xL and the chemotherapeutic agent doxorubicin using polyplexes redirected toward prostate cancer cells using anti-PSMA aptamers (87). They have reported that this construct effectively targets PSMA-expressing prostate cancer cells in a very selective manner (87). These results indicate that co-delivery of chemotherapeutic agents and shRNAs (such as the anti-Bcl-xL shRNA) can selectively target cancer cells and eliminate them with a significant level of specificity (87).

Furthermore, other researchers have generated aptamers harboring affinity and specificity for the HIV integrase (88). They have developed shRNA-aptamer fusions by joining the aptamers as the terminal loop of shRNAs targeting HIV Tat-Rev (88). These researchers have reported that the shRNA-aptamer fusions (using an aptamer named S3R3) can efficiently block HIV replication even in a long-term manner (88). They have also indicated that these shRNA-aptamer fusions exhibit similar suppression properties as those of the integrase inhibitor raltegravir (88). Such data can suggest that aptamer-shRNA fusions may have a bright future ahead of them and may be used for fighting against viral infections that can mediate malignancies such as the human papillomavirus (HPV).

3. Summary and Perspectives

Herein, we discussed the potential and application of aptamers specific for different targets utilized for the targeted delivery of various types of nucleotides with tumor growth suppression characteristics. Broadening the validity of the herein discussed platforms can be achieved through in-depth assessments and preclinical models of malignancies, especially where only in vitro assessments have been reported. Moreover, as we discussed throughout the article, the selective delivery capability of aptamers could be exploited for the treatment of viral infections and many other conditions as well. Special efforts should be made to be able to use innovative platforms for achieving such aims. Furthermore, alongside the types of malignancies popular in the field of aptamer-assisted cargo delivery investigations, other types of less investigated malignancies should also be considered since it is speculated that such outcomes can be achieved for their treatment as well. It is worth mentioning that there are still limitations surrounding this type of therapy. These limitations may include the off-tumor targeting toxicity of these platforms that can overshadow the potential of this type of cancer therapy. Therefore, discovering new strategies for tacking this hurdle is a factor of paramount importance.

Acknowledgements

References

  • 1.

    Safarzadeh Kozani P, Safarzadeh Kozani P, Rahbarizadeh F, Khoshtinat Nikkhoi S. Strategies for dodging the obstacles in CAR T cell therapy. Front Oncol. 2021;11:627549. [PubMed ID: 33869011]. [PubMed Central ID: PMC8047470]. https://doi.org/10.3389/fonc.2021.627549.

  • 2.

    Safarzadeh Kozani P, Safarzadeh Kozani P, Rahbarizadeh F. Novel antigens of CAR T cell therapy: New roads; old destination. Transl Oncol. 2021;14(7):101079. [PubMed ID: 33862524]. [PubMed Central ID: PMC8065293]. https://doi.org/10.1016/j.tranon.2021.101079.

  • 3.

    Hashem Boroojerdi M, Rahbarizadeh F, Safarzadeh Kozani P, Kamali E, Safarzadeh Kozani P. Strategies for having a more effective and less toxic CAR T-cell therapy for acute lymphoblastic leukemia. Med Oncol. 2020;37(11):100. [PubMed ID: 33047234]. [PubMed Central ID: PMC7549730]. https://doi.org/10.1007/s12032-020-01416-3.

  • 4.

    Dunn MR, Jimenez RM, Chaput JC. Analysis of aptamer discovery and technology. Nat Rev Chem. 2017;1(10). https://doi.org/10.1038/s41570-017-0076.

  • 5.

    Zhou J, Rossi JJ. Cell-specific aptamer-mediated targeted drug delivery. Oligonucleotides. 2011;21(1):1-10. [PubMed ID: 21182455]. [PubMed Central ID: PMC3043981]. https://doi.org/10.1089/oli.2010.0264.

  • 6.

    Kaur H, Bruno JG, Kumar A, Sharma TK. Aptamers in the therapeutics and diagnostics pipelines. Theranostics. 2018;8(15):4016-32. [PubMed ID: 30128033]. [PubMed Central ID: PMC6096388]. https://doi.org/10.7150/thno.25958.

  • 7.

    Morita Y, Leslie M, Kameyama H, Volk DE, Tanaka T. Aptamer therapeutics in cancer: Current and future. Cancers (Basel). 2018;10(3). [PubMed ID: 29562664]. [PubMed Central ID: PMC5876655]. https://doi.org/10.3390/cancers10030080.

  • 8.

    Boiziau C, Dausse E, Yurchenko L, Toulme JJ. DNA aptamers selected against the HIV-1 trans-activation-responsive RNA element form RNA-DNA kissing complexes. J Biol Chem. 1999;274(18):12730-7. [PubMed ID: 10212256]. https://doi.org/10.1074/jbc.274.18.12730.

  • 9.

    Cao X, Li S, Chen L, Ding H, Xu H, Huang Y, et al. Combining use of a panel of ssDNA aptamers in the detection of Staphylococcus aureus. Nucleic Acids Res. 2009;37(14):4621-8. [PubMed ID: 19498077]. [PubMed Central ID: PMC2724295]. https://doi.org/10.1093/nar/gkp489.

  • 10.

    Chen F, Zhou J, Luo F, Mohammed AB, Zhang XL. Aptamer from whole-bacterium SELEX as new therapeutic reagent against virulent Mycobacterium tuberculosis. Biochem Biophys Res Commun. 2007;357(3):743-8. [PubMed ID: 17442275]. https://doi.org/10.1016/j.bbrc.2007.04.007.

  • 11.

    Cheung YW, Dirkzwager RM, Wong WC, Cardoso J, D'Arc Neves Costa J, Tanner JA. Aptamer-mediated Plasmodium-specific diagnosis of malaria. Biochimie. 2018;145:131-6. [PubMed ID: 29080831]. https://doi.org/10.1016/j.biochi.2017.10.017.

  • 12.

    Duan N, Wu S, Chen X, Huang Y, Wang Z. Selection and identification of a DNA aptamer targeted to Vibrio parahemolyticus. J Agric Food Chem. 2012;60(16):4034-8. [PubMed ID: 22480209]. https://doi.org/10.1021/jf300395z.

  • 13.

    Dwivedi HP, Smiley RD, Jaykus LA. Selection and characterization of DNA aptamers with binding selectivity to Campylobacter jejuni using whole-cell SELEX. Appl Microbiol Biotechnol. 2010;87(6):2323-34. [PubMed ID: 20582587]. https://doi.org/10.1007/s00253-010-2728-7.

  • 14.

    Fukuda K, Vishinuvardhan D, Sekiya S, Kakiuchi N, Shimotohno K, Kumar PK, et al. Specific RNA aptamers to NS3 protease domain of hepatitis C virus. Nucleic Acids Symp Ser. 1997;(37):237-8. [PubMed ID: 9586087].

  • 15.

    Gopinath SC, Hayashi K, Kumar PK. Aptamer that binds to the gD protein of herpes simplex virus 1 and efficiently inhibits viral entry. J Virol. 2012;86(12):6732-44. [PubMed ID: 22514343]. [PubMed Central ID: PMC3393590]. https://doi.org/10.1128/JVI.00377-12.

  • 16.

    Gopinath SC, Kawasaki K, Kumar PK. Selection of RNA-aptamer against human influenza B virus. Nucleic Acids Symp Ser (Oxf). 2005;(49):85-6. [PubMed ID: 17150645]. https://doi.org/10.1093/nass/49.1.85.

  • 17.

    Guerra-Perez N, Ramos E, Garcia-Hernandez M, Pinto C, Soto M, Martin ME, et al. Molecular and functional characterization of ssDNA aptamers that specifically bind Leishmania infantum PABP. PLoS One. 2015;10(10). e0140048. [PubMed ID: 26457419]. [PubMed Central ID: PMC4601788]. https://doi.org/10.1371/journal.pone.0140048.

  • 18.

    Hamula CL, Zhang H, Guan LL, Li XF, Le XC. Selection of aptamers against live bacterial cells. Anal Chem. 2008;80(20):7812-9. [PubMed ID: 18803393]. https://doi.org/10.1021/ac801272s.

  • 19.

    Jain P, Chakma B, Singh NK, Patra S, Goswami P. Aromatic surfactant as aggregating agent for aptamer-gold nanoparticle-based detection of plasmodium lactate dehydrogenase. Mol Biotechnol. 2016;58(7):497-508. [PubMed ID: 27189484]. https://doi.org/10.1007/s12033-016-9946-x.

  • 20.

    Jang KJ, Lee NR, Yeo WS, Jeong YJ, Kim DE. Isolation of inhibitory RNA aptamers against severe acute respiratory syndrome (SARS) Coronavirus NTPase/Helicase. Biochem Biophys Res Commun. 2008;366(3):738-44. [PubMed ID: 18082623]. [PubMed Central ID: PMC7092905]. https://doi.org/10.1016/j.bbrc.2007.12.020.

  • 21.

    Kumar PK, Machida K, Urvil PT, Kakiuchi N, Vishnuvardhan D, Shimotohno K, et al. Isolation of RNA aptamers specific to the NS3 protein of hepatitis C virus from a pool of completely random RNA. Virology. 1997;237(2):270-82. [PubMed ID: 9356339]. https://doi.org/10.1006/viro.1997.8773.

  • 22.

    Labib M, Zamay AS, Muharemagic D, Chechik AV, Bell JC, Berezovski MV. Aptamer-based viability impedimetric sensor for viruses. Anal Chem. 2012;84(4):1813-6. [PubMed ID: 22303883]. https://doi.org/10.1021/ac203412m.

  • 23.

    Nagarkatti R, de Araujo FF, Gupta C, Debrabant A. Aptamer based, non-PCR, non-serological detection of Chagas disease biomarkers in Trypanosoma cruzi infected mice. PLoS Negl Trop Dis. 2014;8(1). e2650. [PubMed ID: 24454974]. [PubMed Central ID: PMC3894185]. https://doi.org/10.1371/journal.pntd.0002650.

  • 24.

    Zheng CY, Pestilli F, Rokem A. Deconvolution of high dimensional mixtures via boosting, with application to diffusion-weighted MRI of human brain. Adv Neural Inf Process Syst. 2014;27:2699-707. [PubMed ID: 25684972]. [PubMed Central ID: PMC4324561].

  • 25.

    Ababneh N, Alshaer W, Allozi O, Mahafzah A, El-Khateeb M, Hillaireau H, et al. In vitro selection of modified RNA aptamers against CD44 cancer stem cell marker. Nucleic Acid Ther. 2013;23(6):401-7. [PubMed ID: 24171482]. [PubMed Central ID: PMC3868357]. https://doi.org/10.1089/nat.2013.0423.

  • 26.

    McConnell EM, Nguyen J, Li Y. Aptamer-based biosensors for environmental monitoring. Front Chem. 2020;8:434. [PubMed ID: 32548090]. [PubMed Central ID: PMC7272472]. https://doi.org/10.3389/fchem.2020.00434.

  • 27.

    Nguyen VT, Kwon YS, Gu MB. Aptamer-based environmental biosensors for small molecule contaminants. Curr Opin Biotechnol. 2017;45:15-23. [PubMed ID: 28088092]. https://doi.org/10.1016/j.copbio.2016.11.020.

  • 28.

    He F, Wen N, Xiao D, Yan J, Xiong H, Cai S, et al. Aptamer-based targeted drug delivery systems: Current potential and challenges. Curr Med Chem. 2020;27(13):2189-219. [PubMed ID: 30295183]. https://doi.org/10.2174/0929867325666181008142831.

  • 29.

    Bartel DP. MicroRNAs: Genomics, biogenesis, mechanism, and function. Cell. 2004;116(2):281-97. [PubMed ID: 14744438]. https://doi.org/10.1016/s0092-8674(04)00045-5.

  • 30.

    Valencia-Sanchez MA, Liu J, Hannon GJ, Parker R. Control of translation and mRNA degradation by miRNAs and siRNAs. Genes Dev. 2006;20(5):515-24. [PubMed ID: 16510870]. https://doi.org/10.1101/gad.1399806.

  • 31.

    Zamore PD, Haley B. Ribo-gnome: the big world of small RNAs. Science. 2005;309(5740):1519-24. [PubMed ID: 16141061]. https://doi.org/10.1126/science.1111444.

  • 32.

    Sullivan CS, Ganem D. MicroRNAs and viral infection. Mol Cell. 2005;20(1):3-7. [PubMed ID: 16209940]. https://doi.org/10.1016/j.molcel.2005.09.012.

  • 33.

    Croce CM, Calin GA. miRNAs, cancer, and stem cell division. Cell. 2005;122(1):6-7. [PubMed ID: 16009126]. https://doi.org/10.1016/j.cell.2005.06.036.

  • 34.

    Iorio MV, Croce CM. MicroRNAs in cancer: Small molecules with a huge impact. J Clin Oncol. 2009;27(34):5848-56. [PubMed ID: 19884536]. [PubMed Central ID: PMC2793003]. https://doi.org/10.1200/JCO.2009.24.0317.

  • 35.

    Zeng Y, Yi R, Cullen BR. MicroRNAs and small interfering RNAs can inhibit mRNA expression by similar mechanisms. Proc Natl Acad Sci U S A. 2003;100(17):9779-84. [PubMed ID: 12902540]. [PubMed Central ID: PMC187842]. https://doi.org/10.1073/pnas.1630797100.

  • 36.

    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(19):8994-9000. [PubMed ID: 17908999]. https://doi.org/10.1158/0008-5472.CAN-07-1045.

  • 37.

    Chan JA, Krichevsky AM, Kosik KS. MicroRNA-21 is an antiapoptotic factor in human glioblastoma cells. Cancer Res. 2005;65(14):6029-33. [PubMed ID: 16024602]. https://doi.org/10.1158/0008-5472.CAN-05-0137.

  • 38.

    Zhang C, Han L, Zhang A, Yang W, Zhou X, Pu P, et al. Global changes of mRNA expression reveals an increased activity of the interferon-induced signal transducer and activator of transcription (STAT) pathway by repression of miR-221/222 in glioblastoma U251 cells. Int J Oncol. 2010;36(6):1503-12. [PubMed ID: 20428775]. https://doi.org/10.3892/ijo_00000637.

  • 39.

    Kim HJ, Kim YH, Lee DS, Chung JK, Kim S. In vivo imaging of functional targeting of miR-221 in papillary thyroid carcinoma. J Nucl Med. 2008;49(10):1686-93. [PubMed ID: 18794255]. https://doi.org/10.2967/jnumed.108.052894.

  • 40.

    Kim HJ, Chung JK, Hwang DW, Lee DS, Kim S. In vivo imaging of miR-221 biogenesis in papillary thyroid carcinoma. Mol Imaging Biol. 2009;11(2):71-8. [PubMed ID: 19030936]. https://doi.org/10.1007/s11307-008-0188-6.

  • 41.

    Zhang P, Cheng F, Zhou R, Cao J, Li J, Burda C, et al. DNA-hybrid-gated multifunctional mesoporous silica nanocarriers for dual-targeted and microRNA-responsive controlled drug delivery. Angew Chem Int Ed Engl. 2014;53(9):2371-5. [PubMed ID: 24470397]. https://doi.org/10.1002/anie.201308920.

  • 42.

    Esposito CL, Cerchia L, Catuogno S, De Vita G, Dassie JP, Santamaria G, et al. Multifunctional aptamer-miRNA conjugates for targeted cancer therapy. Mol Ther. 2014;22(6):1151-63. [PubMed ID: 24441398]. [PubMed Central ID: PMC4048903]. https://doi.org/10.1038/mt.2014.5.

  • 43.

    Iaboni M, Russo V, Fontanella R, Roscigno G, Fiore D, Donnarumma E, et al. Aptamer-miRNA-212 conjugate sensitizes NSCLC cells to trail. Mol Ther Nucleic Acids. 2016;5. e289. [PubMed ID: 27111415]. [PubMed Central ID: PMC5014461]. https://doi.org/10.1038/mtna.2016.5.

  • 44.

    Russo V, Paciocco A, Affinito A, Roscigno G, Fiore D, Palma F, et al. Aptamer-miR-34c conjugate affects cell proliferation of non-small-cell lung cancer cells. Mol Ther Nucleic Acids. 2018;13:334-46. [PubMed ID: 30340138]. [PubMed Central ID: PMC6197774]. https://doi.org/10.1016/j.omtn.2018.09.016.

  • 45.

    Kanasty R, Dorkin JR, Vegas A, Anderson D. Delivery materials for siRNA therapeutics. Nat Mater. 2013;12(11):967-77. [PubMed ID: 24150415]. https://doi.org/10.1038/nmat3765.

  • 46.

    Shen H, Sun T, Ferrari M. Nanovector delivery of siRNA for cancer therapy. Cancer Gene Ther. 2012;19(6):367-73. [PubMed ID: 22555511]. [PubMed Central ID: PMC3842228]. https://doi.org/10.1038/cgt.2012.22.

  • 47.

    Davis ME, Zuckerman JE, Choi CH, Seligson D, Tolcher A, Alabi CA, et al. Evidence of RNAi in humans from systemically administered siRNA via targeted nanoparticles. Nature. 2010;464(7291):1067-70. [PubMed ID: 20305636]. [PubMed Central ID: PMC2855406]. https://doi.org/10.1038/nature08956.

  • 48.

    Morrissey DV, Lockridge JA, Shaw L, Blanchard K, Jensen K, Breen W, et al. Potent and persistent in vivo anti-HBV activity of chemically modified siRNAs. Nat Biotechnol. 2005;23(8):1002-7. [PubMed ID: 16041363]. https://doi.org/10.1038/nbt1122.

  • 49.

    Okumura A, Pitha PM, Harty RN. ISG15 inhibits Ebola VP40 VLP budding in an L-domain-dependent manner by blocking Nedd4 ligase activity. Proc Natl Acad Sci U S A. 2008;105(10):3974-9. [PubMed ID: 18305167]. [PubMed Central ID: PMC2268823]. https://doi.org/10.1073/pnas.0710629105.

  • 50.

    Whitehead KA, Langer R, Anderson DG. Knocking down barriers: Advances in siRNA delivery. Nat Rev Drug Discov. 2009;8(2):129-38. [PubMed ID: 19180106]. [PubMed Central ID: PMC7097568]. https://doi.org/10.1038/nrd2742.

  • 51.

    Zhou J, Tiemann K, Chomchan P, Alluin J, Swiderski P, Burnett J, et al. Dual functional BAFF receptor aptamers inhibit ligand-induced proliferation and deliver siRNAs to NHL cells. Nucleic Acids Res. 2013;41(7):4266-83. [PubMed ID: 23470998]. [PubMed Central ID: PMC3627597]. https://doi.org/10.1093/nar/gkt125.

  • 52.

    Thiel KW, Hernandez LI, Dassie JP, Thiel WH, Liu X, Stockdale KR, et al. Delivery of chemo-sensitizing siRNAs to HER2+-breast cancer cells using RNA aptamers. Nucleic Acids Res. 2012;40(13):6319-37. [PubMed ID: 22467215]. [PubMed Central ID: PMC3401474]. https://doi.org/10.1093/nar/gks294.

  • 53.

    Hussain AF, Tur MK, Barth S. An aptamer-siRNA chimera silences the eukaryotic elongation factor 2 gene and induces apoptosis in cancers expressing alphavbeta3 integrin. Nucleic Acid Ther. 2013;23(3):203-12. [PubMed ID: 23544955]. https://doi.org/10.1089/nat.2012.0408.

  • 54.

    Dassie JP, Liu XY, Thomas GS, Whitaker RM, Thiel KW, Stockdale KR, et al. Systemic administration of optimized aptamer-siRNA chimeras promotes regression of PSMA-expressing tumors. Nat Biotechnol. 2009;27(9):839-49. [PubMed ID: 19701187]. [PubMed Central ID: PMC2791695]. https://doi.org/10.1038/nbt.1560.

  • 55.

    Ni X, Zhang Y, Ribas J, Chowdhury WH, Castanares M, Zhang Z, et al. Prostate-targeted radiosensitization via aptamer-shRNA chimeras in human tumor xenografts. J Clin Invest. 2011;121(6):2383-90. [PubMed ID: 21555850]. [PubMed Central ID: PMC3104752]. https://doi.org/10.1172/JCI45109.

  • 56.

    Pastor F, Kolonias D, Giangrande PH, Gilboa E. Induction of tumour immunity by targeted inhibition of nonsense-mediated mRNA decay. Nature. 2010;465(7295):227-30. [PubMed ID: 20463739]. [PubMed Central ID: PMC3107067]. https://doi.org/10.1038/nature08999.

  • 57.

    Wang CW, Chung WH, Cheng YF, Ying NW, Peck K, Chen YT, et al. A new nucleic acid-based agent inhibits cytotoxic T lymphocyte-mediated immune disorders. J Allergy Clin Immunol. 2013;132(3):713-722 e11. [PubMed ID: 23791505]. https://doi.org/10.1016/j.jaci.2013.04.036.

  • 58.

    Neff CP, Zhou J, Remling L, Kuruvilla J, Zhang J, Li H, et al. An aptamer-siRNA chimera suppresses HIV-1 viral loads and protects from helper CD4(+) T cell decline in humanized mice. Sci Transl Med. 2011;3(66):66ra6. [PubMed ID: 21248316]. [PubMed Central ID: PMC3138523]. https://doi.org/10.1126/scitranslmed.3001581.

  • 59.

    Wheeler LA, Trifonova R, Vrbanac V, Basar E, McKernan S, Xu Z, et al. Inhibition of HIV transmission in human cervicovaginal explants and humanized mice using CD4 aptamer-siRNA chimeras. J Clin Invest. 2011;121(6):2401-12. [PubMed ID: 21576818]. [PubMed Central ID: PMC3104760]. https://doi.org/10.1172/JCI45876.

  • 60.

    Wheeler LA, Vrbanac V, Trifonova R, Brehm MA, Gilboa-Geffen A, Tanno S, et al. Durable knockdown and protection from HIV transmission in humanized mice treated with gel-formulated CD4 aptamer-siRNA chimeras. Mol Ther. 2013;21(7):1378-89. [PubMed ID: 23629001]. [PubMed Central ID: PMC3702101]. https://doi.org/10.1038/mt.2013.77.

  • 61.

    Yu S, Bi X, Yang L, Wu S, Yu Y, Jiang B, et al. Co-delivery of paclitaxel and PLK1-targeted siRNA using aptamer-functionalized cationic liposome for synergistic anti-breast cancer effects in vivo. J Biomed Nanotechnol. 2019;15(6):1135-48. [PubMed ID: 31072423]. https://doi.org/10.1166/jbn.2019.2751.

  • 62.

    Li J, Hong MJ, Chow JP, Man WY, Mak JP, Ma HT, et al. Co-inhibition of polo-like kinase 1 and Aurora kinases promotes mitotic catastrophe. Oncotarget. 2015;6(11):9327-40. [PubMed ID: 25871386]. [PubMed Central ID: PMC4496220]. https://doi.org/10.18632/oncotarget.3313.

  • 63.

    Gomes-da-Silva LC, Ramalho JS, Pedroso de Lima MC, Simoes S, Moreira JN. Impact of anti-PLK1 siRNA-containing F3-targeted liposomes on the viability of both cancer and endothelial cells. Eur J Pharm Biopharm. 2013;85(3 Pt A):356-64. [PubMed ID: 23659854]. https://doi.org/10.1016/j.ejpb.2013.04.007.

  • 64.

    King SI, Purdie CA, Bray SE, Quinlan PR, Jordan LB, Thompson AM, et al. Immunohistochemical detection of Polo-like kinase-1 (PLK1) in primary breast cancer is associated with TP53 mutation and poor clinical outcom. Breast Cancer Res. 2012;14(2):R40. [PubMed ID: 22405092]. [PubMed Central ID: PMC3446374]. https://doi.org/10.1186/bcr3136.

  • 65.

    Ha GH, Kim DY, Breuer EK, Kim CK. Combination treatment of polo-like kinase 1 and tankyrase-1 inhibitors enhances anticancer effect in triple-negative breast cancer cells. Anticancer Res. 2018;38(3):1303-10. [PubMed ID: 29491053]. https://doi.org/10.21873/anticanres.12352.

  • 66.

    Morry J, Ngamcherdtrakul W, Gu S, Reda M, Castro DJ, Sangvanich T, et al. Targeted treatment of metastatic breast cancer by PLK1 siRNA delivered by an antioxidant nanoparticle platform. Mol Cancer Ther. 2017;16(4):763-72. [PubMed ID: 28138033]. [PubMed Central ID: PMC5445934]. https://doi.org/10.1158/1535-7163.MCT-16-0644.

  • 67.

    Zhou J, Rossi JJ, Shum KT. Methods for assembling B-cell lymphoma specific and internalizing aptamer-siRNA nanoparticles via the sticky bridge. Methods Mol Biol. 2015;1297:169-85. [PubMed ID: 25896003]. https://doi.org/10.1007/978-1-4939-2562-9_12.

  • 68.

    Wullner U, Neef I, Eller A, Kleines M, Tur MK, Barth S. Cell-specific induction of apoptosis by rationally designed bivalent aptamer-siRNA transcripts silencing eukaryotic elongation factor 2. Curr Cancer Drug Targets. 2008;8(7):554-65. [PubMed ID: 18991566]. https://doi.org/10.2174/156800908786241078.

  • 69.

    Liu HY, Yu X, Liu H, Wu D, She JX. Co-targeting EGFR and survivin with a bivalent aptamer-dual siRNA chimera effectively suppresses prostate cancer. Sci Rep. 2016;6:30346. [PubMed ID: 27456457]. [PubMed Central ID: PMC4960556]. https://doi.org/10.1038/srep30346.

  • 70.

    Kanwar JR, Roy K, Kanwar RK. Chimeric aptamers in cancer cell-targeted drug delivery. Crit Rev Biochem Mol Biol. 2011;46(6):459-77. [PubMed ID: 21955150]. [PubMed Central ID: PMC3233271]. https://doi.org/10.3109/10409238.2011.614592.

  • 71.

    Ray P, White RR. Aptamers for targeted drug delivery. Pharmaceuticals (Basel). 2010;3(6):1761-78. [PubMed ID: 27713328]. [PubMed Central ID: PMC4033951]. https://doi.org/10.3390/ph3061761.

  • 72.

    Breaker RR, Joyce GF. A DNA enzyme with Mg(2+)-dependent RNA phosphoesterase activity. Chem Biol. 1995;2(10):655-60. [PubMed ID: 9383471]. https://doi.org/10.1016/1074-5521(95)90028-4.

  • 73.

    Choi BR, Gwak J, Kwon HM, Oh S, Kim KP, Choi WH, et al. Oligodeoxyribozymes that cleave beta-catenin messenger RNA inhibit growth of colon cancer cells via reduction of beta-catenin response transcription. Mol Cancer Ther. 2010;9(6):1894-902. [PubMed ID: 20501807]. https://doi.org/10.1158/1535-7163.MCT-10-0056.

  • 74.

    Gozar MM, Goodchild A, Passioura T, King A, Lai A, Witherington C, et al. Dz13, a DNAzyme targeting c-jun, induces off-target cytotoxicity in endothelial cells with features of nonapoptotic programmed cell death. Oligonucleotides. 2008;18(3):257-68. [PubMed ID: 18699742]. https://doi.org/10.1089/oli.2008.0139.

  • 75.

    Kim JE, Yoon S, Choi BR, Kim KP, Cho YH, Jung W, et al. Cleavage of BCR-ABL transcripts at the T315I point mutation by DNAzyme promotes apoptotic cell death in imatinib-resistant BCR-ABL leukemic cells. Leukemia. 2013;27(8):1650-8. [PubMed ID: 23434731]. https://doi.org/10.1038/leu.2013.60.

  • 76.

    Liang Z, Wei S, Guan J, Luo Y, Gao J, Zhu H, et al. DNAzyme-mediated cleavage of survivin mRNA and inhibition of the growth of PANC-1 cells. J Gastroenterol Hepatol. 2005;20(10):1595-602. [PubMed ID: 16174080]. https://doi.org/10.1111/j.1440-1746.2005.03978.x.

  • 77.

    Xu Z, Yang L, Sun LQ, Cao Y. Use of DNAzymes for cancer research and therapy. Chin Sci Bull. 2012;57(26):3404-8. https://doi.org/10.1007/s11434-012-5380-z.

  • 78.

    Paddison PJ, Caudy AA, Bernstein E, Hannon GJ, Conklin DS. Short hairpin RNAs (shRNAs) induce sequence-specific silencing in mammalian cells. Genes Dev. 2002;16(8):948-58. [PubMed ID: 11959843]. [PubMed Central ID: PMC152352]. https://doi.org/10.1101/gad.981002.

  • 79.

    Wang Z, Rao DD, Senzer N, Nemunaitis J. RNA interference and cancer therapy. Pharm Res. 2011;28(12):2983-95. [PubMed ID: 22009588]. https://doi.org/10.1007/s11095-011-0604-5.

  • 80.

    Abbasi E, Aval SF, Akbarzadeh A, Milani M, Nasrabadi HT, Joo SW, et al. Dendrimers: synthesis, applications, and properties. Nanoscale Res Lett. 2014;9(1):247. [PubMed ID: 24994950]. [PubMed Central ID: PMC4074873]. https://doi.org/10.1186/1556-276X-9-247.

  • 81.

    Askarian S, Abnous K, Taghavi S, Oskuee RK, Ramezani M. Cellular delivery of shRNA using aptamer-conjugated PLL-alkyl-PEI nanoparticles. Colloids Surf B Biointerfaces. 2015;136:355-64. [PubMed ID: 26433348]. https://doi.org/10.1016/j.colsurfb.2015.09.023.

  • 82.

    Mokhtarzadeh A, Alibakhshi A, Hashemi M, Hejazi M, Hosseini V, de la Guardia M, et al. Biodegradable nano-polymers as delivery vehicles for therapeutic small non-coding ribonucleic acids. J Control Release. 2017;245:116-26. [PubMed ID: 27884808]. https://doi.org/10.1016/j.jconrel.2016.11.017.

  • 83.

    Ayatollahi S, Salmasi Z, Hashemi M, Askarian S, Oskuee RK, Abnous K, et al. Aptamer-targeted delivery of Bcl-xL shRNA using alkyl modified PAMAM dendrimers into lung cancer cells. Int J Biochem Cell Biol. 2017;92:210-7. [PubMed ID: 29031805]. https://doi.org/10.1016/j.biocel.2017.10.005.

  • 84.

    Kang MH, Reynolds CP. Bcl-2 inhibitors: Targeting mitochondrial apoptotic pathways in cancer therapy. Clin Cancer Res. 2009;15(4):1126-32. [PubMed ID: 19228717]. [PubMed Central ID: PMC3182268]. https://doi.org/10.1158/1078-0432.CCR-08-0144.

  • 85.

    Kirsh EJ, Baunoch DA, Stadler WM. Expression of bcl-2 and bcl-X in bladder cancer. J Urol. 1998;159(4):1348-53. [PubMed ID: 9507882].

  • 86.

    Krajewski S, Krajewska M, Ehrmann J, Sikorska M, Lach B, Chatten J, et al. Immunohistochemical analysis of Bcl-2, Bcl-X, Mcl-1, and Bax in tumors of central and peripheral nervous system origin. Am J Pathol. 1997;150(3):805-14. [PubMed ID: 9060818]. [PubMed Central ID: PMC1857882].

  • 87.

    Kim E, Jung Y, Choi H, Yang J, Suh JS, Huh YM, et al. Prostate cancer cell death produced by the co-delivery of Bcl-xL shRNA and doxorubicin using an aptamer-conjugated polyplex. Biomaterials. 2010;31(16):4592-9. [PubMed ID: 20206379]. https://doi.org/10.1016/j.biomaterials.2010.02.030.

  • 88.

    Pang KM, Castanotto D, Li H, Scherer L, Rossi JJ. Incorporation of aptamers in the terminal loop of shRNAs yields an effective and novel combinatorial targeting strategy. Nucleic Acids Res. 2018;46(1). e6. [PubMed ID: 29077949]. [PubMed Central ID: PMC5758892]. https://doi.org/10.1093/nar/gkx980.