Abstract
Small-molecule and protein/antibody drugs mainly act on genome-derived proteins to exert pharmacological effects. RNA based therapies hold the promise to expand the range of druggable targets from proteins to RNAs and the genome, as evidenced by several RNA drugs approved for clinical practice and many others under active trials. While chemo-engineered RNA mimics have found their success in marketed drugs and continue dominating basic research and drug development, these molecules are usually conjugated with extensive and various modifications. This makes them completely different from cellular RNAs transcribed from the genome that usually consist of unmodified ribonucleotides or just contain a few posttranscriptional modifications. The use of synthetic RNA mimics for RNA research and drug development is also in contrast with the ultimate success of protein research and therapy utilizing biologic or recombinant proteins produced and folded in living cells instead of polypeptides or proteins synthesized in vitro. Indeed, efforts have been made recently to develop RNA bioengineering technologies for cost-effective and large-scale production of biologic RNA molecules that may better capture the structures, functions, and safety profiles of natural RNAs. In this article, we provide an overview on RNA therapeutics for the treatment of human diseases via RNA interference mechanisms. By illustrating the structural differences between natural RNAs and chemo-engineered RNA mimics, we focus on discussion of a novel class of bioengineered/biologic RNA agents produced through fermentation and their potential applications to RNA research and drug development.
Keywords: therapy, RNAi, miRNA, ncRNA, biotechnology, cancer
1. Introduction
Pharmacotherapy utilizes pharmaceutical compounds for the treatment of human diseases, different from other means such as surgery, radiation, and physical therapy. Pharmaceutical drugs usually interact with particular biologic molecules to exert pharmacological actions for the control of disease. Proteins translated from mRNAs are common targets of current pharmaceutical drugs (Gashaw, Ellinghaus, Sommer, & Asadullah, 2012; Overington, Al-Lazikani, & Hopkins, 2006; Santos, et al., 2017) which are predominantly small molecules (Fig. 1). Although protein therapeutics such as antibodies have revolutionized pharmacotherapy and drug development (Dimitrov, 2012; Secher, et al., 2018; Sliwkowski & Mellman, 2013; Trail, Dubowchik, & Lowinger, 2018), their targets are still mainly protein macromolecules (Fig. 1) while the majority of proteins encoded by the human genome are actually undruggable or non-druggable by conventional modes of therapeutics. Furthermore, genetic or epigenetic changes of an existing protein target can escape from current medications or acquire drug resistance (Brown, Curry, Magnani, Wilhelm-Benartzi, & Borley, 2014; Choi & Yu, 2014). In addition, over 95% of DNA sequences in the human genome are non-protein coding sequences (Mattick, 2004), which many could be transcribed in cells and may be further processed to enormous numbers of functional noncoding RNAs (ncRNAs) (Fig. 1), including transfer RNAs (tRNAs), ribosomal RNAs (rRNAs), microRNAs (miRNAs or miRs), and long noncoding RNAs (lncRNAs) (Esteller, 2011; Liz & Esteller, 2016; Matsui & Corey, 2017). However, such massive families of functional ncRNAs remain as undrugged or unexplored targets for pharmacotherapy, which thus offers new opportunities for drug development.
Fig. 1. Expanding the range of druggable targets with RNA therapeutics.
Proteins derived from the genome remain as favorable targets for pharmacotherapy, whereas the majority of DNA sequences in the human genome are transcribed as non-protein coding transcripts. As current medications are mainly small molecules and proteins (e.g., antibodies) that act on proteins, RNA therapeutics hold great promise to expand druggable genome for the treatment of human diseases, including (1) RNA aptamers that block protein targets, (2) RNAs such as asRNAs, miRNAs, and siRNAs that target mRNAs or various forms of ncRNAs, and (3) gRNAs that directly edit gene sequences.
RNA molecules have emerged as a new class of therapeutics that may permit the re-targeting of mutated targets, which holds great promise to expand the range of druggable targets from proteins to RNAs as well as the genome (Fig. 1). First, the present protein targets as well as previously-undruggable proteins may be inhibited by appropriate RNA molecules, namely aptamers, to elicit the desired pharmacological effects (X. Chen, et al., 2018; Nimjee, White, Becker, & Sullenger, 2017; Zhou & Rossi, 2017). Second, mRNAs and ncRNAs could be directly targeted by a variety of RNA entities such as antisense RNAs (asRNAs), miRNAs, small interfering RNAs (siRNAs), and other forms of small RNAs (sRNAs) to silence target gene expression or function towards the control of disease (X. Chen, et al., 2018; Khorkova & Wahlestedt, 2017; Moschos, Usher, & Lindsay, 2017; Rupaimoole & Slack, 2017). Third, the sequence of a gene dictating disease initiation or progression may be directly altered by using a proper guide RNA (gRNA) and other necessary components to achieve a complete eradication of the disease for a patient (Doudna & Charpentier, 2014; Komor, Badran, & Liu, 2017; O'Day, et al., 2018; H. X. Wang, et al., 2017; F. Zhang, Wen, & Guo, 2014). Indeed, the promise of RNA therapeutics was revealed by the approval of first-of-its-kind mRNA-targeting patisiran (ONPATTRO™) for clinical practice by the United States Food and Drug Administration (FDA) in August, 2018 (Wood, 2018) (https://www.fda.gov/Drugs/DevelopmentApprovalProcess/DrugInnovation/ucm592464.htm, as well as other forms of nucleic acid drugs (Table 1).
Table 1.
Nucleic acid drugs approved by the United States Food and Drug Administration.
| Drug name |
Agent | Target | Clinical application |
Year Approved by FDA |
Reference |
|---|---|---|---|---|---|
| Fomivirsen (Vitravene) | 21mer ASO (phosphorothioate linkages) | Cytomegalovirus (CMV) mRNA | Cytomegalovirus retinitis | 1998 | (Roehr, 1998) |
| Pegaptanib (Macugen) | 27-nt aptamer (pegylated; 2’-fluoro or methoxyl) | Vascular endothelial growth factor (VEGF) protein | Age-related macular degeneration (AMD) | 2004 | (Gryziewicz, 2005) |
| Mipomersen (Kynamro) | 20mer ASO (all phosphorothioate linkages; 2’-O-(2-methoxyethyl); 5-methylcytosine; 5-methyluridine) | Apolipoprotein B mRNA | Familial hyper-cholesterolemia | 2013 | (Robinson, 2013) |
| Eteplirsen (Exondys 51) | 30mer ASO (phosphorodiamidate morpholino oligomer (PMO)) | Exon 51 of dystrophin pre-mRNA | Duchenne muscular dystrophy (DMD) | 2016 | (Aartsma-Rus & Krieg, 2017) |
| Nusinersen (Spinraza) | 18mer ASO (all phosphorothioate linkages plus 2’-O-(2-methoxyethyl) | Survival of motor neuron 2 (SMN2) mRNA | Spinal muscular atrophy (SMA) | 2016 | (Ottesen, 2017) |
| Patisiran (Onpattro) | 21-bp siRNA duplexes (2’-O-methyl; lipid complex) | Transthyretin (TTR) mRNA | Hereditary transthyretin-mediated amyloidosis (hATTR) | 2018 | (Wood, 2018) |
| Inotersen (Tegsedi) | 20mer ASO (all phosphorothioate linkages; 2’-O-(2-methoxyethyl); 5-methyluridine; 5-methylcytosine; sodium salt) | Transthyretin (TTR) mRNA | hATTR | 2018 |
https://www.fda.gov/Drugs/DevelopmentApprovalProcess/DrugInnovation/ucm592464.htm
(Keam, 2018) |
RNA therapeutics not only exhibit different mechanisms of actions but also distinct chemistry and pharmacokinetics properties, when compared to conventional small-molecule and protein therapeutics. As such, the development of novel RNA therapeutics has proven to be highly challenging and the past two decades have witnessed only a limited number of nucleic acid drugs approved for clinical use (Table 1). Since delivery is a well-recognized challenge for RNA research and drug development (Crooke, Wang, Vickers, Shen, & Liang, 2017; Dowdy, 2017; Kanasty, Dorkin, Vegas, & Anderson, 2013), utilization of the right RNA molecules has been overlooked for decades (Duan & Yu, 2016; P. Y. Ho & Yu, 2016). Currently, chemically engineered/synthesized oligonucleotides or RNA “mimics” dominate RNA research and drug development, and some have been successfully approved by the FDA for clinical practices (Table 1). However, these “mimics” are decorated with various and extensive chemical modifications (Bramsen & Kjems, 2012; Khvorova & Watts, 2017; Lundin, et al., 2013; Winkler, 2013), making them totally different from natural RNAs transcribed from the genome and folded in living cells that carry no or minimal posttranscriptional modifications (Cantara, et al., 2011; Limbach, Crain, & McCloskey, 1994; Morena, Argentati, Bazzucchi, Emiliani, & Martino, 2018). The use of chemo-engineered RNA mimics for research and development is also in sharp contrast to protein research and therapy (Leader, Baca, & Golan, 2008; Secher, et al., 2018), and the latter has proven to be highly successful by preferentially using biologic or recombinant proteins produced and folded in living cells rather than polypeptides or proteins synthesized in vitro via peptide chemistry. Therefore, large efforts have been made recently to develop novel biotechnologies for the production of biological/bioengineered RNA agents (BERAs) in living cells (Q. X. Chen, Wang, Zeng, Urayama, & Yu, 2015; P. Y. Ho, et al., 2018; Huang, et al., 2013; M. M. Li, et al., 2015; M. M. Li, Wang, Wu, Huang, & Yu, 2014; P. C. Li, et al., 2018; P. Pereira, et al., 2016; P. A. Pereira, Tomas, Queiroz, Figueiras, & Sousa, 2016), which should represent a new class of tools for RNA research and drug development (Duan & Yu, 2016; P. Y. Ho & Yu, 2016; P. Pereira, Pedro, Queiroz, Figueiras, & Sousa, 2017).
Herein we provide an overview on the promise of RNA therapeutics for the treatment of human diseases via RNA interference (RNAi) mechanisms. By comparing the structural differences between natural RNAs transcribed within cells and chemo-engineered RNAs or mimics made in vitro, we focus on novel biologic RNA agents or BERAs produced via RNA biotechnology and potential applications to research and RNA drug development.
2. RNA interference and relevant RNA agents
2.1. RNA interference
RNAi is an evolutionarily-conserved mechanism among eukaryotes in which ncRNAs control target gene expression at the post-transcriptional level. These RNAi molecules include miRNAs derived from the genome and siRNAs generated from exogenously-introduced double-stranded RNAs (dsRNAs) (Ambros, 2004; Bartel, 2009; Cech & Steitz, 2014; Fire, et al., 1998) (Fig. 2). The miRNA coding genes are initially transcribed by RNA polymerase II within the nucleus as primary miRNA transcripts, namely pri-miRNAs, which are subsequently processed to shorter precursor miRNAs (pre-miRNAs) by the ribonuclease (RNase) III termed Drosha (Fig. 2). On the other hand, some pre-miRNAs are directly excised from introns of protein coding genes. After being exported from nucleus into the cytoplasm by Exportin-5, pre-miRNAs are converted to double-stranded miRNA molecules by the cytoplasmic endoribonuclease Dicer. After being unwound from the miRNA duplexes or siRNAs derived from dsRNAs, the single-stranded guide miRNAs or siRNAs are loaded into the miRNA- or siRNA-induced silencing complexes (RISC or miRISC), and then selectively act on target mRNAs via perfect or imperfect base-pairing interactions, leading to target RNA degradation or translation repression (Fig. 2).
Fig. 2. Cellular RNA interference pathway and manipulation with various types of RNA agents.
Endogenous RNAi pathway involves the initial production of pre-miRNA molecules (1)from longer pri-miRNAs transcribed from miRNA coding genes or directly excised out of introns in the nucleus. Pre-miRNAs are then exported into the cytoplasm and subsequently processed to miRNA duplexes. After unwinding from the duplexes and incorporation into the RISC complex, miRNAs and siRNAs recognize target transcripts via imperfect or perfect base-pairing, leading to RNA cleavage or translation inhibition. The use of synthetic dsRNAs (2), siRNAs (3), asRNAs (4), and miRNA mimics (5), which all consist of various types and degrees of chemical modifications, may exert target gene knockdown via cellular RNAi pathway. Bioengineered, single-stranded miRNA or siRNA or sRNA “prodrugs” (6) produced from bacteria fermentation represent a novel family of “natural” RNAi molecules.
Through silencing or modulating target gene expression, RNAi is involved in almost all cellular processes including the defense against viral infection as well as cell transformation and disease progression (Ambros, 2004; Bartel, 2009; Cech & Steitz, 2014; Mendell &Olson, 2012; Rupaimoole & Slack, 2017). Compared to normal cells, some miRNAs or the whole miRNome profiles are apparently dysregulated in diseased cells to a dramatic degree, which may serve as helpful biomarkers for diagnosis or prognosis (X. Chen, et al., 2008; Landgraf, et al., 2007; Lu, et al., 2005). Furthermore, functional miRNAs critical for disease progression could be utilized for the development of novel therapeutic strategies (Bader, Brown, & Winkler, 2010; Rupaimoole & Slack, 2017; Yu, Tian, Tu, Ho, & Jilek, 2016). On one hand, miRNAs that are overexpressed in the diseased cells and promote disease initiation and development may be inhibited or silenced to achieve the control of disease. On the other hand, miRNAs that are depleted in diseased cells, which are actually able to suppress disease initiation and deterioration, may be re-introduced into the cells to manage disease progression.
2.2. Relevant RNAi agents
The discovery of the RNAi mechanism, genome-derived regulatory miRNAs and their importance in human diseases has allowed investigators to develop various forms of RNA molecules to manipulate the RNAi process or mimic miRNA action towards the intervention of target gene expression, understanding of gene function, and control of cellular processes and disease (Bader, et al., 2010; Lares, Rossi, & Ouellet, 2010; Rupaimoole & Slack, 2017; Yu, et al., 2016). Traditional RNAi modulators include synthetic asRNAs, dsRNAs, siRNAs, and miRNA mimics (Fig. 2), all of which carry various types and degrees of chemical modifications expected to improve overall RNA metabolic stabilities (Bramsen & Kjems, 2012; Khvorova & Watts, 2017; Winkler, 2013). Most recently, “natural” RNAi molecules or BERAs have been introduced into human cells for the control of target gene expression (Fig. 2), which are produced and folded in bacteria through newly-developed RNA bioengineering technologies (Duan & Yu, 2016; P. Y. Ho & Yu, 2016).
BERAs are also distinguished from non-viral or viral-based miRNA or short-hairpin RNA (shRNA) expression systems (Fig. 3) that are widely used for in vitro and in vivo RNAi research and drug development (Brake, et al., 2008; Czauderna, et al., 2003; Y. P. Liu & Berkhout, 2011). Those plasmids and vectors are really DNA molecules. Most importantly, use of such vectors/plasmids may complicate the RNA-based mechanisms because they are dependent upon other factors as well, including the efficiency in the delivery of DNA materials into the nucleus, the integration of target shRNA- or pre-miRNA-coding sequences into the host cell’s genome, and the transcription of coding sequences by host cell into target shRNA or pre-miRNA agents before shRNAs or pre-miRNAs are processed to target RNAi molecules to exert target gene silencing (Fig. 3). Additionally, it is unknown whether and to what degree the DNA materials might cause any side effects.
Fig. 3. Use of viral vectors or DNA plasmids for RNA interference.
DNA plasmids (1) and genetically-modified DNA viral vectors (e.g., adenovirus) (2) or RNA viral vectors (e.g., lentivirus) (3) may be employed for the production of target pre-miRNA or shRNA molecule (1) within the nucleus of host cell, before which the target coding sequence need be integrated into the host cell’s genome. Thus, transcribed shRNAs or pre-miRNAs (1) enter the cellular RNAi pathway; and the resultant siRNAs or miRNAs act on specific transcripts to exert target gene silencing.
3. RNA therapeutics in clinical practice and trials
3.1. RNA therapeutics currently in clinical practice
A number of RNA-based drugs have been approved by the FDA for the treatment of various types of human diseases (Table 1). Fomivirsen is the first antisense oligonucleotide (ASO) drug (literally a DNA molecule; different from other ASO drugs), which was approved by FDA in 1998 for the treatment of cytomegalovirus (CMV) retinitis among immunocompromised patients including those with acquired immune deficiency syndrome (AIDS) (Roehr, 1998), whose antiviral effects are produced via direct targeting of the CMV mRNA (Orr, 2001). Since 2013, a number of other ASO drugs have been successfully marketed in the United States (Aartsma-Rus, 2017; Aartsma-Rus & Krieg, 2017; Keam, 2018; Ottesen, 2017; Robinson, 2013; Stein & Castanotto, 2017; Syed, 2016) (https://www.fda.gov/Drugs/DevelopmentApprovalProcess/DrugInnovation/ucm592464.htm) which all target mRNA molecules in cells. Furthermore, the approval of pegaptanib for the management of neovascular age-related macular degeneration (AMD) (Gryziewicz, 2005) supports the potential of using RNA aptamer to inhibit protein macromolecule for the control of disease. In addition, the most recent approval of first-of-its-kind double-stranded siRNA patisiran by the FDA (Wood, 2018), which was shown to improve multiple clinical manifestations of hereditary transthyretin amyloidosis (Adams, et al., 2018), provides incentives to encourage the development of RNAi based therapeutics.
3.2. RNA therapeutics under clinical investigations
Many RNAi based therapeutics, including ten siRNAs, two ASOs, one aptamer and one siRNA-transfected peripheral blood mononuclear cells, are currently under active recruiting clinical trials for patients with cancers, infectious diseases, genetic or other disorders (Table 2), beyond those trials completed already. For instance, intravenously administered AZD4785, an ASO targeting the mRNA of oncogene KRAS (Ross, et al., 2017), has entered into Phase I trial for the treatment of non-small cell lung cancer (NSCLC) and other advanced solid tumors. On the other hand, siG12D-LODER is a siRNA that specifically targets the mutant KRAS G12D mRNA (Golan, et al., 2015) which is quite common in pancreatic cancers. After the Phase I clinical trial showing that combination of siG12D-LODE and gemcitabine was well tolerated and exhibited a potential efficacy as targeted therapy for locally advanced pancreatic cancer patients (Golan, et al., 2015), siG12D-LODER is currently under Phase II clinical trial for patients with unresectable pancreatic cancers (Table 2). QPI-1002, a synthetic double-stranded siRNA designed to temporarily inhibit the expression of proapototic gene P53 showing no dose-limiting toxicities or safety signals in early trials (Demirjian, et al., 2017), is presently under Phase III clinical trials for the treatment of acute kidney injury and delayed graft function (Table 2). In addition, an antithrombin-targeting siRNA fitusiran, which, administered once-monthly, showed a dose-dependent lowering of the antithrombin level and increased thrombin generation in participants with hemophilia A or B who did not have inhibitory alloantibodies (Pasi, et al., 2017). As such, fitusiran has entered into a Phase III clinical trial for the treatment of hemophilia A and B (Table 2).
Table 2. Some RNAi based therapeutics currently under active recruiting clinical trials.
Data were collected from ClinicalTrials.gov database (https://clinicaltrials.gov/ct2/home) by August 16, 2018. IV, intravenous infusion/injection; SC, subcutaneous; ID, intradermal; EUS, endoscopic ultrasound; PBMC, peripheral blood mononuclear cells; MiHA, minor histocompatibility antigens; CEP290, centrosomal protein of 290 kDa; EphA2, ephrin type-A receptor 2; CBL-B, Casitas B-lineage lymphoma proto-oncogene-b; CTGF, connective tissue growth factor; TRPV1, The transient receptor potential cation channel subfamily V member 1; TGF-β1, Transforming growth factor beta 1; Cox-2, Cyclooxygenase-2; KRAS-G12D, K-ras G12D mutant; NSCLC, non-small cell lung cancer.
| Drug name | Agent/formulation | Target(s) | Condition or disease | Route of administration | Phase | Status | ClinicalTrials.govidentifier | Reference |
|---|---|---|---|---|---|---|---|---|
| QR-1 10 | Antisense oligonucleotide | CEP290 c.2991+1655A>G Mutation (p.Cys998X) | Leber’s Congenital Amaurosis (LCA) | Intravitreal | I/II | Recruiting | NCT03140969 | (Jacobson, et al., 2017) |
| AZD4785 | Antisense oligonucleotide | KRAS | NSCLC; Advanced solid tumors | IV | I | Recruiting | NCT03101839 | (Ross, et al., 2017) |
| Zimura (ARC1905) | Aptamer | Complement 5 (C5) protein | Geographicatrophy secondary to dry age-related macular degeneration | Intravitreous | II | Recruiting | NCT02686658 | (Leung & Landa, 2013) |
| EPHARNA | siRNA (neutral liposomal) | EphA2 | Advanced recurrent solid tumors | IV | I | Recruiting | NCT01591356 | (Wagner, et al., 2017) |
| OLX100010 | siRNA | CTGF | Cicatrix or hypertrophic scars (Safety and PK profiles in healthy subjects) | SC or ID | I | Recruiting | NCT03569267 | (Hwang, et al., 2016) |
| SYL1001 | siRNA | TRPV1 | Moderate to severe dry eye disease (DED) | Topica 1 drop | III | Recruiting | NCT03108664 | (Benitez-Del-Castillo, et al., 2016) |
| DCR-PHXC | siRNA, conjugated to N-acteylgalactosamine (GalNAc) | lactatedehydrogenase A (LDHA) | Primary hyperoxaluria | SC | I | Recruiting | NCT03392896 | http://dicerna.com/pipeline/novel-investigational-drug-ph-dcr-phxc/ |
| STP705 | Two siRNAs (Histidine-Lysine co-Polymer (HKP) peptide) | TGF-β1 and Cox-2 | Hyper trophic scar | ID | I/II | Recruiting | NCT02956317 | (Zhou, et al., 2017) |
| SiG12D LODER | siRNA | KRASG12D | Unresectable pancreatic cancer | Intratumor implantation via EUS | II | Recruiting | NCT01676259 | (Golan, et al., 2015) |
| QPI-1002 | Naked siRNA | P53 | Acute kidney injury following cardiac surgery | IV | III | Recruiting | NCT03510897 | (Demirjian, et al., 2017) |
| QPI-1002 | Naked siRNA | P53 | Delayed graft function | IV | III | Recruiting | NCT02610296 | (Demirjian, et al., 2017) |
| QPI-1007 | siRNA | Caspase-2 | Acutenonar teriticanteri or ischemicoptic neuropathy (NAION) | Intravitreal | II/III | Recruiting | NCT02341560 | (Solano, et al., 2014) |
| Fitusiran | siRNA (lipid nanoparticle) | Antithrombin | Hemophilia A; Hemophilia B | SC | III | Recruiting | NCT03417245 | (Pasi, et al., 2017) |
Clinical investigations also include the evaluation of miRNAs as therapeutics. As an example, numerous preclinical studies have demonstrated miR-34a as a tumor suppressor against various types of cancers through targeting of many oncogenic pathways (see review (Bader, 2012)). A Phase I clinical study on the safety, pharmacokinetics and effectiveness of a liposome-encapsulated synthetic miR-34a mimic, namely MRX34, was conducted in patients with advanced solid tumors (Beg, et al., 2016). While the maximum tolerated dose was 110 mg/m2 in non-hepatocarcinoma patients, that study revealed a rather high incidence of adverse events (e.g., 100% all grades and 38% Grade 3 among all patients) such as fever, fatigue, back pain, nausea, diarrhea, anorexia, and vomiting among patients receiving MRX34 treatment (10-50 mg/m2, i.v., biweekly), which required palliative management with dexamethasone premedication (Beg, et al., 2016). Nevertheless, a blood half-life of over 24 h and dose-dependent exposure (e.g., Cmax and AUC) as well as evidence of antitumor activity among patients with refractory advanced solid tumors was identified for MRX34, which would provide valuable insights into developing miRNA therapeutics (Beg, et al., 2016).
3.3. Challenges in developing RNA therapeutics
RNA therapeutics differs fundamentally from small-molecule and protein drugs in many aspects including the nature of molecules, targets and mechanisms (Fig. 1), while they all need to address the ultimate common questions regarding the efficacy and safety to control disease. Therefore, there are indeed some unique challenges for the development of RNA based therapies. The druggability of a therapeutic target, either protein, RNA or DNA molecule (Fig. 1), as well as the effectiveness, selectivity, and safety of RNA based interventions often require extensive evaluations. Targeting of a transcript within cells with an RNA molecule has proved to be a major challenge in developing RNA therapeutics (Crooke, et al., 2017; Dowdy, 2017; Kanasty, et al., 2013) because that RNA agent has to cross the cellular barriers, in addition to necessary blood stability. By contrast, current macromolecule protein therapeutics mainly act on cell membrane or surface protein targets, and many small-molecule drugs readily cross cell membrane barriers to access cytoplasmic protein targets. Rather, chemical modifications and liposomal formulations (Table 1) are able to improve the stability of RNAs in blood and deliver sufficient RNA molecules to cross cell membranes to elicit the desired pharmacological effects, and thus both approaches remain as popular strategies (Bramsen & Kjems, 2012; X. Chen, et al., 2018; W. Ho, Zhang, & Xu, 2016; Khvorova & Watts, 2017).
Given multiple mechanisms occurring in cellular defense against xenobiotic RNAs (Dalpke & Helm, 2012; Robbins, Judge, & MacLachlan, 2009; Tanji, et al., 2015), pharmacological RNA agents may induce an immunogenic response or cytokine release syndrome, which is unsurprisingly dependent upon the doses and structures (e.g., sizes, sequences, etc.) of RNA molecules. Interestingly, the types of chemical modifications have significant influence on RNA immunogenicity and thus can determine the overall safety profiles (Bramsen & Kjems, 2012; Robbins, et al., 2009). By contrast, some RNA posttranscriptional modifications may suppress immune response (Gehrig, et al., 2012; Kariko & Weissman, 2007; Nallagatla, Toroney, & Bevilacqua, 2008), in addition to their importance in RNA folding, stability and biologic functions. In addition, the complexity of RNA therapy is also increased by the safety of RNA delivery system even if the delivery vehicle itself is not toxic.
4. Natural RNAs versus chemo-engineered RNA mimics
4.1. Natural RNAs carrying no or minimal posttranscriptional modifications
RNAs are polymeric molecules comprised of different numbers and combinations of four major forms of ribonucleotide monomers distinguished by their corresponding nucleobases, adenine (A), guanine (G), cytosine (C), and uracil (U) (Fig. 4). An RNA molecule structurally differs from DNA in the presence of a hydroxyl group at the 2’ position of each ribose, besides the uracil base rather than thymine (T) (Fig. 4). While the primary assemblage of monomeric nucleotides or primary sequence of a particular RNA is undoubtedly critical for the function of that RNA molecule, folding into proper elaborate secondary (e.g., helices or stems, loops, bulges, etc.), tertiary (e.g., junctions, pseudoknot, motifs, etc.), and quaternary (complexes, etc.) structure via Watson-Crick complementary base pairs or other types of physicochemical interactions (Bai, Dai, Harrison, Johnston, & Chen, 2016; Butcher & Pyle, 2011; Jones & Ferre-D'Amare, 2015; Schlick, 2018) ultimately determines RNA global structure that governs its stability, plasticity, interactions with partners or ligands, and biological functions and safety profiles in cells. It is also noteworthy that the ribose and the unconjugated 2’-hydroxyl group are key elements for the higher-order structures of natural RNA molecules (Butcher & Pyle, 2011).
Fig. 4. Chemical structures of unmodified cellular nucleic acids and some postranscriptionally-modified nucleosides.
Different from DNA, an RNA molecule has a hydroxyl group at the 2’ position of each ribose, as well as the uracil base rather than thymine. Natural RNAs made in living cells are generally unmodified nucleic acids, and there is just a small fraction (e.g., < 3%) of ribonucleosides carrying posttranscriptional modifications which actually show a broad chemical diversity. Among these modifications, methylation is the most common form which may occur at both ribose and nucleobases. It is also noteworthy that substantial posttranscriptional modifications are present at the nucleobases.
While RNA molecules produced in living cells are usually comprised of unmodified ribonucleotides, some RNAs (e.g., tRNAs, rRNAs, and mRNAs) do carry a small fraction (e.g., < 3%) of modified nucleosides which exhibit rather a broad chemical diversity (Fig. 4) (Cantara, et al., 2011; Limbach, et al., 1994). Pseudouridine (Ψ), an isomer of uridine, is the first modified natural ribonucleotide discovered and has been recognized as the fifth ribonucleotide because of its ubiquitous prevalence within cellular RNAs (X. Li, Ma, & Yi, 2016; Spenkuch, Motorin, & Helm, 2014). Methylation is revealed as a major form of modification occurring at both ribose (e.g., 2’-O-methyladenosine or Am; 2’-O-methylguanosine or Gm; 2’-O-methylcytidine or Cm; 2’-O-methyluridine or Um; 2’-O-methylpsudouridine or Ψm) and nucleobases (e.g., 1-methylpsudouridine or m1Ψ; 5-methyluridine or m5U; 1-methyladenosine or m1 A; 1-methylguanosine or m1G; 3-methylcytidine or m3C; etc.; Fig. 4). There are also other types of modifications including acetylation and hydroxylation as well as complex modifications, which leads to a variety of minor bases including N4-acetylcytidine (ac4C), 5-hydroxymethylcytidine (hm5C), dihydrouridine (D), N2,7-dimethylguanosine (m2,7G). Interestingly, substantial RNA posttranscriptional modifications occur at the nucleobases while 2’-O-m ethylation and other types of modifications are present at the ribose (Fig. 4). Following posttranscriptional modifications, these RNA molecules exhibit unique chemical structures and physicochemical properties as well as biological functions and interactions with innate immune receptors. Indeed, RNA modification has been recognized as another layer of epigenetic mechanism in biology beyond DNA and protein modifications (El Yacoubi, Bailly, & de Crecy-Lagard, 2012; N. Liu & Pan, 2015; Morena, et al., 2018), although modified miRNAs (Alon, et al., 2012; Kawahara, Zinshteyn, Chendrimada, Shiekhattar, & Nishikura, 2007; Luciano, Mirsky, Vendetti, & Maas, 2004; Shoshan, et al., 2015) are less common in cells, and miRNAs are generally stable within cells and extracellular vesicles in systemic circulation (Mitchell, et al., 2008; Sethi & Lukiw, 2009; Winter & Diederichs, 2011).
4.2. Chemo-engineered RNA agents bearing various types and extensive degrees of chemical modifications
Many types of chemical modifications (Fig. 5) have been developed to improve the diversity of RNAs and enhance the stability of synthetic RNA molecules (Bramsen & Kjems, 2012; Khvorova & Watts, 2017; Winkler, 2013), which have been dominating RNAi research and drug development. In contrast to natural RNA modifications that occur predominantly at the nucleobases (Fig. 4), as described above, synthetic RNA agents produced by chemical synthesis are usually altered at the phosphate linkage and ribose (Fig. 5) that are more susceptible to metabolic degradation. Change of phosphate backbone such as phosphorothioate (PS) (Kawasaki, et al., 1993) may increase the metabolic stability of the resultant chemo-engineered nucleic acid molecule and thus a prolonged half-life. Phosphothioate linkage modification has found its success in fomivirsen, the first antisense ASO drug approved by the FDA for clinical practice, as well as the other four ASOs approved since 2013 (Table 1). Masking or substituting the so-called vulnerable 2’-hydroxy group at the ribose is another popular strategy to improve RNA stability, which includes natural modification at the 2’-O-methyl group and many other types of unnatural alterations such as 2’-fluoro, 2’-O-methoxyethyl, 2'-O-4'-C-methylene bridged or locked nucleic acid (LNA) (Kurreck, Wyszko, Gillen, & Erdmann, 2002; Lundin, et al., 2013) (Fig. 5). Indeed, FDA-approved aptamer (pegaptanib), ASO (mipomersen, nusinersen, and inotersen), and siRNA (patisiran) are all comprised of different types and degrees of sugar modifications (Table 1), supporting the utility of chemical modifications at the ribose.
Fig. 5. Common chemical modifications used for the production of nucleic acid reagents including aptamers, ASOs, siRNAs, miRNA mimics, sgRNAs, tRNA fragments (tRFs), and other types of sRNAs.
Chemical modifications are mainly aimed at improving the pharmacokinetics properties of RNA reagents. These synthetic RNA mimics are characterized by the changes predominantly at the phosphate linkage and ribose vulnerable to RNase-mediated degradation, which is in contrast to cellular RNA modifications mainly occurring at the nucleobases (Fig. 4). Many RNA conjugates are also developed to enhance targeting, and the advanced modifications completely change the linkage moieties and just retain nucleobases for pairing.
Nucleic acids could also be stabilized through 5’-phosphate modifications such as 5’-(E)-vinylphosphoate (E-VP), 5’-methylphosphate and 5’-phosphorothioate (Elkayam, et al., 2017; Haraszti, et al., 2017; Kuimelis & McLaughlin, 1995; Parmar, et al., 2016; Shumyatsky, Wright, & Reddy, 1993) (Fig. 5). There are also increasing interests in the development of 5’- or 3’-conjugates where a particular ligand or drug or targeting molecule (e.g. folate, N-acetylgalactosamine or GalNAc, etc.) is connected to the RNA molecule through a versatile linker (Dohmen, et al., 2012; Foster, et al., 2018; Matsuda, et al., 2015; Nair, et al., 2014; Winkler, 2013) (Fig. 5), which is expected to enhance RNA pharmacokinetics or pharmacodynamics properties. More advanced modifications include phosphorodiamidate morpholine oligonucleotide (PMO) (Iversen, 2001; Summerton & Weller, 1997) and peptide nucleic acid (PNA) (Nielsen, 2005; Nielsen, Egholm, Berg, & Buchardt, 1991; Wu, et al., 2017) where natural bases are retained within oligomer for base pairing and the ribose 5-phosphate is completely replaced (Fig. 5). Eteplirsen is a successful PMO that was approved by FDA in 2016 for clinical management of Duchenne muscular dystrophy (Table 1).
While chemical modifications may lead to favorable pharmacokinetics properties (e.g., longer half-life) which have found their success in FDA-approved drugs (Table 1) and remain as major tools for RNAi research and drug development (Fig. 2), chemo-engineered “RNA mimics” are literally different molecules as compared to natural RNA molecules produced in living cells. With distinct functional groups or moieties, chemo-engineered oligomers undoubtedly have their own secondary and higher order structures as well as intrinsic chemical and biological properties. It is also inadvertently or inevitably overlooked that synthetic RNAi agents obtained from different manufacturers vary largely in their types (e.g., backbone versus sugar; 2’-methoxyethyl versus LNA; etc.), specific sites/positions (e.g., uridine at position 2 versus 3; nucleotides at positions 1-5 versus 11-15; etc.), and degrees (e.g., 100% or parts of the nucleotides; all or partial phosphate linkage; etc.) of chemical modifications. Variable modifications, which may or may not be disclosed, undoubtedly lead to distinct RNAi molecules with different structures and properties, and likely some controversial or variable results despite that they are assumed to have the same “primary sequence” with nitrogenous bases for pairing. Furthermore, depending upon the doses, some chemical modifications might induce immunogenicity (Bramsen & Kjems, 2012; Dalpke & Helm, 2012; Hornung, et al., 2005; Judge, et al., 2005; Robbins, et al., 2009; Tanji, et al., 2015), whereas natural posttranscriptional modifications may suppress immune response (Gehrig, et al., 2012; Kariko & Weissman, 2007; Nallagatla, et al., 2008). In addition, although automated synthesis has facilitated the production, milligrams of oligonucleotides necessary for animal and human trials remain less affordable for common laboratories or investigators, and RNAi drugs for patient care are among the most expensive drugs on the market (Burgart, et al., 2018; Simoens & Huys, 2017; Stein, 2016). The involuntary use of synthetic RNAi agents is also in sharp contrast to protein research and development that preferably utilize biologic or recombinant proteins made and folded in living cells, rather than synthetic proteins, which has led to the ultimate success of protein therapeutics (Dimitrov, 2012; Leader, et al., 2008; Secher, et al., 2018; Sliwkowski & Mellman, 2013; Trail, et al., 2018). Therefore, new approaches to producing biologic RNA agents in living cells may open new avenues for RNA research and drug development (P. Y. Ho & Yu, 2016).
5. Bioengineered RNA molecules
5.1. General features of RNA molecules bioengineered in living cells
RNAi agents may be produced in the same way as the production of various recombinant/bioengineered DNA and protein macromolecules in living cells or organisms (e.g., bacteria, yeast and mammalian cells) by a regular biomedical research laboratory. A high-yield and large-scale RNA bioengineering technology (e.g., mg pure RNA from 1 L fermentation) will make the resultant biologic RNA agents or BERAs more affordable (Table 3). Because recombinant RNAs are produced in living cells, BERAs carry no or minimal posttranscriptional modifications (M. M. Li, et al., 2015; W. P. Wang, et al., 2015), similar as natural RNAs but largely different from chemo-engineered RNA mimics with extensive unnatural modifications. Folded in cells, BERAs may exhibit intrinsic higher order structures necessary for their stabilities and biological functions within cells (Ponchon & Dardel, 2007; Ranaei-Siadat, et al., 2014; W. P. Wang, et al., 2015) despite that naked BERAs were revealed to be still susceptible to metabolism/degradation by serum RNases (W. P. Wang, et al., 2015). In addition, although bioengineered RNA molecules derived from living cells were well tolerated in animal models (P. Y. Ho, et al., 2018; Huang, et al., 2013; Jian, et al., 2017; Tu, et al., 2018; W. P. Wang, et al., 2015; Zhao, et al., 2016; Zhao, et al., 2015), more extensive studies are highly warranted to evaluate their utilities in RNA therapy, whereas a number of chemo-engineered RNA mimics are already in clinical practice (Tables 1 and 3). Nevertheless, the barrier of RNA vulnerability in hosting cells has been overcome to achieve heterogeneous expression of significant levels of target RNA molecules, and BERAs have been successfully produced by novel RNA bioengineering technologies (P. Y. Ho & Yu, 2016). Among these techniques, the tRNA/pre-miRNA-based approach (Q. X. Chen, et al., 2015; P. Y. Ho, et al., 2018) was revealed as a versatile and efficient means for consistent high-yield production of a variety of RNAi agents including miRNAs, siRNAs, aptamers and other types of sRNAs.
Table 3. Comparison of biologic RNA molecules produced in living cells and synthetic RNA agents made by chemical synthesis.
Compared to traditional RNA mimics made by chemical synthesis that are successfully applied to clinical practice, new bioengineered RNAs produced and folded in living cells may better represent the physicochemical and biological properties of natural RNA molecules whose utilities awaits extensive investigations.
| Biologic RNAs (made in living cells) | Synthetic RNAs (by chemical synthesis) |
|---|---|
| Large scale (mg pure RNA per L culture) | Large scale (automated) |
| More affordable | Less affordable; expensive with increased size |
| No or minimal posttranscriptional modifications | Extensive unnatural & some natural modifications |
| Variable lengths (e.g., 20-300 nt) | Variable, with desirable short length (e.g., <60 nt) |
| Folded in living cells | Under chemical environment |
| Tolerated by cells; safety need more studies | Size, sequence and modifications affect safety |
5.2. Fully-processed siRNAs
One way to produce BERAs is to employ RNA-binding protein to protect heterogeneously-expressed RNAs against RNase-mediated degradation. As opposed to the eukaryotic RNAi mechanism against viral infection aforementioned, counter-measures are also evolved among viruses to combat RNAi by producing molecules to repress RNA silencing (Ding & Voinnet, 2007; F. Li & Ding, 2006; Voinnet, Pinto, & Baulcombe, 1999). The p19 protein (Voinnet, et al., 1999) was revealed as an RNAi “suppressor” expressed by the plant RNA virus tombusviruses, which exhibits high affinity and selective binding to double-stranded siRNAs (Silhavy, et al., 2002). Therefore, viral p19 protein functions to sequester sRNAs and disrupt eukaryotic RNAi process. The p19 protein was successfully used for the isolation and detection of sRNAs (Jin, Cid, Poole, & McReynolds, 2010). Overexpression of recombinant p19 protein in E. coli was recently shown to protect double-stranded siRNA species against hydrolysis by RNases and thus allowed enrichment of siRNA molecules (Huang, et al., 2013). Heterogeneous co-expression of p19 and target shRNA in bacteria offered p19 complexed with fully-processed siRNA of interest. Following the isolation of the p19-siRNA complex, the target siRNA was successfully purified by a high-performance liquid chromatography (HPLC) method (Huang, et al., 2013). While p19-enriched, fully-processed siRNAs were active in silencing target gene expression in mammalian cells, the overall yield of target siRNAs was rather low (e.g., about 40 μg from one liter bacterial culture) (Huang, et al., 2013), likely related to the limited levels of p19 protein and shRNA expressed in bacteria and/or low capacity of p19 in loading target siRNA molecules.
5.3. Chimeric RNAs
Another strategy to produce recombinant RNAs is to integrate target sRNA molecules into proper ncRNA scaffolds such as tRNA (Ponchon, Beauvais, Nonin-Lecomte, & Dardel, 2009; Ponchon & Dardel, 2007) that are relatively more stable within cells. Overexpression of recombinant methionine-tRNA (tRNAmet) in E. coli was first reported in 1988 to produce a few mg of tRNAmet from one liter bacterial culture after HPLC purification (Meinnel, Mechulam, & Fayat, 1988). The same levels of production were shown for the transfer-messenger RNA (tmRNA, 10Sa RNA, or ssrA) that exhibit both tRNA and mRNA properties (Gaudin, et al., 2003). Thus, tRNA was established as a scaffold for large-scale production of recombinant RNA molecules (e.g., multi-milligrams of RNAs from one liter bacterial culture) where the anticodon sequence was replaced by target RNA which may be further excised enzymatically from the chimeric RNA (Ponchon, et al., 2009; Ponchon & Dardel, 2007). The tRNA-based strategy was used for the production of various target RNA molecules including a number of RNA aptamers, hammerhead riboswitch RNAs, pre-miRNAs, and RNA-protein complexes for structural and functional studies (M. M. Li, et al., 2015; M. M. Li, et al., 2014; Nelissen, et al., 2012; Paige, Nguyen-Duc, Song, & Jaffrey, 2012; Paige, Wu, & Jaffrey, 2011; Ponchon, et al., 2009; Ponchon, et al., 2013; Ponchon & Dardel, 2007; W. P. Wang, et al., 2015). It was also demonstrated that bioengineered tRNA molecules are consisted of some postranscriptionally-modified nucleosides (M. M. Li, et al., 2015; W. P. Wang, et al., 2015). Nevertheless, the expression levels of chimeric RNAs were revealed to be highly variable (ranging from tens milligrams to micrograms recombinant RNAs per liter bacterial culture), and the majority of target chimeras were not even expressed (Q. X. Chen, et al., 2015; M. M. Li, et al., 2015; M. M. Li, et al., 2014; Nelissen, et al., 2012; Ponchon, et al., 2009; Ponchon & Dardel, 2007), indicating that the structure and stability of the tRNA-bearing chimera determines the expression level (P. Y. Ho, et al., 2018).
Another usable scaffold is 5S rRNA, a member of highly abundant rRNA molecules in cells and relatively larger in size than tRNA, which may accommodate target RNA sequences at different sites (e.g., stem II and III) (Y. Liu, et al., 2010; X. Zhang, et al., 2009). An RNA-based carrier was initially derived from the 5S rRNA harboring an “identifier” nucleic acid that might be utilized as a “biomarker” (Pitulle, Hedenstiema, & Fox, 1995b) due to the lack of specific sequences in rRNA for the monitoring of genetically engineered microorganisms. To further optimize the insert sequence composition and length, a collection of hybrid rRNA/insertion entities consisting of random 13- and 50-base oligonucleotides were created and then introduced into E. coli and several Pseudomonas strains. Almost all of the rRNA chimeras were expressed at detectable levels using Northern blots (D'Souza, Larios-Sanz, Setterquist, Willson, & Fox, 2003), despite that the exact expression levels were not reported. Thus, this rRNA-based system was successfully used for production of functional aptamers by in vivo fermentation (X. Zhang, et al., 2009). The addition of DNAzyme-specific sequences further offered the option of on-demand release of target inserts (Y. Liu, et al., 2010). In particular, DNAzymes could be used to release the desired RNA fragments (e.g., 71 nt) from 5S rRNA chimera (e.g., 160 nt) purified by preparative polyacrylamide gel electrophoresis. This approach yielded 2 mg of target RNA per gram of wet bacterial cells (Y. Liu, et al., 2010). Although there are relatively less research reports, rRNA scaffold holds great promise for the bioengineering of new RNA molecules.
A novel RNA bioengineering strategy was recently developed to offer more consistent, high-yield and large-scale production of biologic RNAi molecules (Q. X. Chen, et al., 2015; P. Y. Ho, et al., 2018; M. M. Li, et al., 2015; W. P. Wang, et al., 2015), based upon stable tRNA/pre-miRNA chimeric carriers (Fig. 6). Although bare pre-miRNAs could be heterogeneously expressed in bacteria, the levels were usually low (e.g., < 2% of recombinant pre-miR-34a among total RNAs or μg from 1 L bacterial culture), which may hinder purification and overall yields (P. Y. Ho, et al., 2018; P. Y. Ho & Yu, 2016; P. Pereira, et al., 2016; P. Pereira, et al., 2014; P. A. Pereira, et al., 2016). On the other hand, tRNA alone could not be overexpressed in bacteria (P. Y. Ho, et al., 2018). As the assembly of tRNA with pre-miRNA was hoped to increase the expression level, the majority of tRNA/pre-miRNA chimeras were still expressed at low levels (e.g., < 2% of total RNAs) or not expressed at all (Q. X. Chen, et al., 2015; M. M. Li, et al., 2014). Following the identification of several hybrid tRNA/pre-miRNAs (e.g., tRNA/pre-miR-34a, tRNA/pre-miR-1291, etc.) that were able to accumulate in bacteria to significantly greater levels (e.g., 5-20% of total RNAs) (Q. X. Chen, et al., 2015; M. M. Li, et al., 2015; W. P. Wang, et al., 2015), these tRNA/pre-miRNA chimeras were established as novel carriers permitting consistent high-yield production of various RNAi agents, where the miRNA sequences may be substituted by target miRNA or siRNA duplexes (Fig. 6). In addition, a single stranded sRNA molecule such as RNA aptamer and asRNA may be directly docked into the stable tRNA/pre-miRNA carrier to achieve overexpression (Q. X. Chen, et al., 2015; P. Y. Ho, et al., 2018; P. C. Li, et al., 2018), and optimization of a particular pre-miR-34a sequence sharply increased the expression yield (e.g., 40-80% of total RNAs).
Fig. 6. The workflow for the production of biologic/bioengineered RNAi agent (BERA) where a target miRNA/siRNA/sRNA is assembled into a tRNA/pre-miRNA carrier.
After a target BERA is designed, corresponding coding sequence is cloned into a vector. Expression of target BERA in fermentation may be verified by RNA gel electrophoresis, and BERA can be purified to a high degree of homogeneity by using different methods (e.g., anion exchange FPLC). Purity of isolated BERA is determined by HPLC analysis and endotoxin pyrogen testing. These BERAs should better capture the properties of cellular RNAs and represent a novel class of RNAi agents for basic research (e.g., structural and functional studies) and drug development.
The workflow of RNA bioengineering using the tRNA/pre-miRNA carrier (Fig. 6) is similar as protein bioengineering. Following the design of a target BERA, its corresponding coding sequence is cloned. Expression and accumulation of target BERA in transformed E. coli may be monitored through denaturing urea polyacrylamide gel electrophoresis analysis of total bacterial RNAs (Fig. 6). It was also demonstrated that a common E. coli strain (e.g., HST08) offered the greatest levels of heterogeneous expression (Q. X. Chen, et al., 2015; W. P. Wang, et al., 2015), and thus BERAs are readily purified to a high degree of homogeneity (> 97%) at a high yield (multiple to tens milligrams from one liter bacterial culture) using a variety of methods including affinity chromatography and anion exchange fast protein liquid chromatography (FPLC) (Q. X. Chen, et al., 2015; P. Y. Ho, et al., 2018; M. M. Li, et al., 2015; P. C. Li, et al., 2018; W. P. Wang, et al., 2015). Pure BERAs may be processed for structural analyses as well as functional studies and translational research such as mechanistic actions in the regulation of target gene expression and potential for therapy.
The tRNA/pre-miRNA-based platform was proven to offer high-yield and large-scale production of many miRNAs (e.g., miR-124, miR-27b, and miR-22, etc.), siRNAs (e.g., siRNA against green fluorescent protein (GFP), etc.), and RNA aptamers (e.g., malachite green aptamer (MGA), and vascular endothelial growth factor (VEGF) aptamer, etc.) (Q. X. Chen, et al., 2015; P. Y. Ho, et al., 2018; P. C. Li, et al., 2018). These miRNAs and siRNAs are very different in size (e.g., 20-22 nt) and arm of origin (5’ or 3’), while single-stranded RNA aptamers or sRNAs may be directly assembled to either 5’ or 3’ of pre-miRNA, demonstrating that the tRNA/pre-miRNA platform is robust and versatile to accommodate various forms RNAi molecules of interest. It is also noteworthy that, different from fully-processed biological siRNAs produced with RNA-binding protein p19 (Huang, et al., 2013), chimeric RNAi agents act as “prodrugs” and utilize mammalian cellular RNAi machinery for processing and actions as discussed below.
6. Applications and promise of bioengineered RNAi molecules
6.1. Structural studies
Distinguished from synthetic RNAi agents bearing extensive chemical modifications that may have distinct physicochemical and biological properties, BERAs made and folded in living cells are “natural” RNA molecules without any chemical modifications or just carrying necessary posttranscriptional modifications critical for higher order structures (M. M. Li, et al., 2015; Ponchon, et al., 2009; Ponchon & Dardel, 2007; W. P. Wang, et al., 2015). A variety of techniques such as mass spectrometry (M. M. Li, et al., 2015; W. P. Wang, et al., 2015), magnetic resonance (Gaudin, et al., 2003; Nelissen, et al., 2012; Ponchon & Dardel, 2007), and X-ray crystallography (Gaudin, et al., 2003) were used to study their primary sequences, natural modifications and secondary and high-order structures. Posttranscriptional modifications learned from BERAs not only illustrate the unique structural characteristics of biologic RNA molecules but also provide insights into RNA epigenetics.
6.2. Functional studies
Bioengineered RNA aptamers and ribozymes have been shown to be biologically active (Q. X. Chen, et al., 2015; Nelissen, et al., 2012; Paige, et al., 2012; Paige, et al., 2011; Pitulle, Hedenstierna, & Fox, 1995a; W. P. Wang, et al., 2015; X. Zhang, et al., 2009). For instance, rRNA-carried VEGF aptamer effectively suppressed angiogenesis, as demonstrated by Chick Chorioallantoic Membrane (CAM) Assay (X. Zhang, et al., 2009). A number of chimeric RNA aptamers were also developed and produced using a tRNA scaffold, which bound to small-molecule fluorophores and then produced fluorescent signals for the detection or imaging of target molecules in the cells (Paige, et al., 2012; Paige, et al., 2011). In addition, upon binding to malachite green, the MGAs produced on a large scale using a tRNA/pre-miR-34a carrier exhibited specific and strong fluorescence (Q. X. Chen, et al., 2015). Because the fluorescence intensity is proportional to the level of MGA, which the latter is susceptible to serum RNases, bioengineered MGA was further developed as a sensor for the quantification of RNase activity that may be useful for diagnosis and prognosis (Q. X. Chen, et al., 2015).
As a new class of RNAs, BERAs are active in suppressing target gene expression and consequently modulating cellular processes. Fully-processed “bacterial” siRNAs produced with p19 RNA-binding protein were effective in repressing target gene expression in mammalian cells without immunogenicity or off-target effects (Huang, et al., 2013). Recombinant pre-miR-29b was able to reduce mRNA and protein levels of its target gene human β-secretase (hBACE1) and then decrease the levels of amyloid-β (Aβ) peptide in N2a695 cells (P. A. Pereira, et al., 2016). Bioengineered tRNA/pre-miR-27b significantly repressed its target gene expression (e.g., cytochrome P450 3A4 (CYP3A4)) and subsequently inhibited cellular drug metabolism capacity (e.g., midazolam 1’-hydroxylation) (M. M. Li, et al., 2014). Recombinant tRNA/pre-miR-1291 produced at high yields from bacteria downregulated the expression of a number of target genes including efflux transporter and cancer related genes and then improved the sensitivity of human carcinoma cells to chemotherapeutics (Jian, et al., 2017; M. M. Li, et al., 2015; Zhao, et al., 2015). In addition, tRNA/pre-miRNA-carried miRNAs/siRNAs were readily released to target miRNAs/siRNAs in human cells and animal models to control gene expression (Q. X. Chen, et al., 2015; P. Y. Ho, et al., 2018; Jian, et al., 2017; P. C. Li, et al., 2018; W. P. Wang, et al., 2015; Zhao, et al., 2015), which were demonstrated by untargeted (e.g., RNA sequencing) and targeted (e.g., qPCR and Western blots, etc.) analyses. Most importantly, bioengineered RNAi molecules were consistently shown to be equally or even more effective than synthetic miRNA/siRNA agents in the regulation of target gene expression and cell functions (Q. X. Chen, et al., 2015; Huang, et al., 2013; M. M. Li, et al., 2015; M. M. Li, et al., 2014; P. A. Pereira, et al., 2016; W. P. Wang, et al., 2015). BERAs should be minimally an addition to current RNAi toolbox, serving as a novel class of RNAi agents for functional studies.
6.3. Therapeutic potential
BERAs as possible therapeutics were demonstrated to be of potential value in diseased animal models (P. Y. Ho, et al., 2018; Jian, et al., 2017; Tu, et al., 2018; W. P. Wang, et al., 2015; Zhao, et al., 2016; Zhao, et al., 2015). Intratumoral injection of bioengineered miR-34a prodrug showed dose dependent effects in the suppression of subcutaneous xenograft tumor growth in mice derived from human NSCLC A549 carcinoma cells (W. P. Wang, et al., 2015). The efficacy of miR-34a for the treatment of NSCLC was further demonstrated with refined miR-34a prodrug molecule administered systemically into metastatic xenograft tumor mouse models (P. Y. Ho, et al., 2018). Meanwhile, intravenous administration of in vivo-jetPEI-formulated miR-34a prodrug dramatically inhibited the growth of orthotopic osteosarcomas in mice engrafted with human 143B cells (Zhao, et al., 2016), as well as lung metastasis (Jian, et al., 2017). Furthermore, the efficacy of chemotherapy for orthotopic 143B xenograft tumors was significantly improved when mice were co-administered with biological miR-34a prodrug (Jian, et al., 2017; Zhao, et al., 2015). In addition, BERAs did not show any severe hepatotoxicity, nephrotoxicity or immunogenicity in mouse models, as indicated by the lack of or minimal changes of mouse body weights, blood chemistries and chemokines (P. Y. Ho, et al., 2018; Jian, et al., 2017; Tu, et al., 2018; W. P. Wang, et al., 2015; Zhao, et al., 2016; Zhao, et al., 2015). These studies exemplify the utility of BERAs for translational research and demonstrate the potential of BERAs as therapeutics.
7. Conclusions and perspectives
The discovery of RNAi mechanisms and the development of RNAi agents not only enhance functional genomics research but also provoke the development of RNA therapeutics that holds the promise to expand the range of druggable targets essentially for all types of diseases. While conventional RNA agents made by chemical synthesis have found their success for clinical practice and continue dominating current RNAi research and drug development, caution should be exercised due to the fact that chemo-engineered RNA mimics are decorated with extensive and various types of chemical modifications, and different manufacturers might offer distinct molecules. By contrast, natural RNA transcribed from the genome in living cells is basically assembled by unmodified ribonucleotides. There are just a minimal fraction of nucleosides with posttranscriptional modifications proven to be critical for RNA folding into higher order structures as well as stability, biological functions and safety profiles. Therefore, bioengineering technologies have been developed very recently for the production of biologic RNAi molecules, which, made and folded in living cells, do exhibit intrinsic structural characteristics while carrying no or minimal posttranscriptional modifications.
The tRNA/pre-miRNA-based technology provides a robust platform for consistent, cost-effective, high-yield, and large-scale production of a variety of biologic RNAi agents including miRNAs, siRNAs and aptamers. The resultant BERAs are biologically active following cellular processing and pairing to their targets in human cells, which consequently modulates target gene expression and controls cellular processes such as metabolism, proliferation, and apoptosis. Furthermore, biologic RNAi molecules are well tolerated in mouse models and highly effective to suppress xenograft tumor growth, whose efficacy and safety (especially immunogenicity) awaits more extensive studies in different model systems. Therefore, bioengineered RNAi molecules represent a novel class of RNA agents that shall better capture the physicochemical, biological and safety properties of natural RNAs and hold great promise for basic and translational research. Nevertheless, while bioengineered RNAs showed favorable stability within human cells, naked BERAs remain vulnerable to large quantity of serum RNases. Proper formulation and delivery system are still needed to increase BERA stability in blood and overcome cellular barriers (Q. Y. Zhang, et al., 2018) to access molecular targets, which is also essential for moving biologic RNAi molecules into clinical investigations and practice. In addition, further studies are warranted to refine current technologies or explore new approaches for RNA bioengineering that may lead to novel tools for RNA research and drug development.
Acknowledgments
This work was supported in part by grants from The National Institute of General Medical Sciences [R01GM133888] and National Cancer Institute [U01CA175315], National Institutes of Health. C. Jian was supported by a fellowship from the Chinese Scholarship Council (No. 201506270112).
Abbreviations
- A
adenine
- ASO
antisense oligonucleotide
- asRNA
antisense RNA
- BERA
bioengineered/biologic RNA agents
- C
cytosine
- CMV
cytomegalovirus
- dsRNA
double-stranded RNA
- FPLC
fast protein liquid chromatography
- FDA
Food and Drug Administration
- G
guanine
- GalNAc
N-acetylgalactosamine
- GFP
green fluorescent protein
- gRNA
guide RNA
- HPLC
high performance liquid chromatography
- lncRNA
long noncoding RNA
- MGA
malachite green aptamer
- miRNA
microRNA
- ncRNA
noncoding RNA
- NSCLC
non-small cell lung cancer
- Ψ
pseudouridine
- PMO
phosphorodiamidate morpholine oligonucleotide
- PNA
peptide nucleic acid
- pre-miRNA
precursor miRNA
- pri-miRNA
primary miRNA
- PS
phosphorothioate
- RNAi
RNA interference
- RNase
ribonuclease
- rRNA
ribosomal RNA
- shRNA
short-hairpin RNA
- siRNA
small interfering RNA
- sRNA
small RNA
- T
thymine
- tRNA
transfer RNA
- U
uracil
- VEGF
vascular endothelial growth factor
Footnotes
Conflict of interest statement
The authors declare no conflict of interests.
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
References
- Aartsma-Rus A, (2017). FDA Approval of Nusinersen for Spinal Muscular Atrophy Makes 2016 the Year of Splice Modulating Oligonucleotides. Nucleic Acid Ther, 27, 67–69. [DOI] [PubMed] [Google Scholar]
- Aartsma-Rus A, & Krieg AM, (2017). FDA Approves Eteplirsen for Duchenne Muscular Dystrophy: The Next Chapter in the Eteplirsen Saga. Nucleic Acid Ther, 27, 1–3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Adams D, Gonzalez-Duarte A, O'Riordan WD, Yang CC, Ueda M, Kristen AV, Tournev I, Schmidt HH, Coelho T, Berk JL, Lin KP, Vita G, Attarian S, Plante-Bordeneuve V, Mezei MM, Campistol JM, Buades J, Brannagan TH 3rd, Kim BJ, Oh J, Parman Y, Sekijima Y, Hawkins PN, Solomon SD, Polydefkis M, Dyck PJ, Gandhi PJ, Goyal S, Chen J, Strahs AL, Nochur SV, Sweetser MT, Garg PP, Vaishnaw AK, Gollob JA, & Suhr OB, (2018). Patisiran, an RNAi Therapeutic, for Hereditary Transthyretin Amyloidosis. N Engl J Med, 379, 11–21. [DOI] [PubMed] [Google Scholar]
- Alon S, Mor E, Vigneault F, Church GM, Locatelli F, Galeano F, Gallo A, Shomron N, & Eisenberg E, (2012). Systematic identification of edited microRNAs in the human brain. Genome Res, 22, 1533–1540. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ambros V, (2004). The functions of animal microRNAs. Nature, 431, 350–355. [DOI] [PubMed] [Google Scholar]
- Bader AG, (2012). miR-34 - a microRNA replacement therapy is headed to the clinic. Front Genet, 3, 120. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bader AG, Brown D, & Winkler M, (2010). The promise of microRNA replacement therapy. Cancer Res, 70, 7027–7030. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bai Y, Dai X, Harrison A, Johnston C, & Chen M, (2016). Toward a next-generation atlas of RNA secondary structure. Brief Bioinform, 17, 63–77. [DOI] [PubMed] [Google Scholar]
- Bartel DP, (2009). MicroRNAs: target recognition and regulatory functions. Cell, 136, 215–233. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Beg MS, Brenner AJ, Sachdev J, Borad M, Kang YK, Stoudemire J, Smith S, Bader AG, Kim S, & Hong DS, (2016). Phase I study of MRX34, a liposomal miR-34a mimic, administered twice weekly in patients with advanced solid tumors. Invest New Drugs. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Benitez-Del-Castillo JM, Moreno-Montanes J, Jimenez-Alfaro I, Munoz-Negrete FJ, Turman K, Palumaa K, Sadaba B, Gonzalez MV, Ruz V, Vargas B, Paneda C, Martinez T, Bleau AM, & Jimenez AI, (2016). Safety and Efficacy Clinical Trials for SYL1001, a Novel Short Interfering RNA for the Treatment of Dry Eye Disease. Invest Ophthalmol Vis Sci, 57, 6447–6454. [DOI] [PubMed] [Google Scholar]
- Brake OT, Hooft K, Liu YP, Centlivre M, Jasmijn von Eije K, & Berkhout B, (2008). Lentiviral Vector Design for Multiple shRNA Expression and Durable HIV-1 Inhibition. Mol Ther, 16, 557–564. [DOI] [PubMed] [Google Scholar]
- Bramsen JB, & Kjems J, (2012). Development of Therapeutic-Grade Small Interfering RNAs by Chemical Engineering. Front Genet, 3, 154. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brown R, Curry E, Magnani L, Wilhelm-Benartzi CS, & Borley J, (2014). Poised epigenetic states and acquired drug resistance in cancer. Nat Rev Cancer, 14, 747–753. [DOI] [PubMed] [Google Scholar]
- Burgart AM, Magnus D, Tabor HK, Paquette ED, Frader J, Glover JJ, Jackson BM, Harrison CH, Urion DK, Graham RJ, Brandsema JF, & Feudtner C, (2018). Ethical Challenges Confronted When Providing Nusinersen Treatment for Spinal Muscular Atrophy. JAMA Pediatr, 172, 188–192. [DOI] [PubMed] [Google Scholar]
- Butcher SE, & Pyle AM, (2011). The molecular interactions that stabilize RNA tertiary structure: RNA motifs, patterns, and networks. Acc Chem Res, 44, 1302–1311. [DOI] [PubMed] [Google Scholar]
- Cantara WA, Crain PF, Rozenski J, McCloskey JA, Harris KA, Zhang X, Vendeix FA, Fabris D, & Agris PF, (2011). The RNA Modification Database, RNAMDB: 2011 update. Nucleic Acids Res, 39, D195–201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cech TR, & Steitz JA, (2014). The noncoding RNA revolution-trashing old rules to forge new ones. Cell, 157, 77–94. [DOI] [PubMed] [Google Scholar]
- Chen QX, Wang WP, Zeng S, Urayama S, & Yu AM, (2015). A general approach to high-yield biosynthesis of chimeric RNAs bearing various types of functional small RNAs for broad applications. Nucleic Acids Res, 43, 3857–3869. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen X, Ba Y, Ma L, Cai X, Yin Y, Wang K, Guo J, Zhang Y, Chen J, Guo X, Li Q, Li X, Wang W, Zhang Y, Wang J, Jiang X, Xiang Y, Xu C, Zheng P, Zhang J, Li R, Zhang H, Shang X, Gong T, Ning G, Wang J, Zen K, Zhang J, & Zhang CY, (2008). Characterization of microRNAs in serum: a novel class of biomarkers for diagnosis of cancer and other diseases. Cell Res, 18, 997–1006. [DOI] [PubMed] [Google Scholar]
- Chen X, Mangala LS, Rodriguez-Aguayo C, Kong X, Lopez-Berestein G, & Sood AK, (2018). RNA interference-based therapy and its delivery systems. Cancer Metastasis Rev, 37, 107–124. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Choi YH, & Yu AM, (2014). ABC transporters in multidrug resistance and pharmacokinetics, and strategies for drug development. Curr Pharm Des, 20, 793–807. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Crooke ST, Wang S, Vickers TA, Shen W, & Liang XH, (2017). Cellular uptake and trafficking of antisense oligonucleotides. Nat Biotechnol, 35, 230–237. [DOI] [PubMed] [Google Scholar]
- Czauderna F, Santel A, Hinz M, Fechtner M, Durieux B, Fisch G, Leenders F, Arnold W, Giese K, Klippel A, & Kaufmann J, (2003). Inducible shRNA expression for application in a prostate cancer mouse model. Nucleic Acids Res, 31, e127. [DOI] [PMC free article] [PubMed] [Google Scholar]
- D'Souza LM, Larios-Sanz M, Setterquist RA, Willson RC, & Fox GE, (2003). Small RNA sequences are readily stabilized by inclusion in a carrier rRNA. Biotechnol Prog, 19, 734–738. [DOI] [PubMed] [Google Scholar]
- Dalpke A, & Helm M, (2012). RNA mediated Toll-like receptor stimulation in health and disease. RNA Biol, 9, 828–842. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Demirjian S, Ailawadi G, Polinsky M, Bitran D, Silberman S, Shernan SK, Burnier M, Hamilton M, Squiers E, Erlich S, Rothenstein D, Khan S, & Chawla LS, (2017). Safety and Tolerability Study of an Intravenously Administered Small Interfering Ribonucleic Acid (siRNA) Post On-Pump Cardiothoracic Surgery in Patients at Risk of Acute Kidney Injury. Kidney Int Rep, 2, 836–843. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dimitrov DS, (2012). Therapeutic proteins. Methods Mol Biol, 899, 1–26. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ding SW, & Voinnet O, (2007). Antiviral immunity directed by small RNAs. Cell, 130, 413–426. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dohmen C, Frohlich T, Lachelt U, Rohl I, Vornlocher HP, Hadwiger P, & Wagner E, (2012). Defined Folate-PEG-siRNA Conjugates for Receptor-specific Gene Silencing. Mol Ther Nucleic Acids, 1, e7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Doudna JA, & Charpentier E, (2014). Genome editing. The new frontier of genome engineering with CRISPR-Cas9. Science, 346, 1258096. [DOI] [PubMed] [Google Scholar]
- Dowdy SF, (2017). Overcoming cellular barriers for RNA therapeutics. Nat Biotechnol, 35, 222–229. [DOI] [PubMed] [Google Scholar]
- Duan Z, & Yu AM, (2016). Bioengineered non-coding RNA agent (BERA) in action. Bioengineered, 7, 411–417. [DOI] [PMC free article] [PubMed] [Google Scholar]
- El Yacoubi B, Bailly M, & de Crecy-Lagard V, (2012). Biosynthesis and function of posttranscriptional modifications of transfer RNAs. Annu Rev Genet, 46, 69–95. [DOI] [PubMed] [Google Scholar]
- Elkayam E, Parmar R, Brown CR, Willoughby JL, Theile CS, Manoharan M, & Joshua-Tor L, (2017). siRNA carrying an (E)-vinylphosphonate moiety at the 5 end of the guide strand augments gene silencing by enhanced binding to human Argonaute-2. Nucleic Acids Res, 45, 3528–3536. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Esteller M, (2011). Non-coding RNAs in human disease. Nat Rev Genet, 12, 861–874. [DOI] [PubMed] [Google Scholar]
- Fire A, Xu S, Montgomery MK, Kostas SA, Driver SE, & Mello CC, (1998). Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature, 391, 806–811. [DOI] [PubMed] [Google Scholar]
- Foster DJ, Brown CR, Shaikh S, Trapp C, Schlegel MK, Qian K, Sehgal A, Rajeev KG, Jadhav V, Manoharan M, Kuchimanchi S, Maier MA, & Milstein S, (2018). Advanced siRNA Designs Further Improve In Vivo Performance of GalNAc-siRNA Conjugates. Mol Ther, 26, 708–717. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gashaw I, Ellinghaus P, Sommer A, & Asadullah K, (2012). What makes a good drug target? Drug Discov Today, 17 Suppl, S24–30. [DOI] [PubMed] [Google Scholar]
- Gaudin C, Nonin-Lecomte S, Tisné C, Corvaisier S, Bordeau V, Dardel F, & Felden B, (2003). The tRNA-like Domains of E.coli and A.aeolicus Transfer–Messenger RNA: Structural and Functional Studies. Journal of Molecular Biology, 331, 457–471. [DOI] [PubMed] [Google Scholar]
- Gehrig S, Eberle ME, Botschen E, Rimbach K, Eberle F, Eigenbrod T, Kaiser S, Holmes WM, Erdmann VA, Sprinzl M, Bec G, Keith G, Dalpke AH, & Helm M, (2012). Identification of modifications in microbial, native tRNA that suppress immunostimulatory activity. J Exp Med, 209, 225–233. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Golan T, Khvalevsky EZ, Hubert A, Gabai RM, Hen N, Segal A, Domb A, Harari G, David EB, Raskin S, Goldes Y, Goldin E, Eliakim R, Lahav M, Kopleman Y, Dancour A, Shemi A, & Galun E, (2015). RNAi therapy targeting KRAS in combination with chemotherapy for locally advanced pancreatic cancer patients. Oncotarget, 6, 24560–24570. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gryziewicz L, (2005). Regulatory aspects of drug approval for macular degeneration. Adv Drug Deliv Rev, 57, 2092–2098. [DOI] [PubMed] [Google Scholar]
- Haraszti RA, Roux L, Coles AH, Turanov AA, Alterman JF, Echeverria D, Godinho B, Aronin N, & Khvorova A, (2017). 5-Vinylphosphonate improves tissue accumulation and efficacy of conjugated siRNAs in vivo. Nucleic Acids Res, 45, 7581–7592. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ho PY, Duan Z, Batra N, Jilek JL, Tu MJ, Qiu JX, Hu Z, Wun T, Lara PN, DeVere White RW, Chen HW, & Yu AM, (2018). Bioengineered Noncoding RNAs Selectively Change Cellular miRNome Profiles for Cancer Therapy. J Pharmacol Exp Ther, 365, 494–506. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ho PY, & Yu AM, (2016). Bioengineering of noncoding RNAs for research agents and therapeutics. Wiley Interdiscip Rev RNA, 7, 186–197. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ho W, Zhang XQ, & Xu X, (2016). Biomaterials in siRNA Delivery: A Comprehensive Review. Adv Healthc Mater, 5, 2715–2731. [DOI] [PubMed] [Google Scholar]
- Hornung V, Guenthner-Biller M, Bourquin C, Ablasser A, Schlee M, Uematsu S, Noronha A, Manoharan M, Akira S, de Fougerolles A, Endres S, & Hartmann G, (2005). Sequence-specific potent induction of IFN-alpha by short interfering RNA in plasmacytoid dendritic cells through TLR7. Nat Med, 11, 263–270. [DOI] [PubMed] [Google Scholar]
- Huang L, Jin J, Deighan P, Kiner E, McReynolds L, & Lieberman J, (2013). Efficient and specific gene knockdown by small interfering RNAs produced in bacteria. Nat Biotechnol, 31, 350–356. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hwang J, Chang C, Kim JH, Oh CT, Lee HN, Lee C, Oh D, Lee C, Kim B, Hong SW, & Lee DK, (2016). Development of Cell-Penetrating Asymmetric Interfering RNA Targeting Connective Tissue Growth Factor. J Invest Dermatol, 136, 2305–2313. [DOI] [PubMed] [Google Scholar]
- Iversen PL, (2001). Phosphorodiamidate morpholino oligomers: favorable properties for sequence-specific gene inactivation. Curr Opin Mol Ther, 3, 235–238. [PubMed] [Google Scholar]
- Jacobson SG, Cideciyan AV, Sumaroka A, Roman AJ, Charng J, Lu M, Choi W, Sheplock R, Swider M, Kosyk MS, Schwartz SB, Stone EM, & Fishman GA, (2017). Outcome Measures for Clinical Trials of Leber Congenital Amaurosis Caused by the Intronic Mutation in the CEP290 Gene. Invest Ophthalmol Vis Sci, 58, 2609–2622. [DOI] [PubMed] [Google Scholar]
- Jian C, Tu MJ, Ho PY, Duan Z, Zhang Q, Qiu JX, DeVere White RW, Wun T, Lara PN, Lam KS, Yu AX, & Yu AM, (2017). Co-targeting of DNA, RNA, and protein molecules provides optimal outcomes for treating osteosarcoma and pulmonary metastasis in spontaneous and experimental metastasis mouse models. Oncotarget, 8, 30742–30755. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jin J, Cid M, Poole CB, & McReynolds LA, (2010). Protein mediated miRNA detection and siRNA enrichment using p19. Biotechniques, 48, xvii–xxiii. [DOI] [PubMed] [Google Scholar]
- Jones CP, & Ferre-D'Amare AR, (2015). RNA quaternary structure and global symmetry. Trends Biochem Sci, 40, 211–220. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Judge AD, Sood V, Shaw JR, Fang D, McClintock K, & MacLachlan I, (2005). Sequence-dependent stimulation of the mammalian innate immune response by synthetic siRNA. Nat Biotechnol, 23, 457–462. [DOI] [PubMed] [Google Scholar]
- Kanasty R, Dorkin JR, Vegas A, & Anderson D, (2013). Delivery materials for siRNA therapeutics. Nat Mater, 12, 967–977. [DOI] [PubMed] [Google Scholar]
- Kariko K, & Weissman D, (2007). Naturally occuring nucleoside modifications suppress the immnostimulatory activity of RNA: implication for therapeutic RNA development. Curr Opin Drug Discov Devel, 10, 523–532. [PubMed] [Google Scholar]
- Kawahara Y, Zinshteyn B, Chendrimada TP, Shiekhattar R, & Nishikura K, (2007). RNA editing of the microRNA-151 precursor blocks cleavage by the Dicer-TRBP complex. EMBO Rep, 8, 763–769. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kawasaki AM, Casper MD, Freier SM, Lesnik EA, Zounes MC, Cummins LL, Gonzalez C, & Cook PD, (1993). Uniformly modified 2'-deoxy-2'-fluoro phosphorothioate oligonucleotides as nuclease-resistant antisense compounds with high affinity and specificity for RNA targets. J Med Chem, 36, 831–841. [DOI] [PubMed] [Google Scholar]
- Keam SJ, (2018). Inotersen: First Global Approval. Drugs, 78, 1371–1376. [DOI] [PubMed] [Google Scholar]
- Khorkova O, & Wahlestedt C, (2017). Oligonucleotide therapies for disorders of the nervous system. Nat Biotechnol, 35, 249–263. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Khvorova A, & Watts JK, (2017). The chemical evolution of oligonucleotide therapies of clinical utility. Nat Biotechnol, 35, 238–248. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Komor AC, Badran AH, & Liu DR, (2017). CRISPR-Based Technologies for the Manipulation of Eukaryotic Genomes. Cell, 168, 20–36. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kuimelis RG, & McLaughlin LW, (1995). Cleavage properties of an oligonucleotide containing a bridged internucleotide 5'-phosphorothioate RNA linkage. Nucleic Acids Res, 23, 4753–4760. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kurreck J, Wyszko E, Gillen C, & Erdmann VA, (2002). Design of antisense oligonucleotides stabilized by locked nucleic acids. Nucleic Acids Res, 30, 1911–1918. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Landgraf P, Rusu M, Sheridan R, Sewer A, Iovino N, Aravin A, Pfeffer S, Rice A, Kamphorst AO, Landthaler M, Lin C, Socci ND, Hermida L, Fulci V, Chiaretti S, Foa R, Schliwka J, Fuchs U, Novosel A, Muller RU, Schermer B, Bissels U, Inman J, Phan Q, Chien M, Weir DB, Choksi R, De Vita G, Frezzetti D, Trompeter HI, Hornung V, Teng G, Hartmann G, Palkovits M, Di Lauro R, Wernet P, Macino G, Rogler CE, Nagle JW, Ju J, Papavasiliou FN, Benzing T, Lichter P, Tam W, Brownstein MJ, Bosio A, Borkhardt A, Russo JJ, Sander C, Zavolan M, & Tuschl T, (2007). A mammalian microRNA expression atlas based on small RNA library sequencing. Cell, 129, 1401–1414. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lares MR, Rossi JJ, & Ouellet DL, (2010). RNAi and small interfering RNAs in human disease therapeutic applications. Trends Biotechnol, 28, 570–579. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Leader B, Baca QJ, & Golan DE, (2008). Protein therapeutics: a summary and pharmacological classification. Nat Rev Drug Discov, 7, 21–39. [DOI] [PubMed] [Google Scholar]
- Leung E, & Landa G, (2013). Update on current and future novel therapies for dry age-related macular degeneration. Expert Rev Clin Pharmacol, 6, 565–579. [DOI] [PubMed] [Google Scholar]
- Li F, & Ding SW, (2006). Virus counterdefense: diverse strategies for evading the RNA-silencing immunity. Annu Rev Microbiol, 60, 503–531. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li MM, Addepalli B, Tu MJ, Chen QX, Wang WP, Limbach PA, LaSalle JM, Zeng S, Huang M, & Yu AM, (2015). Chimeric MicroRNA-1291 Biosynthesized Efficiently in Escherichia coli Is Effective to Reduce Target Gene Expression in Human Carcinoma Cells and Improve Chemosensitivity. Drug Metab Dispos, 43, 1129–1136. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li MM, Wang WP, Wu WJ, Huang M, & Yu AM, (2014). Rapid Production of Novel Pre-MicroRNA Agent hsa-mir-27b in Escherichia coli Using Recombinant RNA Technology for Functional Studies in Mammalian Cells. Drug Metab Dispos, 42, 1791–1795. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li PC, Tu MJ, Ho PY, Jilek JL, Duan Z, Zhang QY, Yu AX, & Yu AM, (2018). Bioengineered NRF2-siRNA Is Effective to Interfere with NRF2 Pathways and Improve Chemosensitivity of Human Cancer Cells. Drug Metab Dispos, 46, 2–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li X, Ma S, & Yi C, (2016). Pseudouridine: the fifth RNA nucleotide with renewed interests. Curr Opin Chem Biol, 33, 108–116. [DOI] [PubMed] [Google Scholar]
- Limbach PA, Crain PF, & McCloskey JA, (1994). Summary: the modified nucleosides of RNA. Nucleic Acids Res, 22, 2183–2196. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu N, & Pan T, (2015). RNA epigenetics. Transl Res, 165, 28–35. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu Y, Stepanov VG, Strych U, Willson RC, Jackson GW, & Fox GE, (2010). DNAzyme-mediated recovery of small recombinant RNAs from a 5S rRNA-derived chimera expressed in Escherichia coli. BMC Biotechnol, 10, 85. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu YP, & Berkhout B, (2011). miRNA cassettes in viral vectors: problems and solutions. Biochim Biophys Acta, 1809, 732–745. [DOI] [PubMed] [Google Scholar]
- Liz J, & Esteller M, (2016). lncRNAs and microRNAs with a role in cancer development. Biochim Biophys Acta, 1859, 169–176. [DOI] [PubMed] [Google Scholar]
- Lu J, Getz G, Miska EA, Alvarez-Saavedra E, Lamb J, Peck D, Sweet-Cordero A, Ebert BL, Mak RH, Ferrando AA, Downing JR, Jacks T, Horvitz HR, & Golub TR, (2005). MicroRNA expression profiles classify human cancers. Nature, 435, 834–838. [DOI] [PubMed] [Google Scholar]
- Luciano DJ, Mirsky H, Vendetti NJ, & Maas S, (2004). RNA editing of a miRNA precursor. RNA, 10, 1174–1177. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lundin KE, Hojland T, Hansen BR, Persson R, Bramsen JB, Kjems J, Koch T, Wengel J, & Smith CI, (2013). Biological activity and biotechnological aspects of locked nucleic acids. Adv Genet, 82, 47–107. [DOI] [PubMed] [Google Scholar]
- Matsuda S, Keiser K, Nair JK, Charisse K, Manoharan RM, Kretschmer P, Peng CG, A V.K. i., Kandasamy P, Willoughby JL, Liebow A, Querbes W, Yucius K, Nguyen T, Milstein S, Maier MA, Rajeev KG, & Manoharan M, (2015). siRNA conjugates carrying sequentially assembled trivalent N-acetylgalactosamine linked through nucleosides elicit robust gene silencing in vivo in hepatocytes. ACS Chem Biol, 10, 1181–1187. [DOI] [PubMed] [Google Scholar]
- Matsui M, & Corey DR, (2017). Non-coding RNAs as drug targets. Nat Rev Drug Discov, 16, 167–179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mattick JS, (2004). RNA regulation: a new genetics? Nat Rev Genet, 5, 316–323. [DOI] [PubMed] [Google Scholar]
- Meinnel T, Mechulam Y, & Fayat G, (1988). Fast purification of a functional elongator tRNAmet expressed from a synthetic gene in vivo. Nucleic Acids Res, 16, 8095–8096. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mendell JT, & Olson EN, (2012). MicroRNAs in stress signaling and human disease. Cell, 148, 1172–1187. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mitchell PS, Parkin RK, Kroh EM, Fritz BR, Wyman SK, Pogosova-Agadjanyan EL, Peterson A, Noteboom J, O'Briant KC, Allen A, Lin DW, Urban N, Drescher CW, Knudsen BS, Stirewalt DL, Gentleman R, Vessella RL, Nelson PS, Martin DB, & Tewari M, (2008). Circulating microRNAs as stable blood-based markers for cancer detection. Proc Natl Acad Sci U S A, 105, 10513–10518. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Morena F, Argentati C, Bazzucchi M, Emiliani C, & Martino S, (2018). Above the Epitranscriptome: RNA Modifications and Stem Cell Identity. Genes (Basel), 9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moschos SA, Usher L, & Lindsay MA, (2017). Clinical potential of oligonucleotide-based therapeutics in the respiratory system. Pharmacol Ther, 169, 83–103. [DOI] [PubMed] [Google Scholar]
- Nair JK, Willoughby JL, Chan A, Charisse K, Alam MR, Wang Q, Hoekstra M, Kandasamy P, Kel'in AV, Milstein S, Taneja N, O'Shea J, Shaikh S, Zhang L, van der Sluis RJ, Jung ME, Akinc A, Hutabarat R, Kuchimanchi S, Fitzgerald K, Zimmermann T, van Berkel TJ, Maier MA, Rajeev KG, & Manoharan M, (2014). Multivalent N-acetylgalactosamine-conjugated siRNA localizes in hepatocytes and elicits robust RNAi-mediated gene silencing. J Am Chem Soc, 136, 16958–16961. [DOI] [PubMed] [Google Scholar]
- Nallagatla SR, Toroney R, & Bevilacqua PC, (2008). A Brilliant Disguise for Self RNA: 5'-end and Internal Modifcations of Primary Transcripts Suppress Elements of Innate Immunity. RNA Biol, 5, 140–144. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nelissen FH, Leunissen EH, van de Laar L, Tessari M, Heus HA, & Wijmenga SS, (2012). Fast production of homogeneous recombinant RNA--towards large-scale production of RNA. Nucleic Acids Res, 40, e102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nielsen PE, (2005). Gene targeting using peptide nucleic acid. Methods Mol Biol, 288, 343–358. [DOI] [PubMed] [Google Scholar]
- Nielsen PE, Egholm M, Berg RH, & Buchardt O, (1991). Sequence-selective recognition of DNA by strand displacement with a thymine-substituted polyamide. Science, 254, 1497–1500. [DOI] [PubMed] [Google Scholar]
- Nimjee SM, White RR, Becker RC, & Sullenger BA, (2017). Aptamers as Therapeutics. Annu Rev Pharmacol Toxicol, 57, 61–79. [DOI] [PMC free article] [PubMed] [Google Scholar]
- O'Day E, Hosta-Rigau L, Oyarzun DA, Okano H, de Lorenzo V, von Kameke C, Alsafar H, Cao C, Chen GQ, Ji W, Roberts RJ, Ronaghi M, Yeung K, Zhang F, & Lee SY, (2018). Are We There Yet? How and When Specific Biotechnologies Will Improve Human Health. Biotechnol J, e1800195. [DOI] [PubMed] [Google Scholar]
- Orr RM, (2001). Technology evaluation: fomivirsen, Isis Pharmaceuticals Inc/CIBA vision. Curr Opin Mol Ther, 3, 288–294. [PubMed] [Google Scholar]
- Ottesen EW, (2017). ISS-N1 makes the First FDA-approved Drug for Spinal Muscular Atrophy. Transl Neurosci, 8, 1–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Overington JP, Al-Lazikani B, & Hopkins AL, (2006). How many drug targets are there? Nat Rev Drug Discov, 5, 993–996. [DOI] [PubMed] [Google Scholar]
- Paige JS, Nguyen-Duc T, Song W, & Jaffrey SR, (2012). Fluorescence imaging of cellular metabolites with RNA. Science, 335, 1194. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Paige JS, Wu KY, & Jaffrey SR, (2011). RNA mimics of green fluorescent protein. Science, 333, 642–646. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Parmar R, Willoughby JL, Liu J, Foster DJ, Brigham B, Theile CS, Charisse K, Akinc A, Guidry E, Pei Y, Strapps W, Cancilla M, Stanton MG, Rajeev KG, Sepp-Lorenzino L, Manoharan M, Meyers R, Maier MA, & Jadhav V, (2016). 5'-(E)-Vinylphosphonate: A Stable Phosphate Mimic Can Improve the RNAi Activity of siRNA-GalNAc Conjugates. Chembiochem, 17, 985–989. [DOI] [PubMed] [Google Scholar]
- Pasi KJ, Rangarajan S, Georgiev P, Mant T, Creagh MD, Lissitchkov T, Bevan D, Austin S, Hay CR, Hegemann I, Kazmi R, Chowdary P, Gercheva-Kyuchukova L, Mamonov V, Timofeeva M, Soh CH, Garg P, Vaishnaw A, Akinc A, Sorensen B, & Ragni MV, (2017). Targeting of Antithrombin in Hemophilia A or B with RNAi Therapy. N Engl J Med, 377, 819–828. [DOI] [PubMed] [Google Scholar]
- Pereira P, Pedro AQ, Queiroz JA, Figueiras AR, & Sousa F, (2017). New insights for therapeutic recombinant human miRNAs heterologous production: Rhodovolum sulfidophilum vs Escherichia coli. Bioengineered, 8, 670–677. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pereira P, Pedro AQ, Tomas J, Maia CJ, Queiroz JA, Figueiras A, & Sousa F, (2016). Advances in time course extracellular production of human pre-miR-29b from Rhodovulum sulfidophilum. Appl Microbiol Biotechnol, 100, 3723–3734. [DOI] [PubMed] [Google Scholar]
- Pereira P, Sousa A, Queiroz J, Correia I, Figueiras A, & Sousa F, (2014). Purification of pre-miR-29 by arginine-affinity chromatography. J Chromatogr B Analyt Technol Biomed Life Sci, 951–952, 16–23. [DOI] [PubMed] [Google Scholar]
- Pereira PA, Tomas JF, Queiroz JA, Figueiras AR, & Sousa F, (2016). Recombinant pre-miR-29b for Alzheimer s disease therapeutics. Sci Rep, 6, 19946. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pitulle C, Hedenstierna KO, & Fox GE, (1995a). A novel approach for monitoring genetically engineered microorganisms by using artificial, stable RNAs. Appl Environ Microbiol, 61, 3661–3666. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pitulle C, Hedenstierna KO, & Fox GE, (1995b). A novel approach for monitoring genetically engineered microorganisms by using artificial, stable RNAs. Appl Environ Microbiol, 61, 3661–3666. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ponchon L, Beauvais G, Nonin-Lecomte S, & Dardel F, (2009). A generic protocol for the expression and purification of recombinant RNA in Escherichia coli using a tRNA scaffold. Nat Protoc, 4, 947–959. [DOI] [PubMed] [Google Scholar]
- Ponchon L, Catala M, Seijo B, El Khouri M, Dardel F, Nonin-Lecomte S, & Tisne C, (2013). Co-expression of RNA-protein complexes in Escherichia coli and applications to RNA biology. Nucleic Acids Res, 41, e150. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ponchon L, & Dardel F, (2007). Recombinant RNA technology: the tRNA scaffold. Nat Methods, 4, 571–576. [DOI] [PubMed] [Google Scholar]
- Ranaei-Siadat E, Merigoux C, Seijo B, Ponchon L, Saliou JM, Bernauer J, Sanglier-Cianferani S, Dardel F, Vachette P, & Nonin-Lecomte S, (2014). In vivo tmRNA protection by SmpB and pre-ribosome binding conformation in solution. RNA, 20, 1607–1620. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Robbins M, Judge A, & MacLachlan I, (2009). siRNA and innate immunity. Oligonucleotides, 19, 89–102. [DOI] [PubMed] [Google Scholar]
- Robinson JG, (2013). Management of familial hypercholesterolemia: a review of the recommendations from the National Lipid Association Expert Panel on Familial Hypercholesterolemia. J Manag Care Pharm, 19, 139–149. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Roehr B, (1998). Fomivirsen approved for CMV retinitis. J Int Assoc Physicians AIDS Care, 4, 14–16. [PubMed] [Google Scholar]
- Ross SJ, Revenko AS, Hanson LL, Ellston R, Staniszewska A, Whalley N, Pandey SK, Revill M, Rooney C, Buckett LK, Klein SK, Hudson K, Monia BP, Zinda M, Blakey DC, Lyne PD, & Macleod AR, (2017). Targeting KRAS-dependent tumors with AZD4785, a high-affinity therapeutic antisense oligonucleotide inhibitor of KRAS. Sci Transl Med, 9. [DOI] [PubMed] [Google Scholar]
- Rupaimoole R, & Slack FJ, (2017). MicroRNA therapeutics: towards a new era for the management of cancer and other diseases. Nat Rev Drug Discov, 16, 203–222. [DOI] [PubMed] [Google Scholar]
- Santos R, Ursu O, Gaulton A, Bento AP, Donadi RS, Bologa CG, Karlsson A, Al-Lazikani B, Hersey A, Oprea TI, & Overington JP, (2017). A comprehensive map of molecular drug targets. Nat Rev Drug Discov, 16, 19–34. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schlick T, (2018). Adventures with RNA graphs. Methods, 143, 16–33. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Secher T, Guilleminault L, Reckamp K, Amanam I, Plantier L, & Heuze-Vourc'h N, (2018). Therapeutic antibodies: A new era in the treatment of respiratory diseases? Pharmacol Ther, 189, 149–172. [DOI] [PubMed] [Google Scholar]
- Sethi P, & Lukiw WJ, (2009). Micro-RNA abundance and stability in human brain: specific alterations in Alzheimer's disease temporal lobe neocortex. Neurosci Lett, 459, 100–104. [DOI] [PubMed] [Google Scholar]
- Shoshan E, Mobley AK, Braeuer RR, Kamiya T, Huang L, Vasquez ME, Salameh A, Lee HJ, Kim SJ, Ivan C, Velazquez-Torres G, Nip KM, Zhu K, Brooks D, Jones SJ, Birol I, Mosqueda M, Wen YY, Eterovic AK, Sood AK, Hwu P, Gershenwald JE, Robertson AG, Calin GA, Markel G, Fidler IJ, & Bar-Eli M, (2015). Reduced adenosine-to-inosine miR-455–5p editing promotes melanoma growth and metastasis. Nat Cell Biol, 17, 311–321. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shumyatsky G, Wright D, & Reddy R, (1993). Methylphosphate cap structure increases the stability of 7SK, B2 and U6 small RNAs in Xenopus oocytes. Nucleic Acids Res, 21, 4756–4761. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Silhavy D, Molnár A, Lucioli A, Szittya G, Hornyik C, Tavazza M, & Burgyán J, (2002). A viral protein suppresses RNA silencing and binds silencing- generated, 21- to 25- nucleotide double- stranded RNAs. EMBO J, 21, 3070–3080. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Simoens S, & Huys I, (2017). Market access of Spinraza (Nusinersen) for spinal muscular atrophy: intellectual property rights, pricing, value and coverage considerations. Gene Ther, 24, 539–541. [DOI] [PubMed] [Google Scholar]
- Sliwkowski MX, & Mellman I, (2013). Antibody therapeutics in cancer. Science, 341, 1192–1198. [DOI] [PubMed] [Google Scholar]
- Solano EC, Kornbrust DJ, Beaudry A, Foy JW, Schneider DJ, & Thompson JD, (2014). Toxicological and pharmacokinetic properties of QPI-1007, a chemically modified synthetic siRNA targeting caspase 2 mRNA, following intravitreal injection. Nucleic Acid Ther, 24, 258–266. [DOI] [PubMed] [Google Scholar]
- Spenkuch F, Motorin Y, & Helm M, (2014). Pseudouridine: still mysterious, but never a fake (uridine)! RNA Biol, 11, 1540–1554. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stein CA, (2016). Eteplirsen Approved for Duchenne Muscular Dystrophy: The FDA Faces a Difficult Choice. Mol Ther, 24, 1884–1885. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stein CA, & Castanotto D, (2017). FDA-Approved Oligonucleotide Therapies in 2017. Mol Ther, 25, 1069–1075. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Summerton J, & Weller D, (1997). Morpholino antisense oligomers: design, preparation, and properties. Antisense Nucleic Acid Drug Dev, 7, 187–195. [DOI] [PubMed] [Google Scholar]
- Syed YY, (2016). Eteplirsen: First Global Approval. Drugs, 76, 1699–1704. [DOI] [PubMed] [Google Scholar]
- Tanji H, Ohio U, Shibata T, Taoka M, Yamauchi Y, Isobe T, Miyake K, & Shimizu T, (2015). Toll-like receptor 8 senses degradation products of single-stranded RNA. Nat Struct Mol Biol, 22, 109–115. [DOI] [PubMed] [Google Scholar]
- Trail PA, Dubowchik GM, & Lowinger TB, (2018). Antibody drug conjugates for treatment of breast cancer: Novel targets and diverse approaches in ADC design. Pharmacol Ther, 181, 126–142. [DOI] [PubMed] [Google Scholar]
- Tu MJ, Ho PY, Zhang QY, Jian C, Qiu JX, Kim EJ, Bold RJ, Gonzalez FJ, Bi H, & Yu AM, (2018). Bioengineered miRNA-1291 prodrug therapy in pancreatic cancer cells and patient-derived xenograft mouse models. Cancer Lett, 442, 82–90. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Voinnet O, Pinto YM, & Baulcombe DC, (1999). Suppression of gene silencing: a general strategy used by diverse DNA and RNA viruses of plants. Proc Natl Acad Sci USA, 96, 14147–14152. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wagner MJ, Mitra R, McArthur MJ, Baze W, Barnhart K, Wu SY, Rodriguez-Aguayo C, Zhang X, Coleman RL, Lopez-Berestein G, & Sood AK, (2017). Preclinical Mammalian Safety Studies of EPHARNA (DOPC Nanoliposomal EphA2-Targeted siRNA). Mol Cancer Ther, 16, 1114–1123. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang HX, Li M, Lee CM, Chakraborty S, Kim HW, Bao G, & Leong KW, (2017). CRISPR/Cas9-Based Genome Editing for Disease Modeling and Therapy: Challenges and Opportunities for Nonviral Delivery. Chem Rev, 117, 9874–9906. [DOI] [PubMed] [Google Scholar]
- Wang WP, Ho PY, Chen QX, Addepalli B, Limbach PA, Li MM, Wu WJ, Jilek JL, Qiu JX, Zhang HJ, Li T, Wun T, White RD, Lam KS, & Yu AM, (2015). Bioengineering Novel Chimeric microRNA-34a for Prodrug Cancer Therapy: High-Yield Expression and Purification, and Structural and Functional Characterization. J Pharmacol Exp Ther, 354, 131–141. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Winkler J, (2013). Oligonucleotide conjugates for therapeutic applications. Ther Deliv, 4, 791–809. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Winter J, & Diederichs S, (2011). Argonaute proteins regulate microRNA stability: Increased microRNA abundance by Argonaute proteins is due to microRNA stabilization. RNA Biol, 8, 1149–1157. [DOI] [PubMed] [Google Scholar]
- Wood H, (2018). FDA approves patisiran to treat hereditary transthyretin amyloidosis. Nat Rev Neurol. [DOI] [PubMed] [Google Scholar]
- Wu JC, Meng QC, Ren HM, Wang HT, Wu J, & Wang Q, (2017). Recent advances in peptide nucleic acid for cancer bionanotechnology. Acta Pharmacol Sin, 38, 798–805. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yu AM, Tian Y, Tu MJ, Ho PY, & Jilek JL, (2016). MicroRNA Pharmacoepigenetics: Posttranscriptional Regulation Mechanisms behind Variable Drug Disposition and Strategy to Develop More Effective Therapy. Drug Metab Dispos, 44, 308–319. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang F, Wen Y, & Guo X, (2014). CRISPR/Cas9 for genome editing: progress, implications and challenges. Hum Mol Genet, 23, R40–46. [DOI] [PubMed] [Google Scholar]
- Zhang QY, Ho PY, Tu MJ, Jilek JL, Chen QX, Zeng S, & Yu AM, (2018). Lipidation of polyethylenimine-based polyplex increases serum stability of bioengineered RNAi agents and offers more consistent tumoral gene knockdown in vivo. Int J Pharm, 547, 537–544. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang X, Potty AS, Jackson GW, Stepanov V, Tang A, Liu Y, Kourentzi K, Strych U, Fox GE, & Willson RC, (2009). Engineered 5S ribosomal RNAs displaying aptamers recognizing vascular endothelial growth factor and malachite green. J Mol Recognit, 22, 154–161. [DOI] [PubMed] [Google Scholar]
- Zhao Y, Tu MJ, Wang WP, Qiu JX, Yu AX, & Yu AM, (2016). Genetically engineered pre-microRNA-34a prodrug suppresses orthotopic osteosarcoma xenograft tumor growth via the induction of apoptosis and cell cycle arrest. Sci Rep, 6, 26611. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhao Y, Tu MJ, Yu YF, Wang WP, Chen QX, Qiu JX, Yu AX, & Yu AM, (2015). Combination therapy with bioengineered miR-34a prodrug and doxorubicin synergistically suppresses osteosarcoma growth. Biochem Pharmacol, 98, 602–613. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhou J, & Rossi J, (2017). Aptamers as targeted therapeutics: current potential and challenges. Nat Rev Drug Discov, 16, 181–202. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhou J, Zhao Y, Simonenko V, Xu JJ, Liu K, Wang D, Shi J, Zhong T, Zhang L, Zeng L, Huang B, Tang S, Lu AY, Mixson AJ, Sun Y, Lu PY, & Li Q, (2017). Simultaneous silencing of TGF-beta1 and COX-2 reduces human skin hypertrophic scar through activation of fibroblast apoptosis. Oncotarget, 8, 80651–80665. [DOI] [PMC free article] [PubMed] [Google Scholar]






