Next Article in Journal
Blackgrass (Alopecurus myosuroides Huds.) Multiple Resistance to ACCase- and ALS-Inhibitors and Its Competition with Winter Wheat
Previous Article in Journal
Beyond the Curtains: Identification of the Genetic Cause of Foetal Developmental Abnormalities Through the Application of Molecular Autopsy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

RNA-Targeting Techniques: A Comparative Analysis of Modern Approaches for RNA Manipulation in Cancer Research and Therapeutics

by
Michaela A. Boti
,
Marios A. Diamantopoulos
and
Andreas Scorilas
*
Department of Biochemistry and Molecular Biology, Faculty of Biology, National and Kapodistrian University of Athens, Panepistimiopolis, 15701 Athens, Greece
*
Author to whom correspondence should be addressed.
Genes 2025, 16(10), 1168; https://doi.org/10.3390/genes16101168
Submission received: 13 August 2025 / Revised: 6 September 2025 / Accepted: 17 September 2025 / Published: 2 October 2025
(This article belongs to the Section RNA)

Abstract

RNA-targeting techniques have emerged as powerful tools in cancer research and therapeutics, offering precise and programmable control over gene expression at the post-transcriptional level. Once viewed as passive intermediates in the central dogma, RNA molecules are now recognized as dynamic regulators of cellular function, capable of influencing transcription, translation, and epigenetic regulation. Advances in high-throughput sequencing technologies, transcriptomics, and structural RNA biology have uncovered a diverse landscape of coding and non-coding RNAs involved in oncogenesis, drug resistance, and tumor progression. In response, several RNA-targeting strategies have been developed to modulate these transcripts, including antisense oligonucleotides (ASOs), RNA interference (RNAi), CRISPR-Cas13 systems, small molecules, and aptamers. This review provides a comparative analysis of these technologies, highlighting their molecular mechanisms, therapeutic potential, and current limitations. Emphasis is placed on the translational progress of RNA-targeting agents, including recent FDA approvals and ongoing clinical trials for cancer indications. Through a critical comparison of these strategies, this review underscores the growing significance of RNA-targeting technologies as a foundation for next-generation cancer therapeutics and precision oncology.

1. Introduction

For decades, the central dogma of molecular biology—describing the flow of genetic information from DNA to RNA to protein—has served as the foundation for understanding a variety of cellular processes and for replicating the related mechanisms in vitro [1,2,3]. RNA was long considered a short-lived molecule whose primary function was to act as messenger RNA (mRNA), produced during DNA transcription and subsequently translated into polypeptides by ribosomes [4,5,6]. However, advances in transcriptomics, as well as transcriptome analysis methods and technologies, have made clear that RNA molecules are not just passive intermediates of the gene expression process [7].
One of the most significant breakthroughs in RNA biology has been the development of high-throughput sequencing technologies, particularly RNA sequencing (RNA-seq) [8,9,10]. This approach enabled the comprehensive, unbiased, and quantitative analysis of the transcriptome with unprecedented depth and precision. Compared to earlier hybridization-based techniques, RNA-seq allows for the detection of novel transcripts, alternative splicing events, RNA modifications [11,12], and numerous non-coding RNAs (ncRNAs) [13], revealing that the majority of the human genome is transcribed into RNA, much of which carries regulatory functions rather than encoding proteins [7,14]. Additionally, the emergence of single-cell RNA sequencing has further advanced our understanding by revealing cell-to-cell variability and transcriptomic heterogeneity across tissues, developmental stages, and disease states [15,16]. These technological breakthroughs gave insights into RNA research, broadening the recognized roles of RNA well beyond protein translation, and changing our perception regarding RNA and its functions.
Beyond acting as template for protein synthesis, RNA plays diverse and essential roles within the cell. These molecules are key regulators of gene expression at both transcriptional and post-transcriptional levels, controlling a wide range of processes including chromatin remodeling, transcriptional activation or repression, mRNA splicing, editing, transport, localization, stability, and translation [17,18,19,20,21,22,23,24]. In addition to regulatory roles, some RNAs catalyze biochemical reactions, acting as ribozymes, while others serve as structural and functional components of large ribonucleoprotein complexes like the spliceosome and ribosome [25,26]. The functional diversity of the transcriptome is reflected by the various classes of ncRNAs, including microRNAs (miRNAs), small interfering RNAs (siRNAs), long non-coding RNAs (lncRNAs), and circular RNAs (circRNAs), as well as catalytically active RNAs [27,28]. These molecules play crucial roles in maintaining cellular homeostasis [29,30,31], while dysregulation of their biogenesis, processing, or activity has been implicated in various diseases, particularly in cancer [32,33,34,35,36,37]. More precisely, it is well-established that ncRNAs can promote tumor growth, invasion, and resistance to therapy or act as tumor suppressors. Due to their specific expression patterns and regulatory roles, they are increasingly recognized as promising therapeutic targets, as well as valuable tools in both basic research and therapeutics.
Despite the introduction of the discussed high-throughput approaches, functional characterization of RNA molecules has required the development of complementary technologies capable of targeting and manipulating RNA directly [38,39,40]. RNA-targeting technologies offer the ability to modulate gene expression post-transcriptionally, often with greater specificity and reversibility than DNA-targeting methods [41,42]. The growing recognition of RNA’s diverse roles in gene regulation, catalysis, and structural organization has further catalyzed the development of these technologies for both basic research and clinical applications. They have enabled the in-depth investigation of RNA functions, interactions, and localization, and they serve as powerful tools for studying the effects of RNA dysregulation, a hallmark of cancer [43,44]. The introduction of these approaches has transformed our ability to manipulate RNA molecules, offering new insights into cellular biology and enabling the development of innovative therapeutic strategies [45,46].
Several distinct RNA-targeting strategies have been developed, each exploiting unique molecular mechanisms to achieve post-transcriptional modulation. Among the most well-established are antisense oligonucleotides (ASOs) and RNA interference (RNAi), which utilize sequence complementarity to inhibit translation, affect splicing, or induce transcript degradation [47,48,49]. These approaches have already been translated into clinical settings and are used in FDA-approved therapies for genetic and neurological disorders. More recent innovations include CRISPR-Cas13 systems, which enable programmable RNA cleavage and editing, offering high specificity and modularity for both research and therapeutic applications [41,50]. Additionally, small molecules that bind structured RNA motifs have emerged as a strategy to alter RNA stability or function, especially for ncRNAs or pathogenic repeat expansions [51].
Despite these advances, several significant challenges remain in the field. The structural complexity and dynamic expression patterns of RNA, along with the difficulty of efficient intracellular delivery, continue to hinder the widespread application of RNA-targeting technologies [52,53]. Additional obstacles include off-target effects, unintended immune activation, and variable efficacy depending on cellular or disease context [52,53,54]. Moreover, many current strategies are constrained to specific cell types or pathological conditions, highlighting the need for continued innovation to improve specificity, potency, and delivery mechanisms. Nonetheless, the ability to directly and precisely manipulate RNA holds great potential for studying gene function, dissecting disease mechanisms, and developing next-generation therapeutics with unprecedented precision.
This review examines the expanding toolkit of RNA-targeting technologies—from antisense oligonucleotides and RNA interference to CRISPR-Cas13 and RNA-binding molecules—and their growing relevance in medical oncology. We focus on their underlying molecular mechanisms, preclinical and clinical progress, therapeutic applications, and the specific opportunities they present for treating complex diseases such as cancer. We also address the key challenges that continue to limit their widespread clinical adoption and discuss their potential advances for overcoming these limitations. Through a critical analysis of recent developments, this review aims to highlight the current impact and the future potential of RNA-targeted approaches in cancer research and therapeutics.

2. Diverse Strategies for Post-Transcriptional RNA Modulation

RNA-targeting technologies have become essential tools in cancer research, offering precise and versatile modulation of gene expression at the post-transcriptional level [55]. This section provides a comparative overview of both established and emerging RNA-targeting strategies that have been developed and optimized to investigate RNA functions and regulatory mechanisms in cancer biology. Traditional methods such as antisense oligonucleotides (ASOs) and RNA interference (RNAi) are discussed alongside more recent innovations like the CRISPR-Cas13 system and aptamers, which have significantly expanded the toolkit for transcriptome manipulation. The present section focuses on the molecular mechanisms underlying each approach, their optimization for specificity and efficiency, and their functional applications in cancer research. In addition, we discuss the fundamental distinctions among these strategies, highlighting their respective advantages and limitations in experimental processes.

2.1. Antisense Oligonucleotides: Sequence-Specific RNA Targeting

One of the earliest methods developed for RNA manipulation is antisense oligonucleotides (ASOs) technology. First proposed in the late 1970s (Figure 1) [56], this approach involved the design of short, synthetic strands of DNA or RNA that can bind to a target RNA in a sequence-specific manner via Watson–Crick base pairing, enabling selective modulation of gene expression [57]. Over the decades, chemical modifications have been introduced to enhance ASOs’ stability, binding affinity, and cellular uptake, making them powerful tools for both basic research and therapeutic applications, especially in oncology.
Based on the type of chemical modifications they have been subjected to, ASOs are generally classified into three generations. First-generation chemistries include the widely used phosphate backbone modifications, e.g., phosphorothioate (PS). These modifications confer resistance to endonucleases and improve bioavailability but reduce the affinity for the target RNA [58]. Second-generation ASOs introduce modifications to the ribose sugar, particularly at the 2′-O position in RNA or the 2′ position in DNA. The most common examples of this generation are 2′-O-methyl (2′-OMe), 2′-O-methoxy-ethyl (2′-MOE) and 2′-fluoro (2′-F). These modifications enhance RNA binding affinity and further improve resistance to nuclease degradation [59]. Additionally, conformationally constrained DNA analogues like locked nucleic acid (LNA) and tricyclo-DNA (tcDNA) exhibit even greater binding affinity. LNAs contain a methyl bridge connecting the 2′-O and 4′ carbon atoms of the ribose ring [60], while tcDNAs introduce an ethylene bridge with a cyclopropane ring between the 3′ and 5′ carbon positions [61]. The presence of a chemical bridge locks the ribose ring in a favorable conformation for RNA hybridization, thereby enhancing binding efficiency. Third-generation chemistries involve more extensive modifications to the nucleobase or sugar-phosphate backbone. Typical examples include phosphonodiamidate morpholino oligomers (PMOs) and peptide nucleic acids (PNAs) [62,63]. Both PMOs and PNAs are uncharged, highly resistant to nucleases, and exhibit variable binding affinity to target RNA sequences, providing unique pharmacological properties ideal for therapeutic applications. [62,64].

2.1.1. Dual Strategies: Degradation and Steric Blockade by ASOs

Based on their mechanism of action, ASOs are categorized into two main functional classes: gapmer ASOs, which induce degradation of target RNA transcripts, and steric-blocking ASOs (SBOs), which bind to their target RNA and sterically hinder its interaction with RNA-binding proteins or components of the cellular machinery (Figure 2) [65]. Both approaches enable precise modulation of gene expression at post-transcriptional level, but through distinct molecular pathways.
Gapmer ASOs are chimeric oligonucleotides designed to induce RNA degradation via RNase H1-mediated cleavage (Figure 3) [66,67]. Structurally, they consist of a central region of deoxyribonucleotides—the “gap”—flanked by chemically modified ribonucleotides on both the 5′ and 3′ ends [68]. These modifications enhance target binding affinity and prevent recognition by RNase H1, increasing nuclease resistance. Upon sequence-specific binding to the target RNA, the central DNA region forms a DNA/RNA duplex that is recognized by RNase H1, an endogenous endonuclease that selectively cleaves the RNA strand of the hybrid duplex while leaving the ASO intact for subsequent target recognition and cleavage [69]. This mechanism results in degradation of the target transcript and has been extensively applied in cancer research to downregulate oncogenic transcripts, disrupt tumor survival pathways, and overcome drug resistance. The incorporation of nuclease-resistant flanks improves pharmacokinetic stability and minimizes off-target interactions, while the DNA “gap” maintains the functional compatibility required for efficient RNase H1 recruitment and catalytic activity.
In contrast, SBOs exert their function by directly binding to specific sequences within target RNA molecules without inducing degradation (Figure 3) [65]. Unlike gapmer ASOs, which recruit RNase H1 to cleave the target RNA, SBOs act through a physical, occupancy-based mechanism. Upon binding, the oligonucleotide sterically hinders the access of RNA-binding proteins, components of the spliceosome or ribosome, and other elements of the RNA-processing machinery [70,71]. This interference can disrupt essential interactions involved in RNA splicing, translation initiation, or RNA stability, thereby modulating the fate and function of the RNA transcript. By selectively targeting cis-acting regulatory elements, SBOs can redirect alternative splicing events, alter mRNA maturation, or prevent the recruitment of translational machinery [57]. A notable subclass of steric-blocking ASOs is the splice-switching ASOs (ssASOs), which specifically bind to pre-mRNA splice sites and modulate splicing [68,72].
In addition to modulating mRNA processing, SBOs have been effectively used to target ncRNAs, particularly miRNAs. Anti-miRNAs (AMOs) are a class of SBOs designed to bind mature miRNAs, preventing their interaction with target mRNAs and thereby inhibiting miRNA-mediated repression. Through these diverse, non-degradative mechanisms, SBOs enable precise post-transcriptional regulation of gene expression while preserving RNA integrity, making them valuable therapeutic tools for correcting splicing defects, modulating protein expression, and inhibiting oncogenic miRNA activity.

2.1.2. ssASOs: Splicing Modulation Using Steric-Blocking ASOs

In contrast to the gapmer-based mechanism that results in transcript degradation, ssASOs modulate pre-mRNA processing without inducing RNA cleavage (Figure 3) [73]. These oligonucleotides bind to specific splice sites or cis-regulatory elements like exonic splicing enhancers (ESEs), silencers (ESSs), or intronic elements, and direct alternative splicing [74]. Depending on the target site and the chemistry of the oligonucleotide, ssASOs can induce exon skipping or exon inclusion, thereby restoring or modifying the coding potential of the mature mRNA. ssASOs are commonly used to restore or disrupt reading frames, leading to increased protein levels for loss-of-function (LoF) variants or decreased protein levels for toxic gain-of-function (GoF) variants, respectively, dysregulations that are commonly observed in cancer [75,76,77]. ssASOs are typically composed of fully modified nucleotides, such as PMOs or LNAs, which are resistant to RNase H activity. This ensures the integrity of the RNA while allowing selective modulation of splicing events [78,79].
AMOs are synthetic ASOs designed to bind with great specificity to mature miRNAs (Figure 3) [80]. Once bound, they sterically hinder the miRNA from interacting with its mRNA targets and from being loaded into RISC, effectively neutralizing its gene-silencing function [81]. Mechanistically, AMOs resemble the action of gapmer ASOs but differ from them, since they typically do not induce degradation of their targets; instead, they function through competitive inhibition of miRNA–mRNA interactions [81,82,83].
Much like ASOs, AMOs are often subjected to chemical modification such as 2′-OMe, LNAs, or PS backbones to enhance nuclease resistance, binding affinity, and cellular uptake. These modifications improve the stability and efficacy of AMOs in vivo [83]. However, due to the transient nature of AMOs, inhibition of miRNA function is short-lived, necessitating multiple administrations to achieve a long-term knockdown effect.

2.2. Beyond AMOs: Using miRNA Sponges for miRNA Modulation

In cancer, numerous miRNAs are dysregulated, with oncogenic miRNAs (oncomiRs) often being upregulated. The overexpression of these miRNAs leads to the post-transcriptional repression of tumor suppressor genes, thereby promoting cancer development, proliferation, and metastasis [84]. While AMOs have been widely used to counteract these effects, additional RNA-based strategies such as miRNA sponges have also been developed. In cancer research, AMOs and sponges have been extensively used to investigate the functional relevance of well-characterized oncomiRs such as miR-21, miR-155, and the miR-221/222 cluster, which are frequently overexpressed across multiple tumor types. Functional inhibition of these miRNAs has revealed their critical roles in regulating apoptosis, cell cycle progression, metastasis, and immune modulation, underscoring their therapeutic potential and importance in cancer biology [85,86,87].

miRNA Sponges: Decoy Transcripts for Sustained miRNA Inhibition

In contrast to synthetic AMOs, miRNA sponges are typically DNA-encoded RNA transcripts engineered to contain multiple tandem binding sites complementary to one or more miRNAs [83,88]. These binding sites are often designed to match the miRNA seed region—critical for target recognition—allowing the sponge to sequester not just a single miRNA, but entire families sharing the same seed sequence. By competing with endogenous mRNA targets, sponges reduce the availability of free miRNAs for RISC loading and activity [83,89].
miRNA sponges are usually delivered via plasmid or viral vectors, resulting in their stable expression within the cell [89]. This enables long-term inhibition of miRNA function, making sponges particularly suited for in vivo cancer studies. Compared to AMOs, sponges offer prolonged miRNA suppression but often exhibit lower specificity and may cause off-target effects due to broader miRNA family inhibition. Moreover, their efficacy can vary depending on promoter strength, high endogenous miRNA concentration, and cell type [83,90].

2.3. RNA Interference: Harnessing the Cell’s Machinery for Precise RNA Degradation

RNA interference (RNAi) is a naturally occurring, sequence-specific post-transcriptional gene silencing mechanism, first described in Caenorhabditis elegans (C. elegans) in 1998 (Figure 1) [91]. It is mediated by small double-stranded RNA (dsRNA) molecules that direct intracellular machinery to selectively degrade complementary mRNA targets, thereby preventing their translation into protein [92]. Since its discovery, RNAi has become an indispensable tool in functional genomics and a promising therapeutic strategy, especially in cancer where aberrant gene expression drives tumor growth and progression [93].
The primary effectors of RNAi-based approaches are small interfering RNAs (siRNAs) and short hairpin RNAs (shRNAs) [94,95]. More precisely, siRNAs are synthetic RNA duplexes, typically 21–23 base-pairs in length, with characteristic 2 nucleotides 3′ overhangs that allow them to be recognized by the enzymatic machinery of RNAi. These molecules are designed to be fully complementary to their target RNA, enabling precise, sequence-specific silencing [95]. To enhance the efficacy of RNAi-based molecules, siRNAs are chemically modified similarly to ASOs, most commonly with 2’-OMe, PS, and 2’-F. These modifications improve nuclease resistance, immune evasion and target specificity, making them more suitable for in vivo approaches and clinical applications [96]. In contrast, shRNAs are generated intracellularly through the transcription of DNA vectors (e.g., plasmids or viral systems) that have been previously transfected into the cell [94,97]. The produced shRNAs are folding and form stem-loop structures. After being exported to the cytoplasm, these structures are processed by the RNase III enzyme Dicer into siRNA-like duplexes [98].

2.3.1. Mechanistic Insights into RNAi: How siRNAs and shRNAs Achieve Targeted Gene Silencing

RNAi-mediated gene silencing occurs through a multistep process involving recognition, loading, and cleavage of target transcripts. For synthetic siRNAs, the duplex is introduced into the cytoplasm where it is directly loaded into the RNA-induced silencing complex (RISC) (Figure 3). The sense (passenger) strand of siRNA is degraded, and the antisense (guide) strand directs RISC to complementary mRNA transcripts [99]. Subsequently, Argonaute 2 (AGO2) mediates precise cleavage between the 10th and 11th nucleotides relative to the guide strand’s 5′ end [100], resulting in rapid degradation of the transcript by exonucleases [99].
In the case of shRNAs, the RNAi pathway begins in the nucleus, where the hairpin-structured RNA is generated via transcription and exported to the cytoplasm by Exportin-5 [101], a protein that is not required for the activity of synthetic siRNAs. Once in the cytoplasm, Dicer processes the hairpin into a siRNA-like duplex, which enters the same downstream pathway as exogenous siRNA (Figure 3) [97]. This approach offers a more sustained knockdown due to the continuous intracellular production of the shRNA; however, it may also lead to acute cytotoxicity, particularly at high expression levels or with certain delivery systems.

2.3.2. Limitations of siRNA and shRNA: Choosing the Optimal RNAi-Based Tool

While both siRNAs and shRNAs harness the cell’s endogenous RNA interference machinery to silence gene expression, they differ significantly in design, delivery, and duration of effect (Table 1). siRNAs are synthetic molecules delivered directly into the cytoplasm, offering rapid and transient silencing with minimal cellular processing requirements [95]. Their use allows for precise control over dosage and timing, and they are often favored for short-term experiments or applications where regulated, transient expression is desirable.
In contrast, shRNAs are expressed intracellularly from DNA vectors, enabling long-term and stable gene knockdown due to continuous transcription [94]. This makes them particularly suitable for applications requiring prolonged silencing, such as in vivo disease models or stable cell lines. However, shRNA-based approaches involve nuclear transcription, export, and cytoplasmic processing by Dicer, introducing additional steps that may increase complexity and variability [97]. Moreover, high or unregulated expression of shRNAs—particularly when driven by strong promoters—has been associated with cytotoxicity [102], while their use can also lead to target-independent dysregulation of endogenous miRNAs and global transcriptomic changes [103]. The choice between siRNA and shRNA depends on the specific experimental objectives, the required duration and reversibility of gene silencing, and the biological context in which the RNAi is applied. Careful consideration of these factors is essential during both experimental design and therapeutic development.
Table 1. Comparative overview of siRNA and shRNA, two widely used molecules for RNAi-mediated RNA degradation. This table summarizes the key characteristics of small interfering RNA (siRNA) and short hairpin RNA (shRNA), highlighting their respective advantages and limitations in research and therapeutics.
Table 1. Comparative overview of siRNA and shRNA, two widely used molecules for RNAi-mediated RNA degradation. This table summarizes the key characteristics of small interfering RNA (siRNA) and short hairpin RNA (shRNA), highlighting their respective advantages and limitations in research and therapeutics.
FeaturesiRNAshRNA
GenerationSynthetic RNA duplexes delivered into
cytoplasm [95,104]
Vector-expressed hairpin RNA transcribed in the nucleus [97,104]
Mechanism of ActionDirect incorporation into RISC; AGO2-mediated mRNA cleavage [95]Transcription, export from nucleus, processed by Dicer into siRNA duplex, RISC loading [98]
Duration of EffectTransient and rapid [104,105]Long-lasting and stable due to continuous intracellular expression [97]
DeliveryRequires delivery vehicles (e.g., lipid nanoparticles) [106]Delivered via plasmid or viral vectors [97,104,107]
AdvantagesPrecise control over dosage and timing; minimal processing [107]Sustained knockdown; suitable for stable gene silencing [94,97]
LimitationsSusceptible to degradation; transient effect [104,108]Risk of cytotoxicity; insertional mutagenesis; delayed onset [102,103]
ApplicationsShort-term gene silencing for in vivo functional studies [104,105]In vivo disease models, stable cell line generation [108]

2.4. Emergence of CRISPR-Cas13 Systems

While RNAi represented a significant advancement in post-transcriptional gene regulation, the discovery of CRISPR-Cas system in the 2010s marked the next major leap in RNA manipulation technologies (Figure 1) [109,110]. Originally identified as an adaptive immune defense in bacteria [111], CRISPR-Cas systems were initially developed and adapted for genome editing applications [109,112,113,114,115]. However, the subsequent identification of Cas13 protein, an RNA-targeting effector, expanded the scope of CRISPR technologies beyond DNA editing [116,117,118]. Unlike Cas9, which induces double-stranded DNA breaks, Cas13 associates with a guide RNA (gRNA) that directs it to cleave single-stranded RNA (ssRNA) in a sequence-specific manner, without affecting genomic DNA [115,116]. This feature has enabled precise transcriptome modulation, facilitating functional RNA studies in cancer biology and laying the foundation for the development of novel RNA-targeting therapeutic strategies [119,120,121,122,123].
The Cas13 family is classified into 11 distinct subtypes: Cas13a, Cas13b, Cas13c, Cas13d, Cas13e, Cas13f, Cas13g, Cas13h, Cas13i, Cas13x, and Cas13y [124,125,126,127,128]. While all Cas13 proteins function as RNA-guided RNases, they differ in size, domain architecture, gRNA requirements, and functional efficiency in eukaryotic systems. Among the diverse Cas13 subtypes, Cas13a, Cas13b, and Cas13d have been most extensively characterized and optimized for use in mammalian cells, due to their robust activity and adaptability for transcriptome engineering applications [129,130,131,132].

2.4.1. Cas13-Mediated Mechanism: From Recognition to Cleavage

Cas13-mediated RNA cleavage proceeds through a multistep process involving gRNA recognition, engagement, and catalytic cleavage [116]. Once introduced into the cell, Cas13 forms a ribonucleoprotein complex with the gRNA, which recruits Cas13 to the target RNA (Figure 3) [133]. In contrast to RNAi, which relies on host components like Dicer and Argonaute, Cas13 functions as a self-sufficient RNA-guided endonuclease [100].
Upon base-pairing between the gRNA and the target ssRNA, Cas13 undergoes a conformational rearrangement that activates its two Higher Eukaryotes and Prokaryotes Nucleotide-binding (HEPN) domains [134]. These domains come into close proximity, forming an active catalytic center that cleaves the target RNA, typically at flexible, unstructured regions near the hybridization site [134,135]. Unlike AGO2-mediated RNAi, Cas13 does not require a specific sequence motif for cleavage, allowing broader target flexibility.

2.4.2. Current Limitations of the CRISPR-Cas13 System

CRISPR-Cas13 offers several advantages over traditional RNA-targeting technologies, including high specificity and efficient editing. Moreover, this system is highly flexible, allowing for the design of gRNAs that target a broad range of RNA molecules, enabling precise post-transcriptional modulation.
Despite its promise, the CRISPR-Cas13 system also faces notable limitations. One of the major challenges is the potential collateral activity that results in off-target effects [136]. Specifically, upon recognition and cleavage of the intended RNA target, Cas13 undergoes a conformational change that activates its HEPN nuclease domains. This activation can result in non-specific cleavage of nearby, non-target RNAs, potentially disrupting normal cellular function and confounding experimental results [137,138]. Collateral activity differs from traditional off-target effects, which also have been reported in CRISPR-Cas13 systems. More precisely, off-target effects are independent of the presence of the intended on-target, whereas collateral activity is reported only upon recognition of on-target RNA that perfectly matches the gRNA [133]. Recent studies have shown that expressing Cas13 or gRNA alone can trigger off-target effects, intrinsic targeting of host RNAs, and cellular toxicity [139]. Furthermore, while efficient and cell-type-specific delivery of CRISPR-Cas13 components in vivo remains an area of active development and is still challenging for specific tissues [140], recent studies have demonstrated that it can be effectively achieved in various tissues using adeno-associated viruses (AAVs) and polymer-based systems [141,142]. Furthermore, although Cas13 does not require protospacer adjacent motifs (PAMs) for target recognition, some orthologs (Cas13a, Cas13b, Cas13c) recognize protospacer flanking sequences (PFSs) [143], a restriction that should be taken under consideration when designing gRNAs.

2.4.3. Optimizing CRISPR-Cas13 Systems: Recent Advances in Cas13-Based RNA Cleavage

Importantly, Cas13 variants have recently been engineered to overcome collateral activity [144]. A dual-fluorescence reporter system in mammalian cells was used to screen over 200 engineered Cas13 variants to assess collateral effects and on-target efficiency. The Cas13d and Cas13X variants demonstrated robust on-target activity with significantly reduced collateral effects. These high-fidelity variants exhibited minimal transcriptome-wide off-target effects, did not induce detectable cytotoxicity or cell growth arrest, and maintained RNA knockdown efficacy comparable to that of the wild-type enzyme. Importantly, in vivo assessments using transgenic mouse models and adeno-associated virus (AAV)-mediated delivery confirmed the absence of measurable collateral RNA cleavage, highlighting their safety and functional precision [42]. These high-fidelity Cas13 variants represent a major step forward in the development of more precise RNA-targeting technologies, offering enhanced specificity and safety profiles for both research and potential therapeutic applications.

2.4.4. Beyond Cleavage: Expanding the Potential of Cas13-Based Systems in RNA Biology

A highly versatile advancement of the CRISPR-Cas13 platform is the engineering of catalytically inactive Cas13, commonly known as dead Cas13 (dCas13) [116]. This variant is generated through point mutations in the enzyme’s HEPN domains, effectively deactivating its intrinsic ribonuclease activity while retaining its sequence-specific RNA-binding capacity [145]. As a result, dCas13 functions as a programmable RNA-binding scaffold, capable of being fused to diverse effector domains to modulate RNA fate and function in a sequence-specific manner [146].
One of the most powerful applications of dCas13 is RNA base editing, particularly adenosine-to-inosine (A-to-I) editing. Fusion of dCas13 to catalytically active deaminase domains, such as adenosine deaminase acting on RNA 2 (ADAR2), enables site-specific editing of endogenous RNA transcripts (Figure 3) [147,148,149]. This technology underpins systems such as REPAIR (RNA Editing for Programmable A-to-I Replacement) and RESCUE (RNA Editing for Specific C-to-U Exchange) [145,150], which can allow transient and reversible correction of pathogenic point mutations at the RNA level, without altering the genome. This capability holds significant potential for modeling cancer-associated mutations, investigating their functional impact, and correcting oncogenic variants, contributing to the development of innovative RNA-targeted therapeutic strategies [149]. In addition to base editing, dCas13 has been exploited for performing epitranscriptomic changes by fusing it to RNA-modifying enzymes such as methyltransferases or demethylases, enzymes that alter N6-methyladenosine (m6A) marks [151,152,153]. These engineered systems enable site-specific addition or removal of specific chemical groups to ribonucleotides, offering new avenues for investigating how m6A and other epitranscriptomic marks regulate RNA stability, translation efficiency, and degradation [154,155,156,157]. Given the growing recognition of epitranscriptomic dysregulation in cancer, dCas13-based tools provide a powerful platform to explore the functional impact of RNA modifications on tumorigenesis, cancer progression, and therapeutic resistance [154].
Beyond RNA editing, dCas13 has proven effective in post-transcriptional gene regulation. For example, when fused to initiation factors, dCas13 can selectively enhance the translation of target mRNAs by promoting translational initiation [158]. In contrast, when dCas13 binds to the target mRNA near its 5′-end, it acts as a physical roadblock for scanning ribosomes, resulting in translational repression without affecting mRNA stability (Figure 3) [159]. Similarly, fusion of dCas13 to RNA-binding splicing factors (e.g., hnRNPa1 domains) enables programmable modulation of alternative splicing patterns at specific pre-mRNA sites [158,160]. This approach is particularly valuable for investigating and modulating cancer-associated splice variants implicated in cancer development, progression, and therapeutic resistance [161,162,163].
Additionally, dCas13 is a powerful tool for real-time RNA imaging in live cells [164]. When fused to fluorescent proteins (e.g., dCas13-GFP), it enables sequence-specific visualization of RNA localization, trafficking, and dynamic expression patterns without influencing endogenous RNA integrity (Figure 3) [165]. This application provides critical insight into the real-time subcellular localization of oncogenic transcripts, RNA–protein interactions, and transcriptomic reprogramming during cancer progression, cellular stress responses, and treatment-induced transcriptional remodeling [165,166,167].
dCas13 is a versatile and precise tool for studying and manipulating RNA molecules in a sequence-specific manner. Its expanding applications—including epitranscriptomic editing, splicing modulation, and live-cell RNA imaging—have made it especially valuable in cancer research. These capabilities provide powerful means to investigate RNA functions, the impacts of its dysregulation and to develop targeted RNA-based strategies for both diagnostics and potential therapies.

2.5. Small Molecules as RNA-Targeting Tools in Cancer Research

Small molecules have traditionally served as the foundation of drug discovery, primarily targeting proteins to modulate their function [168,169]. However, recent advances in RNA structural biology, high-throughput technologies, and computational modeling have broadened their application to include RNA molecules as direct targets [8,170]. This expansion is particularly significant in cancer research, where small molecules offer promising strategies to modulate or degrade dysregulated RNA transcripts—including oncogenic mRNAs, lncRNAs, and miRNAs—that contribute to tumor progression and resistance to therapy [171,172,173]. In contrast to ASOs, siRNA/shRNA, and CRISPR-Cas13 systems that rely on Watson–Crick base pairing, small molecules interact with RNA via non-covalent forces including hydrogen bonding, π–π stacking, and hydrophobic interactions [174]. These interactions enable the recognition of complex and often transient higher-order RNA structural motifs—such as internal loops, bulges, apical loops, pseudoknots, and G-quadruplexes (G4s)—thereby offering a structurally driven strategy for modulating RNA function in cells [174,175]. One particularly attractive class of RNA structures in oncology is G4s, motifs that can form in the untranslated regions (UTRs) of mRNAs and influence their stability and translational efficiency. G4s have been identified in the 5′ UTRs of several oncogenic transcripts, including those encoding NRAS and KRAS, where they act as translational repressors [176]. Stabilization of these structures using small-molecule ligands has been shown to inhibit translation of the associated oncogenic transcripts, offering a strategy to suppress cancer drivers at the RNA level [177,178]. Several small-molecule G4 ligands have demonstrated the ability to selectively bind and stabilize these motifs, leading to the suppression of oncogene expression in cancer cells [178,179].

2.5.1. Functional Mechanisms of Small-Molecule RNA Targeting

RNA-targeting small molecules exert their therapeutic effects through diverse structure-dependent mechanisms, making them particularly promising in cancer treatment [174,180]. One major strategy involves allosteric inhibition of translation or ribonucleoprotein (RNP) complex formation. In this mechanism, small molecules bind to highly structured regions such as internal ribosome entry sites (IRES) or G4s within oncogenic mRNAs or viral mRNAs that are linked to cancer development, thereby blocking cap-independent translation and disrupting critical RNA–protein interactions required for mRNP assembly (Figure 4A) [181,182,183].
Another mechanism targets miRNA biogenesis, where compounds that bind to the stem-loop structures of primary or precursor miRNAs can interfere with processing by Drosha or Dicer [184], leading to decreased levels of oncogenic miRNAs in tumor cells [185,186]. Additionally, small molecules can modulate alternative splicing by stabilizing interactions between pre-mRNA and components of the spliceosome, such as U1 snRNP (Figure 4A) [187]. This mechanism can be utilized to promote exon inclusion and highlights the potential of splicing modulation in malignancies driven by aberrant splice variants [188].
Although classical small molecules do not typically induce RNA degradation through pathways like RNaseH1 or RISC, emerging approaches have adapted targeted degradation strategies for RNA. One such strategy is the development of ribonuclease-targeting chimeras (RiboTACs), which are bifunctional molecules consisting of an RNA-binding domain linked to a recruiter of endogenous ribonucleases like RNase L [189], that have been designed to selectively degrade RNAs associated with cancer [190,191]. More precisely, these chimeras can induce selective cleavage of oncogenic transcripts by leveraging proximity-induced enzymatic degradation [192], highlighting their significant potential in cancer treatment development. Compared to oligonucleotide-based therapeutics like ASOs and RNAi-based approaches, RIBOTACs are characterized by enhanced pharmacokinetic properties since—just like a small molecule—they are likely to have access to a broad range of tissues when administered to animals (Table 2) [193].
Finally, some small molecules function by stabilizing tumor-suppressive RNAs, including certain lncRNAs, enhancing their structural integrity and biological activity, and thus prolonging their half-life and tumor-suppressive functions within cancer cells [194]. These mechanisms underscore the versatility of small molecules as a platform for modulating RNA function in cancer research and therapeutics.
Table 2. Comparative summary of RNA-targeting approaches. This table provides a comparative overview of major RNA-targeting strategies applied in cancer research and therapeutics, including their mechanisms of action, primary molecular targets, advantages, and limitations.
Table 2. Comparative summary of RNA-targeting approaches. This table provides a comparative overview of major RNA-targeting strategies applied in cancer research and therapeutics, including their mechanisms of action, primary molecular targets, advantages, and limitations.
TechniqueMechanism of ActionPrimary
Target
AdvantagesLimitations
Gapmer ASOs
[57,65,70]
Bind complementary RNA and recruit RNase H for RNA cleavage and degradationmRNAs, lncRNAsEnhanced binding affinity; nuclease resistance; reduced immunogenicity; high stabilitySequence- and structure-dependent hepatotoxicity and possible nephrotoxicity (driven by off-target RNase H1-mediated cleavage of pre-mRNAs); requires nuclear delivery; off-target effects; tumor heterogeneity;
SBOs
[70,71,72]
Bind target RNA to sterically block translation or modulate splicingPre-mRNAs, mRNAsEnables splicing or translation modulation without degrading RNA; lower risk of degradation-associated toxicityLimited to steric interference; nuclear delivery still challenging; not suitable for complete gene knockdown; often requires nuclear delivery;
RNAi/
siRNAs
[49,104]
Duplex siRNAs incorporated into RISC guide AGO2 to cleave perfectly matched RNAsmRNAsPotent knockdown; high sequence specificity; exploits endogenous machineryExtracellular and intracellular barriers (e.g., nuclease degradation, poor membrane permeability, renal clearance, endosomal entrapment); risk of immune activation; off-target silencing; cytotoxicity;
RNAi/shRNAs
[49,104,195]
Expressed shRNAs are processed by Dicer into siRNAs, loaded into RISC to degrade RNAsmRNAsLong-lasting effect via plasmid or viral delivery; uses endogenous RNAi machineryRisk of insertional mutagenesis (viral vectors); potential competition with endogenous miRNA pathways; off-target effects; immune response;
AMOS
[83,90]
Bind mature miRNAs to block interaction with endogenous mRNAsMature
miRNAs
Specific inhibition of miRNA functionPossible off-target effects across related miRNA family members; challenges in delivery specificity and stability;
Aptamers
[196,197,198]
Structured nucleic acid ligands bind RNA or other targets with high affinityStructured RNA motifsHigh specificity & affinity; low immunogenicity; customizable for diverse targetsDependent on well-defined RNA structure; limited high-resolution structural data;
CRISPR–Cas13
Systems
[50,112,119,129,133]
Cas13–gRNA complex binds and cleaves target ssRNA via HEPN domainsmRNAs, ncRNAsHighly programmable, specific, reversible;Challenging delivery to specific tissues; collateral cleavage can cause off-target RNA degradation and cytotoxicity; often requires PFS; immunogenicity concerns;
CRISPR–dCas13
Systems
[158,159,160]
Catalytically inactive Cas13 binds RNA and directs fused effectors (editing, splicing)mRNAs, pre-mRNAs, ncRNAsProgrammable modulation (splicing, editing, translation) without cleavageLarge size complicates delivery; immunogenicity; long-term safety and efficacy not yet fully evaluated; collateral effects remain a concern;
Small
Molecules
[51,168,170]
Bind RNA structural motifs to modulate interactions, splicing, translation, miRNA biogenesisStructured RNA motifsChemically tunable; potential for diverse RNA function modulationRNA structural variability; limited high-resolution structural data hinders rational design; risk of off-target binding

2.5.2. Challenges in Targeting RNA with Small Molecules

Despite growing progress in targeting RNA with small molecules, significant barriers remain. One of the major challenges is the limited availability of high-resolution, experimentally validated RNA three-dimensional structures [180]. Since the efficacy of small molecules in modulating RNA function is dependent on recognizing specific secondary or tertiary structural motifs, a detailed understanding of the target RNA conformation is essential [174]. Unlike the extensive and diverse databases available for RNA sequences, current RNA structure repositories are disproportionately populated by a narrow subset of elements such as ribosomes and riboswitches [180], without significant structural data for cancer-associated RNAs, including lncRNAs, oncogenic mRNAs, and aberrantly spliced transcripts. This gap in knowledge impedes rational drug design and limits the application of structure-based modeling techniques.
Furthermore, despite major advances in protein structure prediction through deep learning algorithms—such as AlphaFold, which has revolutionized de novo protein modeling for drug discovery—analogous progress in RNA structural prediction remains restricted [199,200]. This gap stems primarily from the limited quantity and compositional diversity of experimentally determined RNA tertiary structures, which constrain the training of machine learning algorithms and impedes the generation of accurate, predictive models for RNA folding [201]. As a result, current computational tools lack the robustness required to reliably predict the complex and dynamic architectures of functional RNAs, particularly those implicated in cancer pathogenesis.
In addition, the chemical space for targeting RNA remains poorly investigated, as traditional drug design principles—optimized for proteins—do not apply to RNA [180]. Thus, designing compounds with appropriate physicochemical properties to selectively and potently bind structured RNA elements demands new, data-driven frameworks. Effective binding of a small molecule to the RNA target does not guarantee functional modulation, as high-affinity interactions may fail to induce the necessary conformational changes or disrupt critical RNA-mediated processes required for biological activity [180].
Another critical limitation in the field pertains to the development of heterobifunctional RNA-targeting compounds like RiboTACs [193]. While these molecules represent a promising strategy by harnessing endogenous ribonucleases to achieve selective RNA degradation, the current repertoire of effector modules (e.g., RNase recruiters) remains limited, thereby restricting the applicability of this platform across different RNA classes. Additionally, robust methods for target validation and transcriptome-wide selectivity profiling are not yet well-established. This lack of comprehensive profiling raises concerns about off-target effects and unintended interactions with structurally similar RNA elements, proteins, or even DNA [180].
Collectively, these limitations highlight the urgent need for integrative approaches that combine experimental RNA structural elucidation, advanced computational modeling, and transcriptome-wide functional screening to refine RNA-targeting strategies and enhance their therapeutic utility in cancer (Table 2).

2.6. Aptamers: Nucleic Acid Ligands for RNA Targeting in Oncology

Aptamers are synthetic single-stranded nucleic acids capable of folding into complex three-dimensional conformations that bind to protein targets with high specificity and affinity [202]. Aptamers with affinity for a desired target are selected from a large oligonucleotide library through a process called SELEX, which stands for Sequential Evolution of Ligands by Exponential Enrichment [203,204,205]. During this process, non-binding aptamers are discarded and aptamers binding to the proposed target are expanded. Initial positive selection rounds are sometimes followed by negative selection, a strategy that improves the selectivity of the resulting aptamer candidates. After multiple selection rounds, high-affinity aptamer candidates with optimal binding characteristics are isolated for downstream applications (Figure 4B) [203,206].
A recent breakthrough study has provided the first direct experimental evidence that aptamers can modulate RNA function by binding structured RNA targets without inducing degradation. Specifically, L-DNA aptamers were shown to recognize and bind with high affinity to precursor miR-155, a structured RNA critical in oncogenic pathways. Unlike traditional Watson–Crick base pairing used by antisense oligonucleotides or siRNAs, these L-DNA aptamers engage in cross-chiral interactions with D-RNA, demonstrating a unique mode of molecular recognition. Importantly, the bound aptamers effectively inhibited Dicer-mediated processing of pre-mir-155, thus blocking its maturation into the functional miRNA [207]. Notably, the sequences and binding modes of these L-DNA aptamers are distinct from previously reported L-RNA aptamers, suggesting a diverse landscape of aptamer–RNA recognition strategies. This study not only establishes the feasibility of direct RNA targeting by aptamers but also highlights the potential of L-DNA as a chemically robust and biologically stable platform for therapeutic applications. Despite this promise, further work is needed to validate the generalizability of this approach to other structured RNAs and to evaluate its efficacy in vivo.
Aptamers are also emerging as biosensors for RNA detection in diagnostic applications. Their high affinity, specificity, and ability to undergo conformational changes upon target binding make them ideal molecular recognition elements in biosensor platforms. Aptamer-based biosensors (aptasensors) have been developed for the sensitive and selective detection of oncogenic RNAs (Figure 2), including dysregulated miRNAs commonly associated with cancer [208,209]. These biosensors can be integrated with various signal transduction modalities, such as electrochemical, optical, or fluorescence-based systems, enabling real-time and label-free detection. Their chemical stability, ease of modification, and low immunogenicity offer key advantages over traditional antibody-based detection methods [210]. In oncology, aptasensors hold promise for early disease detection, monitoring of treatment response, and real-time quantification of therapeutic targets, thus providing a complementary diagnostic modality to aptamer-based therapeutics within an integrated precision medicine framework.

Exploiting Aptamers for the Generation of Aptamer-Oligonucleotide Chimeras in RNA Therapeutics

Their programmable nature, small size, and low immunogenicity make aptamers particularly attractive not only for targeting RNA molecules directly but also for delivering RNA-based therapeutics to specific cell types. In recent years, aptamer-oligonucleotide chimeras have emerged as a powerful strategy to enhance the precision, efficacy, and safety of RNA-based therapeutics. [198,211]. In detail, aptamers can be chemically or enzymatically fused to therapeutic oligonucleotides, forming aptamer-oligonucleotide chimeras that combine the cell-targeting capabilities of aptamers with the previously described functions of siRNAs, shRNAs, AMOs, or ASOs [196,197,212,213,214].
Aptamer-siRNA chimeras (AsiCs) enable receptor-specific gene silencing and have been evaluated in preclinical models for cancer therapy and immune modulation. For instance, an EpCAM-targeting aptamer fused to a siRNA against Polo-like kinase 1 (PLK1) effectively reduced tumor burden in breast cancer models [215]. Modifications such as the addition of 3′ di-uridine overhangs to enhance DICER recognition, or incorporation of guide strand–aptamer fusions with optimized stem-loop structures, have been shown to improve processing and silencing efficiency [215,216,217]. Furthermore, AsiCs have been engineered to sensitize immune-refractory tumors to checkpoint blockade therapy, as demonstrated by a chimera consisting of the nucleolin-targeting aptamer AS1411 conjugated to a siRNA against SMG1. This construct enhanced tumor antigenicity and promoted responsiveness to immunotherapy without inducing immune activation in healthy tissues [218].
Furthermore, shRNAs can offer prolonged gene silencing by mimicking endogenous precursor miRNAs and being transcribed from DNA templates [107,196]. Aptamer-shRNA chimeras can be synthesized as single molecules using PCR-based or co-transcriptional approaches, improving delivery efficiency and simplifying production workflows [219,220]. Therapeutic efficacy has been demonstrated in models targeting RORGT in Th17 cells, where a CD30-targeting aptamer-shRNA chimera effectively silenced RORGT and reduced inflammatory cytokine expression [219], as well as in prostate cancer, where a PSMA-targeting aptamer-shRNA against DNA-PK enhanced tumor radiosensitivity and extended survival in preclinical models [196]. These findings highlight the promising potential of aptamer-shRNA constructs in autoimmune and cancer therapies.
Aptamer-anti-miRNA chimeras are constructs that enable cell-specific inhibition of oncogenic miRNAs, offering a targeted approach to modulate dysregulated miRNA pathways in disease. A notable example is GL21.T aptamer, which targets the AXL receptor tyrosine kinase and is conjugated to an anti-miR-222. This chimera (GL21.T–anti-miR-222) demonstrated therapeutic efficacy in glioblastoma models by reducing miR-222 levels, suppressing AXL activity, and significantly inhibiting tumor growth following intravenous administration [213]. Both linear and bivalent designs have been investigated to improve silencing efficiency and enable multitargeting capabilities, such as incorporating multiple anti-miRNAs within a single construct [196,213]. These strategies enhance therapeutic flexibility and broaden the potential application of aptamer–anti-miRNAs chimeras across different miRNA-driven cancer types.
Finally, aptamer-ASO chimeras provide a targeted strategy to inhibit specific transcripts with enhanced specificity and stability. For example, a PD-L1-targeting aptamer-ASO chimera effectively suppressed PD-L1 expression in a murine colorectal cancer model, reducing tumor burden to levels comparable to anti-PD-L1 antibody therapy. Importantly, this construct also decreased levels of soluble PD-L1, overcoming key mechanisms of immune evasion [212]. Additionally, circularized aptamer-ASO constructs, generated by ligating the 5′ end of the aptamer to the 3′ end of the ASO, demonstrated improved serum stability and durable silencing activity. These circular chimeras exhibited potent inhibition of SARS-CoV-2 N gene transcripts in vitro, highlighting their therapeutic potential in infectious disease contexts beyond oncology [221].

3. The Significance of RNA-Targeting Techniques in Cancer Research and Therapy

RNA-targeting strategies have emerged as powerful and versatile tools in cancer research and therapy, offering novel insights into the molecular mechanisms underlying tumor development, progression, and drug resistance [222,223,224,225]. The recognition that RNA molecules play critical regulatory roles in a wide array of biological processes has markedly expanded their potential as therapeutic targets, while advances in RNA biology, high-throughput sequencing technologies, and transcriptomic profiling have further transformed our understanding of cancer pathogenesis and paved the way for innovative therapeutic interventions [77,226,227]. Unlike traditional approaches that primarily target proteins, RNA-targeting techniques enable the direct and selective modulation of gene expression at the post-transcriptional level, offering the ability to intervene more precisely in disease pathways.
One of the most promising contributions of RNA-targeting strategies in oncology is the ability to degrade or inhibit specific mRNAs that encode oncogenic proteins, thereby mitigating the disease phenotype. ASOs, CRISPR-Cas13 systems, and siRNAs have shown efficacy in silencing mRNAs involved in key cancer-related pathways, such as those regulating cell proliferation, apoptosis, and angiogenesis [223,228,229]. For example, RNA-based approaches that target mutant KRAS mRNA—commonly implicated in pancreatic, lung, and colorectal cancers—are being explored as potential therapeutics to inhibit one of the most elusive oncogenic drivers in solid tumors [228,230,231]. By reducing the production of pathogenic proteins directly at the RNA level, these therapies can offer enhanced specificity and reduced off-target toxicity compared to traditional small-molecule inhibitors that target proteins [232].
Beyond targeting mRNAs, RNA-targeting techniques have also advanced our ability to study and therapeutically modulate ncRNAs, which are increasingly recognized as key regulators of cancer biology [233,234,235,236]. For instance, miRNAs can act as oncogenic or tumor suppressor molecules by regulating gene expression post-transcriptionally. Dysregulation of miRNA expression profiles is a hallmark of many cancers, contributing to uncontrolled cell growth, metastasis, and resistance to therapy [237,238]. Therapeutic strategies involving miRNA inhibitors like AMOs or miRNA sponges are being developed to restore normal regulatory function and suppress tumor progression [239,240]. Similarly, lncRNAs and circRNAs, which control chromatin architecture, transcription, and signaling networks, are emerging as novel targets for RNA-based therapeutics in cancer [228,241].
Small molecules and aptamers are expanding the RNA-targeting toolkit, as they are increasingly being recognized as powerful RNA-targeting modalities in oncology. Small molecules offer unique advantages by modulating RNA structure and function, enabling the selective targeting of complex RNA motifs such as G4s, IRES elements, and splice sites involved in cancer pathogenesis [176,180,181,188]. Emerging platforms like RiboTACs extend this capability by coupling RNA binding to catalytic RNA degradation, providing enhanced therapeutic precision and bioavailability [193]. Aptamers, by contrast, act as highly specific nucleic acid ligands that can bind RNA targets or deliver RNA therapeutics directly to cancer cells via aptamer-oligonucleotide chimeras [196,197,212,215]. Their ability to block miRNA processing or serve as tumor-selective delivery vehicles significantly expands the utility of RNA-targeting approaches for both therapeutic and diagnostic applications [197,208].
Importantly, RNA-targeting approaches also offer promise in overcoming drug resistance—one major challenge in cancer therapy. By selectively silencing resistance-associated genes or modifying splicing patterns that give rise to resistant isoforms, RNA-based approaches can enhance the efficacy of existing treatments and prevent relapse [242,243]. Additionally, RNA-editing tools and ssASOs provide a means to correct aberrant splicing events commonly seen in cancer, such as those affecting tumor suppressor genes like TP53 or BRCA2 [242,244,245].
As cancer research continues to evolve, RNA-targeting techniques are emerging as a critical modality in the development of next-generation molecular therapies. By enabling precise modulation of both coding and non-coding RNAs involved in tumorigenesis, these approaches offer innovative ways to overcome the limitations of traditional treatments, facilitating the advancement of personalized therapeutic strategies, and potentially improving clinical outcomes in cancer patients.

4. Current Landscape of FDA-Approved RNA-Targeting Therapies: A Focus on Cancer

RNA-targeting therapies have emerged as transformative tools in modern medicine, particularly in the treatment of genetic disorders [246,247,248]. By manipulating RNA molecules directly, these therapies enable precise modulation of gene expression either by altering pre-mRNA splicing, degrading oncogenic mRNAs, or inhibiting the translation of disease-causing transcripts (Figure 2) [74,249,250]. ASOs and siRNA-based approaches have been introduced as leading RNA-targeting modalities, with several agents achieving FDA approval—primarily for non-cancerous conditions.
Notable examples include nusinersen (Spinraza), a ssASO approved in 2016 for the treatment of spinal muscular atrophy (SMA), which corrects SMN2 splicing to increase the production of the survival motor neuron (SMN) protein [251,252]. Similarly, eteplirsen, approved for Duchenne muscular dystrophy, induces exon skipping in the dystrophin transcript to partially restore protein function [253,254]. In siRNA-based therapeutics, patisiran (Onpattro) became the first FDA-approved siRNA drug in 2018, targeting TTR mRNA to treat hereditary transthyretin-mediated amyloidosis (hATTR) by reducing toxic protein accumulation [195,248]. Additionally, risdiplam (Evrysdi), approved in 2020 for SMA, became the first non-ribosomal small molecule that was shown to directly modulate RNA splicing [255]. It binds to the SMN2 pre-mRNA, enhancing exon 7 inclusion and restoring SMN protein levels [256]. Together, these agents highlight the therapeutic versatility of RNA-targeting approaches across a range of genetic disorders and have paved the way for their application in more complex diseases like cancer.
Despite these successes, the utilization of RNA-targeted therapies in oncology has faced significant challenges. Unlike the described disorders, which typically involve a single, well-defined genetic defect, cancer is a multifaceted and highly heterogeneous disease, often driven by a complex interplay of multiple genetic and epigenetic factors [257,258]. A major challenge in RNA-targeted cancer therapy is the identification of RNA targets that are indispensable for tumor cell survival while exhibiting high specificity to minimize off-target effects on normal cells. Furthermore, the inherent genetic heterogeneity and rapid evolution of cancer cells frequently diminish the durability of treatments aimed at a single target, necessitating strategies that can address tumor adaptability and resistance.
Until recently, no RNA-targeting therapy had been approved for cancer, despite extensive preclinical efforts involving ASOs, siRNAs, CRISPR-Cas13, aptamers, and RNA-modulating small molecules. However, in mid-2024, imetelstat became the first FDA-approved RNA-targeting drug for cancer indication (Figure 1 and Table 3). Imetelstat is an ASO-based telomerase inhibitor that binds to the RNA component of the enzyme (TERC) and was approved for the treatment of lower-risk myelodysplastic syndromes [259,260]. This milestone not only validates the clinical potential of RNA-targeting strategies in oncology but also underscores the feasibility of targeting ncRNAs involved in essential cancer pathways.
Although imetelstat currently stands as the only FDA-approved RNA-targeting therapy for cancer, numerous other RNA-targeting agents are undergoing clinical evaluation. These ongoing efforts highlight the therapeutic potential of these approaches in cancer treatment and disease management, while expanding the scope of precision medicine by enabling the modulation of previously inaccessible targets.

5. Targeting RNA in Cancer: Insights from Current Clinical Trials

The application of RNA-targeting technologies in oncology is advancing rapidly, driven by their ability to modulate gene expression at the transcriptomic level with high specificity and functional precision. Various therapeutic platforms—including ASOs and siRNAs—are currently being evaluated in clinical trials for diverse cancer types [261,262]. These approaches are designed to selectively degrade oncogenic transcripts, modulate alternative splicing events, or disrupt pathological RNA–protein interactions (Figure 2), thereby targeting key mechanisms underlying cancer cell survival, proliferation, and metastasis.

5.1. Advancing ASOs in Human Clinical Trials

Although ASO-based therapies have not yet achieved broad regulatory approval in oncology, several candidates have advanced to mid- and late-stage clinical trials, demonstrating growing interest in their therapeutic potential. One prominent example is danvatirsen (AZD9150), a next-generation ASO that targets STAT3 mRNA, a transcription factor implicated in tumor growth and immune evasion [263]. Danvatirsen is currently being investigated in an ongoing Phase II clinical trial (NCT02983578) in combination with the PDL1 inhibitor durvalumab, in patients with non-small cell lung cancer (NSCLC), advanced pancreatic cancer, and mismatch repair-deficient (dMMR) colorectal cancer (Table 3). While Phase I trials have indicated favorable safety and preliminary antitumor activity, definitive efficacy data are pending, and the therapy remains in the investigational stage [264]. Another candidate, BP1002, is an ASO targeting BCL2 mRNA, a key anti-apoptotic factor frequently overexpressed in hematologic malignancies. BP1002 is currently under evaluation in a Phase I clinical trial (NCT04072458) for patients with relapsed or refractory lymphoid malignancies [265]. Additional ASO candidates, including OT-101 (targeting TGFB2 mRNA), WGI-0301 (targeting AKT1 mRNA), and BP1001 (GRB2), are currently undergoing early-phase clinical evaluation in solid tumors [266,267,268]. These efforts highlight the versatility of ASOs in targeting coding RNAs implicated in oncogenesis, even as challenges related to delivery, specificity, and durability of response remain under investigation.

5.2. siRNA-Based Therapies in Clinical Trials

Although no siRNA therapy has received FDA approval for cancer treatment thus far, multiple siRNA candidates targeting oncogenic transcripts have advanced into clinical trials, underscoring the therapeutic potential of RNAi in oncology. A promising agent is EPHA2-targeting siRNA, developed for treating patients with advanced solid tumors that often exhibit resistance to standard therapies. This siRNA functions by specifically silencing EPHA2 mRNA, a receptor tyrosine kinase overexpressed in many cancers and associated with tumor growth, angiogenesis, and metastasis [269]. This agent is currently being evaluated in a Phase I clinical trial to assess its safety profile and determine the optimal dosing regimen [270]. A novel siRNA therapeutic targeting NUDT21 (NCT06424301) has been developed for the treatment of retinoblastoma (RB) [271]. While current approaches such as intra-arterial and intravitreal chemotherapy have improved outcomes, they remain limited by toxicity and drug resistance [272]. This siRNA selectively silences NUDT21, a regulator of 3′UTR processing that promotes tumor proliferation through SMC1A stabilization. Preclinical studies have shown that NUDT21 inhibition suppresses tumor growth with minimal off-target toxicity. Delivered via intraocular injection, this agent represents a first-in-class approach for RB and is now being evaluated in an early-phase clinical trial to assess safety and preliminary efficacy in refractory cases [271]. Additionally, STP705, a first-in-class siRNA therapeutic simultaneously targeting TGFB1 and COX2 mRNAs, has shown remarkable efficacy in a Phase II trial (NCT04669808) for cutaneous basal cell carcinoma, achieving a 100% complete response rate with an excellent safety profile (Table 3) [273]. This dual-targeting approach inhibits two critical drivers of tumor progression and inflammation, offering a novel therapeutic avenue in oncology.

5.3. CRISPR-Cas13-Based Approaches: Still in Preclinical Stage for Cancer

Although CRISPR-Cas13 systems represent a relatively recent advance in RNA-targeting technology—first characterized less than a decade ago—they have already entered clinical trials for non-cancer indications, highlighting their exceptional translational potential. While no Cas13-based therapies for cancer treatment have yet been tested in humans, early-phase clinical programs in non-oncologic diseases have demonstrated both feasibility and safety. Notably, HuidaGene Therapeutics has launched two first-in-human Cas13 programs: the HERO trial, evaluating HG204 for MECP2 Duplication Syndrome (MDS), and the SIGHT-I and BRIGHT trials, testing HG202 for neovascular age-related macular degeneration (nAMD). These therapies harness Cas13′s ability to selectively degrade pathogenic RNAs—MECP2 mRNA in MDS and VEGFA mRNA in nAMD—offering a programmable, transcript-specific approach for treating these diseases [274,275]. Importantly, HG202 received FDA clearance in 2024, becoming the first Cas13-based RNA-editing therapy authorized for human testing in the USA [275].
Although still in preclinical stages for oncology, several studies have already demonstrated the potential of Cas13 systems in cancer models. For instance, in glioma models, lentiviral overexpression of LwCas13a induced a collateral RNA cleavage effect upon targeting EGFP and the tumor-specific EGFRvIII mutant, resulting in selective tumor cell death and inhibition of intracranial tumor growth. Bulk and single-cell RNA sequencing confirmed this effect, highlighting Cas13a’s tumor-eliminating capability [276]. In bladder cancer, a targeted liposome delivery system transported Cas13a gene circuits controlled by tumor-specific promoters and activated by near-infrared light. This system enabled multiplexed crRNA targeting and receptor-specific delivery without genomic disruption, offering a versatile gene therapy approach [277]. Furthermore, to overcome challenges in spatial and temporal regulation of Cas13a activity, innovative delivery systems such as hierarchical RNA nanococoons (RNCOs-D) have been developed. These nanococoons encapsulate Cas13a ribonucleoproteins along with chemotherapeutic agents and remain inactive until exposed to the acidic tumor microenvironment, which triggers their disassembly and reactivates Cas13a. This activation is further refined by the specific cleavage of EGFRvIII mRNA, enabling targeted silencing of this oncogenic transcript. highlighting a promising strategy for precise, controlled, and multimodal cancer therapy using CRISPR-Cas13a [278].
These milestones strongly support the clinical viability of Cas13 technologies and pave the way for future cancer applications. With ongoing progress in delivery systems and tumor-selective targeting, CRISPR-Cas13 holds a significant promise to overcome current therapeutic limitations in cancer, offering precise, reversible control over oncogenic transcripts.

5.4. Small Molecules and Aptamers Targeting RNA: Current Landscape and Prospects

Despite growing interest in RNA as a therapeutic target, there are currently no small-molecule drugs in clinical trials for cancer that directly bind and target RNA molecules. While several small molecules have been investigated for their ability to modulate RNA-related processes—such as H3B-8800 and E7107—these agents act by targeting proteins involved in RNA processing, rather than RNA itself [279,280,281]. Similarly, PRMT5 inhibitors, such as JNJ-64619178 currently in Phase I clinical trials for cancer (NCT03573310), modulate RNA splicing by altering the methylation status of splicing-related proteins rather than through direct RNA binding [282]. Although these compounds influence RNA biology and transcriptome dynamics, their mechanisms are indirect, relying on disruption of RNA-associated proteins rather than structural or sequence-specific interaction with RNA. Similarly, while several aptamer therapeutics have entered clinical trials for cancer, none function by binding RNA molecules directly to exert their anticancer effects. A notable agent, NOX-A12 (olaptesed pegol), is a Spiegelmer L-RNA aptamer that binds the chemokine CXCL12 [283], modulating the tumor microenvironment. Currently undergoing Phase I/II trials in glioblastoma (NCT04121455) [284], NOX-A12 highlights the clinical potential of aptamers in cancer therapeutics. To date, no aptamer in active clinical evaluation in oncology binds directly to RNA targets such as oncogenic mRNAs or structured ncRNAs, since the study of these molecules remain restricted to preclinical research.
Nonetheless, recent advances in RNA structural biology, high-throughput screening, and computational drug design are now enabling the development of small molecules and aptamers that directly bind RNA with increasing specificity and affinity. These efforts are leading to the identification of preclinical candidates that bind defined RNA motifs, offering a promising new direction for transcriptome-targeted cancer therapeutics.
Table 3. Clinical status of RNA-targeting therapies for cancer. This table summarizes key RNA-targeting therapeutic agents and their current status in clinical settings. Non-cancer indications are included for comparison, emphasizing the translational potential of RNA-targeting modalities.
Table 3. Clinical status of RNA-targeting therapies for cancer. This table summarizes key RNA-targeting therapeutic agents and their current status in clinical settings. Non-cancer indications are included for comparison, emphasizing the translational potential of RNA-targeting modalities.
Therapeutic AgentTargetRNA-Targeting StrategyClinical StatusFDA Approval & Indication
Imetelstat (Rytelo)TERCASOFDA-approved (2024)Approved for adult patients with low- to intermediate-1 risk myelodysplastic syndromes (MDS) [259,260].
danvatirsen (AZD9150)STAT3 mRNAASOPhase II clinical trialInvestigated in combination with durvalumab for NSCLC, advanced pancreatic cancer, and dMMR colorectal cancer; safety promising, efficacy pending [264].
BP1002BCL2 mRNAASOPhase I clinical trialEvaluated in patients with relapsed or refractory lymphoid malignancies [265].
OT-101TGFB2 mRNAASOEarly-phase clinical trialsUnder evaluation in various solid tumor types [268].
WGI-0301AKT1 mRNAASOEarly-phase clinical trialsUnder evaluation [267].
BP1001GRB2 mRNAASOEarly-phase clinical trialsUnder evaluation [266].
EPHA2-targeting siRNAEPHA2 mRNARNAi/siRNAPhase I clinical trialTargeting EPHA2 mRNA in advanced solid tumors; safety profile and dosing under assessment [270].
NUDT21-targeting siRNANUDT21 mRNARNAi/siRNAEarly-phase clinical trialFirst-in-class siRNA for retinoblastoma; delivered intraocularly; safety and preliminary efficacy under evaluation [271].
STP705 (dual-target siRNA)TGFΒ1 & COX2 mRNAsRNAi/siRNAPhase II clinical trialAchieved 100% complete response in cutaneous basal cell carcinoma with excellent safety profile [273].
HG202VEGFA mRNA
(non-cancer)
CRISPR-Cas13Early human clinical trialsFDA-cleared for clinical trials in neovascular age-related macular degeneration (nAMD); first Cas13-based RNA-editing therapy in human testing [275].

6. RNA-Based Therapeutics in Cancer: Current Advances and Future Perspectives

RNA-targeting strategies are at the forefront of molecular oncology, marking a significant advancement in how drugs are developed to treat cancer. While no siRNA or Cas13-based therapies have yet received FDA approval for cancer indications, the growing number of clinical trials highlights increasing confidence in RNA molecules as effective therapeutic targets. A major milestone was achieved in 2024 with the approval of imetelstat, marking the first FDA-approved ASO therapy for cancer and demonstrating regulatory acceptance of RNA-based approaches in cancer therapy [260]. Numerous other ASO and siRNA candidates are currently in Phase I/II trials, addressing hematologic and solid malignancies with encouraging preliminary results.
At the same time, the RNA-targeting field is moving beyond traditional approaches. CRISPR-Cas13 systems, while still in preclinical development for cancer [112,116,145,146], have reached clinical testing in non-cancer diseases, laying the groundwork for future oncology applications [285]. Small molecules that directly bind RNA, although not yet entered clinical trials for cancer, are increasingly identified from high-throughput screens and structure-guided design efforts. These efforts focus on targeting well-defined structural motifs within non-coding RNAs and oncogenic mRNA that play critical roles in cancer. Additionally, advances in delivery systems are helping to overcome critical challenges in RNA stability, biodistribution, and intracellular uptake [286].
Looking forward, future research will likely focus on combining therapies that use RNA-targeting agents alongside immunotherapies, kinase inhibitors, or chemotherapies to overcome treatment resistance and enhance the longevity of patient responses. Personalized RNA-targeting approaches, guided by detailed analysis of tumor transcriptomes, alternative splicing patterns, and RNA modifications, are expected to play an increasingly important role in cancer treatment. As this field advances, RNA-targeting therapies have the potential to transform oncology by enabling precise, versatile, and tumor-specific molecular interventions.

7. Conclusions

RNA-targeting strategies are rapidly emerging as powerful tools in cancer research and therapy, enabling precise modulation of gene expression at the RNA level. Unlike traditional treatments that act on proteins, these approaches can selectively silence oncogenic transcripts or correct aberrant splicing, offering high specificity with reduced toxicity. The recent approval of imetelstat, which targets the RNA component of telomerase, marks a major milestone and validates the clinical potential of RNA-targeting strategies in oncology. Advances in RNA biology, high-throughput sequencing techniques, and transcriptome profiling are ushering in a new era of innovation in the field. Future efforts will likely focus on combining RNA-targeted therapies with existing cancer treatments and developing personalized approaches based on tumor-specific RNA signatures. Overall, RNA-based therapeutics represent a versatile and promising approach to overcoming the complexity and resistance mechanisms inherent in cancer.
This review offers novel contributions compared to existing literature:
  • Comprehensive comparative analysis of key RNA-targeting platforms, including antisense oligonucleotides, RNA interference, CRISPR-Cas13 systems, and RNA-binding small molecules, is performed, emphasizing their distinct mechanisms and applications in oncology.
  • Discussion of recent clinical and preclinical advancements is presented with an emphasis on translational relevance, highlighting how RNA-targeting approaches are progressing toward real-world cancer treatments.
  • A forward-looking perspective that outlines future research directions, including personalized RNA therapeutics based on tumor-specific transcriptomic signatures, is presented.

Author Contributions

Writing—original draft preparation, M.A.B.; writing—reviewing and editing, M.A.D. and M.A.B.; visualization, M.A.D.; supervision, A.S.; project administration, A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Beckert, B.; Masquida, B. Synthesis of RNA by in vitro transcription. Methods Mol. Biol. 2011, 703, 29–41. [Google Scholar] [CrossRef] [PubMed]
  2. Van Oijen, A.M.; Loparo, J.J. Single-Molecule Studies of the Replisome. Annu. Rev. Biophys. 2010, 39, 429–448. [Google Scholar] [CrossRef] [PubMed]
  3. Libicher, K.; Hornberger, R.; Heymann, M.; Mutschler, H. In vitro self-replication and multicistronic expression of large synthetic genomes. Nat. Commun. 2020, 11, 904. [Google Scholar] [CrossRef]
  4. Jacob, F.; Monod, J. Genetic regulatory mechanisms in the synthesis of proteins. J. Mol. Biol. 1961, 3, 318–356. [Google Scholar] [CrossRef] [PubMed]
  5. Crick, F. Central Dogma of Molecular Biology. Nature 1970, 227, 561–563. [Google Scholar] [CrossRef]
  6. Jia, X.; He, X.; Huang, C.; Li, J.; Dong, Z.; Liu, K. Protein translation: Biological processes and therapeutic strategies for human diseases. Signal Transduct. Target. Ther. 2024, 9, 44. [Google Scholar] [CrossRef]
  7. Okazaki, Y.; Furuno, M.; Kasukawa, T.; Adachi, J.; Bono, H.; Kondo, S.; Nikaido, I.; Osato, N.; Saito, R.; Suzuki, H.; et al. Analysis of the mouse transcriptome based on functional annotation of 60,770 full-length cDNAs. Nature 2002, 420, 563–573. [Google Scholar] [CrossRef]
  8. Athanasopoulou, K.; Boti, M.A.; Adamopoulos, P.G.; Skourou, P.C.; Scorilas, A. Third-Generation Sequencing: The Spearhead towards the Radical Transformation of Modern Genomics. Life 2021, 12, 30. [Google Scholar] [CrossRef]
  9. Levin, J.Z.; Berger, M.F.; Adiconis, X.; Rogov, P.; Melnikov, A.; Fennell, T.; Nusbaum, C.; Garraway, L.A.; Gnirke, A. Targeted next-generation sequencing of a cancer transcriptome enhances detection of sequence variants and novel fusion transcripts. Genome Biol. 2009, 10, R115. [Google Scholar] [CrossRef]
  10. Hrdlickova, R.; Toloue, M.; Tian, B. RNA-Seq methods for transcriptome analysis. WIREs RNA 2017, 8, e1364. [Google Scholar] [CrossRef]
  11. Yu, F.; Qi, H.; Gao, L.; Luo, S.; Njeri Damaris, R.; Ke, Y.; Wu, W.; Yang, P. Identifying RNA Modifications by Direct RNA Sequencing Reveals Complexity of Epitranscriptomic Dynamics in Rice. Genom. Proteom. Bioinform. 2023, 21, 788–804. [Google Scholar] [CrossRef]
  12. Xie, Y.-Y.; Zhong, Z.-D.; Chen, H.-X.; Ren, Z.-H.; Qiu, Y.-T.; Lan, Y.-L.; Wu, F.; Kong, J.-W.; Luo, R.-J.; Zhang, D.; et al. Single-molecule direct RNA sequencing reveals the shaping of epitranscriptome across multiple species. Nat. Commun. 2025, 16, 5119. [Google Scholar] [CrossRef]
  13. Micheel, J.; Safrastyan, A.; Wollny, D. Advances in Non-Coding RNA Sequencing. Non-Coding RNA 2021, 7, 70. [Google Scholar] [CrossRef]
  14. Liu, S.; Huang, J.; Zhou, J.; Chen, S.; Zheng, W.; Liu, C.; Lin, Q.; Zhang, P.; Wu, D.; He, S.; et al. NAP-seq reveals multiple classes of structured noncoding RNAs with regulatory functions. Nat. Commun. 2024, 15, 2425. [Google Scholar] [CrossRef]
  15. Wang, S.; Sun, S.T.; Zhang, X.Y.; Ding, H.R.; Yuan, Y.; He, J.J.; Wang, M.S.; Yang, B.; Li, Y.B. The Evolution of Single-Cell RNA Sequencing Technology and Application: Progress and Perspectives. Int. J. Mol. Sci. 2023, 24, 2943. [Google Scholar] [CrossRef]
  16. Cheng, C.; Chen, W.; Jin, H.; Chen, X. A Review of Single-Cell RNA-Seq Annotation, Integration, and Cell-Cell Communication. Cells 2023, 12, 1970. [Google Scholar] [CrossRef]
  17. Chen, L.L.; Kim, V.N. Small and long non-coding RNAs: Past, present, and future. Cell 2024, 187, 6451–6485. [Google Scholar] [CrossRef]
  18. Iwakawa, H.-O.; Tomari, Y. Life of RISC: Formation, action, and degradation of RNA-induced silencing complex. Mol. Cell 2022, 82, 30–43. [Google Scholar] [CrossRef]
  19. Bartel, D.P. Metazoan MicroRNAs. Cell 2018, 173, 20–51. [Google Scholar] [CrossRef]
  20. Yao, R.-W.; Wang, Y.; Chen, L.-L. Cellular functions of long noncoding RNAs. Nat. Cell Biol. 2019, 21, 542–551. [Google Scholar] [CrossRef]
  21. Statello, L.; Guo, C.-J.; Chen, L.-L.; Huarte, M. Gene regulation by long non-coding RNAs and its biological functions. Nat. Rev. Mol. Cell Biol. 2021, 22, 96–118. [Google Scholar] [CrossRef]
  22. Nadhan, R.; Isidoro, C.; Song, Y.S.; Dhanasekaran, D.N. Signaling by LncRNAs: Structure, Cellular Homeostasis, and Disease Pathology. Cells 2022, 11, 2517. [Google Scholar] [CrossRef]
  23. Lafontaine, D.L. Noncoding RNAs in eukaryotic ribosome biogenesis and function. Nat. Struct. Mol. Biol. 2015, 22, 11–19. [Google Scholar] [CrossRef] [PubMed]
  24. Kaliatsi, E.G.; Giarimoglou, N.; Stathopoulos, C.; Stamatopoulou, V. Non-Coding RNA-Driven Regulation of rRNA Biogenesis. Int. J. Mol. Sci. 2020, 21, 9738. [Google Scholar] [CrossRef] [PubMed]
  25. Gongadze, G.M. 5S rRNA and ribosome. Biochemistry 2011, 76, 1450–1464. [Google Scholar] [CrossRef]
  26. Kondo, Y.; Oubridge, C.; van Roon, A.M.; Nagai, K. Crystal structure of human U1 snRNP, a small nuclear ribonucleoprotein particle, reveals the mechanism of 5′ splice site recognition. eLife 2015, 4, e04986. [Google Scholar] [CrossRef] [PubMed]
  27. Alonso, D.; Mondragón, A. Mechanisms of catalytic RNA molecules. Biochem. Soc. Trans. 2021, 49, 1529–1535. [Google Scholar] [CrossRef]
  28. Westhof, E. Ribozymes, catalytically active RNA molecules. Introd. Methods Mol. Biol. 2012, 848, 1–4. [Google Scholar] [CrossRef]
  29. Cesana, M.; Cacchiarelli, D.; Legnini, I.; Santini, T.; Sthandier, O.; Chinappi, M.; Tramontano, A.; Bozzoni, I. A Long Noncoding RNA Controls Muscle Differentiation by Functioning as a Competing Endogenous RNA. Cell 2011, 147, 358–369. [Google Scholar] [CrossRef]
  30. Ounzain, S.; Micheletti, R.; Arnan, C.; Plaisance, I.; Cecchi, D.; Schroen, B.; Reverter, F.; Alexanian, M.; Gonzales, C.; Ng, S.Y.; et al. CARMEN, a human super enhancer-associated long noncoding RNA controlling cardiac specification, differentiation and homeostasis. J. Mol. Cell. Cardiol. 2015, 89, 98–112. [Google Scholar] [CrossRef]
  31. Jie, M.; Feng, T.; Huang, W.; Zhang, M.; Feng, Y.; Jiang, H.; Wen, Z. Subcellular Localization of miRNAs and Implications in Cellular Homeostasis. Genes 2021, 12, 856. [Google Scholar] [CrossRef]
  32. Müller, S.; Raulefs, S.; Bruns, P.; Afonso-Grunz, F.; Plötner, A.; Thermann, R.; Jäger, C.; Schlitter, A.M.; Kong, B.; Regel, I.; et al. Next-generation sequencing reveals novel differentially regulated mRNAs, lncRNAs, miRNAs, sdRNAs and a piRNA in pancreatic cancer. Mol. Cancer 2015, 14, 94. [Google Scholar] [CrossRef]
  33. Law, P.T.Y.; Qin, H.; Ching, A.K.K.; Lai, K.P.; Co, N.N.; He, M.; Lung, R.W.M.; Chan, A.W.H.; Chan, T.-F.; Wong, N. Deep sequencing of small RNA transcriptome reveals novel non-coding RNAs in hepatocellular carcinoma. J. Hepatol. 2013, 58, 1165–1173. [Google Scholar] [CrossRef] [PubMed]
  34. Li, M.; Wang, J.; Liu, D.; Huang, H. High-throughput sequencing reveals differentially expressed lncRNAs and circRNAs, and their associated functional network, in human hypertrophic scars. Mol. Med. Rep. 2018, 18, 5669–5682. [Google Scholar] [CrossRef]
  35. Chuang, T.-D.; Xie, Y.; Yan, W.; Khorram, O. Next-generation sequencing reveals differentially expressed small noncoding RNAs in uterine leiomyoma. Fertil. Steril. 2018, 109, 919–929. [Google Scholar] [CrossRef]
  36. Yan, H.; Bu, P. Non-coding RNA in cancer. Essays Biochem. 2021, 65, 625–639. [Google Scholar] [CrossRef] [PubMed]
  37. Ghafouri-Fard, S.; Hussen, B.M.; Taheri, M.; Ayatollahi, S.A. Emerging role of circular RNAs in breast cancer. Pathol. Res. Pract. 2021, 223, 153496. [Google Scholar] [CrossRef] [PubMed]
  38. O’Connell, M.R. Molecular Mechanisms of RNA Targeting by Cas13-containing Type VI CRISPR-Cas Systems. J. Mol. Biol. 2019, 431, 66–87. [Google Scholar] [CrossRef]
  39. Terns, M.P. CRISPR-Based Technologies: Impact of RNA-Targeting Systems. Mol. Cell 2018, 72, 404–412. [Google Scholar] [CrossRef]
  40. Yang, H.; Patel, D.J. Structures, mechanisms and applications of RNA-centric CRISPR-Cas13. Nat. Chem. Biol. 2024, 20, 673–688. [Google Scholar] [CrossRef]
  41. Burmistrz, M.; Krakowski, K.; Krawczyk-Balska, A. RNA-Targeting CRISPR-Cas Systems and Their Applications. Int. J. Mol. Sci. 2020, 21, 1122. [Google Scholar] [CrossRef]
  42. Tong, H.; Huang, J.; Xiao, Q.; He, B.; Dong, X.; Liu, Y.; Yang, X.; Han, D.; Wang, Z.; Wang, X.; et al. High-fidelity Cas13 variants for targeted RNA degradation with minimal collateral effects. Nat. Biotechnol. 2023, 41, 108–119. [Google Scholar] [CrossRef]
  43. Bradley, R.K.; Anczuków, O. RNA splicing dysregulation and the hallmarks of cancer. Nat. Rev. Cancer 2023, 23, 135–155. [Google Scholar] [CrossRef]
  44. Goodall, G.J.; Wickramasinghe, V.O. RNA in cancer. Nat. Rev. Cancer 2021, 21, 22–36. [Google Scholar] [CrossRef]
  45. Taniguchi, H.; Suzuki, Y.; Imai, K.; Adachi, Y. Antitumoral RNA-targeted oligonucleotide therapeutics: The third pillar after small molecule inhibitors and antibodies. Cancer Sci. 2022, 113, 2952–2961. [Google Scholar] [CrossRef] [PubMed]
  46. Çakan, E.; Lara, O.D.; Szymanowska, A.; Bayraktar, E.; Chavez-Reyes, A.; Lopez-Berestein, G.; Amero, P.; Rodriguez-Aguayo, C. Therapeutic Antisense Oligonucleotides in Oncology: From Bench to Bedside. Cancers 2024, 16, 2940. [Google Scholar] [CrossRef]
  47. Watts, J.K.; Corey, D.R. Silencing disease genes in the laboratory and the clinic. J. Pathol. 2012, 226, 365–379. [Google Scholar] [CrossRef] [PubMed]
  48. Esau, C.C. Inhibition of microRNA with antisense oligonucleotides. Methods 2008, 44, 55–60. [Google Scholar] [CrossRef] [PubMed]
  49. Aigner, A. Gene silencing through RNA interference (RNAi) in vivo: Strategies based on the direct application of siRNAs. J. Biotechnol. 2006, 124, 12–25. [Google Scholar] [CrossRef]
  50. Kordyś, M.; Sen, R.; Warkocki, Z. Applications of the versatile CRISPR-Cas13 RNA targeting system. Wiley Interdiscip. Rev. RNA 2022, 13, e1694. [Google Scholar] [CrossRef]
  51. Childs-Disney, J.L.; Yang, X.; Gibaut, Q.M.R.; Tong, Y.; Batey, R.T.; Disney, M.D. Targeting RNA structures with small molecules. Nat. Rev. Drug Discov. 2022, 21, 736–762. [Google Scholar] [CrossRef]
  52. Paul, A.; Muralidharan, A.; Biswas, A.; Kamath, B.V.; Joseph, A.; Alex, A.T. siRNA therapeutics and its challenges: Recent advances in effective delivery for cancer therapy. OpenNano 2022, 7, 100063. [Google Scholar] [CrossRef]
  53. Winkle, M.; El-Daly, S.M.; Fabbri, M.; Calin, G.A. Noncoding RNA therapeutics—Challenges and potential solutions. Nat. Rev. Drug Discov. 2021, 20, 629–651. [Google Scholar] [CrossRef]
  54. Moreno, P.M.; Pêgo, A.P. Therapeutic antisense oligonucleotides against cancer: Hurdling to the clinic. Front. Chem. 2014, 2, 87. [Google Scholar] [CrossRef] [PubMed]
  55. Mikutis, S.; Bernardes, G.J.L. Technologies for Targeted RNA Degradation and Induced RNA Decay. Chem. Rev. 2024, 124, 13301–13330. [Google Scholar] [CrossRef]
  56. Zamecnik, P.C.; Stephenson, M.L. Inhibition of Rous sarcoma virus replication and cell transformation by a specific oligodeoxynucleotide. Proc. Natl. Acad. Sci. USA 1978, 75, 280–284. [Google Scholar] [CrossRef]
  57. Bennett, C.F.; Swayze, E.E. RNA targeting therapeutics: Molecular mechanisms of antisense oligonucleotides as a therapeutic platform. Annu. Rev. Pharmacol. Toxicol. 2010, 50, 259–293. [Google Scholar] [CrossRef] [PubMed]
  58. Eckstein, F. Phosphorothioates, essential components of therapeutic oligonucleotides. Nucleic Acid Ther. 2014, 24, 374–387. [Google Scholar] [CrossRef] [PubMed]
  59. Yu, R.Z.; Grundy, J.S.; Geary, R.S. Clinical pharmacokinetics of second generation antisense oligonucleotides. Expert Opin. Drug Metab. Toxicol. 2013, 9, 169–182. [Google Scholar] [CrossRef]
  60. Obika, S.; Nanbu, D.; Hari, Y.; Andoh, J.-I.; Morio, K.-I.; Doi, T.; Imanishi, T. Stability and structural features of the duplexes containing nucleoside analogues with a fixed N-type conformation, 2′-O,4′-C-methyleneribonucleosides. Tetrahedron Lett. 1998, 39, 5401–5404. [Google Scholar] [CrossRef]
  61. Renneberg, D.; Leumann, C.J. Watson-Crick base-pairing properties of tricyclo-DNA. J. Am. Chem. Soc. 2002, 124, 5993–6002. [Google Scholar] [CrossRef]
  62. Summerton, J.; Weller, D. Morpholino antisense oligomers: Design, preparation, and properties. Antisense Nucleic Acid Drug Dev. 1997, 7, 187–195. [Google Scholar] [CrossRef]
  63. Hanvey, J.C.; Peffer, N.J.; Bisi, J.E.; Thomson, S.A.; Cadilla, R.; Josey, J.A.; Ricca, D.J.; Hassman, C.F.; Bonham, M.A.; Au, K.G.; et al. Antisense and antigene properties of peptide nucleic acids. Science 1992, 258, 1481–1485. [Google Scholar] [CrossRef]
  64. Smulevitch, S.V.; Simmons, C.G.; Norton, J.C.; Wise, T.W.; Corey, D.R. Enhancement of strand invasion by oligonucleotides through manipulation of backbone charge. Nat. Biotechnol. 1996, 14, 1700–1704. [Google Scholar] [CrossRef]
  65. Dhuri, K.; Bechtold, C.; Quijano, E.; Pham, H.; Gupta, A.; Vikram, A.; Bahal, R. Antisense Oligonucleotides: An Emerging Area in Drug Discovery and Development. J. Clin. Med. 2020, 9, 2004. [Google Scholar] [CrossRef] [PubMed]
  66. Stein, H.; Hausen, P. Enzyme from calf thymus degrading the RNA moiety of DNA-RNA Hybrids: Effect on DNA-dependent RNA polymerase. Science 1969, 166, 393–395. [Google Scholar] [CrossRef] [PubMed]
  67. Wu, H.; Lima, W.F.; Zhang, H.; Fan, A.; Sun, H.; Crooke, S.T. Determination of the role of the human RNase H1 in the pharmacology of DNA-like antisense drugs. J. Biol. Chem. 2004, 279, 17181–17189. [Google Scholar] [CrossRef] [PubMed]
  68. Wahlestedt, C.; Salmi, P.; Good, L.; Kela, J.; Johnsson, T.; Hökfelt, T.; Broberger, C.; Porreca, F.; Lai, J.; Ren, K.; et al. Potent and nontoxic antisense oligonucleotides containing locked nucleic acids. Proc. Natl. Acad. Sci. USA 2000, 97, 5633–5638. [Google Scholar] [CrossRef]
  69. Monia, B.P.; Lesnik, E.A.; Gonzalez, C.; Lima, W.F.; McGee, D.; Guinosso, C.J.; Kawasaki, A.M.; Cook, P.D.; Freier, S.M. Evaluation of 2′-modified oligonucleotides containing 2′-deoxy gaps as antisense inhibitors of gene expression. J. Biol. Chem. 1993, 268, 14514–14522. [Google Scholar]
  70. Lundin, K.E.; Gissberg, O.; Smith, C.I.E. Oligonucleotide Therapies: The Past and the Present. Hum. Gene Ther. 2015, 26, 475–485. [Google Scholar] [CrossRef]
  71. Scharner, J.; Aznarez, I. Clinical Applications of Single-Stranded Oligonucleotides: Current Landscape of Approved and In-Development Therapeutics. Mol. Ther. 2021, 29, 540–554. [Google Scholar] [CrossRef] [PubMed]
  72. Hammond, S.M.; Aartsma-Rus, A.; Alves, S.; Borgos, S.E.; Buijsen, R.A.M.; Collin, R.W.J.; Covello, G.; Denti, M.A.; Desviat, L.R.; Echevarría, L.; et al. Delivery of oligonucleotide-based therapeutics: Challenges and opportunities. EMBO Mol. Med. 2021, 13, e13243. [Google Scholar] [CrossRef]
  73. Kole, R.; Krainer, A.R.; Altman, S. RNA therapeutics: Beyond RNA interference and antisense oligonucleotides. Nat. Rev. Drug Discov. 2012, 11, 125–140. [Google Scholar] [CrossRef]
  74. Havens, M.A.; Hastings, M.L. Splice-switching antisense oligonucleotides as therapeutic drugs. Nucleic Acids Res. 2016, 44, 6549–6563. [Google Scholar] [CrossRef]
  75. Liang, X.H.; Sun, H.; Shen, W.; Wang, S.; Yao, J.; Migawa, M.T.; Bui, H.H.; Damle, S.S.; Riney, S.; Graham, M.J.; et al. Antisense oligonucleotides targeting translation inhibitory elements in 5′ UTRs can selectively increase protein levels. Nucleic Acids Res. 2017, 45, 9528–9546. [Google Scholar] [CrossRef]
  76. Sasaki, S.; Sun, R.; Bui, H.H.; Crosby, J.R.; Monia, B.P.; Guo, S. Steric Inhibition of 5′ UTR Regulatory Elements Results in Upregulation of Human CFTR. Mol. Ther. 2019, 27, 1749–1757. [Google Scholar] [CrossRef]
  77. Ma, W.K.; Voss, D.M.; Scharner, J.; Costa, A.S.H.; Lin, K.T.; Jeon, H.Y.; Wilkinson, J.E.; Jackson, M.; Rigo, F.; Bennett, C.F.; et al. ASO-Based PKM Splice-Switching Therapy Inhibits Hepatocellular Carcinoma Growth. Cancer Res. 2022, 82, 900–915. [Google Scholar] [CrossRef]
  78. Summerton, J. Morpholino antisense oligomers: The case for an RNase H-independent structural type. Biochim. Biophys. Acta 1999, 1489, 141–158. [Google Scholar] [CrossRef]
  79. Rigo, F.; Seth, P.P.; Bennett, C.F. Antisense oligonucleotide-based therapies for diseases caused by pre-mRNA processing defects. Adv. Exp. Med. Biol. 2014, 825, 303–352. [Google Scholar] [CrossRef] [PubMed]
  80. Diener, C.; Keller, A.; Meese, E. Emerging concepts of miRNA therapeutics: From cells to clinic. Trends Genet. 2022, 38, 613–626. [Google Scholar] [CrossRef] [PubMed]
  81. Saenz-Pipaon, G.; Dichek, D.A. Targeting and delivery of microRNA-targeting antisense oligonucleotides in cardiovascular diseases. Atherosclerosis 2023, 374, 44–54. [Google Scholar] [CrossRef]
  82. Davis, S.; Lollo, B.; Freier, S.; Esau, C. Improved targeting of miRNA with antisense oligonucleotides. Nucleic Acids Res. 2006, 34, 2294–2304. [Google Scholar] [CrossRef]
  83. Lima, J.F.; Cerqueira, L.; Figueiredo, C.; Oliveira, C.; Azevedo, N.F. Anti-miRNA oligonucleotides: A comprehensive guide for design. RNA Biol. 2018, 15, 338–352. [Google Scholar] [CrossRef]
  84. Otmani, K.; Rouas, R.; Berehab, M.; Lewalle, P. The regulatory mechanisms of oncomiRs in cancer. Biomed. Pharmacother. 2024, 171, 116165. [Google Scholar] [CrossRef]
  85. Ding, T.; Cui, P.; Zhou, Y.; Chen, C.; Zhao, J.; Wang, H.; Guo, M.; He, Z.; Xu, L. Antisense Oligonucleotides against miR-21 Inhibit the Growth and Metastasis of Colorectal Carcinoma via the DUSP8 Pathway. Mol. Ther. Nucleic Acids 2018, 13, 244–255. [Google Scholar] [CrossRef] [PubMed]
  86. Bautista-Sánchez, D.; Arriaga-Canon, C.; Pedroza-Torres, A.; De La Rosa-Velázquez, I.A.; González-Barrios, R.; Contreras-Espinosa, L.; Montiel-Manríquez, R.; Castro-Hernández, C.; Fragoso-Ontiveros, V.; Álvarez-Gómez, R.M.; et al. The Promising Role of miR-21 as a Cancer Biomarker and Its Importance in RNA-Based Therapeutics. Mol. Ther. Nucleic Acids 2020, 20, 409–420. [Google Scholar] [CrossRef]
  87. Cai, Z.K.; Chen, Q.; Chen, Y.B.; Gu, M.; Zheng, D.C.; Zhou, J.; Wang, Z. microRNA-155 promotes the proliferation of prostate cancer cells by targeting annexin 7. Mol. Med. Rep. 2015, 11, 533–538. [Google Scholar] [CrossRef]
  88. Ebert, M.S.; Neilson, J.R.; Sharp, P.A. MicroRNA sponges: Competitive inhibitors of small RNAs in mammalian cells. Nat. Methods 2007, 4, 721–726. [Google Scholar] [CrossRef]
  89. Tay, F.C.; Lim, J.K.; Zhu, H.; Hin, L.C.; Wang, S. Using artificial microRNA sponges to achieve microRNA loss-of-function in cancer cells. Adv. Drug Del. Rev. 2015, 81, 117–127. [Google Scholar] [CrossRef] [PubMed]
  90. Kluiver, J.; Slezak-Prochazka, I.; Smigielska-Czepiel, K.; Halsema, N.; Kroesen, B.-J.; van den Berg, A. Generation of miRNA sponge constructs. Methods 2012, 58, 113–117. [Google Scholar] [CrossRef] [PubMed]
  91. Fire, A.; Xu, S.; Montgomery, M.K.; Kostas, S.A.; Driver, S.E.; Mello, C.C. Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature 1998, 391, 806–811. [Google Scholar] [CrossRef]
  92. Hannon, G.J.; Rossi, J.J. Unlocking the potential of the human genome with RNA interference. Nature 2004, 431, 371–378. [Google Scholar] [CrossRef]
  93. Traber, G.M.; Yu, A.M. RNAi-Based Therapeutics and Novel RNA Bioengineering Technologies. J. Pharmacol. Exp. Ther. 2023, 384, 133–154. [Google Scholar] [CrossRef] [PubMed]
  94. Paddison, P.J.; Caudy, A.A.; Bernstein, E.; Hannon, G.J.; Conklin, D.S. Short hairpin RNAs (shRNAs) induce sequence-specific silencing in mammalian cells. Genes Dev. 2002, 16, 948–958. [Google Scholar] [CrossRef]
  95. López-Fraga, M.; Martínez, T.; Jiménez, A. RNA interference technologies and therapeutics: From basic research to products. Biodrugs 2009, 23, 305–332. [Google Scholar] [CrossRef] [PubMed]
  96. Chernikov, I.V.; Ponomareva, U.A.; Chernolovskaya, E.L. Structural Modifications of siRNA Improve Its Performance In Vivo. Int. J. Mol. Sci. 2023, 24, 956. [Google Scholar] [CrossRef] [PubMed]
  97. Brummelkamp, T.R.; Bernards, R.; Agami, R. A system for stable expression of short interfering RNAs in mammalian cells. Science 2002, 296, 550–553. [Google Scholar] [CrossRef]
  98. Zhang, H.; Kolb, F.A.; Jaskiewicz, L.; Westhof, E.; Filipowicz, W. Single processing center models for human Dicer and bacterial RNase III. Cell 2004, 118, 57–68. [Google Scholar] [CrossRef]
  99. Elbashir, S.M.; Lendeckel, W.; Tuschl, T. RNA interference is mediated by 21- and 22-nucleotide RNAs. Genes Dev. 2001, 15, 188–200. [Google Scholar] [CrossRef]
  100. Liu, J.; Carmell, M.A.; Rivas, F.V.; Marsden, C.G.; Thomson, J.M.; Song, J.J.; Hammond, S.M.; Joshua-Tor, L.; Hannon, G.J. Argonaute2 is the catalytic engine of mammalian RNAi. Science 2004, 305, 1437–1441. [Google Scholar] [CrossRef]
  101. Yi, R.; Qin, Y.; Macara, I.G.; Cullen, B.R. Exportin-5 mediates the nuclear export of pre-microRNAs and short hairpin RNAs. Genes Dev. 2003, 17, 3011–3016. [Google Scholar] [CrossRef]
  102. McBride, J.L.; Boudreau, R.L.; Harper, S.Q.; Staber, P.D.; Monteys, A.M.; Martins, I.; Gilmore, B.L.; Burstein, H.; Peluso, R.W.; Polisky, B.; et al. Artificial miRNAs mitigate shRNA-mediated toxicity in the brain: Implications for the therapeutic development of RNAi. Proc. Natl. Acad. Sci. USA 2008, 105, 5868–5873. [Google Scholar] [CrossRef]
  103. Kohen, R.; Baldwin, K.T.; Garay, P.M.; Tsukahara, T.; Chen, A.; Flynn, C.G.; Johnson, C.; Zhao, X.-F.; Sutton, M.A.; Iwase, S.; et al. Small-hairpin RNAs cause target-independent microRNA dysregulation in neurons and elicit global transcriptomic changes. bioRxiv 2020. [Google Scholar] [CrossRef]
  104. Rao, D.D.; Vorhies, J.S.; Senzer, N.; Nemunaitis, J. siRNA vs. shRNA: Similarities and differences. Adv. Drug Deliv. Rev. 2009, 61, 746–759. [Google Scholar] [CrossRef]
  105. Mansoori, B.; Sandoghchian Shotorbani, S.; Baradaran, B. RNA interference and its role in cancer therapy. Adv. Pharm. Bull. 2014, 4, 313–321. [Google Scholar] [CrossRef] [PubMed]
  106. Leung, A.K.; Tam, Y.Y.; Cullis, P.R. Lipid nanoparticles for short interfering RNA delivery. Adv. Genet. 2014, 88, 71–110. [Google Scholar] [CrossRef]
  107. McAnuff, M.A.; Rettig, G.R.; Rice, K.G. Potency of siRNA versus shRNA mediated knockdown in vivo. J. Pharm. Sci. 2007, 96, 2922–2930. [Google Scholar] [CrossRef]
  108. Aguiar, S.; van der Gaag, B.; Cortese, F.A.B. RNAi mechanisms in Huntington’s disease therapy: siRNA versus shRNA. Transl. Neurodegener. 2017, 6, 30. [Google Scholar] [CrossRef]
  109. Jinek, M.; Chylinski, K.; Fonfara, I.; Hauer, M.; Doudna, J.A.; Charpentier, E. A programmable dual-RNA-guided DNA endonuclease in adaptive bacterial immunity. Science 2012, 337, 816–821. [Google Scholar] [CrossRef] [PubMed]
  110. Mali, P.; Yang, L.; Esvelt, K.M.; Aach, J.; Guell, M.; DiCarlo, J.E.; Norville, J.E.; Church, G.M. RNA-guided human genome engineering via Cas9. Science 2013, 339, 823–826. [Google Scholar] [CrossRef]
  111. Ishino, Y.; Shinagawa, H.; Makino, K.; Amemura, M.; Nakata, A. Nucleotide sequence of the iap gene, responsible for alkaline phosphatase isozyme conversion in Escherichia coli, and identification of the gene product. J. Bacteriol. 1987, 169, 5429–5433. [Google Scholar] [CrossRef]
  112. Navarro, C.; Díaz, M.P.; Duran, P.; Castro, A.; Díaz, A.; Cano, C.; Carbonell-Zabaleta, A.-K.; Solano-Jimenez, D.-S.; Rivera-Porras, D.; Contreras-Velásquez, J.C.; et al. CRISPR-Cas Systems: A Functional Perspective and Innovations. Int. J. Mol. Sci. 2025, 26, 3645. [Google Scholar] [CrossRef] [PubMed]
  113. Ishino, Y.; Krupovic, M.; Forterre, P. History of CRISPR-Cas from Encounter with a Mysterious Repeated Sequence to Genome Editing Technology. J. Bacteriol. 2018, 200, e00580-17. [Google Scholar] [CrossRef] [PubMed]
  114. Wright, A.V.; Nuñez, J.K.; Doudna, J.A. Biology and Applications of CRISPR Systems: Harnessing Nature’s Toolbox for Genome Engineering. Cell 2016, 164, 29–44. [Google Scholar] [CrossRef]
  115. Boti, M.A.; Athanasopoulou, K.; Adamopoulos, P.G.; Sideris, D.C.; Scorilas, A. Recent Advances in Genome-Engineering Strategies. Genes 2023, 14, 129. [Google Scholar] [CrossRef]
  116. Abudayyeh, O.O.; Gootenberg, J.S.; Essletzbichler, P.; Han, S.; Joung, J.; Belanto, J.J.; Verdine, V.; Cox, D.B.T.; Kellner, M.J.; Regev, A.; et al. RNA targeting with CRISPR-Cas13. Nature 2017, 550, 280–284. [Google Scholar] [CrossRef]
  117. Abudayyeh, O.O.; Gootenberg, J.S.; Konermann, S.; Joung, J.; Slaymaker, I.M.; Cox, D.B.; Shmakov, S.; Makarova, K.S.; Semenova, E.; Minakhin, L.; et al. C2c2 is a single-component programmable RNA-guided RNA-targeting CRISPR effector. Science 2016, 353, aaf5573. [Google Scholar] [CrossRef] [PubMed]
  118. East-Seletsky, A.; O’Connell, M.R.; Knight, S.C.; Burstein, D.; Cate, J.H.; Tjian, R.; Doudna, J.A. Two distinct RNase activities of CRISPR-C2c2 enable guide-RNA processing and RNA detection. Nature 2016, 538, 270–273. [Google Scholar] [CrossRef]
  119. Palaz, F.; Kalkan, A.K.; Can, Ö.; Demir, A.N.; Tozluyurt, A.; Özcan, A.; Ozsoz, M. CRISPR-Cas13 System as a Promising and Versatile Tool for Cancer Diagnosis, Therapy, and Research. ACS Synth. Biol. 2021, 10, 1245–1267. [Google Scholar] [CrossRef]
  120. Durán-Vinet, B.; Araya-Castro, K.; Calderón, J.; Vergara, L.; Weber, H.; Retamales, J.; Araya-Castro, P.; Leal-Rojas, P. CRISPR/Cas13-Based Platforms for a Potential Next-Generation Diagnosis of Colorectal Cancer through Exosomes Micro-RNA Detection: A Review. Cancers 2021, 13, 4640. [Google Scholar] [CrossRef]
  121. Xu, D.; Cai, Y.; Tang, L.; Han, X.; Gao, F.; Cao, H.; Qi, F.; Kapranov, P. A CRISPR/Cas13-based approach demonstrates biological relevance of vlinc class of long non-coding RNAs in anticancer drug response. Sci. Rep. 2020, 10, 1794. [Google Scholar] [CrossRef]
  122. Liang, W.-W.; Müller, S.; Hart, S.K.; Wessels, H.-H.; Méndez-Mancilla, A.; Sookdeo, A.; Choi, O.; Caragine, C.M.; Corman, A.; Lu, L.; et al. Transcriptome-scale RNA-targeting CRISPR screens reveal essential lncRNAs in human cells. Cell 2024, 187, 7637–7654.e7629. [Google Scholar] [CrossRef]
  123. Li, S.; Li, X.; Xue, W.; Zhang, L.; Yang, L.-Z.; Cao, S.-M.; Lei, Y.-N.; Liu, C.-X.; Guo, S.-K.; Shan, L.; et al. Screening for functional circular RNAs using the CRISPR–Cas13 system. Nat. Methods 2021, 18, 51–59. [Google Scholar] [CrossRef]
  124. Smargon, A.A.; Cox, D.B.T.; Pyzocha, N.K.; Zheng, K.; Slaymaker, I.M.; Gootenberg, J.S.; Abudayyeh, O.A.; Essletzbichler, P.; Shmakov, S.; Makarova, K.S.; et al. Cas13b Is a Type VI-B CRISPR-Associated RNA-Guided RNase Differentially Regulated by Accessory Proteins Csx27 and Csx28. Mol. Cell 2017, 65, 618–630.e617. [Google Scholar] [CrossRef]
  125. Yan, W.X.; Chong, S.; Zhang, H.; Makarova, K.S.; Koonin, E.V.; Cheng, D.R.; Scott, D.A. Cas13d Is a Compact RNA-Targeting Type VI CRISPR Effector Positively Modulated by a WYL-Domain-Containing Accessory Protein. Mol. Cell 2018, 70, 327–339.e325. [Google Scholar] [CrossRef]
  126. Hu, Y.; Chen, Y.; Xu, J.; Wang, X.; Luo, S.; Mao, B.; Zhou, Q.; Li, W. Metagenomic discovery of novel CRISPR-Cas13 systems. Cell Discov. 2022, 8, 107. [Google Scholar] [CrossRef] [PubMed]
  127. Xu, C.; Zhou, Y.; Xiao, Q.; He, B.; Geng, G.; Wang, Z.; Cao, B.; Dong, X.; Bai, W.; Wang, Y.; et al. Programmable RNA editing with compact CRISPR-Cas13 systems from uncultivated microbes. Nat. Methods 2021, 18, 499–506. [Google Scholar] [CrossRef]
  128. He, B.; Peng, W.; Huang, J.; Zhang, H.; Zhou, Y.; Yang, X.; Liu, J.; Li, Z.; Xu, C.; Xue, M.; et al. Modulation of metabolic functions through Cas13d-mediated gene knockdown in liver. Protein Cell 2020, 11, 518–524. [Google Scholar] [CrossRef] [PubMed]
  129. Basit, A.; Liu, A.; Zheng, W.; Zhu, J. A review on the mechanism and potential diagnostic application of CRISPR/Cas13a system. Mamm. Genome 2025, 36, 709–726. [Google Scholar] [CrossRef]
  130. McCoullough, L.C.; Fareh, M.; Hu, W.; Sozzi, V.; Makhlouf, C.; Droungas, Y.; Lee, C.L.; Takawy, M.; Fabb, S.A.; Payne, T.J.; et al. CRISPR-Cas13b-mediated suppression of HBV replication and protein expression. J. Hepatol. 2024, 81, 794–805. [Google Scholar] [CrossRef] [PubMed]
  131. Gupta, R.; Ghosh, A.; Chakravarti, R.; Singh, R.; Ravichandiran, V.; Swarnakar, S.; Ghosh, D. Cas13d: A New Molecular Scissor for Transcriptome Engineering. Front. Cell Dev. Biol. 2022, 10, 866800. [Google Scholar] [CrossRef]
  132. Tieu, V.; Sotillo, E.; Bjelajac, J.R.; Chen, C.; Malipatlolla, M.; Guerrero, J.A.; Xu, P.; Quinn, P.J.; Fisher, C.; Klysz, D.; et al. A versatile CRISPR-Cas13d platform for multiplexed transcriptomic regulation and metabolic engineering in primary human T cells. Cell 2024, 187, 1278–1295.e1220. [Google Scholar] [CrossRef]
  133. Shi, P.; Wu, X. Programmable RNA targeting with CRISPR-Cas13. RNA Biol. 2024, 21, 1–9. [Google Scholar] [CrossRef]
  134. Zhang, B.; Ye, Y.; Ye, W.; Perčulija, V.; Jiang, H.; Chen, Y.; Li, Y.; Chen, J.; Lin, J.; Wang, S.; et al. Two HEPN domains dictate CRISPR RNA maturation and target cleavage in Cas13d. Nat. Commun. 2019, 10, 2544. [Google Scholar] [CrossRef]
  135. Zhang, B.; Ye, W.; Ye, Y.; Zhou, H.; Saeed, A.; Chen, J.; Lin, J.; Perčulija, V.; Chen, Q.; Chen, C.J.; et al. Structural insights into Cas13b-guided CRISPR RNA maturation and recognition. Cell Res. 2018, 28, 1198–1201. [Google Scholar] [CrossRef] [PubMed]
  136. Wan, Y.; Huang, C.; Feng, D.; Wang, L.; Lin, X.; Zhao, X.; Han, L.; Zhu, Y.; Hao, L.; Du, H.; et al. Characterizing the collateral activity of CRISPR/Cas13 in mammalian cells: Implications for RNA editing and therapeutic applications. Int. J. Biol. Macromol. 2024, 283, 137861. [Google Scholar] [CrossRef]
  137. Shi, P.; Murphy, M.R.; Aparicio, A.O.; Kesner, J.S.; Fang, Z.; Chen, Z.; Trehan, A.; Guo, Y.; Wu, X. Collateral activity of the CRISPR/RfxCas13d system in human cells. Commun. Biol. 2023, 6, 334. [Google Scholar] [CrossRef] [PubMed]
  138. Li, Y.; Xu, J.; Guo, X.; Li, Z.; Cao, L.; Liu, S.; Guo, Y.; Wang, G.; Luo, Y.; Zhang, Z.; et al. The collateral activity of RfxCas13d can induce lethality in a RfxCas13d knock-in mouse model. Genome Biol. 2023, 24, 20. [Google Scholar] [CrossRef]
  139. Ai, Y.; Liang, D.; Wilusz, J.E. CRISPR/Cas13 effectors have differing extents of off-target effects that limit their utility in eukaryotic cells. Nucleic Acids Res. 2022, 50, e65. [Google Scholar] [CrossRef] [PubMed]
  140. Tao, J.; Bauer, D.E.; Chiarle, R. Assessing and advancing the safety of CRISPR-Cas tools: From DNA to RNA editing. Nat. Commun. 2023, 14, 212. [Google Scholar] [CrossRef]
  141. Powell, J.E.; Lim, C.K.W.; Krishnan, R.; McCallister, T.X.; Saporito-Magriña, C.; Zeballos, M.A.; McPheron, G.D.; Gaj, T. Targeted gene silencing in the nervous system with CRISPR-Cas13. Sci. Adv. 2022, 8, eabk2485. [Google Scholar] [CrossRef]
  142. Blanchard, E.L.; Vanover, D.; Bawage, S.S.; Tiwari, P.M.; Rotolo, L.; Beyersdorf, J.; Peck, H.E.; Bruno, N.C.; Hincapie, R.; Michel, F.; et al. Treatment of influenza and SARS-CoV-2 infections via mRNA-encoded Cas13a in rodents. Nat. Biotechnol. 2021, 39, 717–726. [Google Scholar] [CrossRef] [PubMed]
  143. Huynh, N.; Depner, N.; Larson, R.; King-Jones, K. A versatile toolkit for CRISPR-Cas13-based RNA manipulation in Drosophila. Genome Biol. 2020, 21, 279. [Google Scholar] [CrossRef] [PubMed]
  144. Engineered Cas13 variants with minimal collateral RNA targeting. Nat. Biotechnol. 2023, 41, 29–30. [CrossRef]
  145. Cox, D.B.T.; Gootenberg, J.S.; Abudayyeh, O.O.; Franklin, B.; Kellner, M.J.; Joung, J.; Zhang, F. RNA editing with CRISPR-Cas13. Science 2017, 358, 1019–1027. [Google Scholar] [CrossRef]
  146. Villiger, L.; Joung, J.; Koblan, L.; Weissman, J.; Abudayyeh, O.O.; Gootenberg, J.S. CRISPR technologies for genome, epigenome and transcriptome editing. Nat. Rev. Mol. Cell Biol. 2024, 25, 464–487. [Google Scholar] [CrossRef]
  147. Nishikura, K. A-to-I editing of coding and non-coding RNAs by ADARs. Nat. Rev. Mol. Cell Biol. 2016, 17, 83–96. [Google Scholar] [CrossRef]
  148. Vesely, C.; Jantsch, M.F. An I for an A: Dynamic Regulation of Adenosine Deamination-Mediated RNA Editing. Genes 2021, 12, 1026. [Google Scholar] [CrossRef]
  149. Pecori, R.; Di Giorgio, S.; Paulo Lorenzo, J.; Nina Papavasiliou, F. Functions and consequences of AID/APOBEC-mediated DNA and RNA deamination. Nat. Rev. Genet. 2022, 23, 505–518. [Google Scholar] [CrossRef] [PubMed]
  150. Abudayyeh, O.O.; Gootenberg, J.S.; Franklin, B.; Koob, J.; Kellner, M.J.; Ladha, A.; Joung, J.; Kirchgatterer, P.; Cox, D.B.T.; Zhang, F. A cytosine deaminase for programmable single-base RNA editing. Science 2019, 365, 382–386. [Google Scholar] [CrossRef]
  151. Jia, G.; Fu, Y.; Zhao, X.; Dai, Q.; Zheng, G.; Yang, Y.; Yi, C.; Lindahl, T.; Pan, T.; Yang, Y.G.; et al. N6-methyladenosine in nuclear RNA is a major substrate of the obesity-associated FTO. Nat. Chem. Biol. 2011, 7, 885–887. [Google Scholar] [CrossRef]
  152. Liu, J.; Yue, Y.; Han, D.; Wang, X.; Fu, Y.; Zhang, L.; Jia, G.; Yu, M.; Lu, Z.; Deng, X.; et al. A METTL3-METTL14 complex mediates mammalian nuclear RNA N6-adenosine methylation. Nat. Chem. Biol. 2014, 10, 93–95. [Google Scholar] [CrossRef]
  153. Sun, X.; Wang, D.O.; Wang, J. Targeted manipulation of m(6)A RNA modification through CRISPR-Cas-based strategies. Methods 2022, 203, 56–61. [Google Scholar] [CrossRef] [PubMed]
  154. Jiang, X.; Liu, B.; Nie, Z.; Duan, L.; Xiong, Q.; Jin, Z.; Yang, C.; Chen, Y. The role of m6A modification in the biological functions and diseases. Signal Transduct. Target. Ther. 2021, 6, 74. [Google Scholar] [CrossRef]
  155. Shi, H.; Wei, J.; He, C. Where, When, and How: Context-Dependent Functions of RNA Methylation Writers, Readers, and Erasers. Mol. Cell 2019, 74, 640–650. [Google Scholar] [CrossRef] [PubMed]
  156. Yang, Y.; Hsu, P.J.; Chen, Y.S.; Yang, Y.G. Dynamic transcriptomic m(6)A decoration: Writers, erasers, readers and functions in RNA metabolism. Cell Res. 2018, 28, 616–624. [Google Scholar] [CrossRef] [PubMed]
  157. Zhao, B.S.; Roundtree, I.A.; He, C. Post-transcriptional gene regulation by mRNA modifications. Nat. Rev. Mol. Cell Biol. 2017, 18, 31–42. [Google Scholar] [CrossRef]
  158. Otoupal, P.B.; Cress, B.F.; Doudna, J.A.; Schoeniger, J.S. CRISPR-RNAa: Targeted activation of translation using dCas13 fusions to translation initiation factors. Nucleic Acids Res. 2022, 50, 8986–8998. [Google Scholar] [CrossRef]
  159. Apostolopoulos, A.; Kawamoto, N.; Chow, S.Y.A.; Tsuiji, H.; Ikeuchi, Y.; Shichino, Y.; Iwasaki, S. dCas13-mediated translational repression for accurate gene silencing in mammalian cells. Nat. Commun. 2024, 15, 2205. [Google Scholar] [CrossRef]
  160. Núñez-Álvarez, Y.; Espie-Caullet, T.; Buhagiar, G.; Rubio-Zulaika, A.; Alonso-Marañón, J.; Luna-Pérez, E.; Blazquez, L.; Luco, R.F. A CRISPR-dCas13 RNA-editing tool to study alternative splicing. Nucleic Acids Res. 2024, 52, 11926–11939. [Google Scholar] [CrossRef]
  161. Deng, K.; Yao, J.; Huang, J.; Ding, Y.; Zuo, J. Abnormal alternative splicing promotes tumor resistance in targeted therapy and immunotherapy. Transl. Oncol. 2021, 14, 101077. [Google Scholar] [CrossRef] [PubMed]
  162. Zhang, Y.; Qian, J.; Gu, C.; Yang, Y. Alternative splicing and cancer: A systematic review. Signal Transduct. Target. Ther. 2021, 6, 78. [Google Scholar] [CrossRef]
  163. Chen, H.; Tang, J.; Xiang, J. Alternative Splicing in Tumorigenesis and Cancer Therapy. Biomolecules 2025, 15, 789. [Google Scholar] [CrossRef]
  164. Yang, L.-Z.; Wang, Y.; Li, S.-Q.; Yao, R.-W.; Luan, P.-F.; Wu, H.; Carmichael, G.G.; Chen, L.-L. Dynamic Imaging of RNA in Living Cells by CRISPR-Cas13 Systems. Mol. Cell 2019, 76, 981–997.e7. [Google Scholar] [CrossRef]
  165. Cao, H.; Wang, Y.; Zhang, N.; Xia, S.; Tian, P.; Lu, L.; Du, J.; Du, Y. Progress of CRISPR-Cas13 Mediated Live-Cell RNA Imaging and Detection of RNA-Protein Interactions. Front. Cell Dev. Biol. 2022, 10, 866820. [Google Scholar] [CrossRef] [PubMed]
  166. Chen, M.; Sui, T.; Yang, L.; Qian, Y.; Liu, Z.; Liu, Y.; Wang, G.; Lai, L.; Li, Z. Live imaging of RNA and RNA splicing in mammalian cells via the dcas13a-SunTag-BiFC system. Biosens. Bioelectron. 2022, 204, 114074. [Google Scholar] [CrossRef] [PubMed]
  167. Tian, Y.-F.; Zhang, Y.-P.; Wu, Q.-M.; Pang, D.-W.; Liu, S.-L.; Wang, Z.-G. Real-Time Imaging of Single Viral mRNA Translation in Live Cells Using CRISPR/dCas13. Anal. Chem. 2023, 95, 16298–16304. [Google Scholar] [CrossRef]
  168. Scott, D.E.; Bayly, A.R.; Abell, C.; Skidmore, J. Small molecules, big targets: Drug discovery faces the protein-protein interaction challenge. Nat. Rev. Drug Discov. 2016, 15, 533–550. [Google Scholar] [CrossRef]
  169. Stockwell, B.R. Exploring biology with small organic molecules. Nature 2004, 432, 846–854. [Google Scholar] [CrossRef]
  170. Donlic, A.; Hargrove, A.E. Targeting RNA in mammalian systems with small molecules. Wiley Interdiscip. Rev. RNA 2018, 9, e1477. [Google Scholar] [CrossRef]
  171. Chen, B.; Dragomir, M.P.; Yang, C.; Li, Q.; Horst, D.; Calin, G.A. Targeting non-coding RNAs to overcome cancer therapy resistance. Signal Transduct. Target. Ther. 2022, 7, 121. [Google Scholar] [CrossRef]
  172. Xu, K.; Guo, H.; Xia, A.; Wang, Z.; Wang, S.; Wang, Q. Non-coding RNAs in radiotherapy resistance: Roles and therapeutic implications in gastrointestinal cancer. Biomed. Pharmacother. 2023, 161, 114485. [Google Scholar] [CrossRef] [PubMed]
  173. Shen, C.; Cao, R.; Zhou, Q.; Zhou, T.; Wu, T.; Gao, W.; Wang, G.; Feng, G.; Qiao, L.; Wang, T. LINC00654 promotes ovarian cancer progression by facilitating nuclear export of HuR and stabilizing oncogenic mRNAs. Oncogene 2025, 44, 3422–3436. [Google Scholar] [CrossRef]
  174. Falese, J.P.; Donlic, A.; Hargrove, A.E. Targeting RNA with small molecules: From fundamental principles towards the clinic. Chem. Soc. Rev. 2021, 50, 2224–2243. [Google Scholar] [CrossRef] [PubMed]
  175. Blythe, A.J.; Fox, A.H.; Bond, C.S. The ins and outs of lncRNA structure: How, why and what comes next? Biochim. Biophys. Acta 2016, 1859, 46–58. [Google Scholar] [CrossRef]
  176. Miglietta, G.; Cogoi, S.; Marinello, J.; Capranico, G.; Tikhomirov, A.S.; Shchekotikhin, A.; Xodo, L.E. RNA G-Quadruplexes in Kirsten Ras (KRAS) Oncogene as Targets for Small Molecules Inhibiting Translation. J. Med. Chem. 2017, 60, 9448–9461. [Google Scholar] [CrossRef]
  177. Chan, K.H.; Zheng, B.X.; Leung, A.S.; Long, W.; Zhao, Y.; Zheng, Y.; Wong, W.L. A NRAS mRNA G-quadruplex structure-targeting small-molecule ligand reactivating DNA damage response in human cancer cells for combination therapy with clinical PI3K inhibitors. Int. J. Biol. Macromol. 2024, 279, 135308. [Google Scholar] [CrossRef] [PubMed]
  178. Sanchez-Martin, V.; Soriano, M.; Garcia-Salcedo, J.A. Quadruplex Ligands in Cancer Therapy. Cancers 2021, 13, 3156. [Google Scholar] [CrossRef]
  179. Savva, L.; Georgiades, S.N. Recent Developments in Small-Molecule Ligands of Medicinal Relevance for Harnessing the Anticancer Potential of G-Quadruplexes. Molecules 2021, 26, 841. [Google Scholar] [CrossRef] [PubMed]
  180. Tong, Y.; Childs-Disney, J.L.; Disney, M.D. Targeting RNA with small molecules, from RNA structures to precision medicines: IUPHAR review: 40. Br. J. Pharmacol. 2024, 181, 4152–4173. [Google Scholar] [CrossRef]
  181. Davila-Calderon, J.; Patwardhan, N.N.; Chiu, L.Y.; Sugarman, A.; Cai, Z.; Penutmutchu, S.R.; Li, M.L.; Brewer, G.; Hargrove, A.E.; Tolbert, B.S. IRES-targeting small molecule inhibits enterovirus 71 replication via allosteric stabilization of a ternary complex. Nat. Commun. 2020, 11, 4775. [Google Scholar] [CrossRef]
  182. Carnevali, M.; Parsons, J.; Wyles, D.L.; Hermann, T. A modular approach to synthetic RNA binders of the hepatitis C virus internal ribosome entry site. Chembiochem 2010, 11, 1364–1367. [Google Scholar] [CrossRef]
  183. Dibrov, S.M.; Parsons, J.; Carnevali, M.; Zhou, S.; Rynearson, K.D.; Ding, K.; Garcia Sega, E.; Brunn, N.D.; Boerneke, M.A.; Castaldi, M.P.; et al. Hepatitis C virus translation inhibitors targeting the internal ribosomal entry site. J. Med. Chem. 2014, 57, 1694–1707. [Google Scholar] [CrossRef]
  184. Shang, R.; Lee, S.; Senavirathne, G.; Lai, E.C. microRNAs in action: Biogenesis, function and regulation. Nat. Rev. Genet. 2023, 24, 816–833. [Google Scholar] [CrossRef]
  185. Velagapudi, S.P.; Cameron, M.D.; Haga, C.L.; Rosenberg, L.H.; Lafitte, M.; Duckett, D.R.; Phinney, D.G.; Disney, M.D. Design of a small molecule against an oncogenic noncoding RNA. Proc. Natl. Acad. Sci. USA 2016, 113, 5898–5903. [Google Scholar] [CrossRef]
  186. Maucort, C.; Vo, D.D.; Aouad, S.; Charrat, C.; Azoulay, S.; Di Giorgio, A.; Duca, M. Design and Implementation of Synthetic RNA Binders for the Inhibition of miR-21 Biogenesis. ACS Med. Chem. Lett. 2021, 12, 899–906. [Google Scholar] [CrossRef]
  187. Naryshkin, N.A.; Weetall, M.; Dakka, A.; Narasimhan, J.; Zhao, X.; Feng, Z.; Ling, K.K.; Karp, G.M.; Qi, H.; Woll, M.G.; et al. Motor neuron disease. SMN2 splicing modifiers improve motor function and longevity in mice with spinal muscular atrophy. Science 2014, 345, 688–693. [Google Scholar] [CrossRef]
  188. Chang, J.G.; Yang, D.M.; Chang, W.H.; Chow, L.P.; Chan, W.L.; Lin, H.H.; Huang, H.D.; Chang, Y.S.; Hung, C.H.; Yang, W.K. Small molecule amiloride modulates oncogenic RNA alternative splicing to devitalize human cancer cells. PLoS ONE 2011, 6, e18643. [Google Scholar] [CrossRef]
  189. Palacino, J.; Swalley, S.E.; Song, C.; Cheung, A.K.; Shu, L.; Zhang, X.; Van Hoosear, M.; Shin, Y.; Chin, D.N.; Keller, C.G.; et al. Corrigendum: SMN2 splice modulators enhance U1-pre-mRNA association and rescue SMA mice. Nat. Chem. Biol. 2016, 12, 304. [Google Scholar] [CrossRef]
  190. Meyer, S.M.; Williams, C.C.; Akahori, Y.; Tanaka, T.; Aikawa, H.; Tong, Y.; Childs-Disney, J.L.; Disney, M.D. Small molecule recognition of disease-relevant RNA structures. Chem. Soc. Rev. 2020, 49, 7167–7199. [Google Scholar] [CrossRef]
  191. Tong, Y.; Gibaut, Q.M.R.; Rouse, W.; Childs-Disney, J.L.; Suresh, B.M.; Abegg, D.; Choudhary, S.; Akahori, Y.; Adibekian, A.; Moss, W.N.; et al. Transcriptome-Wide Mapping of Small-Molecule RNA-Binding Sites in Cells Informs an Isoform-Specific Degrader of QSOX1 mRNA. J. Am. Chem. Soc. 2022, 144, 11620–11625. [Google Scholar] [CrossRef]
  192. Costales, M.G.; Aikawa, H.; Li, Y.; Childs-Disney, J.L.; Abegg, D.; Hoch, D.G.; Pradeep Velagapudi, S.; Nakai, Y.; Khan, T.; Wang, K.W.; et al. Small-molecule targeted recruitment of a nuclease to cleave an oncogenic RNA in a mouse model of metastatic cancer. Proc. Natl. Acad. Sci. USA 2020, 117, 2406–2411. [Google Scholar] [CrossRef]
  193. Dey, S.K.; Jaffrey, S.R. RIBOTACs: Small Molecules Target RNA for Degradation. Cell Chem. Biol. 2019, 26, 1047–1049. [Google Scholar] [CrossRef]
  194. Ganser, L.R.; Kelly, M.L.; Patwardhan, N.N.; Hargrove, A.E.; Al-Hashimi, H.M. Demonstration that Small Molecules can Bind and Stabilize Low-abundance Short-lived RNA Excited Conformational States. J. Mol. Biol. 2020, 432, 1297–1304. [Google Scholar] [CrossRef]
  195. Weng, Y.; Xiao, H.; Zhang, J.; Liang, X.J.; Huang, Y. RNAi therapeutic and its innovative biotechnological evolution. Biotechnol. Adv. 2019, 37, 801–825. [Google Scholar] [CrossRef]
  196. Driscoll, J.; Gondaliya, P.; Zinn, D.A.; Jain, R.; Yan, I.K.; Dong, H.; Patel, T. Using aptamers for targeted delivery of RNA therapies. Mol. Ther. 2025, 33, 1344–1367. [Google Scholar] [CrossRef]
  197. Shigdar, S.; Schrand, B.; Giangrande, P.H.; de Franciscis, V. Aptamers: Cutting edge of cancer therapies. Mol. Ther. 2021, 29, 2396–2411. [Google Scholar] [CrossRef]
  198. Sun, H.; Zhu, X.; Lu, P.Y.; Rosato, R.R.; Tan, W.; Zu, Y. Oligonucleotide aptamers: New tools for targeted cancer therapy. Mol. Ther. Nucleic Acids 2014, 3, e182. [Google Scholar] [CrossRef]
  199. Vázquez Torres, S.; Leung, P.J.Y.; Venkatesh, P.; Lutz, I.D.; Hink, F.; Huynh, H.H.; Becker, J.; Yeh, A.H.; Juergens, D.; Bennett, N.R.; et al. De novo design of high-affinity binders of bioactive helical peptides. Nature 2024, 626, 435–442. [Google Scholar] [CrossRef]
  200. Jumper, J.; Evans, R.; Pritzel, A.; Green, T.; Figurnov, M.; Ronneberger, O.; Tunyasuvunakool, K.; Bates, R.; Žídek, A.; Potapenko, A.; et al. Highly accurate protein structure prediction with AlphaFold. Nature 2021, 596, 583–589. [Google Scholar] [CrossRef]
  201. Townshend, R.J.L.; Eismann, S.; Watkins, A.M.; Rangan, R.; Karelina, M.; Das, R.; Dror, R.O. Geometric deep learning of RNA structure. Science 2021, 373, 1047–1051. [Google Scholar] [CrossRef]
  202. Zhu, G.; Chen, X. Aptamer-based targeted therapy. Adv. Drug Del. Rev. 2018, 134, 65–78. [Google Scholar] [CrossRef]
  203. Chen, M.; Yu, Y.; Jiang, F.; Zhou, J.; Li, Y.; Liang, C.; Dang, L.; Lu, A.; Zhang, G. Development of Cell-SELEX Technology and Its Application in Cancer Diagnosis and Therapy. Int. J. Mol. Sci. 2016, 17, 2079. [Google Scholar] [CrossRef]
  204. Fang, X.; Tan, W. Aptamers generated from cell-SELEX for molecular medicine: A chemical biology approach. Acc. Chem. Res. 2010, 43, 48–57. [Google Scholar] [CrossRef] [PubMed]
  205. Tuerk, C.; Gold, L. Systematic evolution of ligands by exponential enrichment: RNA ligands to bacteriophage T4 DNA polymerase. Science 1990, 249, 505–510. [Google Scholar] [CrossRef]
  206. Kim, M.; Kim, D.-M.; Kim, K.-S.; Jung, W.; Kim, D.-E. Applications of Cancer Cell-Specific Aptamers in Targeted Delivery of Anticancer Therapeutic Agents. Molecules 2018, 23, 830. [Google Scholar] [CrossRef]
  207. Dey, S.; Sczepanski, J.T. In vitro selection of l-DNA aptamers that bind a structured d-RNA molecule. Nucleic Acids Res. 2020, 48, 1669–1680. [Google Scholar] [CrossRef]
  208. Wang, H.; Wang, H.; Zhang, M.; Jia, Y.; Li, Z. A label-free aptamer-based biosensor for microRNA detection by the RNA-regulated fluorescence of malachite green. RSC Adv. 2019, 9, 32906–32910. [Google Scholar] [CrossRef] [PubMed]
  209. Wang, L.; Hast, K.; Aggarwal, T.; Baci, M.; Hong, J.; Izgu, E.C. MicroRNA detection in biologically relevant media using a split aptamer platform. Bioorg. Med. Chem. 2022, 69, 116909. [Google Scholar] [CrossRef]
  210. Napit, R.; Jaysawal, S.K.; Chowdhury, R.; Catague, J.; Melke, H.; Pham, C.V.; Xu, H.; Jia, L.; Lin, J.; Hou, Y.; et al. Aptasensors and Advancement in Molecular Recognition Technology. Adv. Mater. Technol. 2025, 10, 2400504. [Google Scholar] [CrossRef]
  211. Tesmer, V.M.; Lennarz, S.; Mayer, G.; Tesmer, J.J. Molecular mechanism for inhibition of g protein-coupled receptor kinase 2 by a selective RNA aptamer. Structure 2012, 20, 1300–1309. [Google Scholar] [CrossRef]
  212. Luo, F.; Yang, G.; Bai, X.; Yuan, D.; Li, L.; Wang, D.; Lu, X.; Cheng, Y.; Wang, Y.; Song, X.; et al. Anti-tumor effect of PD-L1-targeting antagonistic aptamer-ASO delivery system with dual inhibitory function in immunotherapy. Cell Chem. Biol. 2023, 30, 1390–1401.e1396. [Google Scholar] [CrossRef]
  213. Catuogno, S.; Rienzo, A.; Di Vito, A.; Esposito, C.L.; de Franciscis, V. Selective delivery of therapeutic single strand antimiRs by aptamer-based conjugates. J. Control. Release 2015, 210, 147–159. [Google Scholar] [CrossRef]
  214. Russo, V.; Paciocco, A.; Affinito, A.; Roscigno, G.; Fiore, D.; Palma, F.; Galasso, M.; Volinia, S.; Fiorelli, A.; Esposito, C.L.; et al. Aptamer-miR-34c Conjugate Affects Cell Proliferation of Non-Small-Cell Lung Cancer Cells. Mol. Ther. Nucleic Acids 2018, 13, 334–346. [Google Scholar] [CrossRef]
  215. Gilboa-Geffen, A.; Hamar, P.; Le, M.T.N.; Wheeler, L.A.; Trifonova, R.; Petrocca, F.; Wittrup, A.; Lieberman, J. Gene Knockdown by EpCAM Aptamer–siRNA Chimeras Suppresses Epithelial Breast Cancers and Their Tumor-Initiating Cells. Mol. Cancer Ther. 2015, 14, 2279–2291. [Google Scholar] [CrossRef]
  216. Dassie, J.P.; Liu, X.-y.; Thomas, G.S.; Whitaker, R.M.; Thiel, K.W.; Stockdale, K.R.; Meyerholz, D.K.; McCaffrey, A.P.; McNamara, J.O.; Giangrande, P.H. Systemic administration of optimized aptamer-siRNA chimeras promotes regression of PSMA-expressing tumors. Nat. Biotechnol. 2009, 27, 839–846. [Google Scholar] [CrossRef]
  217. Berezhnoy, A.; Brenneman, R.; Bajgelman, M.; Seales, D.; Gilboa, E. Thermal Stability of siRNA Modulates Aptamer- conjugated siRNA Inhibition. Mol. Ther. Nucleic Acids 2012, 1, e51. [Google Scholar] [CrossRef] [PubMed]
  218. Meraviglia-Crivelli, D.; Villanueva, H.; Menon, A.P.; Zheleva, A.; Moreno, B.; Villalba-Esparza, M.; Pastor, F. A pan-tumor-siRNA aptamer chimera to block nonsense-mediated mRNA decay inflames and suppresses tumor progression. Mol. Ther. Nucleic Acids 2022, 29, 413–425. [Google Scholar] [CrossRef] [PubMed]
  219. Shi, X.; Song, P.; Tao, S.; Zhang, X.; Chu, C.-Q. Silencing RORγt in Human CD4+ T cells with CD30 aptamer-RORγt shRNA Chimera. Sci. Rep. 2019, 9, 10375. [Google Scholar] [CrossRef] [PubMed]
  220. Song, P.; Chou, Y.K.; Zhang, X.; Meza-Romero, R.; Yomogida, K.; Benedek, G.; Chu, C.-Q. CD4 aptamer–RORγt shRNA chimera inhibits IL-17 synthesis by human CD4+ T cells. Biochem. Biophys. Res. Commun. 2014, 452, 1040–1045. [Google Scholar] [CrossRef]
  221. Yang, G.; Zhang, S.; Song, W.; Bai, X.; Li, L.; Luo, F.; Cheng, Y.; Wang, D.; Wang, Y.; Chen, J.; et al. Efficient Targeted Delivery of Bifunctional Circular Aptamer-ASO Chimera to Suppress the SARS-CoV-2 Proliferation and Inflammation. Small 2023, 19, 2207066. [Google Scholar] [CrossRef]
  222. Zhang, Z.; Chen, J.; Zhu, Z.; Zhu, Z.; Liao, X.; Wu, J.; Cheng, J.; Zhang, X.; Mei, H.; Yang, G. CRISPR-Cas13-Mediated Knockdown of lncRNA-GACAT3 Inhibited Cell Proliferation and Motility, and Induced Apoptosis by Increasing p21, Bax, and E-Cadherin Expression in Bladder Cancer. Front. Mol. Biosci. 2020, 7, 627774. [Google Scholar] [CrossRef]
  223. Wang, Y.; Zhu, J.; Liu, S.; Sun, Z.; Wen, G.; Huang, D.; Chen, M.; Liu, Y.; Lin, F. CRISPR-Cas13a Targeting the FGFR3-TACC3 Fusion Gene Inhibits Proliferation of Bladder Cancer Cells in vitro and in vivo. Onco Targets Ther. 2024, 17, 1197–1207. [Google Scholar] [CrossRef]
  224. Salguero-Aranda, C.; Sancho-Mensat, D.; Canals-Lorente, B.; Sultan, S.; Reginald, A.; Chapman, L. STAT6 knockdown using multiple siRNA sequences inhibits proliferation and induces apoptosis of human colorectal and breast cancer cell lines. PLoS ONE 2019, 14, e0207558. [Google Scholar] [CrossRef]
  225. Shahbazi, R.; Asik, E.; Kahraman, N.; Turk, M.; Ozpolat, B.; Ulubayram, K. Modified gold-based siRNA nanotherapeutics for targeted therapy of triple-negative breast cancer. Nanomedicine 2017, 12, 1961–1973. [Google Scholar] [CrossRef] [PubMed]
  226. Zhang, Q.; Yang, L.; Liu, Y.H.; Wilkinson, J.E.; Krainer, A.R. Antisense oligonucleotide therapy for H3.3K27M diffuse midline glioma. Sci. Transl. Med. 2023, 15, eadd8280. [Google Scholar] [CrossRef] [PubMed]
  227. Raguraman, P.; Balachandran, A.A.; Chen, S.; Diermeier, S.D.; Veedu, R.N. Antisense Oligonucleotide-Mediated Splice Switching: Potential Therapeutic Approach for Cancer Mitigation. Cancers 2021, 13, 5555. [Google Scholar] [CrossRef] [PubMed]
  228. Jiang, W.; Li, H.; Liu, X.; Zhang, J.; Zhang, W.; Li, T.; Liu, L.; Yu, X. Precise and efficient silencing of mutant Kras(G12D) by CRISPR-CasRx controls pancreatic cancer progression. Theranostics 2020, 10, 11507–11519. [Google Scholar] [CrossRef]
  229. Shembrey, C.; Yang, R.; Casan, J.; Hu, W.; Chen, H.; Singh, G.J.; Sadras, T.; Prasad, K.; Shortt, J.; Johnstone, R.W.; et al. Principles of CRISPR-Cas13 mismatch intolerance enable selective silencing of point-mutated oncogenic RNA with single-base precision. Sci. Adv. 2024, 10, eadl0731. [Google Scholar] [CrossRef] [PubMed]
  230. Zhao, X.; Liu, L.; Lang, J.; Cheng, K.; Wang, Y.; Li, X.; Shi, J.; Wang, Y.; Nie, G. A CRISPR-Cas13a system for efficient and specific therapeutic targeting of mutant KRAS for pancreatic cancer treatment. Cancer Lett. 2018, 431, 171–181. [Google Scholar] [CrossRef]
  231. Wang, P.; Zhang, H.; Yang, J.; Li, Z.; Wang, Y.; Leng, X.; Ganapathy, S.; Isakson, P.; Chen, C.; Zhu, T. Mu-KRAS attenuates Hippo signaling pathway through PKCι to sustain the growth of pancreatic cancer. J. Cell. Physiol. 2020, 235, 408–420. [Google Scholar] [CrossRef]
  232. Skoulidis, F.; Li, B.T.; Dy, G.K.; Price, T.J.; Falchook, G.S.; Wolf, J.; Italiano, A.; Schuler, M.; Borghaei, H.; Barlesi, F.; et al. Sotorasib for Lung Cancers with KRAS p.G12C Mutation. N. Engl. J. Med. 2021, 384, 2371–2381. [Google Scholar] [CrossRef] [PubMed]
  233. Bhat, A.A.; Afzal, O.; Afzal, M.; Gupta, G.; Thapa, R.; Ali, H.; Hassan Almalki, W.; Kazmi, I.; Alzarea, S.I.; Saleem, S.; et al. MALAT1: A key regulator in lung cancer pathogenesis and therapeutic targeting. Pathol. Res. Pract. 2024, 253, 154991. [Google Scholar] [CrossRef]
  234. Thangavelu, L.; Moglad, E.; Gupta, G.; Menon, S.V.; Gaur, A.; Sharma, S.; Kaur, M.; Chahar, M.; Sivaprasad, G.V.; Deorari, M. GAS5 lncRNA: A biomarker and therapeutic target in breast cancer. Pathol. Res. Pract. 2024, 260, 155424. [Google Scholar] [CrossRef]
  235. He, B.; Zhao, Z.; Cai, Q.; Zhang, Y.; Zhang, P.; Shi, S.; Xie, H.; Peng, X.; Yin, W.; Tao, Y.; et al. miRNA-based biomarkers, therapies, and resistance in Cancer. Int. J. Biol. Sci. 2020, 16, 2628–2647. [Google Scholar] [CrossRef]
  236. Chakrabortty, A.; Patton, D.J.; Smith, B.F.; Agarwal, P. miRNAs: Potential as Biomarkers and Therapeutic Targets for Cancer. Genes 2023, 14, 1375. [Google Scholar] [CrossRef]
  237. Singh, A.; Singh, A.K.; Giri, R.; Kumar, D.; Sharma, R.; Valis, M.; Kuca, K.; Garg, N. The role of microRNA-21 in the onset and progression of cancer. Future Med. Chem. 2021, 13, 1885–1906. [Google Scholar] [CrossRef] [PubMed]
  238. Fu, J.; Imani, S.; Wu, M.Y.; Wu, R.C. MicroRNA-34 Family in Cancers: Role, Mechanism, and Therapeutic Potential. Cancers 2023, 15, 4723. [Google Scholar] [CrossRef] [PubMed]
  239. Kang, E.; Kortylewski, M. Lipid Nanoparticle-Mediated Delivery of miRNA Mimics to Myeloid Cells. Methods Mol. Biol. 2023, 2691, 337–350. [Google Scholar] [CrossRef]
  240. Han, M.; Liu, M.; Wang, Y.; Chen, X.; Xu, J.; Sun, Y.; Zhao, L.; Qu, H.; Fan, Y.; Wu, C. Antagonism of miR-21 reverses epithelial-mesenchymal transition and cancer stem cell phenotype through AKT/ERK1/2 inactivation by targeting PTEN. PLoS ONE 2012, 7, e39520. [Google Scholar] [CrossRef]
  241. Lai, X.J.; Cheng, H.F. LncRNA colon cancer-associated transcript 1 (CCAT1) promotes proliferation and metastasis of ovarian cancer via miR-1290. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 322–328. [Google Scholar] [CrossRef]
  242. Wang, B.D.; Lee, N.H. Aberrant RNA Splicing in Cancer and Drug Resistance. Cancers 2018, 10, 458. [Google Scholar] [CrossRef]
  243. Berman, E.; Jhanwar, S.; Hedvat, C.; Arcila, M.E.; Wahab, O.A.; Levine, R.; Maloy, M.; Ma, W.; Albitar, M. Resistance to imatinib in patients with chronic myelogenous leukemia and the splice variant BCR-ABL1(35INS). Leuk. Res. 2016, 49, 108–112. [Google Scholar] [CrossRef]
  244. Li, Z.; Li, Q.; Han, L.; Tian, N.; Liang, Q.; Li, Y.; Zhao, X.; Du, C.; Tian, Y. Pro-apoptotic effects of splice-switching oligonucleotides targeting Bcl-x pre-mRNA in human glioma cell lines. Oncol. Rep. 2016, 35, 1013–1019. [Google Scholar] [CrossRef] [PubMed]
  245. Meyer, S.; Stevens, A.; Paredes, R.; Schneider, M.; Walker, M.J.; Williamson, A.J.K.; Gonzalez-Sanchez, M.B.; Smetsers, S.; Dalal, V.; Teng, H.Y.; et al. Acquired cross-linker resistance associated with a novel spliced BRCA2 protein variant for molecular phenotyping of BRCA2 disruption. Cell Death Dis. 2017, 8, e2875. [Google Scholar] [CrossRef]
  246. Hill, S.F.; Meisler, M.H. Antisense Oligonucleotide Therapy for Neurodevelopmental Disorders. Dev. Neurosci. 2021, 43, 247–252. [Google Scholar] [CrossRef] [PubMed]
  247. Lauffer, M.C.; van Roon-Mom, W.; Aartsma-Rus, A.; N =1 Collaborative. Possibilities and limitations of antisense oligonucleotide therapies for the treatment of monogenic disorders. Commun. Med. 2024, 4, 6. [Google Scholar] [CrossRef]
  248. Hu, B.; Zhong, L.; Weng, Y.; Peng, L.; Huang, Y.; Zhao, Y.; Liang, X.-J. Therapeutic siRNA: State of the art. Signal Transduct. Target. Ther. 2020, 5, 101. [Google Scholar] [CrossRef] [PubMed]
  249. Crooke, S.T.; Witztum, J.L.; Bennett, C.F.; Baker, B.F. RNA-Targeted Therapeutics. Cell Metab. 2018, 27, 714–739. [Google Scholar] [CrossRef]
  250. Khvorova, A.; Watts, J.K. The chemical evolution of oligonucleotide therapies of clinical utility. Nat. Biotechnol. 2017, 35, 238–248. [Google Scholar] [CrossRef]
  251. Bennett, C.F.; Krainer, A.R.; Cleveland, D.W. Antisense Oligonucleotide Therapies for Neurodegenerative Diseases. Annu. Rev. Neurosci. 2019, 42, 385–406. [Google Scholar] [CrossRef] [PubMed]
  252. Hoy, S.M. Nusinersen: First Global Approval. Drugs 2017, 77, 473–479. [Google Scholar] [CrossRef]
  253. Torres-Masjoan, L.; Aguti, S.; Zhou, H.; Muntoni, F. Clinical applications of exon-skipping antisense oligonucleotides in neuromuscular diseases. Mol. Ther. 2025, 33, 2689–2704. [Google Scholar] [CrossRef]
  254. Saad, F.A.; Siciliano, G.; Angelini, C. Advances in Dystrophinopathy Diagnosis and Therapy. Biomolecules 2023, 13, 1319. [Google Scholar] [CrossRef]
  255. Dhillon, S. Risdiplam: First Approval. Drugs 2020, 80, 1853–1858. [Google Scholar] [CrossRef]
  256. Paik, J. Risdiplam: A Review in Spinal Muscular Atrophy. CNS Drugs 2022, 36, 401–410. [Google Scholar] [CrossRef]
  257. Macaluso, M.; Paggi, M.G.; Giordano, A. Genetic and epigenetic alterations as hallmarks of the intricate road to cancer. Oncogene 2003, 22, 6472–6478. [Google Scholar] [CrossRef] [PubMed]
  258. Caiado, F.; Silva-Santos, B.; Norell, H. Intra-tumour heterogeneity—Going beyond genetics. FEBS J. 2016, 283, 2245–2258. [Google Scholar] [CrossRef]
  259. Keam, S.J. Imetelstat: First Approval. Drugs 2024, 84, 1149–1155. [Google Scholar] [CrossRef]
  260. U.S. Food and Drug Administration. FDA Approves Imetelstat for Low- to Intermediate-1 Risk Myelodysplastic Syndromes with Transfusion-Dependent Anemia. Available online: https://www.fda.gov/drugs/resources-information-approved-drugs/fda-approves-imetelstat-low-intermediate-1-risk-myelodysplastic-syndromes-transfusion-dependent (accessed on 15 September 2025).
  261. Khuu, A.; Verreault, M.; Colin, P.; Tran, H.; Idbaih, A. Clinical Applications of Antisense Oligonucleotides in Cancer: A Focus on Glioblastoma. Cells 2024, 13, 1869. [Google Scholar] [CrossRef] [PubMed]
  262. Narasipura, E.A.; VanKeulen-Miller, R.; Ma, Y.; Fenton, O.S. Ongoing Clinical Trials of Nonviral siRNA Therapeutics. Bioconjugate Chem. 2023, 34, 1177–1197. [Google Scholar] [CrossRef]
  263. Zou, S.; Tong, Q.; Liu, B.; Huang, W.; Tian, Y.; Fu, X. Targeting STAT3 in Cancer Immunotherapy. Mol. Cancer 2020, 19, 145. [Google Scholar] [CrossRef]
  264. National Library of Medicine. Danvatirsen and Durvalumab in Treating Patients with Advanced and Refractory Pancreatic, Non-Small Cell Lung Cancer, and Mismatch Repair Deficient Colorectal Cancer. Available online: https://www.clinicaltrials.gov/study/NCT02983578#study-record-dates (accessed on 15 September 2025).
  265. National Library of Medicine. A Clinical Trial of BP1002 in Patients with Advanced Lymphoid Malignancies. Available online: https://clinicaltrials.gov/study/NCT04072458?term=antisense%20oligonucleotides,%20cancer&rank=1 (accessed on 15 September 2025).
  266. National Library of Medicine. BP1001-A in Patients with Advanced or Recurrent Solid Tumors. Available online: https://clinicaltrials.gov/study/NCT04196257?term=antisense%20oligonucleotides,%20cancer&rank=5 (accessed on 15 September 2025).
  267. National Library of Medicine. A Phase I First in Human Study to Evaluate the Safety, Tolerability, and Pharmacokinetics of WGI-0301 in Patients with Advanced Solid Tumors. Available online: https://clinicaltrials.gov/study/NCT05267899?term=antisense%20oligonucleotides,%20cancer&limit=100&sort=StudyFirstPostDate&rank=4 (accessed on 15 September 2025).
  268. National Library of Medicine. A Study of OT-101 with mFOLFIRINOX in Patients with Advanced and Unresectable or Metastatic Pancreatic Cancer (STOP-PC). Available online: https://clinicaltrials.gov/study/NCT06079346?term=antisense%20oligonucleotides,%20cancer&limit=100&sort=StudyFirstPostDate&rank=2 (accessed on 15 September 2025).
  269. Okuyama, T.; Sakamoto, R.; Kumagai, K.; Nishizawa, M.; Kimura, T.; Sugie, T.; Kimura, T. EPHA2 antisense RNA modulates EPHA2 mRNA levels in basal-like/triple-negative breast cancer cells. Biochimie 2020, 179, 169–180. [Google Scholar] [CrossRef]
  270. National Library of Medicine. EphA2 siRNA in Treating Patients with Advanced or Recurrent Solid Tumors. Available online: https://clinicaltrials.gov/study/NCT01591356?term=sirna,%20cancer&rank=28 (accessed on 15 September 2025).
  271. National Library of Medicine. Targeting NUDT21 siRNA Drugs for Patients with Refractory Retinoblastoma. Available online: https://clinicaltrials.gov/study/NCT06424301?term=sirna,%20cancer&rank=4 (accessed on 15 September 2025).
  272. Francis, J.H.; Marr, B.P.; Brodie, S.E.; Abramson, D.H. Anterior Ocular Toxicity of Intravitreous Melphalan for Retinoblastoma. JAMA Ophthalmol. 2015, 133, 1459–1463. [Google Scholar] [CrossRef] [PubMed]
  273. National Library of Medicine. Open Label, Dose Escalation Study for the Safety and Efficacy of STP705 in Adult Patients with Basal Cell Carcinoma. Available online: https://clinicaltrials.gov/study/NCT04669808?term=NCT04669808&rank=1 (accessed on 15 September 2025).
  274. HuidaGene Therapeutics. HuidaGene Therapeutics Announced First Patient Dosed in the HERO Clinical Trial of HG204 for MECP2 Duplication Syndrome. Available online: https://www.huidagene.com/new/news/71.html (accessed on 15 September 2025).
  275. HuidaGene Therapeutics. HuidaGene Therapeutics Receives the First-Ever FDA Clearance of CRISPR/Cas13 RNA-Editing HG202 for Macular Degeneration. Available online: https://www.huidagene.com/new/news/70 (accessed on 15 September 2025).
  276. Wang, Q.; Liu, X.; Zhou, J.; Yang, C.; Wang, G.; Tan, Y.; Wu, Y.; Zhang, S.; Yi, K.; Kang, C. The CRISPR-Cas13a Gene-Editing System Induces Collateral Cleavage of RNA in Glioma Cells. Adv. Sci. 2019, 6, 1901299. [Google Scholar] [CrossRef]
  277. Fan, J.; Liu, Y.; Liu, L.; Huang, Y.; Li, X.; Huang, W. A Multifunction Lipid-Based CRISPR-Cas13a Genetic Circuit Delivery System for Bladder Cancer Gene Therapy. ACS Synth. Biol. 2020, 9, 343–355. [Google Scholar] [CrossRef] [PubMed]
  278. Fan, N.; Bian, X.; Li, M.; Chen, J.; Wu, H.; Peng, Q.; Bai, H.; Cheng, W.; Kong, L.; Ding, S.; et al. Hierarchical self-uncloaking CRISPR-Cas13a–customized RNA nanococoons for spatial-controlled genome editing and precise cancer therapy. Sci. Adv. 2022, 8, eabn7382. [Google Scholar] [CrossRef]
  279. Steensma, D.P.; Wermke, M.; Klimek, V.M.; Greenberg, P.L.; Font, P.; Komrokji, R.S.; Yang, J.; Brunner, A.M.; Carraway, H.E.; Ades, L.; et al. Phase I First-in-Human Dose Escalation Study of the oral SF3B1 modulator H3B-8800 in myeloid neoplasms. Leukemia 2021, 35, 3542–3550. [Google Scholar] [CrossRef] [PubMed]
  280. Hong, D.S.; Kurzrock, R.; Naing, A.; Wheler, J.J.; Falchook, G.S.; Schiffman, J.S.; Faulkner, N.; Pilat, M.J.; O’Brien, J.; LoRusso, P. A phase I, open-label, single-arm, dose-escalation study of E7107, a precursor messenger ribonucleic acid (pre-mRNA) splicesome inhibitor administered intravenously on days 1 and 8 every 21 days to patients with solid tumors. Investig. New Drugs 2014, 32, 436–444. [Google Scholar] [CrossRef]
  281. Seiler, M.; Yoshimi, A.; Darman, R.; Chan, B.; Keaney, G.; Thomas, M.; Agrawal, A.A.; Caleb, B.; Csibi, A.; Sean, E.; et al. H3B-8800, an orally available small-molecule splicing modulator, induces lethality in spliceosome-mutant cancers. Nat. Med. 2018, 24, 497–504. [Google Scholar] [CrossRef]
  282. National Library of Medicine. A Study of JNJ-64619178, an Inhibitor of PRMT5 in Participants with Advanced Solid Tumors, NHL, and Lower Risk MDS. Available online: https://clinicaltrials.gov/study/NCT03573310 (accessed on 15 September 2025).
  283. Ouyang, W.; Yu, Z.; Zhao, X.; Lu, S.; Wang, Z. Aptamers in hematological malignancies and their potential therapeutic implications. Crit. Rev. Oncol. Hematol. 2016, 106, 108–117. [Google Scholar] [CrossRef] [PubMed]
  284. National Library of Medicine. Glioblastoma Treatment with Irradiation and Olaptesed Pegol (NOX-A12) in MGMT Unmethylated Patients (GLORIA). Available online: https://clinicaltrials.gov/study/NCT04121455 (accessed on 15 September 2025).
  285. National Library of Medicine. A Study in Subjects with Otoferlin Mutation-Related Hearing Loss Using RNA Base-eDiting Therapy (SOUND). Available online: https://clinicaltrials.gov/study/NCT06025032?term=NCT06025032&rank=1 (accessed on 15 September 2025).
  286. Ezike, T.C.; Okpala, U.S.; Onoja, U.L.; Nwike, C.P.; Ezeako, E.C.; Okpara, O.J.; Okoroafor, C.C.; Eze, S.C.; Kalu, O.L.; Odoh, E.C.; et al. Advances in drug delivery systems, challenges and future directions. Heliyon 2023, 9, e17488. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Key milestones in the development and application of RNA-targeting approaches. This figure summarizes major scientific and clinical milestones in the advancement of RNA-targeting technologies, spanning from the introduction of antisense oligonucleotides (ASOs) in 1978 to the initiation of the first clinical trial involving CRISPR-Cas13 in 2025. It presents significant discoveries in RNA biology and key regulatory approvals that underscore the growing potential of RNA-targeting strategies in cancer therapy.
Figure 1. Key milestones in the development and application of RNA-targeting approaches. This figure summarizes major scientific and clinical milestones in the advancement of RNA-targeting technologies, spanning from the introduction of antisense oligonucleotides (ASOs) in 1978 to the initiation of the first clinical trial involving CRISPR-Cas13 in 2025. It presents significant discoveries in RNA biology and key regulatory approvals that underscore the growing potential of RNA-targeting strategies in cancer therapy.
Genes 16 01168 g001
Figure 2. Applications for RNA-targeting technologies. A schematic diagram summarizing the major categories of RNA-targeting strategies and their respective applications. These include RNA degraders (e.g., siRNA/shRNA, ASOs, Cas13d) for transcript knockdown, RNA trackers and detectors for RNA detection or localization, RNA base editors for modulating RNA base modifications, splicing regulators for modulating alternative splicing events, translation regulators to control protein synthesis, and non-coding RNA (ncRNA) inhibitors targeting oncogenic or regulatory RNAs.
Figure 2. Applications for RNA-targeting technologies. A schematic diagram summarizing the major categories of RNA-targeting strategies and their respective applications. These include RNA degraders (e.g., siRNA/shRNA, ASOs, Cas13d) for transcript knockdown, RNA trackers and detectors for RNA detection or localization, RNA base editors for modulating RNA base modifications, splicing regulators for modulating alternative splicing events, translation regulators to control protein synthesis, and non-coding RNA (ncRNA) inhibitors targeting oncogenic or regulatory RNAs.
Genes 16 01168 g002
Figure 3. Mechanisms of RNA-targeting systems for RNA manipulation. This figure illustrates the molecular mechanisms of major RNA-targeting platforms used in cancer research and therapeutics. It includes antisense oligonucleotides (ASOs) subclasses, which act through RNase H1-mediated degradation or modulation of pre-mRNA splicing and protein interactions. RNA interference (RNAi) mechanisms via siRNA (1A→5) and shRNA (1B→5) are depicted, highlighting Dicer processing, RISC incorporation, and mRNA cleavage. CRISPR-Cas13 and catalytically inactive dCas13 systems are shown as programmable tools for targeted RNA cleavage, transcript editing, or RNA localization.
Figure 3. Mechanisms of RNA-targeting systems for RNA manipulation. This figure illustrates the molecular mechanisms of major RNA-targeting platforms used in cancer research and therapeutics. It includes antisense oligonucleotides (ASOs) subclasses, which act through RNase H1-mediated degradation or modulation of pre-mRNA splicing and protein interactions. RNA interference (RNAi) mechanisms via siRNA (1A→5) and shRNA (1B→5) are depicted, highlighting Dicer processing, RISC incorporation, and mRNA cleavage. CRISPR-Cas13 and catalytically inactive dCas13 systems are shown as programmable tools for targeted RNA cleavage, transcript editing, or RNA localization.
Genes 16 01168 g003
Figure 4. Small molecules and aptamers as tools for RNA targeting. (A) Schematic representation of the primary mechanisms by which small molecules target RNA. These compounds can modulate pre-mRNA splicing by inhibiting or stabilizing interactions with the splicing machinery, inhibit translation by binding to internal ribosome entry sites (IRES) and blocking ribosome binding, correct structural abnormalities (like expansions of CUG repeats) in mRNAs, and interfere with pre-miRNA disrupting its biogenesis. (B) Overview of the Systematic Evolution of Ligands by EXponential enrichment (SELEX) process, illustrating the iterative selection of aptamers with high affinity and specificity for a given RNA target.
Figure 4. Small molecules and aptamers as tools for RNA targeting. (A) Schematic representation of the primary mechanisms by which small molecules target RNA. These compounds can modulate pre-mRNA splicing by inhibiting or stabilizing interactions with the splicing machinery, inhibit translation by binding to internal ribosome entry sites (IRES) and blocking ribosome binding, correct structural abnormalities (like expansions of CUG repeats) in mRNAs, and interfere with pre-miRNA disrupting its biogenesis. (B) Overview of the Systematic Evolution of Ligands by EXponential enrichment (SELEX) process, illustrating the iterative selection of aptamers with high affinity and specificity for a given RNA target.
Genes 16 01168 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Boti, M.A.; Diamantopoulos, M.A.; Scorilas, A. RNA-Targeting Techniques: A Comparative Analysis of Modern Approaches for RNA Manipulation in Cancer Research and Therapeutics. Genes 2025, 16, 1168. https://doi.org/10.3390/genes16101168

AMA Style

Boti MA, Diamantopoulos MA, Scorilas A. RNA-Targeting Techniques: A Comparative Analysis of Modern Approaches for RNA Manipulation in Cancer Research and Therapeutics. Genes. 2025; 16(10):1168. https://doi.org/10.3390/genes16101168

Chicago/Turabian Style

Boti, Michaela A., Marios A. Diamantopoulos, and Andreas Scorilas. 2025. "RNA-Targeting Techniques: A Comparative Analysis of Modern Approaches for RNA Manipulation in Cancer Research and Therapeutics" Genes 16, no. 10: 1168. https://doi.org/10.3390/genes16101168

APA Style

Boti, M. A., Diamantopoulos, M. A., & Scorilas, A. (2025). RNA-Targeting Techniques: A Comparative Analysis of Modern Approaches for RNA Manipulation in Cancer Research and Therapeutics. Genes, 16(10), 1168. https://doi.org/10.3390/genes16101168

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop