Fact-checked by Grok 2 weeks ago

RNA editing

RNA editing is a co- or post-transcriptional modification process in which the nucleotide sequence of an RNA molecule is altered from that encoded by the corresponding DNA template, resulting in changes to the RNA's base composition and potential functional outcomes. The two predominant types of RNA editing in eukaryotes are adenosine-to-inosine (A-to-I) editing, which converts adenosine (A) to inosine (I) via hydrolytic deamination and is recognized as guanosine (G) during translation, and cytidine-to-uridine (C-to-U) editing, which deaminates cytosine (C) to uridine (U). These modifications are catalyzed by specific enzyme families: the adenosine deaminase acting on RNA (ADAR) proteins and ADATs for A-to-I events, primarily on double-stranded RNA or transfer RNA (tRNA), and the apolipoprotein B mRNA editing enzyme catalytic polypeptide-like (APOBEC) family for C-to-U events, often on single-stranded RNA substrates. A-to-I editing is the most widespread form in metazoans, with enzymes (including ADAR1, ADAR2, and the catalytically inactive ADAR3) targeting structured regions to generate transcriptomic diversity, recode in proteins, and modulate stability or splicing. For instance, ADAR1 plays a in innate immunity by editing self double-stranded RNAs to prevent their recognition by sensors like , thus averting autoimmune responses, while ADAR2 facilitates essential recoding events in neuronal transcripts such as the glutamate receptor subunit GRIA2, changing a to codon for proper channel function. In contrast, C-to-U editing is less prevalent but vital in specific contexts, such as APOBEC1-mediated editing of apolipoprotein B mRNA in the intestine to produce a truncated , or APOBEC3 family members' roles in antiviral defense by hypermutating viral genomes. Beyond these core mechanisms, RNA editing expands the without altering the , influences evolutionary adaptation through conserved recoding sites under positive selection, and has implications for when dysregulated, including neurological disorders, cancer, and immune pathologies. Recent advances in sequencing and bioinformatics have revealed millions of editing sites across transcriptomes, predominantly in non-coding regions like Alu elements, highlighting its role in fine-tuning gene regulation and inspiring therapeutic applications, such as site-directed editing tools for correcting pathogenic mutations.

Definition and Biological Significance

Definition of RNA Editing

RNA editing refers to a set of enzymatic processes that modify the nucleotide sequence of primary RNA transcripts after transcription, resulting in RNA molecules that encode proteins or perform functions distinct from those directly specified by the genomic DNA. These modifications can introduce changes that expand transcriptomic and proteomic diversity without altering the underlying DNA sequence. Unlike standard RNA processing events such as splicing, capping, or polyadenylation, RNA editing specifically involves targeted alterations to the RNA sequence itself. The phenomenon was first discovered in 1986 in the mitochondria of trypanosomes, where extensive insertion of (U) residues into mitochondrial mRNAs was found to be necessary for proper coding of proteins, resolving apparent discrepancies between sequences and translated products. This unexpected finding challenged the and introduced the concept of as a malleable information carrier. In the late , editing was expanded to mammalian systems with the identification of cytidine-to- (C-to-U) editing in (apoB) mRNA in the intestine, which generates a truncated essential for . Adenosine-to-inosine (A-to-I) editing in mammals was subsequently recognized in the early through studies of brain transcripts, particularly in mRNAs like those encoding subunits, where it modulates neuronal excitability. The scope of RNA editing encompasses several types of sequence alterations, primarily base substitutions such as A-to-I and C-to-U, as well as insertion and deletion of nucleotides, particularly U residues in kinetoplastid protists. These processes occur across diverse RNA species, including mRNAs, non-coding RNAs, and viral RNAs, but are distinct from routine posttranscriptional modifications that do not change the sequence information. Major enzymes driving RNA editing include the acting on RNA () family, which catalyzes A-to-I substitutions by deaminating to in double-stranded RNA regions, and the mRNA editing enzyme catalytic polypeptide-like () family, particularly APOBEC1, which performs C-to-U editing via deamination. These enzymes are highly conserved and play pivotal roles in regulating across eukaryotes.

Role in Gene Expression and Diversity

RNA editing plays a pivotal role in expanding proteomic diversity by enabling post-transcriptional modifications that alter the coding potential of transcripts beyond the constraints of the genomic sequence. Through recoding events, such as codon changes that substitute in proteins, RNA editing generates protein isoforms with distinct functions, thereby increasing the complexity of the without requiring genomic mutations. Additionally, editing influences gene expression by regulating alternative splicing, mRNA stability, and translational efficiency; for instance, adenosine-to-inosine (A-to-I) modifications can introduce or remove splice sites, leading to diverse transcript variants, while edited sequences may affect miRNA binding and mRNA decay rates. In humans, these processes contribute to editing in approximately 1-3% of transcripts, primarily through A-to-I changes mediated by enzymes. Biologically, RNA editing exhibits tissue-specific patterns that fine-tune cellular functions, particularly in dynamic environments like the . In the , elevated editing levels support by modulating neurotransmitter receptor properties and neuronal excitability, allowing adaptive responses to neural activity. also facilitates rapid to environmental es, such as oxidative or hypoxic conditions, by altering transcripts involved in metabolic pathways or response genes, thereby enhancing cellular without permanent genetic alterations. In organelles like plant mitochondria and chloroplasts, editing serves as an error-correction mechanism, converting mismatched (e.g., C-to-U) to restore functional protein sequences that would otherwise be defective due to mutational biases in organellar genomes. A prominent example is the Q/R site editing in the glutamate receptor subunit GluA2 (GRIA2), where A-to-I deamination changes a glutamine codon to arginine, rendering AMPA receptors impermeable to calcium ions and preventing excitotoxicity in neurons. This nearly complete editing (>99%) in mature neurons exemplifies how precise modifications safeguard synaptic function. Genome-wide, ADAR-mediated A-to-I editing targets millions of sites in humans, with over 100 million potential sites in Alu repetitive elements alone, with editing occurring in nearly all adenosines within these editable Alu repeats, though often at low levels (<1%), and affecting a majority of genes containing such elements (67.4% of RefSeq genes). From an evolutionary perspective, RNA editing confers adaptive advantages by providing a flexible layer of plasticity, enabling organisms to respond to changing environments or developmental cues without the fixation of potentially deleterious in the . This is particularly evident in species facing variable conditions, where edited transcripts support physiological acclimation and increased at the RNA level.

Detection Methods

Next-Generation Sequencing

Next-generation sequencing (NGS) enables the genome-wide identification and quantification of RNA editing sites by aligning reads to a or matched genomic DNA sequence, detecting mismatches that indicate editing events. For A-to-I editing, is reverse-transcribed and sequenced as , appearing as A-to-G mismatches, while C-to-U editing manifests as C-to-T mismatches. This comparative approach distinguishes editing from genomic variants by requiring RNA-DNA discordance in the same sample. However, C-to-U editing detection faces additional challenges due to its lower prevalence and potential overlap with common SNPs, often requiring higher stringency filters. Standard protocols involve performing on poly(A)-selected or total libraries with deep sequencing coverage, typically exceeding 100x per site to ensure sufficient read depth for reliable variant calling and to minimize stochastic errors. Specialized bioinformatics tools, such as REDItools and GIREMI, aligned BAM files to annotate sites, apply filters for , mapping quality, and allelic fraction thresholds (e.g., >0.01 for minor alleles), and output editing events while excluding repetitive regions or low-complexity sequences. These tools integrate statistical tests, like in GIREMI, to prioritize true positives without requiring multiple replicates initially. NGS offers high-throughput , allowing cost-effective discovery of thousands to millions of editing sites across transcriptomes in a single experiment, and supports single-cell resolution through scRNA-seq to reveal cell-type-specific patterns. However, challenges include distinguishing true from single nucleotide polymorphisms (SNPs), sequencing artifacts, or errors, particularly in polymorphic or repetitive genomic regions. Solutions involve using matched DNA-seq for validation, biological replicates to assess reproducibility, and statistical models—such as those estimating false discovery rates or employing generalized linear models—to correct for biases and filter false positives. The widespread adoption of NGS for RNA editing detection accelerated post-2010, driven by Illumina platforms that enabled scalable RNA-seq experiments, leading to the identification of over 2 million human A-to-I sites by 2013 through refined pipelines. NGS is often complemented by mass spectrometry for orthogonal validation.

Mass Spectrometry

Mass spectrometry (MS) serves as a powerful orthogonal method for validating RNA editing events by directly analyzing either modified nucleosides in RNA or the resulting altered amino acids in proteins derived from edited transcripts. At the nucleotide level, the principle involves enzymatic digestion of RNA into individual nucleosides, followed by liquid chromatography coupled with tandem mass spectrometry (LC-MS/MS), where inosine (arising from A-to-I editing) is identified and quantified based on its unique mass-to-charge ratio (m/z) and chromatographic retention time relative to standard curves. This approach distinguishes inosine from adenosine, enabling precise measurement of editing efficiency in specific RNA populations, such as mRNA or viral RNA. At the protein level, MS targets tryptic peptides from proteomes, detecting recoding events like the Q/R site in AMPA receptor subunit GluA2, where A-to-I editing converts a glutamine codon to arginine, producing a detectable mass shift in the peptide spectrum. Key techniques include nanoLC-MS/MS for proteome-wide scans, which employs high-resolution instruments like the to sequence peptides from complex samples, often using tandem mass tag (TMT) labeling for multiplexed quantification across multiple samples. For inosine-specific quantification, methods like nucleic acid isotope labeling coupled mass spectrometry (NAIL-MS) incorporate stable isotopes during to track modification dynamics, allowing differentiation of newly synthesized versus pre-existing edited RNAs with . These approaches have been refined post-2015 for higher throughput; for instance, custom proteogenomic databases integrate RNA editing variants to match MS spectra, identifying edited peptides in tissue from over 170 subjects. Similarly, advanced LC-MS/MS protocols with chemical derivatization have enabled detection of low-abundance inosine in single cells or mRNA. The primary advantages of MS-based validation lie in its ability to confirm the functional consequences of RNA editing, such as substitutions that alter protein properties (e.g., calcium permeability in GluA2-edited receptors), providing direct evidence beyond nucleotide-level observations. It complements next-generation sequencing by verifying editing at the translational output, with high specificity for modified species. However, limitations include low sensitivity for rare editing events, often necessitating enrichment strategies like of specific RNAs or peptides, as seen in the sparse detection of only 10 out of 294 predicted recoding events in proteomes. Sample preparation challenges, such as RNA instability and ion suppression in complex matrices, further require rigorous controls. Post-2015 multiplexing innovations, including TMT and improved , have enhanced scalability for validating editing impacts in aging or disease contexts, such as Alzheimer's tissue.

Emerging Techniques

Recent advances in RNA editing detection have introduced chemically-assisted sequencing methods that enable enzyme-free detection of A-to-I edits through bisulfite-like assays, which chemically modify bases to distinguish them from without relying on enzymatic treatments that can introduce biases. These approaches, detailed in a 2025 review, improve specificity for non-canonical edits by leveraging reactivity differences, allowing for more accurate mapping in complex transcriptomes. Building on foundations from next-generation sequencing and , such methods address longstanding gaps in sensitivity for low-frequency events. Nanopore direct sequencing has emerged as a key tool for real-time detection of inosine reads, preserving native structure to identify A-to-I sites without reverse transcription artifacts. The Dinopore algorithm, developed for platforms, uses signal patterns to pinpoint with high precision across organisms, facilitating the study of dynamics in . A 2025 analysis highlights how these advancements in direct enhance the detection of modifications, including events, by integrating to classify base-level signals. In 2024, developments in AI-driven tools have advanced site annotation for RNA editing, with models like those using fine-tuned architectures predicting edit sites by analyzing sequence context and secondary structure features. These tools, building on convolutional neural networks such as EditPredict, achieve superior annotation accuracy for potential editing hotspots, aiding in the prioritization of candidates from large datasets. Such predictive frameworks are particularly valuable for identifying non-model organism edits where experimental validation is resource-intensive. Complementing this, 2025 multi-omics integration efforts have enhanced understanding of RNA editing by combining with epigenomic and proteomic data, revealing roles in tumor evolution. These emerging techniques offer substantial benefits, including enhanced accuracy for detecting low-abundance edits in cancer and brain tissues, where traditional methods often struggle with noise. By minimizing artifacts, they have been shown to substantially reduce false positives, enabling more reliable profiling of editing landscapes in disease states.

Core Mechanisms of RNA Editing

Deamination-Based Substitution

Deamination-based substitution represents the predominant mechanism of RNA editing in metazoans, involving the enzymatic removal of an amino group from specific nucleotide bases, which alters their base-pairing properties and can change codon meaning during translation. This process primarily targets cytidine (C) or adenosine (A) residues, converting them to uridine (U) or inosine (I), respectively, through hydrolytic deamination catalyzed by members of the metazoan-specific APOBEC family for C-to-U editing and the ADAR family for A-to-I editing. These modifications expand the proteome diversity without altering the genomic sequence, influencing gene expression, protein function, and cellular responses. The core reaction proceeds via , where water acts as the to deaminate the : for example, is converted to plus ( + H₂O → + NH₃). This enzymatic activity requires double-stranded (dsRNA) substrates, as the editing enzymes recognize structured regions to ensure specificity and avoid off-target effects. enzymes, in particular, feature double-stranded -binding domains (dsRBDs) that facilitate substrate recognition, followed by a catalytic deaminase domain that executes the through a two-step mechanism involving a tetrahedral . enzymes similarly rely on structure for positioning but often require auxiliary factors for efficient activity . Deamination-based editing accounts for the vast majority of known RNA editing events in eukaryotic transcriptomes, far outnumbering other or modification types. efficiency and site selection are tightly regulated by factors such as the enzyme's subcellular localization—ADARs are primarily , though ADAR1 has a cytoplasmic isoform, while certain APOBECs exhibit variable or cytoplasmic distribution—and the local RNA secondary structure, which can enhance or inhibit access to target sites. These regulatory elements ensure that editing occurs in a controlled manner, often in response to cellular stress or developmental cues. Specific instances of C-to-U and A-to-I editing, such as those mediated by APOBEC1 in mRNA or ADAR2 in transcripts, exemplify this mechanism's role in fine-tuning protein isoforms.

Insertion and Deletion Editing

Insertion and deletion editing represents a form of RNA modification that alters the length of the transcript by adding or removing (U) residues, in contrast to substitution that changes individual bases without affecting length. This process is primarily observed in the mitochondria of kinetoplastid protists, such as trypanosomes, where it is directed by guide RNAs (gRNAs). The mechanism involves the formation of a chimeric duplex between the pre-mRNA and a gRNA, which base-pairs with the target mRNA to specify editing sites through complementary regions. An endonuclease cleaves the mRNA at mismatched sites within this duplex, creating 5' and 3' fragments. For U insertion, the RET1 enzyme (a 3' terminal uridylyl , or TUTase) adds U residues to the 3' end of the upstream fragment in a non-templated manner, with the number of Us determined by the gRNA anchor sequence to restore complementarity; the fragments are then joined by REL2. In U deletion, mismatched Us are removed from the 3' fragment by the 3'-5' activity associated with the REH1 , followed by re-ligation via REL1. This cycle progresses in a 3' to 5' direction along the mRNA, often requiring multiple gRNAs for extensive editing, and can involve the addition of poly-U tails during insertion steps that are subsequently trimmed to match the gRNA template. Up to over 200 Us may be inserted in a single transcript, such as in the subunit III (COIII) mRNA of , to generate functional open reading frames from cryptic pre-mRNAs. This type of editing is rare among eukaryotes and is essential for the expression of most mitochondrial genes in trypanosomes, where it corrects frameshifts and restores translatable sequences in up to 12 maxicircle-encoded mRNAs. No direct homologs of the editing machinery exist in mammals or other higher eukaryotes. Exceptions include limited U insertion events reported in mitochondria, which primarily feature substitutional editing but occasionally add internal Us to refine transcripts. Additionally, some cases of physical editing occur via trans-splicing in certain protists, where RNA fragments are joined to effectively insert sequences, though this differs from the gRNA-templated U modifications in kinetoplastids.00468-0)

Deamination-Based Editing

C-to-U Editing

C-to-U RNA editing is a in which residues in transcripts are deaminated to , altering the without changing the underlying DNA sequence. This process occurs predominantly in nuclear-encoded messenger RNAs in animals and in organellar transcripts in , contributing to protein diversity and functional optimization. The reaction is catalyzed by cytidine deaminases and is highly site-specific, often guided by auxiliary factors that recognize particular RNA motifs. In mammalian systems, the primary responsible for C-to-U editing is APOBEC1, a cytoplasmic cytidine deaminase that requires the cofactor APOBEC1 complementation factor (ACF) to form an active editing complex. APOBEC1 deaminates specific s in target RNAs, with editing efficiency reaching over 90% in tissues like the . Site specificity is achieved through an 11-nucleotide mooring located 3-5 nucleotides downstream of the edited , typically featuring a core of 5'-UUUN-3' that positions the correctly. A canonical example is the editing of (APOB) mRNA, where of 6666 (C6666) to creates a premature (UAA), truncating the full-length ApoB100 protein into the shorter ApoB48 isoform essential for lipid transport in the intestine. In plant mitochondria and plastids, C-to-U editing is mediated by pentatricopeptide repeat (PPR) proteins of the PLS subclass, which contain a C-terminal DYW deaminase domain responsible for the deamination activity. These DYW-type enzymes recognize cis-elements in organellar RNAs via the PPR array and catalyze at hundreds of sites, often restoring evolutionarily conserved or creating functional codons. For instance, in plastids, editing converts an ACG initiation codon to AUG in the psbL mRNA of , enabling translation of the subunit PsbL. In mitochondria, similar editing events optimize codons for protein function, with over 400 sites identified in species like . C-to-U editing is not typically observed in animal mitochondria, including humans, unlike the extensive editing in plant mitochondria and plastids. Recent research (as of 2025) has revealed cancer-specific hyper- patterns driven by family members, such as elevated C-to-U modifications in hematologic malignancies that promote tumor progression through altered protein isoforms and increased mutational load. As of 2025, has been shown to catalyze site-specific C-to-U editing of transfer RNAs (tRNAs), with implications for cancer and immune responses.

A-to-I Editing

A-to-I RNA editing, the conversion of (A) to (I) in , is the most prevalent form of editing in animals and is catalyzed by the acting on (ADAR) family of enzymes. These enzymes include ADAR1, ADAR2, and ADAR3, each with distinct expression patterns and functions. ADAR1 exists in two isoforms: the constitutively expressed nuclear p110 form and the interferon-inducible cytoplasmic p150 form, which is upregulated during immune responses. ADAR2 is primarily nuclear and responsible for constitutive editing events in specific transcripts. In contrast, ADAR3 lacks deaminase activity and acts as an inhibitor of editing by competing for double-stranded (dsRNA) substrates. The editing process involves hydrolytic deamination of within dsRNA structures, where is subsequently recognized as (G) during and splicing, potentially altering codon meaning or RNA stability. For instance, in messenger RNAs (mRNAs), this can lead to amino acid recoding; a prominent example is the Q/R site in the GRIA2 transcript encoding the subunit GluA2, where editing changes CAG () to CIG (read as CGG, ), modulating calcium permeability in neuronal synapses. Humans harbor over 100 confirmed recoding sites in mRNAs, many affecting receptors and channels critical for . In non-coding RNAs, particularly those with Alu repetitive elements, ADARs perform hyper-editing, extensively modifying adenosines to stabilize structures or prevent immune sensing of dsRNA. Functionally, A-to-I editing diversifies the proteome and regulates innate immunity; ADAR1 p150 edits viral dsRNA to evade interferon responses, thereby inhibiting excessive inflammation. In cancer, recent transcriptome-wide mapping has revealed dysregulated editing patterns, such as elevated ADAR1 activity promoting tumor immune evasion and progression in hematologic malignancies, highlighting therapeutic potential through ADAR modulation. As of 2024, advances show A-to-I editing's role in oncogenesis, including non-synonymous mutations in tumor progression.

Editing in Specific RNA Types

Messenger RNA Editing

Messenger RNA (mRNA) editing, predominantly mediated by deamination-based mechanisms such as A-to-I conversion, enables post-transcriptional diversification of the by altering codon sequences in protein-coding transcripts. This process occurs in a majority of protein-coding genes, with editing events being particularly prevalent in the , where they contribute to neuronal and function. For instance, the serotonin receptor 2C (HTR2C) transcript undergoes extensive A-to-I editing at five sites within its , generating up to 24 distinct isoforms that modulate receptor signaling and G-protein coupling efficiency. The primary functions of mRNA editing include recoding to produce protein isoforms with modified properties, such as altered selectivity or in voltage-gated channels. For example, editing in subunits of channels like KCNA1 changes an to , altering recovery from inactivation and fine-tuning neuronal excitability. Additionally, exonic edits can regulate by introducing or disrupting exonic splicing enhancer motifs, thereby influencing isoform production and patterns. A canonical example is the / (Q/R) site in the GluA2 subunit of receptors, where nearly complete A-to-I editing converts a codon to CIG (read as CGG), replacing with in the channel pore and preventing calcium influx, which is essential for synaptic transmission and . In diseases like (), reduced ADAR2 activity leads to incomplete editing at this site, resulting in calcium-permeable receptors that promote degeneration through .

Transfer RNA Editing

Transfer RNA (tRNA) editing primarily involves post-transcriptional modifications that ensure accurate by refining anticodon functionality and wobble base pairing. One key type is C-to-U editing in anticodon loops, as observed in mitochondrial tRNAs where it alters codon ; for instance, in mitochondria, C-to-U editing at position 35 of the anticodon converts a tRNA to decode aspartate codons instead of a mismatched sequence. Another prominent type is A-to-I editing at the wobble position (position 34), which expands codon by allowing to pair with U, C, or A in the third codon position, thereby enabling a single tRNA to decode multiple synonymous codons. This A-to-I process is analogous to deamination-based in but is specialized for tRNA anticodon integrity. The enzymes mediating these edits are highly specific. For A-to-I editing, the ADAT complex—comprising 1 (homodimeric, akin to Tad1), ADAT2, and ADAT3 (forming a heterodimer)—catalyzes at wobble and adjacent positions in various tRNAs, with ADAT2/3 targeting up to eight tRNA species in eukaryotes. These enzymes act independently or in complexes, recognizing tRNA structural motifs without requiring guide RNAs. tRNA editing serves critical functions, including correction of genomic encoding errors in tRNA and facilitation of the for degenerate codon decoding. By introducing at the wobble position, editing compensates for limitations in tRNA gene diversity, allowing efficient of the full codon repertoire without requiring additional tRNA isoacceptors. This process is essential for viability in , where Tad1/2/3 mutants exhibit severe growth defects due to impaired tRNA function, and in mammals, where disruptions lead to translational inefficiencies. Illustrative examples highlight tRNA editing's biological impact. In , the TadA enzyme (a bacterial homolog) performs A-to-I editing at the wobble position of tRNA^Arg(ACG), forming inosine-34 essential for codon decoding and cell viability. In humans, defects in mitochondrial tRNA modification due to mutations in tRNA genes, such as impaired wobble base modification in tRNA^Lys and tRNA^Leu, underlie diseases like myoclonic epilepsy with ragged-red fibers (MERRF) and mitochondrial encephalomyopathy, , and stroke-like episodes (MELAS), where disruptions affect anticodon integrity and . These cases underscore tRNA editing's and modifications' role in preventing translational errors that propagate to cellular dysfunction.

Ribosomal RNA Editing

In eukaryotes, sequence-altering RNA in (rRNA) is rare compared to the abundant post-transcriptional chemical modifications that support and function. While modifications such as pseudouridylation ( of to ) and 2'-O-methylation enhance rRNA stability, folding, and translational efficiency, they do not change the nucleotide sequence and are distinct from . Rare instances of A-to-I have been suggested in rRNA expansion segments, but this is not a primary mechanism and lacks widespread confirmation in cytoplasmic rRNA. Sequence is more prominent in organellar rRNAs, such as C-to-U changes in mitochondria and plastids (covered in the organelle editing section). The enzymes for rRNA modifications, though not editing, are organized into small nucleolar ribonucleoprotein (snoRNP) complexes. Box H/ACA snoRNPs guide pseudouridylation via the dyskerin (Cbf5 in ) synthase, and box C/D snoRNPs direct 2'-O-methylation by fibrillarin (Nop1 in ), using complementary snoRNA sequences for site specificity. Defects in these modification processes are linked to ribosomopathies, such as dyskeratosis congenita from DKC1 mutations impairing pseudouridylation, leading to telomere shortening, failure, and premature aging.

Editing in Organelles and Viruses

Plant Mitochondrial and Plastid Editing

In plant mitochondria and plastids, RNA editing primarily involves cytidine-to-uridine (C-to-U) deamination, serving as a post-transcriptional mechanism to compensate for mutations in organellar genomes. In Arabidopsis thaliana, over 400 such sites have been identified in mitochondrial transcripts, with 456 C-to-U conversions reported exclusively in mRNAs, including 441 within open reading frames (ORFs), eight in introns, and seven in untranslated regions. Chloroplasts exhibit far fewer sites, with approximately 41 C-to-U edits detected in A. thaliana transcripts. While C-to-U editing predominates, rare U-to-C reversals occur in some land plants, though none were found in A. thaliana organelles. These edits are highly specific, occurring almost exclusively at conserved positions to restore evolutionarily conserved protein sequences. The editing machinery relies on pentatricopeptide repeat (PPR) proteins, particularly those of the PLS subclass containing DYW domains, which act as site-specific trans-factors by binding target RNAs via their RNA-recognition motifs. These PPR-DYW proteins recruit deaminases to catalyze the C-to-U conversion, often in complex with cofactors such as multiple organellar RNA-editing factors (MORFs) and RNA-editing interacting proteins (RIPs), which enhance editing efficiency through protein-protein interactions. For instance, MORF proteins stabilize PPR-RNA complexes and facilitate deaminase activity, while RIPs may bridge interactions between editing factors. This coordinated system ensures precise editing without off-target effects, reflecting the nuclear-encoded control over organellar RNA processing. Functionally, these edits correct accumulated in organellar genomes over , often restoring canonical essential for protein function; notable examples include the creation of initiation codons or elimination of premature stop codons (UAG to UGG, ). Editing is also tissue-specific, with varying efficiencies observed across developmental stages or organs, potentially fine-tuning under environmental cues. In (Zea mays) chloroplasts, RNA is integral to the trans-splicing of the rps12 pre-mRNA, where PPR protein ZmPPR4 facilitates intron assembly and subsequent editing to produce functional ribosomal protein S12. The loss of editing sites in carnivorous plants, such as Drosera rotundifolia and Nepenthes ventricosa, correlates with organellar genome restructuring, including losses and reduced editing to only six sites in plastids, suggesting compensatory genomic changes in nutrient-stressed lineages.

Viral RNA Editing

Viral RNA editing primarily involves modifications to the genome or transcripts by cellular enzymes, enabling viruses to adapt during replication and interact with defenses. One key type is A-to-I editing mediated by adenosine deaminases acting on RNA (), particularly ADAR1, which targets double-stranded regions in RNAs. This editing introduces , read as during , potentially altering sequences and reducing the of double-stranded RNA intermediates. In contrast, C-to-U editing in coronaviruses like is driven by APOBEC family deaminases, such as APOBEC3A, which convert to uracil in single-stranded RNA, leading to hypermutation patterns observed in emerging variants. These editing events serve critical functions in , including the generation of protein isoforms that enhance replication or evade immunity. A-to-I editing can disrupt to limit over-replication or create variants with altered antigenicity, while C-to-U hypermutation promotes akin to antigenic variation in other RNA viruses. Both types contribute to immune evasion by attenuating the host response; for instance, ADAR1-mediated editing of viral RNAs prevents excessive activation of pattern recognition receptors like , thereby dampening type I production and allowing persistent infection. In some cases, viruses exploit this to reduce antiviral signaling, as seen in hyper-edited defective interfering RNAs that sequester host sensors without productive replication. Notable examples illustrate host-virus dynamics in RNA editing. In measles virus, persistent infections in the brain feature extensive A-to-I hyper-editing by ADAR1, which suppresses and but can be hijacked to facilitate long-term persistence by editing non-structural proteins. Similarly, in HIV-1, ADAR1 edits adenosines in the 5' untranslated region, as well as the Tat and coding sequences, influencing nuclear export of unspliced viral transcripts and promoting virion production through modified protein isoforms. For , 2024 analyses of variants revealed recurrent C-to-U editing sites biased toward single-stranded regions, driven by host APOBEC3A, which accelerates mutational instability and contributes to the evolution of immune-escape variants like sublineages. This interplay highlights ADAR1's dual role: it restricts viral spread by editing to inactivate genomes but enables evasion when viruses induce or tolerate editing, underscoring editing as a battleground in host-virus conflict.

Evolutionary Origins and Implications

Proposed Origins

One prominent hypothesis posits that RNA editing represents a vestige of the ancient , where self-replicating RNA molecules required post-transcriptional correction mechanisms to mitigate errors from inherently error-prone replication processes. In this view, enzymes like adenosine deaminases acting on RNA (ADARs) repurposed ancient RNA-modifying activities—originally involved in RNA or repair—for modern editing functions, allowing persistence of these "old players" in eukaryotic genomes. This theory aligns with the broader paradigm, suggesting that editing evolved as a legacy of primordial fidelity challenges before the emergence of DNA-based genomes. A complementary proposal frames RNA editing as a compensatory to alleviate mutational load in organelles, particularly mitochondria and plastids, which exhibit high mutation rates due to limited recombination and exposure to . In plant organelles, C-to-U editing frequently restores conserved amino acids at critical codon positions, counteracting deleterious genomic changes and maintaining protein functionality, such as hydrophobicity in membrane-bound complexes. This adaptive role is supported by the selective fixation of editing sites over evolutionary time, driven by functional constraints rather than neutral drift. Evidence for the deep antiquity of RNA editing comes from its presence in kinetoplastids, a basal eukaryotic lineage, where U-insertion/deletion editing likely originated in a common ancestor following the divergence from euglenoids over a billion years ago. Comparative analyses reveal that guide RNA-encoding minicircles and productively edited transcripts, such as those for ND8 and oxidase subunits, are conserved across kinetoplastid subgroups, indicating an early innovation predating more specialized forms like extensive pan-editing. In , C-to-U editing shows signs of , independently arising in mitochondrial and genomes to address similar mutational pressures, despite phylogenetic distance from kinetoplastid systems. Key proposals from the further suggest that A-to-I editing by ADARs arose via a "mistaken identity" mechanism, where enzymes, evolved for antiviral defense against double-stranded , inadvertently targeted cellular transcripts, leading to functional recoding events. The antiquity of RNA editing is underscored by conserved sites across eukaryotic clades, with estimates placing the last eukaryotic common ancestor approximately 1.8–2 billion years ago.

Evolutionary Conservation and Diversity

RNA editing mechanisms exhibit varying degrees of evolutionary conservation across eukaryotic lineages, reflecting their ancient origins while adapting to lineage-specific pressures. The core acting on RNA () enzymes, responsible for A-to-I editing, are highly conserved among metazoans, with structural and functional similarities preserved from to vertebrates, underscoring their essential role in neural function and development. In contrast, pentatricopeptide repeat (PPR) protein-mediated C-to-U editing in plant organelles shows significant loss in certain angiosperm lineages, where editing sites have diminished over time, accompanied by a reduction in PPR numbers, suggesting relaxed selective constraints in these advanced flowering . Diversity in RNA editing is evident in unique forms restricted to specific eukaryotic groups, highlighting independent evolutionary innovations. U-insertion/deletion editing, which extensively modifies mitochondrial mRNAs by adding or removing uridines, is characteristic of kinetoplastids within the excavate protists, enabling the restoration of functional coding sequences from highly derived genomes and distinguishing this mechanism from point-mutation editing prevalent elsewhere. Recent discoveries have revealed non-canonical A-to-I editing in fungi, independent of metazoan ADARs and mediated by tRNA-specific deaminases during , allowing for adaptive responses such as antiviral defense through modulation of nearby . Evolutionary drivers of RNA editing vary by context and genomic region, balancing adaptive benefits with stochastic processes. In brain-expressed genes, nonsynonymous A-to-I editing sites show signatures of positive selection, particularly in , where they enhance protein diversity and neural adaptability, contributing to lineage-specific traits. Conversely, the abundance of editing in non-coding regions, such as introns and repetitive elements, is often attributed to neutral drift, where off-target deaminase activity accumulates without strong purifying selection, potentially serving as a reservoir for future functional innovations. This duality may facilitate , as editing-generated transcriptomic variability can promote and ecological divergence, exemplified by elevated editing rates in species with recent adaptive radiations. Comparative genomic analyses further illuminate how editing sites correlate with , influencing their distribution and evolution. In , a substantial proportion of A-to-I editing events occur within Alu transposon insertions, where inverted Alu pairs create double-stranded RNA substrates that attract enzymes, leading to hyper-editing that shapes diversity and potentially drives primate-specific adaptations.

Therapeutic Applications

Enzyme-Based RNA Editing Therapies

Enzyme-based RNA editing therapies harness endogenous cellular enzymes, such as for adenosine-to-inosine (A-to-I) editing and for cytidine-to-uridine (C-to-U) editing, by delivering synthetic guide RNAs or engineered recruiters to direct site-specific modifications in target mRNAs. These approaches avoid direct alteration, focusing instead on transient RNA-level corrections that can restore protein function in genetic disorders. For A-to-I editing, platforms recruit the cell's own enzymes using partially double-stranded guide RNAs that anchor to the target transcript, enabling precise without introducing foreign proteins. Wave Life Sciences' platform exemplifies this, utilizing chemically optimized delivered subcutaneously to activate endogenous for editing the SERPINA1 mRNA in (AATD), a causing and liver due to misfolded protein accumulation. Therapeutic targets include monogenic diseases amenable to single-nucleotide corrections or modulation. In AATD, Wave's WVE-006 achieved the first demonstration of therapeutic RNA editing in humans during the phase 1b/2a RestorAATion-2 , where a 200 mg multidose regimen resulted in 7.2 μM wild-type M-AAT (64.4% of total AAT) with 60.3% reduction in mutant Z-AAT, sustained for at least 2 months; a single 400 mg dose yielded 5.3 μM M-AAT (47.2% of total AAT) with 49% Z-AAT reduction. As of September 2025, the therapy was well-tolerated with no serious adverse events; multidose expansion data expected in Q1 2026, and as of Q3 2025, the trial supports monthly or less frequent subcutaneous dosing with a favorable safety profile. For (DMD), preclinical recruitment has corrected point mutations in the DMD gene, such as the missense 1682G>A , restoring expression in mouse models via A-to-I editing to revert the codon, potentially addressing cases caused by such mutations. These therapies offer key advantages, including reversibility—edits last only as long as the RNA persists, typically days to weeks—and avoidance of DNA off-target effects, minimizing risks of permanent mutations or seen in DNA editing. Delivery systems like LNPs for or adeno-associated viruses (AAV) for tissue-specific targeting enhance accessibility, with LNPs enabling non-invasive subcutaneous dosing as in WVE-006. However, challenges persist, notably variable editing efficiencies due to suboptimal ADAR recruitment and guide RNA stability, though recent trials show levels up to 64%. Immune responses to synthetic guides, including innate activation via Toll-like receptors, can reduce efficacy and cause , as observed in early oligonucleotide trials. Specificity remains a hurdle, with off-target A-to-I edits potentially altering unintended transcripts, necessitating advanced chemical modifications like 2'-O-methyl substitutions to improve precision. Ongoing optimizations, such as AI-guided guide design, aim to address these to advance clinical viability.

CRISPR-Based RNA Editing Tools

CRISPR-based RNA editing tools leverage the programmable RNA-targeting capabilities of Cas13 enzymes, which are fused to deaminase domains to enable precise base conversions without altering the genomic DNA. These systems emerged as a safer alternative to DNA editing by allowing transient, reversible modifications at the RNA level, building briefly on native enzyme mechanisms like ADAR for adenine deamination. Unlike traditional CRISPR-Cas9, which cleaves DNA, Cas13 binds and edits single-stranded RNA transcripts guided by CRISPR RNAs (crRNAs), minimizing off-target effects and immunogenicity risks. Key systems include the RNA Editing for Programmable A to I Replacement (REPAIR), which uses a catalytically dead Cas13b (dCas13b) fused to the deaminase domain of ADAR2 to achieve adenosine-to-inosine (A-to-I) editing. This , guided by crRNA, recruits ADAR2 to target sites without sequence constraints beyond the protospacer flanking sequence, enabling correction of disease-associated mutations in transcripts like those for β-thalassemia. For combined knockdown and editing, Cas13d variants, such as those in the CasRx system, provide robust RNA cleavage alongside deaminase fusions; Cas13d's smaller size (~775 amino acids) facilitates delivery, and it has been adapted for isoform-specific targeting in mammalian cells, achieving up to 90% knockdown efficiency in primary T cells. Systems like LEAPER advance A-to-I editing using RNA-templated recruitment of ADAR without Cas13 dependency; for U-to-C, RESCUE uses Cas13b with engineered APOBEC1, demonstrating high specificity . Recent developments focus on engineering more compact and versatile Cas13 orthologs to improve delivery and multiplexing. Evolved variants like Cas13X and Cas13Y, discovered from uncultivated microbes, are the smallest in the family at 775–800 amino acids, reducing payload size for viral vectors and enabling RNA interference or base editing in hard-to-transfect cells. These compact enzymes support multiplex editing in vivo, as shown in a 2024 Cas13d platform that simultaneously regulates dozens of transcripts in mouse models via arrayed crRNAs, achieving coordinated knockdown of immune checkpoints without genomic integration. Such advancements postdate earlier Cas7-11 explorations, filling gaps in scalable, tissue-specific applications. In therapeutic contexts, these tools show promise in by tuning immune responses; for instance, Cas13-mediated editing of PD-1 transcripts in T cells reduces inhibitory signaling, enhancing antitumor activity in preclinical models. A 2024 Stanford-developed Cas13d platform enables metabolic tuning in immune cells by regulating glycolytic enzyme transcripts, boosting CAR-T cell persistence and efficacy against solid tumors without permanent DNA changes. For neurological applications, similar Cas13 tools target metabolic pathways in neurons, with efficiencies reaching 50% in mouse brain tissues for correcting metabolic disorders like those in models. The 2025 platform represents a breakthrough, integrating a hairpin-structured with ADAR2 deaminase to displace non-target strands, achieving up to 67% editing efficiency and minimal bystander edits in cells. This system enhances CRISPR-Cas13 fusions for therapeutic , particularly , by reducing off-target rates to below 1% in multiplex settings.

Comparison to DNA Editing

RNA editing therapies offer a transient modification to RNA transcripts, typically lasting hours to days, in stark contrast to the permanent genomic changes induced by DNA editing techniques such as . This reversibility stems from RNA's short lifespan and natural turnover, allowing effects to dissipate without lasting genomic impact, whereas DNA edits persist across cell divisions and generations. Furthermore, RNA editing circumvents the need for double-strand breaks (DSBs) required in many DNA editing methods, thereby minimizing risks of off-target mutations, insertions, or deletions that could lead to unintended genomic instability. These approaches are especially advantageous for non-dividing or post-mitotic cells, like neurons or mature muscle cells, where DNA editing efficiency is limited due to poor access or integration challenges. A key benefit of RNA editing lies in its capacity to address splicing defects—common in diseases like spinal muscular atrophy—by precisely altering pre-mRNA sequences to restore functional isoforms, all without modifying the underlying DNA. This preserves the genome's integrity, reducing long-term risks such as oncogenesis associated with DNA alterations. Ethically, RNA editing avoids germline transmission of changes, making it preferable for applications where heritable modifications raise concerns, unlike DNA therapies that could affect future generations. Despite these advantages, RNA editing requires repeated administrations to maintain therapeutic effects, given the ephemeral nature of RNA modifications, which contrasts with the one-time dosing potential of DNA edits. Delivery remains a hurdle, mirroring challenges in mRNA-based , including , , and tissue-specific targeting via nanoparticles or vectors. Illustrative examples highlight these distinctions: CRISPR-Cas9 DNA editing has yielded approved therapies like Casgevy for (SCD), involving editing of hematopoietic stem cells to permanently reactivate fetal hemoglobin production. Conversely, ADAR-based RNA editing platforms are advancing for SCD and β-thalassemia, enabling reversible correction of aberrant transcripts in blood cells without genomic cuts. Looking ahead, the RNA editing technologies market is projected to expand from USD 262 million in 2024 to USD 397 million by 2030, driven by growing clinical pipelines.

References

  1. [1]
    RNA editing enzymes: structure, biological functions and applications
    Mar 16, 2024 · This article provides an overview of the structure, function, and applications of RNA editing enzymes.
  2. [2]
    A-to-I RNA editing — immune protector and transcriptome diversifier
    Apr 24, 2018 · This Review focuses on the most abundant form of RNA editing in Metazoans, adenosine-to-inosine (A-to-I) editing, in which enzymes encoded by ...Missing: article | Show results with:article
  3. [3]
    RNA editing enzymes: structure, biological functions and applications
    Mar 16, 2024 · This review article aims to investigate the mechanism of RNA editing, its relationship to disease, and the potential application of tools based ...
  4. [4]
    Expanding the proteome: A-to-I RNA editing provides an adaptive ...
    Apr 10, 2023 · Widespread application of DNA- and RNA-sequencing has revealed that A-to-I editing is prevalent in mammals, cephalopods, insects, and other ...
  5. [5]
    Functional Impact of RNA editing and ADARs on regulation of gene ...
    Aug 19, 2014 · RNA editing imposes a layer of reciprocal regulation on miRNA biology. In addition to messenger RNAs, editing is also known to take place on ...
  6. [6]
    Widespread RNA Editing of Embedded Alu Elements in the Human ...
    We find that editing of transcripts with embedded Alu sequences is a global phenomenon in the human transcriptome, observed in 2674 (∼2%) of all publicly ...
  7. [7]
    The emerging role of RNA editing in plasticity
    Jun 1, 2015 · Because RNA editing can be regulated, it is an ideal tool for increasing genetic diversity, adaptation and environmental acclimation.<|control11|><|separator|>
  8. [8]
    Mitochondrial hypoxic stress induces widespread RNA editing by ...
    Feb 21, 2019 · In response to acute hypoxia, A3G-mediated RNA editing may promote Warburg effect by preferring glycolysis over mitochondrial respiration and ...
  9. [9]
    C-to-U and U-to-C: RNA editing in plant organelles and beyond
    Apr 9, 2023 · The genomes in the two energy-converting organelles of plant cells, chloroplasts and mitochondria, contain numerous 'errors' that are ...
  10. [10]
    The essential role of AMPA receptor GluR2 subunit RNA editing in ...
    This review focuses on the critical role of glutamine/arginine (Q/R) site RNA editing of the GluA2 subunit in regulating AMPA receptor Ca2+-permeability. AMPA ...Abstract · Introduction · The GluA2 Subunit of the... · Novel Insights into the...
  11. [11]
    A-to-I RNA editing occurs at over a hundred million genomic sites ...
    Most sites are edited at low levels (<1% of transcripts), with the median editing level being 0.475%. Notably, the distribution for other types of mismatches ...
  12. [12]
    The evolution and adaptation of A-to-I RNA editing | PLOS Genetics
    Nov 28, 2017 · RNA editing may be advantageous for adaptation because it contributes to transcriptome diversity, generates plasticity in genomic regions of ...
  13. [13]
  14. [14]
  15. [15]
    Identification of human RNA editing sites: a historical perspective
    May 18, 2016 · Although next-generation sequencing was shown to be a potent tool in validation of editing sites predicted using ESTs, the most promising ...
  16. [16]
  17. [17]
  18. [18]
    Advances in Detection Methods for A‐to‐I RNA Editing - PMC - NIH
    Apr 14, 2025 · This review presents the evolution of methodologies for RNA editing detection and examines recent advances, including chemically‐assisted, enzyme‐assisted, and ...
  19. [19]
  20. [20]
    Atlas of RNA editing events affecting protein expression in aged and ...
    Dec 2, 2021 · We present an extended reference set of brain RNA editing events, identify a subset that are found to be expressed at the protein level,
  21. [21]
    Cell culture NAIL-MS allows insight into human tRNA and ... - Nature
    Jan 15, 2021 · We present a fast and reliable stable isotope labeling strategy which allows in-depth study of RNA modification dynamics in human cell culture.
  22. [22]
  23. [23]
    Advances in Detection Methods for A‐to‐I RNA Editing
    Apr 14, 2025 · The first A-to-I RNA editing was detected using Sanger sequencing (Sanger et al. 1977) by comparing nucleotide differences between genomic DNA ...Missing: mammals | Show results with:mammals
  24. [24]
    Transfer learning enables identification of multiple types of RNA ...
    May 14, 2024 · Here, we develop TandemMod, a transferable deep learning framework capable of detecting multiple types of RNA modifications in single DRS data.
  25. [25]
    Nanopore sequencing to detect A-to-I editing sites - PubMed
    We have developed, named Dinopore (Detection of inosine with nanopore sequencing), to interrogate the A-to-I editome of any organism.Missing: real- time 2023-2025
  26. [26]
    (PDF) Advances in Detecting RNA Modifications Using Direct RNA ...
    Oct 30, 2025 · Schematic overview of the Oxford Nanopore Technologies (ONT) direct RNA sequencing workflow and RNA modification detection. Schematic ...
  27. [27]
    Detection of RNA Editing Sites by GPT Fine-tuning - YouTube
    Dec 9, 2024 · This study introduces a novel methodology leveraging advanced AI techniques, specifically OpenAI's GPT-3.5, to predict both the occurrence ...Missing: driven annotation<|separator|>
  28. [28]
    Prediction of RNA editable sites with convolutional neural network
    In this work, we constructed Convolutional Neural Network (CNN) models to predict human RNA editing events in both Alu regions and non-Alu regions.Missing: EditomeDB | Show results with:EditomeDB
  29. [29]
    [PDF] Detection of RNA Editing Sites by GPT Fine-tuning - OpenReview
    In RNA editing prediction, the goal is not to locate regions where the binding will occur but to predict whether a specific site will undergo editing. 2. Page 3 ...Missing: EditomeDB | Show results with:EditomeDB
  30. [30]
  31. [31]
    Sequencing accuracy and systematic errors of nanopore direct RNA ...
    May 28, 2024 · We present the first comprehensive evaluation of sequencing accuracy and characterisation of systematic errors in dRNA-seq data from diverse organisms.Missing: real- | Show results with:real-
  32. [32]
    Structural perspectives on adenosine to inosine RNA editing by ...
    The hydrolytic deamination proceeds through a two-step mechanism with a tetrahedral intermediate. (B) Domain architecture of human ADARs. ADAR1 and ADAR 2 ...
  33. [33]
    RNA Editing—Systemic Relevance and Clue to Disease Mechanisms?
    In contrast to ADAR-dependent A-to-I RNA editing, APOBEC needs auxiliary proteins for deamination as detailed in Table 1 and exemplified in the bottom-left ...The Adar Gene Family · The Apobec Gene Family · Apobec-Dependent Rna Editing...
  34. [34]
    Targeted RNA editing: novel tools to study post-transcriptional ...
    ADAR enzymes contain double-stranded RNA-binding motifs (dsRBMs) and a catalytic deaminase domain (referred to hereafter as. ADARcd, with cd indicating ...<|control11|><|separator|>
  35. [35]
    Nucleocytoplasmic Distribution of Human RNA-editing Enzyme ...
    The human RNA-editing enzyme adenosine deaminase that acts on RNA (ADAR1) is expressed in two versions. A longer 150-kDa protein is interferon inducible and ...Pyruvate Kinase (pk) Fusions... · Xenopus Adar1. 1 Seemingly... · Dsrbd Cross Talk Requires...
  36. [36]
  37. [37]
  38. [38]
  39. [39]
    Uridine insertion/deletion RNA editing in trypanosome mitochondria
    The RNA modification phenomenon of uridine (U) insertion/deletion RNA editing was discovered in mitochondria of trypanosomatid protists more than 15 years ago.Missing: seminal papers
  40. [40]
    Evolution of RNA editing in trypanosome mitochondria - PNAS
    The term RNA editing describes several types of posttranscriptional modifications of RNAs that involve either specific insertion/deletion or modifications of ...Evolution Of Rna Editing In... · Minicircle-Encoded Grnas In... · C To U Editing Of Trna<|control11|><|separator|>
  41. [41]
    Apobec-1 and apolipoprotein B mRNA editing - PubMed - NIH
    Apobec-1, the catalytic component of the apoB mRNA editing enzyme complex, has been cloned. This article begins with an overview of the general biology of apoB ...Missing: paper | Show results with:paper
  42. [42]
    RNA Editing Enzyme APOBEC1 and Some of Its Homologs Can Act ...
    Here, we show that APOBEC1 and its homologs APOBEC3C and APOBEC3G exhibit potent DNA mutator activity in an E. coli assay.Missing: paper | Show results with:paper
  43. [43]
    A sequence-specific RNA-binding protein complements apobec-1 ...
    The specificity of editing is conferred by an 11-nucleotide mooring sequence ... A sequence-specific RNA-binding protein complements apobec-1 To edit ...
  44. [44]
    Low Expression of the Apolipoprotein B mRNA–Editing Transgene ...
    The apoB mRNA–editing catalytic polypeptide 1 (APOBEC-1) deaminates the cytidine in codon 2153 (C6666AA) to form a uridine. The resulting in-frame stop codon (U ...
  45. [45]
    DYW domain structures imply an unusual regulation principle in ...
    Jun 21, 2021 · Plant RNA editing specifically converts several hundreds of cytidines to uridines in mitochondrial and chloroplast transcripts. The RNA editing ...
  46. [46]
    Sequences directing C to U editing of the plastid psbL mRNA ... - NIH
    In plastids, editing of an ACG codon to an AUG codon creates the translation initiation codon for the psbL and ndhD transcripts in tobacco.
  47. [47]
    RNA editing in Arabidopsis mitochondria effects 441 C to U ... - PNAS
    A total of 456 C to U, but no U to C, conversions were identified exclusively in mRNAs, 441 in ORFs, 8 in introns, and 7 in leader and trailer sequences.
  48. [48]
    Advances in RNA editing in hematopoiesis and associated ...
    May 22, 2025 · Adenosine-to-inosine (A-to-I) RNA editing is a prevalent RNA modification essential for cell survival. The process is catalyzed by the ...
  49. [49]
    A third member of the RNA-specific adenosine deaminase gene ...
    ADAR3 inhibited in vitro the activities of RNA editing enzymes of the ADAR gene family, raising the possibility of a regulatory role in RNA editing.
  50. [50]
  51. [51]
  52. [52]
    Landscape of adenosine-to-inosine RNA recoding across human ...
    Mar 4, 2022 · ... sites). Fig. 2: The set of 1517 detected sites exhibits ADAR-dependent RNA editing features. figure 2. a Distribution of the 12 possible ...
  53. [53]
    RNA Editing of the Human Serotonin 5-HT 2C Receptor - Nature
    May 1, 2001 · RNA encoding the human serotonin 5-HT 2C receptor (5-HT 2C R) undergoes adenosine-to-inosine RNA editing events at five positions, resulting in an alteration ...
  54. [54]
    Regulation of serotonin-2C receptor G-protein coupling by RNA ...
    Here we show that transcripts encoding the 2C subtype of serotonin receptor (5-HT(2C)R) undergo RNA editing events in which genomically encoded adenosine ...
  55. [55]
    A-to-I RNA Editing: Effects on Proteins Key to Neural Excitability - PMC
    May 10, 2012 · RNA editing by adenosine-to-inosine conversion (A-to-I editing) can introduce codon changes in mRNAs and hence generate structurally and functionally different ...
  56. [56]
    Transcript diversification in the nervous system: A to I RNA-editing in ...
    Jul 8, 2012 · A to I editing is most abundant in the central nervous system (CNS). Here, targets for this type of nucleotide modification frequently encode ...
  57. [57]
    Cis- and trans-regulations of pre-mRNA splicing by RNA editing ...
    Feb 7, 2020 · As proposed by previous studies, A-to-I editing may modulate splicing through disrupting branch point sequence (BPS)21, creating novel 3′ ...
  58. [58]
    Profound downregulation of the RNA editing enzyme ADAR2 in ALS ...
    These results indicate that ADAR2 downregulation is a profound pathological change relevant to death of motor neurons in ALS.<|control11|><|separator|>
  59. [59]
    MiREDiBase, a manually curated database of validated and putative ...
    Aug 4, 2021 · unexpectedly found that a consistent percentage of miRNA editing events are non-canonical, especially C-to-A and G-to-U. Similar data were ...
  60. [60]
    C to U Editing Stimulates A to I Editing in the Anticodon Loop of a ...
    In marsupial mitochondria, a single C to U editing event at the second position of the anticodon (C35) changes a tRNA such that it recognizes aspartate in ...
  61. [61]
    A-to-I editing on tRNAs: Biochemical, biological and evolutionary ...
    It is present mainly at three positions on tRNAs: position 34, which is the first nucleotide of the anticodon (wobble-position), position 37 (following the ...
  62. [62]
    Unveiling the A-to-I mRNA editing machinery and its regulation and ...
    May 10, 2024 · A-to-I editing also occurs at positions 34 (the wobble position of the anticodon) and 37 of tRNAs. While A37 editing has been found only in ...
  63. [63]
    [PDF] ADATs: roles in tRNA editing and relevance to disease - HAL
    Jul 22, 2024 · The most studied example is the TadA- or ADAT2/3-mediated A-to-I conversion of the tRNA wobble position in the anticodon of prokaryotic or ...
  64. [64]
    The molecular basis of tRNA selectivity by human pseudouridine ...
    Jul 11, 2024 · Most PUS are “stand-alone” enzymes that autonomously bind target RNAs and catalyze the isomerization reaction without additional factors, with ...
  65. [65]
    Sequential action of a tRNA base editor in conversion of cytidine to ...
    Oct 11, 2022 · We show that an enzyme, TrcP, mediates the editing of C-to-U followed by the conversion of U to Ψ, consecutively.<|separator|>
  66. [66]
    Determinants of tRNA editing and modification - PubMed Central - NIH
    Inosine could thus allow a single tRNA to base pair with multiple codons for the same amino acid, expanding a tRNA's decoding capacity and obviating the need ...
  67. [67]
    An adenosine-to-inosine tRNA-editing enzyme that can perform C-to ...
    Here we show that down-regulation of the Trypanosoma brucei tRNA-editing enzyme by RNAi leads to a reduction in both C-to-U and A-to-I editing of tRNA in vivo.<|separator|>
  68. [68]
    tadA, an essential tRNA-specific adenosine deaminase from ... - NIH
    We show that tadA is an essential gene in E.coli, underscoring the critical function of inosine at the wobble position in prokaryotes.Missing: queuosine | Show results with:queuosine
  69. [69]
    Human mitochondrial diseases associated with tRNA wobble ...
    These findings suggest that the wobble modification deficiency plays a primary role in the molecular pathogenesis of the MELAS and MERRF mitochondrial diseases.Missing: editing | Show results with:editing
  70. [70]
    CO2-sensitive tRNA modification associated with human ... - Nature
    May 14, 2018 · Loss of tRNA modifications or tRNA-modifying enzymes is strongly associated with human disease caused by mitochondrial dysfunction8.
  71. [71]
    Tuning the ribosome: The influence of rRNA modification on ...
    In eukaryotes, the most abundant rRNA modifications are 2′-O-methylation ... sites of partial 2′-O-methylation and pseudouridylation have been identified.
  72. [72]
    snoRNAs: functions and mechanisms in biological processes, and ...
    May 12, 2022 · The former two types of snoRNAs participate in the processing of ribosomal RNA (rRNA) by adding 2′-O-methylation and pseudouridylation ...
  73. [73]
    Profiling of 2′-O-Me in human rRNA reveals a subset of fractionally ...
    Jun 1, 2016 · The box C/D and H/ACA snoRNPs: key players in the modification, processing and the dynamic folding of ribosomal RNA. Wiley Interdiscip. Rev. RNA.
  74. [74]
    2′-O-Methylation of Ribosomal RNA: Towards an Epitranscriptomic ...
    Oct 3, 2018 · 2′-O-Methylation protects RNA from hydrolysis and modifies RNA strand flexibility but does not contribute to Watson-Crick base pairing.
  75. [75]
    Regulation of translation by ribosomal RNA pseudouridylation
    Aug 18, 2023 · In eukaryotes, ribosomal RNA (rRNA) is decorated with 1 to 2% pseudouridine mostly in conserved functional domains (10). Loss of just a few ...
  76. [76]
    Ribosomal RNA 2′-O-methylations regulate translation by ... - PNAS
    We show that rRNA 2′-O-methylations affect the function and fidelity of the ribosome and change the balance between different ribosome conformational states.
  77. [77]
    Control of protein synthesis through mRNA pseudouridylation by ...
    Jul 28, 2023 · Inherited mutations in dyskerin, resulting in defective rRNA pseudouridylation, are associated with the X-linked syndrome dyskeratosis congenita ...
  78. [78]
    RNA-specific adenosine deaminase ADAR1 suppresses measles ...
    Oct 23, 2009 · Extensive A-to-I editing has been described in viral RNAs isolated from the brains of patients persistently infected with measles virus, ...
  79. [79]
    RNA editing by ADAR1 regulates innate and antiviral immune ...
    Oct 17, 2017 · ADAR1-dependent A-to-I editing has recently been recognized as a key process for marking dsRNA as self, therefore, preventing innate immune ...
  80. [80]
    Rampant C→U Hypermutation in the Genomes of SARS-CoV-2 and ...
    Jun 24, 2020 · The key findings in the study were the combined evidence for an APOBEC-like editing process driving initial sequences changes in SARS-CoV-2 and ...
  81. [81]
    The role of A-to-I RNA editing in infections by RNA viruses
    Feb 25, 2021 · A-to-I RNA editing has been shown to directly affect the genome/transcriptome of RNA viruses with significant repercussions for viral protein synthesis, ...
  82. [82]
    Host-mediated RNA editing in viruses | Biology Direct | Full Text
    Mar 28, 2023 · Here we synthesize the current knowledge of host-mediated RNA editing in a variety of viruses by considering two enzyme families, viz., ADARs and APOBECs.
  83. [83]
    Host RNA editor restricts measles | Nature Reviews Microbiology
    Jan 13, 2011 · A host RNA-editing enzyme (ADAR1) inhibits the replication and cytotoxicity of measles virus and other respiratory RNA viruses.
  84. [84]
    Editing of HIV-1 RNA by the double-stranded RNA deaminase ...
    Finally, we report that ADAR1 associates with HIV-1 RNAs and edits adenosines in the 5' untranslated region (UTR) and the Rev and Tat coding sequence.
  85. [85]
    The ADAR1 editing enzyme is encapsidated into HIV-1 virions
    In particular, A-to-I editing was identified in the 5′ untranslated region (5′UTR) and within the Tat and Rev sequences and in the vicinity of Rev responsive ...
  86. [86]
    C→U transition biases in SARS-CoV-2: still rampant 4 years from ...
    Dec 11, 2024 · Future functional studies are required to determine editing preferences, impacts on replication fitness in vivo of SARS-CoV-2 and other RNA ...
  87. [87]
    Did RNA editing in plant organellar genomes originate under natural ...
    Oct 21, 2008 · Background. The C↔U substitution types of RNA editing have been observed frequently in organellar genomes of land plants.Missing: compensatory load
  88. [88]
  89. [89]
    Ancient origin of RNA editing in kinetoplastid protozoa - ScienceDirect
    The mechanism of editing is still uncertain, but multiple ribonucleoprotein complexes have been identified that contain components of the enzymatic machinery.
  90. [90]
    A late origin of the extant eukaryotic diversity: divergence time ...
    May 19, 2011 · Estimates of the age of Last Eukaryotic Common Ancestor (LECA) vary approximately twofold, from ~1,100 million years ago (Mya) to ~2,300 Mya.
  91. [91]
    The evolution of RNA editing and pentatricopeptide repeat genes
    May 9, 2011 · Loss of RNA editing in angiosperms. Over 100 chloroplast editing ... PPR proteins in angiosperms (c. 200–250) have other functions. This ...
  92. [92]
    U-insertion/deletion mRNA editing holoenzyme: definition in sight
    Abstract. RNA editing is a process that alters DNA-encoded sequence and is distinct from splicing, 5′ capping and 3′ additions.
  93. [93]
    Adaptation of A-to-I RNA editing in Drosophila | PLOS Genetics
    We detected strong signatures of positive selection on the nonsynonymous editing sites in Drosophila brains, and the beneficial editing sites were ...
  94. [94]
    The preponderance of nonsynonymous A-to-I RNA editing ... - Nature
    Nov 27, 2019 · For instance, in humans, the fraction of sites subject to nonsynonymous editing is lower than that subject to synonymous editing and the editing ...
  95. [95]
    RNA editing: a driving force for adaptive evolution? - PMC
    RNA editing changes a codon, which in turn leads to a protein with amino acid substitution and altered functional properties. Examples: neurotransmitter ...
  96. [96]
    Adenosine-to-inosine RNA editing shapes transcriptome diversity in ...
    Adenosine-to-inosine (A-to-I) RNA editing is a widespread modification of the transcriptome. ... GM Borchert, et al., Adenosine deamination in human transcripts ...Adenosine-To-Inosine Rna... · Results · Rna Editing Level In...<|control11|><|separator|>
  97. [97]
    Wave Life Sciences Announces First-Ever Therapeutic RNA Editing ...
    Oct 16, 2024 · The first-ever therapeutic RNA editing in humans was achieved with WVE-006, restoring wild-type M-AAT levels in AATD patients, with a single ...Missing: ADAR- | Show results with:ADAR-
  98. [98]
    Emerging clinical applications of ADAR based RNA editing - PMC
    May 26, 2025 · Wave Life Sciences has initiated RNA editing therapies with their lead drug candidate WVE-006, currently undergoing Phase1 (NCT06186492) and ...
  99. [99]
  100. [100]
    RNA Editing Therapeutics: Advances, Challenges and Perspectives ...
    Oct 14, 2022 · In this review, we discuss the main changes that RNA editing affects in cardiovascular gene expression and biology, as well as we introduce the ...
  101. [101]
    Site-directed RNA editing: recent advances and open challenges
    Common problems of all RNA-based therapeutics are their stability, antigenicity, specificity and deliverability. For instance, phosphorothioate linkage and 2ʹO ...Crispr-Cas13 Approaches For... · Endogenous Adar Approaches · Optimizing Rna Design
  102. [102]
    RNA editing with CRISPR-Cas13 - Science
    Oct 25, 2017 · This system, referred to as RNA Editing for Programmable A to I Replacement (REPAIR), which has no strict sequence constraints, can be used to ...
  103. [103]
    Cas13d-mediated isoform-specific RNA knockdown with a unified ...
    Jul 29, 2025 · We show that EEJs are broadly accessible to the Cas13d complex–and therefore targetable–for RNA knockdown. This expands the diversity of RNAs ...
  104. [104]
    A versatile CRISPR-Cas13d platform for multiplexed transcriptomic ...
    Feb 29, 2024 · MEGA enables quantitative, reversible, and massively multiplexed gene knockdown in primary human T cells without targeting or cutting genomic DNA.
  105. [105]
    Leveraging CRISPR gene editing technology to optimize the efficacy ...
    Oct 25, 2024 · In this review we summarize potential uses of the CRISPR system to improve results of CAR T-cells therapy including optimizing efficacy and safety.
  106. [106]
    A new RNA editing tool could enhance cancer treatment
    Feb 21, 2024 · Cell therapies for cancer can be potentially enhanced using a CRISPR RNA-editing platform, according to a new study published Feb. 21 in Cell.Missing: accuracy abundance false positives 50-70% 2023
  107. [107]
    New CRISPR technology could help repair damaged neurons
    May 21, 2025 · “We've added a new tool to the CRISPR toolbox, using it to control RNA localization inside the cell. This has never been achieved before and, ...Missing: 2024 metabolic tuning
  108. [108]
    The Past, Present, and Future of RNA Editing Therapeutics
    Apr 18, 2025 · Shape Therapeutics has pioneered AI-guided design of RNA payloads and developed programmable ADAR systems delivered via AAV vectors. Wave Life ...Molecular Mechanisms · Clinical Applications And... · Emerging Frontiers And...<|separator|>
  109. [109]
    Assessing and advancing the safety of CRISPR-Cas tools: from DNA ...
    Jan 13, 2023 · Unlike DNA editing, RNA editing offers an alternative to genome editing in certain therapeutic applications in a reversible and tunable ...
  110. [110]
    No Cuts, Just Edits: Therapeutic A-to-I RNA Editing as a Safer ...
    In 2023, the FDA approved Casgevy, the first clinically approved CRISPR-based gene-editing therapy for sickle cell disease. Casgevy utilizes CRISPR/Cas9 ...Missing: enrichment rare detection
  111. [111]
    Gene editing: DNA versus RNA | Drug Discovery News
    Feb 13, 2023 · With RNA editing, you've got a lot more targets to take care of,” he said. Extending the lifespan of the RNA editor to reach all of those ...Gene Editing: Dna Versus Rna · A Renewable Target · Two And Done<|separator|>
  112. [112]
    RNA editing set to take off: could it outperform gene editing?
    Feb 12, 2024 · RNA editing has recently made significant progress as a potentially safer alternative to gene editing for treating a number of genetic diseases.
  113. [113]
    RNA-editing drugs advance into clinical trials - Nature
    Apr 18, 2024 · The drug aims to treat AATD, a genetic lung and liver disease that is caused by mutation in the SERPINA1 gene, which encodes the AAT protein.Table 1 | Adar-Based Editors... · An Endogenous Editor · Common Ground<|separator|>
  114. [114]
    RNA editing: Expanding the potential of RNA therapeutics - PMC
    ADAR-based RNA editing has emerged as a powerful tool to engineer RNAs, enable correction of disease-causing mutations, and modulate protein functions.
  115. [115]
    FDA Approves First Gene Therapies to Treat Patients with Sickle ...
    Dec 8, 2023 · CRISPR/Cas9 can be directed to cut DNA in targeted areas, enabling the ability to accurately edit (remove, add, or replace) DNA where it was cut ...
  116. [116]
    Gene and RNA Editing: Methods, Enabling Technologies ... - arXiv
    The purpose of gene and RNA editing is to modulate the root of disease by fixing the gene's DNA mutation [6] . Report issue for preceding element.
  117. [117]
    Global RNA Editing Technologies Market By Share, Size and ...
    The Global RNA Editing Technologies Market, valued at USD 262.28 Million in 2024, is projected to experience a CAGR of 7.16% to reach USD 397.16 Million by 2030 ...