Skip to main content
Wiley Open Access Collection logoLink to Wiley Open Access Collection
. 2024 Jun 22;15(3):e1863. doi: 10.1002/wrna.1863

Revealing the hidden RBP–RNA interactions with RNA modification enzyme‐based strategies

Hua Jin 1,2, Chong Li 1, Yunxiao Jia 1, Yuxuan Qi 3, Weilan Piao 1,2,
PMCID: PMC11469752  PMID: 39392204

Abstract

RNA‐binding proteins (RBPs) are powerful and versatile regulators in living creatures, playing fundamental roles in organismal development, metabolism, and various diseases by the regulation of gene expression at multiple levels. The requirements of deep research on RBP function have promoted the rapid development of RBP–RNA interplay detection methods. Recently, the detection method of fusing RNA modification enzymes (RME) with RBP of interest has become a hot topic. Here, we reviewed RNA modification enzymes in adenosine deaminases that act on RNA (ADAR), terminal nucleotidyl transferase (TENT), and activation‐induced cytosine deaminase/ApoB mRNA editing enzyme catalytic polypeptide‐like (AID/APOBEC) protein family, regarding the biological function, biochemical activity, and substrate specificity originated from enzyme selves, their domains and partner proteins. In addition, we discussed the RME activity screening system, and the RME mutations with engineered enzyme activity. Furthermore, we provided a systematic overview of the basic principles, advantages, disadvantages, and applications of the RME‐based and cross‐linking and immunopurification (CLIP)‐based RBP target profiling strategies, including targets of RNA‐binding proteins identified by editing (TRIBE), RNA tagging, surveying targets by APOBEC‐mediated profiling (STAMP), CLIP‐seq, and their derivative technology.

This article is categorized under:

  • RNA Interactions with Proteins and Other Molecules > Protein‐RNA Recognition

  • RNA Processing > RNA Editing and Modification

Keywords: RNA modification, RNA tagging, RNA‐binding protein, STAMP, TRIBE


The timeline for development of RNA‐binding proteins' target RNA profiling methods.

graphic file with name WRNA-15-e1863-g001.jpg

1. INTRODUCTION

RNA‐binding proteins (RBPs) interact with RNA inside cells, and ensure the precisely‐regulated RNA life cycle such as co‐transcriptional processing, nuclear export, functioning, and degradation (Gerstberger et al., 2014). Considering that RBPs are expected to account for over 10% of human genes (Mukherjee et al., 2019), the protein–RNA interaction forms an extremely intricate network within the cell and underlies fundamental cellular activity (Darnell, 2010; Fecko et al., 2007; McHugh et al., 2014; Wang, Tidei, et al., 2015).

Defects in RBPs affect various biological processes including organismal development, viral infection, and cellular defense, thus causing a variety of human diseases (Klattenhoff et al., 2013; Ule et al., 2003). Several RBPs are associated with neurological disorders and cancer. For instance, Nova proteins are target antigens in the autoimmune disorder paraneoplastic opsoclonus–myoclonus ataxia (POMA), which gives rise to a neurodegenerative syndrome (Musunuru & Darnell, 2001). FMRP is an RBP highly expressed in brain. Absence or mutation of FMRP leads to the Fragile X syndrome, the most frequent cause of inherited mental retardation (Zalfa et al., 2003). Dementia and motor‐neuron diseases are associated with accumulation of disordered RBPs, such as TDP‐43 in amyotrophic lateral sclerosis (ALS) and ataxin‐1(ATXN1) in Ataxia (Bossy‐Wetzel et al., 2004). Also, perturbations in RBP‐RNA networks are causally related to cancer development. RBMS1 is ubiquitously expressed in triple negative breast cancer (TNBC). Researchers have found that RBMS1 regulates anti‐tumor T cell immunity in TNBC by regulating programmed death ligand 1 (PD‐L1) levels. As a new immune checkpoint regulator, RBMS1 has great potential to be the diagnostic marker and therapeutic target for TNBC (Zhang et al., 2022). The elucidation of RBP–RNA interaction is significant to understand underlying mechanisms in these biological and pathological processes, which has promoted the rapid development of methodology detecting RBP's target transcripts.

Currently, there are many methods for studying RBP–RNA interactions and they are mainly divided into two categories. The first category is RNA‐centric methods (Grawe et al., 2021; Ramanathan et al., 2019), which are designed to identify proteins that interact with a specific RNA of interest. This category includes both in vitro and in vivo methods. In vitro methods employ streptavidin‐binding biotinylated RNA (Zheng et al., 2016) or the S1 RNA aptamer (Leppek & Stoecklin, 2014), or use an engineered CRISPR endoribonuclease Csy4 (Lee et al., 2013) to immobilize bait RNA. A fluorescence‐labeled specific RNA can also be hybridized with a recombinant protein microarray chip to profile proteins binding to a given RNA (Kretz et al., 2013; Siprashvili et al., 2016). Among methods in vivo, chromatin isolation by RNA purification (ChIRP; Chu et al., 2011; Chu et al., 2015) and capture hybridization analysis of RNA targets (CHART; Simon et al., 2011; West et al., 2014) are based on a formaldehyde cross‐linking strategy, RNA antisense purification (RAP; McHugh & Guttman, 2018), peptide nucleic acid (PNA)‐assisted identification of RBPs (PAIR; Zeng et al., 2006), MS2 in vivo biotin tagged RNA affinity purification (MS2‐BioTRAP; Tsai et al., 2011), and tandem RNA isolation procedure (TRIP; Matia‐Gonzalez et al., 2017) rely on a UV cross‐linking strategy, while RNA–protein interaction detection (RaPID; Ramanathan et al., 2018) is a noncross‐linking method. The second category is protein‐centric methods, which focus on a specific protein of interest to identify its interacting RNAs. Cross‐linking and immunopurification (CLIP)‐based methods and RNA modification enzyme (RME)‐based methods are available in this category. Each method in the protein‐centric category has its unique advantages and limitations, making the choice of appropriate methods crucial for addressing specific biological questions.

2. THE DEVELOPMENT OF METHODS PROFILING RBPs TARGET RNAs

Over the last 25 years, a series of techniques with different accuracy and resolution have emerged for profiling RBP's target RNAs, and they can be roughly classified into CLIP‐based and RME‐based methods (Figure 1; Hafner et al., 2021; Lee & Ule, 2018; Ramanathan et al., 2019; Zarnegar et al., 2016). In 2000, Tenenbaum and colleagues developed RNA immunoprecipitation‐microarray (RIP‐Chip; Tenenbaum et al., 2000), and successfully profiled the mRNA compositions in the messenger ribonucleoprotein complexes (mRNPs) at a transcriptomic level for the first time. The method has been widely used, however, it has very obvious weaknesses like low target specificity and absence of binding site information, leading to difficulty in discovering RBP‐binding motif (Dahm et al., 2012). To overcome the weaknesses, several improved methods were generated. Ule and Darnell pioneered the ultraviolet (UV) crosslinking and immunoprecipitation (CLIP)‐based methodology. In 2003, they established an in vivo method CLIP (Ule et al., 2003, 2005), which makes covalent protein–RNA‐linking by irradiation of cells with UV and then purifies a specific protein–RNA complex using immunoprecipitation (IP) followed by SDS–PAGE separation. CLIP revealed dozens of mRNAs targeted and regulated by Nova protein in the brain, laying a foundation for the later exploration of the network of protein and RNA. In 2008, to accurately illustrate the RNA‐binding sites of an any given RBP on a large scale, Licatalosi and colleagues introduced the next‐generation sequencing (NGS) technology to CLIP method, and invented high‐throughput sequencing of CLIP cDNA library (HITS‐CLIP; Licatalosi et al., 2008). In 2010, the use of nucleotide analogs 4‐thiouridine (4‐SU) and 6‐thioguanosine (6‐SG) further improved the resolution and UV‐crosslinking efficiency of CLIP, and launched photocatalytivable ribonucleoside enhanced crosslinking and immunoprecipitation (PAR‐CLIP; Hafner, Landthaler, Burger, Khorshid, Hausser, Berninger, Rothballer, Ascano, et al., 2010; Hafner, Landthaler, Burger, Khorshid, Hausser, Berninger, Rothballer, Ascano, et al., 2010). In the same year, König and colleagues ingeniously utilized the characteristics that UV‐crosslinked RNA sites frequently induce mutations during reverse transcription (RT), and thereby developed individual‐nucleotide resolution CLIP (iCLIP; Konig et al., 2010, 2011), achieving nucleotide resolution in binding site identification.

FIGURE 1.

FIGURE 1

The timeline for development of RBP's target RNA profiling methods.

Even though CLIP‐based methods had been optimized for several years, previously developed CLIP methods still had problems like high experimental failure rate, low reproducibility, high false‐positive signals, and low complexity of generated CLIP‐seq libraries. Thus, in 2016, Gene W. Yeo group developed the method enhanced CLIP (eCLIP; van Nostrand et al., 2016) based on the theory that reverse transcription is usually terminated at UV‐crosslinked RNA sites. eCLIP further improved the library preparation by ligating two linear adapters separately. As the second adapter was ligated to cDNA fragments at right after RT‐terminated 3′ ends, eCLIP could maintain the single nucleotide resolution in RBP‐binding site identification. Additionally, efficient enzymatic reactions, parallel IgG and size‐matched input controls, and using adaptors with unique molecular identifiers (UMIs) not only greatly increased library complexity and PCR duplicate identification accuracy, but also reduced background noise and adaptor ligation bias, resulting in more reliable results. Afterward, eCLIP was used to draw the large‐scale RBP–RNA maps and study the functions of 356 human RBPs systematically (van Nostrand, Freese, et al., 2020). To accomplish efficient large‐scale eCLIP experiments, Gene W. Yeo group developed a multiplexing eCLIP method, antibody‐barcode eCLIP (ABC), hiring antibodies with DNA barcoded in 2023 (Lorenz et al., 2023). In the ABC method, the antibody against a given RBP was linked with a barcoded adaptor, which is latter ligated to the RBP‐crosslinked RNA fragments after immunoprecipitation step. As immunoprecipitation of several RBPs could be conducted in a single tube, and the time‐consuming steps of SDS–PAGE running and membrane transfer were also removed in the ABC method, it is highly suitable for large‐scale experiments. The above advancements have ameliorated CLIP into a robust, multiplex, and scalable methodology for RBP target discovery.

Although CLIP and its derivative technology are powerful, they still have intrinsic features, such as depending on antibodies, requiring a large number of materials, UV crosslinking‐induced bias, and lacking isoform sensitivity. To develop methods with a totally different principle, scientists paid attention to a fusion protein system between RBPs and RNA modification enzymes (RMEs). As well known, RNA undergoes a variety of biochemical modifications, including methylation, 3′‐end tailing, and deamination (Boccaletto et al., 2022). In the RBP–RME fusion protein system, the RNA modification domain can make marks on RBP‐bound RNAs, and the marks can be identified by RNA sequencing. Wickens's lab and Rosbash's lab pioneered the RME‐based RBP's target RNA identification (Lapointe et al., 2015; McMahon et al., 2016).

In late 2015, RNA tagging method was developed in yeasts to reveal target RNAs bound by proteins, even like Bfr1p that do not have typical RNA‐binding domains (Lapointe et al., 2015). RNA tagging involves the fusion of a given RBP with the poly(U) polymerase (PUP) from Caenorhabditis elegans. Because PUP alone is not able to tether to mRNAs, it hardly uridylates mRNAs without assistance of other partners. While the RBP–PUP fusion protein can specifically add multiple uridines to 3′ terminals of target mRNAs which are bound by the RBP in the fusion protein. At the similar time, TRIBE was invented in flies (McMahon et al., 2016), which expresses the fusion protein of a RBP and the catalytic domain of Drosophila‐originated RNA editing enzyme ADAR (dADARcd) in vivo. The resulting A‐to‐I (read as G) RNA base‐editing events reveal the RBP's targets. Then, HyperTRIBE was established in flies in 2018 by introducing hyperactive E488Q mutation to dADARcd (Xu et al., 2018). HyperTRIBE was adapted in mammals in 2020 by employing catalytic domain of human ADAR2 (hADAR2cd) E488Q (Herzog et al., 2020; Jin et al., 2020). Later, HyperTRIBE was used in plants with dADARcd E488Q (Arribas‐Hernandez, Rennie, Koster, et al., 2021; Arribas‐Hernandez, Rennie, Schon, et al., 2021; Zhou et al., 2021) and in yeasts with hADAR2cd E488Q (Piao et al., 2023). In 2019, Meyer developed deamination adjacent to RNA modification targets (DART‐Seq) that can detect global m6A modification without using antibodies (Meyer, 2019). In 2021, Brannan et al. applied Rattus norvegicus‐derived C‐to‐U RNA editing enzyme rAPOBEC1 to development of surveying targets by APOBEC‐mediated profiling (STAMP) in mammals (Brannan et al., 2021). The tactics mentioned above will shed light on RBP's interactome in various organisms (van Nostrand, Pratt, et al., 2020).

Here, we focus on the RME‐based methods and introduce the RNA modification enzymes that are used in developing these methods. Also, we review the basic principles, advantages and disadvantages, and applications of the representative techniques, TRIBE, RNA tagging, STAMP, and their derivative methods.

3. ADENOSINE DEAMINASES THAT ACT ON RNA (ADAR) AND ITS DERIVATIVE TRIBE

3.1. The A‐to‐I RNA deaminase ADAR family

The A‐to‐I (G) editing is one of the most common RNA modification events in higher eukaryotes, and it is catalyzed by ADARs. ADAR family acts on double‐stranded RNA (dsRNA) regions and employs a base‐flipping mechanism to site‐specifically deaminate adenosines within duplex RNAs (Matthews et al., 2016). The mRNA editing can alter pre‐mRNA splicing, coded protein sequences, base pairing between mRNAs and miRNAs, circRNA biogenesis, and so on, thereby regulating gene expression and contributing to protein diversity (Beghini et al., 2000; Chen et al., 2013; Ivanov et al., 2015; Kawahara et al., 2007; Rosenthal & Seeburg, 2012; Wang et al., 2013). RNA editing can alter the structure and function of RNA, thus playing crucial roles in biological systems. Therefore, ADAR‐mediated RNA editing is associated with many human diseases (Chen et al., 2023; Nishikura, 2010; Slotkin & Nishikura, 2013; Song et al., 2022), including cancer (Chen et al., 2013), neurological disorders (Nakahama et al., 2021; Yang et al., 2021), metabolic disorders (Knebel et al., 2024), and immune diseases (Cai et al., 2023; Quin et al., 2021).

Most species have more than one ADAR homolog, however, ADAR family is not found in several nonmetazoan eukaryotes, such as yeasts, fungi, and plants (Savva et al., 2012). ADAR1 and ADAR2 are universally expressed in almost all tissues in mammals, whereas the expression of mammalian ADAR3, Drosophila ADAR, and C. elegans ADAR1 are restricted mainly to nervous systems (Chen et al., 2000; Melcher et al., 1996; Palladino et al., 2000; Tonkin et al., 2002). Three members of ADAR family have been identified in human: hADAR1, hADAR2 (Hajji et al., 2022), and hADAR3(Wang et al., 2017).

The hADAR1 gene produces abundant protein isoforms of ADAR1 in human, hADAR1 p80, p110, and p150. ADAR‐1p80 localizes in the nucleolus in mouse and human while little is known about its editing capability (Figure 2; Lu et al., 2023; Yang et al., 2003). The short constitutively expressed isoform ADAR1–p110 is generated from multiple promoters, while the large protein isoform ADAR1–p150 can be induced by interferon through acting on the upstream interferon‐inducible promoter of ADAR1 (Figure 2; George & Samuel, 1999; Patterson & Samuel, 1995; Xing et al., 2023). The ADAR1–p150 is believed to be involved in cellular defense against viruses and other pathogens (Sarkis et al., 2018). ADAR1–p110 is located in the nucleus due to the absence of nuclear export signal (NES), and p150 (Hayashi & Suzuki, 2013) is mainly localized in the cytoplasm (Poulsen et al., 2001). Furthermore, mADAR1 has a conceptually similar gene organization and expression profile as the hADAR1 (George et al., 2005).

FIGURE 2.

FIGURE 2

The characteristics of RNA modification enzymes. The main orthologues of ADAR, TENT, and AID/APOBEC family are listed. The same type of domains in uniform colors are designated for once in the figure. dsRBD, double‐strand RNA binding domain; LIM, Lin28‐interacting module; NES, nuclear export signal; NLS, nuclear localization signal; NTD, nucleotidyl transferase domain; PAP‐assoc, PAP‐associated domain; ZF C2H2, C2H2‐type zinc finger domain; ZK, CCHC‐type zinc knuckle domain; ZK1 CCHC, CCHC‐type zinc knuckle domain 1.

The human ADAR2 produces four protein isoforms of hADAR2a, 2b, 2c, and 2d with slight differences in C‐terminal deaminase regions, and hADAR2a and 2b are the major isoforms among them (Hajji et al., 2022; Lai et al., 1997). In vitro editing assay using GluR‐B transcripts as substrates has shown that hADAR2a and 2b have higher editing activity than hADAR2c and 2d (Lai et al., 1997).

The hADAR2 is localized to the nucleus, which is mediated by its N‐terminal nuclear localization sequence (NLS; Figure 2). Also, Pin1 binds to the phosphorylated Ser/Thr‐Pro motif in hADAR2 to isomerize the proline at 33 aa, assisting hADAR2 nuclear import (Marcucci et al., 2011). In Pin1 knockout mouse fibroblasts, mislocalized hADAR2 is poly‐ubiquitinated and degraded in the cytoplasm by E3 ubiquitin ligase WWP2 mainly through Pro‐Pro‐x‐Tyr (PPxY) motif in both N‐terminus and C‐terminus of hADAR2 (Deffit & Hundley, 2016; Marcucci et al., 2011). So, the N‐terminus is important for hADAR2 nuclear localization and responsible for its instability in the cytoplasm. In addition, the change in subcellular localization of ADAR2 protein between the nucleoplasm and the nucleolus may serve as a mechanism to regulate endogenous ADAR2 substrate editing (Desterro et al., 2003; Sansam et al., 2003).

The hADAR3 was supposed to originate from the duplication of the hADAR2 gene, given the similarity in sequence and domain organization between ADAR2p and ADAR3p (Barraud & Allain, 2012; Chen et al., 2000). The ADAR3p is highly expressed in the brain, retains the ability to interact with dsRNA and ssRNA (single‐stranded RNA) but does not bring about any editing (Figure 2; Chen et al., 2000). Nonetheless, mADAR3 was illuminated to have function in memory and learning in mice (Mladenova et al., 2018).

3.2. The substrate binding and editing specificity of ADAR

ADAR‐editing adenosines are typically situated within complete or nearly‐complete duplex structures (Phelps et al., 2015), both dsRBD (double‐strand RNA binding domain) and deaminase domain engage in base grabbing and adjustment of these substrates (Figure 2 and 3a; Deffit & Hundley, 2016; Wang et al., 2017). The dsRBD specifically recognizes its binding sites by both shape and sequence of RNA (Stefl et al., 2006, 2010; Stephens et al., 2004). Deletion experiment proved the nonessential role of N‐terminus and dsRBD1 of ADAR2 in editing 15 bp RNA substrate (Macbeth et al., 2004). On top of that, ZBD (Z‐DNA binding domain) of ADAR1 is related to Z‐conformation duplexes (Herbert & Rich, 2001), while ADAR3 carries an N‐terminal arginine‐rich domain for picking up ssRNA (Chen et al., 2000).

FIGURE 3.

FIGURE 3

The detail of RME‐based RBP's target RNA profiling methods. (a) TRIBE, RNA tagging, and STAMP can determine different classes of target RNA by blending RBP with a special RNA modification enzyme (RME), comprising CePUP‐2, rAPOBEC1, or the catalytic domain of ADAR (ADARcd). (b) The model that hADAR2 binds to target under natural situations. The homodimer of hADAR2 binds to dsRNA, then local dsRNA structure is distorted, facilitating base‐flipping and deamination. (c) The model that rAPOBEC1 binds to ApoB mRNA under natural situations. A1CF homodimer (blue) together with rAPOBEC1 homodimer (pink) edit ApoB mRNA in the nucleus. ApoB mRNA includes regulatory sequence (black line), edited C (orange line), spacer sequence (blue line) and mooring sequence (gray line). N represents N‐terminal, C represents C‐terminal. (d) The preferred secondary structure and sequence in TRIBE/hyperTRIBE editing. (e) The nearest‐neighbor sequence preference for editing in different methods is showed. HyperTRIBE (dADAR E488Q) shows lower sequence bias for editing than in TRIBE (dADAR).

Although these nucleic acid‐binding domains primarily deposit ADARs near editing sites, the binding and editing are separate events. In other words, the binding isn't always followed by efficient editing, which only happens when appropriate editing sites are near the binding sites. Different deaminase domains, for example from hADAR1 and hADAR2, have their own substrate preference for editing (Kallman et al., 2003; Wang et al., 2018). ADAR2‐specific RNA‐binding loop and enzyme catalytic core in deaminase domain are in charge of the nearest‐neighbor preference for editing (Matthews et al., 2016; Wang & Beal, 2016). Homodimerization of ADAR on its target dsRNA enhances editing efficiency (Figure 3b; Barraud & Allain, 2012; Jaikaran et al., 2002; Poulsen et al., 2006; Stefl et al., 2010; Valente & Nishikura, 2007).

ADARs exhibit a multitude of editing sites due to nonstrict sequence specificity of reactions. There are the rules of ADAR catalysis: (1) The deaminase domain first catalyzes flipping‐out of adenosine from dsRNA. So, imperfect duplex structure is enough for deamination and A–C mismatch is dominant compared to A–U pair at the editing position (Kallman et al., 2003; Levanon et al., 2004; Wong et al., 2001). (2) Adenosines favor a 5′ neighbor of U or C and a 3′ neighbor of G for editing (Bazak et al., 2014; St Laurent et al., 2013; Wheeler et al., 2015). (3) Varied secondary structures of RNA, like bulges, hairpins, loops, and stems, have plus or minus effects on editing (Lehmann & Bass, 1999), which partly arises from specific recognition patterns of dsRBD (Stefl et al., 2006). (4) RNA tertiary structure is one of the origins of substrate specificity (Enstero et al., 2009).

To date, many amino acid mutants of ADARs have been identified in vitro (Matthews et al., 2016; Stefl et al., 2006) and in vivo screening (Kuttan & Bass, 2012; Poulsen et al., 2006; Wang et al., 2018, 2019; Wang & Beal, 2016; Wang, Havel, & Beal, 2015). The hyper active mutant hADAR1 E1008Q (Wang, Havel, & Beal, 2015) and hADAR2 E488Q (Kuttan & Bass, 2012) took the top spot, even though their catalytic efficiency depends on sequence‐context. Further engineering of ADAR proteins and inquiry into the edit‐aiding partner proteins will contribute to understanding of base editors and progress of RBP–RME methodology (Abudayyeh et al., 2019; Schneider et al., 2014; Stroppel et al., 2021).

3.3. TRIBE and HyperTRIBE identify RBP target transcripts based on ADAR2

TRIBE (McMahon et al., 2016) in flies entails the fusion of an RBP to the catalytic domain of Drosophila ADAR (RBP‐dADARcd), the fusion protein expressed in cells can put A‐to‐I editing tags on RBP‐bound mRNA in vivo (Figure 3a and 4a). Subsequently, these editing events can be pinpointed through high‐throughput mRNA sequencing and bioinformatics analysis. The samples of dADARcd‐only expression and no expression should be included in parallel so dADARcd‐mediated random editing and endogenous editing can be monitored and later be removed from targets sites. As mentioned above, N‐terminus of ADAR leads to its nuclear localization. In TRIBE, the N‐terminal dsRBD region of ADAR is substituted by RBP, so the localization of fusion protein depends on RBP. TRIBE makes 1–2 edits per target transcript on average, ~50% of edited sites are located within 100 nt from a CLIP peak, and ~80% are less than 500 nt from a CLIP peak (McMahon et al., 2016). Although TRIBE is not able to exactly map RBP‐binding positions, RBP‐binding motif can be identified by motif search with sequences around editing sites, for example with sequences ±100 nt of editing sites (McMahon et al., 2016). If the binding sites of a RBP are enriched in a specific region of mRNA, like 5′ or 3′ untranslated region (UTR), the sequences of enriched region from edited transcripts can be used for motif search (Jin et al., 2020). The motif search also needs to include a negative control composed of random sequences from non‐targets.

FIGURE 4.

FIGURE 4

Schematic representation of TRIBE, TRIBE‐ID, STAMP, RNA tagging, PIE‐seq, and HITS‐CLIP. (a and b) TRIBE and its variant TRIBE‐ID can study RNA–protein interactions based on A‐to‐I (G) RNA editing. In TRIBE‐ID, the treatment of cells with rapamycin can induce FRP‐FKBP protein hetero‐dimerization inside cells, thus RBP–ADARcd complex is transiently formed and edits RBP's target RNA. (c) STAMP determines RNA targets with C‐to‐U edit. (d) RNA‐tagging expresses the fusion protein of RBP with tailing enzyme PUP‐2 (up) and employs a unique procedure of library preparation (down). (e) PIE‐seq expresses APOBEC1‐RBP‐ADARcd fusion protein and the target RNA molecules should have both A‐to‐I (G) and C‐to‐U edits. (f) HITS‐CLIP combines CLIP with high‐throughput sequencing for the first time. The same type of elements in the figure are lettered for once. ssDNA and ssRNA are arranged from 5′ to 3′ if there is no comment. ADARcd, ADAR catalytic domain.

The RBP domain in TRIBE guides the fusion protein binding to RBP target mRNA, but TRIBE editing remains ADARcd‐derived strong preference for double‐stranded RNA region (Eggington et al., 2011; Macbeth et al., 2005; McMahon et al., 2016; Montiel‐Gonzalez et al., 2013; Phelps et al., 2015; Vogel et al., 2014; Vogel & Stafforst, 2014) especiclly a bulged A within dsRNA region (Eifler et al., 2013) even in the absence of ADAR dsRBDs (Figure 3d). In addition, the nearest‐neighbor of 5′ U/C and 3′ G are favored for editing (Figure 3d,e) while the exclusively single‐stranded regions and 5′ G are unsavory (Eggington et al., 2011; Kuttan & Bass, 2012; Porath et al., 2014).

Since TRIBE is limited by the low editing efficiency, as well as sequence and structural biases in editing, HyperTRIBE was developed by introducing hyperactive E488Q mutation into dADARcd (Rahman et al., 2018; Xu et al., 2018) in fly. The labeling efficiency improved more than 10‐fold in HyperTRIBE, sequence and structural biases were relieved, and false‐negative rate was dramatically cut. Subsequently, HyperTRIBE was adapted in mammals and yeasts with hADAR2cd‐E488Q (Herzog et al., 2020; Jin et al., 2020; Piao et al., 2023), in plants with dADARcd‐E488Q (Arribas‐Hernandez, Rennie, Koster, et al., 2021; Arribas‐Hernandez, Rennie, Schon, et al., 2021; Zhou et al., 2021).

In mammals, the catalytic domain from the protein isoform a of human ADAR2‐E488Q (hADAR2cd‐E488Q) was employed to make HyperTRIBE fusion protein (Herzog et al., 2020; Jin et al., 2020), beacuse the protein isoforms of hADAR2a and hADAR2b have more robust activity than those of hADAR2c and hADAR2d. Jin and colleagues employed HyperTRIBE to prove that eIF4E‐binding protein (4E‐BP) is associated with a specific set of mRNAs through eIF4E, under mTOR‐inhibited conditions in vivo in both Drosophila and mammals (Jin et al., 2020). Although some in vitro evidence had previously supported that 4E‐BP and eIF4E can bind to capped mRNA, 4E‐BP‐hyperTRIBE provided the first evidence confirming this association in vivo and explained how 4E‐BP preferentially inhibits translation of a subset of mRNA with TOP motif (Jin et al., 2020; Yang et al., 2022). Also, it is worth noting that TRIBE is a semi‐quantitative method and can be used to estimate relative protein–RNA interaction under different conditions (Jin et al., 2020).

With hyperTRIBE and other methods, Herzog and colleagues specified that TDP‐43 targets the mRNAs of several upstream regulators of CREB, thereby regulating CREB activity and dendritic branching. TDP‐43 dysfunction reduces transcription factor CREB activity and dendritic branching, so causes neurodegenerative diseases (Herzog et al., 2020).

Interestingly, MS2–TRIBE system could monitor spatial organization of transcription. In this system, MS2‐binding sites were inserted into an endogenous gene locus. During the transcription of the gene, ectopically‐expressed MCP (MS2 coat protein)‐ADARcd fusion protein can bind to the MS2 stem‐loop RNA region and make tags on spatially adjacent RNA, which gives the information about nuclear organization of transcription (Biswas et al., 2020). Anyway, even though hyper‐dADARcd can be used in mammalian HyperTRIBE (Biswas et al., 2020), its editing efficiency is weaker than hyper‐hADAR2cd in mammals (Jin et al., 2020). Thus, it is highly recommended to use hADAR2cd E488Q in mammalian HyperTRIBE.

In plants, several editing enzyme activities were compared for setting up HyperTRIBE, including adenosine deaminase (dADARcd E488Q and TadA of E.coli origin) and cytidine deaminase (rAPOBEC1, PmCDA1, and AtCDA1). Among them, the dADARcd‐E488Q performed well (Zhou et al., 2021) and was used to identify target mRNAs of a RBP AtUBP1c, and the m6A‐binding protein ECT2 and ECT3 in Arabidopsis (Arribas‐Hernandez, Rennie, Koster, et al., 2021; Arribas‐Hernandez, Rennie, Schon, et al., 2021; Martinez‐Perez et al., 2023; Zhou et al., 2021). In recent studies,the targets of Khd1p and Bfr1p were identified in S. cerevisiae by HyperTRIBE and yeast HyperTRIBE was successful with hADAR2cd‐E488Q (Piao et al., 2023).

The TRIBE approach facilitates the identification of cell‐specific (e.g., neurons) RBP–RNA interactions with minimal amounts of RNA (McMahon et al., 2016), which can be obtained by manual cell sorting, fluorescence‐activated cell sorting (FACS), or laser microdissection for plant tissues (Burjoski & Reddy, 2021). The transient expression of the fusion protein, for example, by inducible transcription system, is a better choice since long‐term expression will affect cell physiology too much and the edited targets are easily degraded by nonsense‐mediated mRNA decay (NMD) pathway. Another choice can be inducible protein dimerization between RBP and ADARcd like in TRIBE‐ID (Seo & Kleiner, 2023), which requires the expression of two fusion proteins, RBP–FRB (FKBP and rapamycin‐binding) and FKBP‐ADARcd (Figure 4b). FRB and FKBP interaction, or hetero‐dimerization, can be induced by rapamycin, so RBP's target RNA is only edited after adding rapamycin to cells in TRIBE‐ID. This method can avoid side effects from long‐term RBP‐ADARcd expression and quantify dynamic RNA–protein interactions in the form of inducible.

In summary, HyperTRIBE is a convenient, semi‐quantitative method, and is independent of antibodies. It has the advantage of examining RBP's targets in a small number of cells, being good to be used in the tissues with heterogenous cell types, as well as in organisms with low UV‐crosslinking efficiency because of cell wall and others. In addition, HyperTRIBE can determine indirect RBP–RNA interactions and be applicable in several representative organisms, including mammals, flies, and plants, by employing different origins of ADARcd‐E488Q.

4. TERMINAL NUCLEOTIDYL TRANSFERASE AND ITS DERIVATIVE RNA TAGGING

4.1. Terminal nucleotidyl transferase family

RNA tailing, nontemplated nucleotide addition in RNA 3′ termini, is a major type of mRNA modification (Doma & Parker, 2007; Martin & Keller, 2007; Munroe & Jacobson, 1990; West et al., 2006) and a common and conserved RNA processing pathway. Canonical poly(A) polymerase (PAP) adds poly(A) tails to the 3′ ends of mRNA during mRNA maturation in a transcription‐coupled manner in the nuclei of eukaryotic cells. Canonical PAP consists of an N‐terminal nucleotidyl transferase domain (NTD) and a C‐terminal RNA‐binding domain (RBD) with a nuclear localization signal (Liudkovska & Dziembowski, 2021). Terminal nucleotidyl transferases (TENTs) catalyze adenosylation, uridine acidification, guanylation, and cytidine acidification at the 3′ ends in the nucleus or cytoplasm (Norbury, 2013). TENT genes are well conserved in metazoan, including vertebrates, flies, and worms. Vertebrate TENT family can be sub‐classified into non‐canonical poly(A) polymerase (ncPAP) and terminal uridylyltransferase (TUTase) based on substrate preference for ATP or UTP. Some ncPAPs, like hTENT4A/4B, have broader substrate tolerance so generate A/G mixed‐tailing (Lim et al., 2018; Wen et al., 2023). It was suggested to refer to all vertebrate ncPAPs and TUTases by their respective TENT family names in the updated nomenclature and they typically include NTD and PAP‐associated domain (PAPd) but not RBD. The NTD and PAPd together comprise the core catalytic domain of TENT (Liudkovska & Dziembowski, 2021; Yu & Kim, 2020). In addition, CCA‐adding enzymes, poly(UG) polymerase (CeMUT‐2), and the enzyme with broad specificity (CeF31C3.2) were found in worms and fungi and all these RNA tailing enzymes other than canonical PAP were also called as noncanonical ribonucleotidyl transferases (rNTases; Preston et al., 2019).

4.2. Terminal uridylyltransferase subfamily

Terminal uridylyltransferases (TUTases) catalyze mono‐ and oligo‐uridylation of aberrant snRNA, miRNA, snoRNA, rRNA, 7SL, tRNA, Y and vault RNA, mRNA, and viral RNA in vivo, thereby playing key roles in their biogenesis, regulation, stability, and function (Liudkovska & Dziembowski, 2021; Warkocki et al., 2018; Yashiro & Tomita, 2018). They exist widely in mammals, plants, trypanosomes, worms, fruit flies, fission yeast, and other organisms. The hTUT1 (hTENT1) and hTUT4/7 (hTENT3A/3B) in human, Tailor (dTENT3) in fly, USIP1 (CeTENT1), and PUP‐1/2/3 (CeTENT3) in C. elegans, Cid1 and Cid16 in fission yeast have uridine acidification activity (Liudkovska & Dziembowski, 2021; Pisacane & Halic, 2017; Yu & Kim, 2020). Human TUT4/7 and CePUP‐1 (CDE‐1) are large multi‐domain enzymes, while the others, except for hTUT1, are simple and small proteins (Figure 2). PUP‐1/2/3 are essential for C. elegans germline development, PUP‐1 and PUP‐2 promote germline development redundantly under conditions of heat stress (Li & Maine, 2018). Similarly, TUT4/7 are important for gametogenesis and early embryonic development in mice (Morgan et al., 2017, 2019).

4.3. The processivity and substrate specificity of TUTases

The hTUT1 and its C. elegans ortholog USIP1 function in spliceosomal U6 snRNA biogenesis and regulation by shaping oligo(U) tails of U6. The hTUT1 carries RNA‐recognition motif (RRM), which possibly decides substrate specificity and gives enzyme processivity so binding to its substrates enough time for oligo‐uridylation (Liudkovska & Dziembowski, 2021; Ruegger et al., 2015; Trippe et al., 2006).

hTUT4/7 and its worm ortholog PUP‐2 (Poly(U) Polymerases‐2) participate in pre‐miRNA let‐7 degradation by Lin28‐dependent oligo‐uridylation of pre‐let‐7, making sure clearance of let‐7 in early animal development (Heo et al., 2009; Lehrbach et al., 2009). On the contrary, in the absence of Lin28 like in differentiated cells from late development stages, hTUT4/7 could only add mono(U) to pre‐let‐7 and promote the processing of pre‐let‐7 by Dicer (Heo et al., 2012). PUP‐2 possesses typical NTD and PAPd, but lacks any RNA‐binding domains, while in addition to NTD and PAPd, hTUT4/7 also consist of CCHC‐type Zink knuckle domain (ZK) and C2H2‐type Zink finger domain (ZF). The mono‐uridylation conformation requires NTD and PAPd of hTUT4/7 while oligo‐uridylation conformation needs not only NTD, PAPd, ZK2 of hTUT4/7 but also ZK of Lin28 (Figure 2; Faehnle et al., 2017). Thus, Lin28 is essential for hTUT4/7 and PUP2 processivity and oligo‐uridylation.

Interestingly, Tailor in fly oligo‐uridylates and degrades mirtron pre‐miRNA, which ends with 3′‐AG (Bortolamiol‐Becet et al., 2015; Reimao‐Pinto et al., 2015). The structural evidence indicates that R327 and N347 in Tailor core catalytic domain cooperatively contribute to substrate preference for 3′‐G and R327 possibly facilitates oligo‐uridylation (Cheng et al., 2019; Kroupova et al., 2019). So, in the case of Tailor, the core catalytic domain can support its processivity at least partly.

The mRNA and viral RNA are also the targets of TUTases in several organisms (Wen et al., 2023). The cellular mRNAs (Lim et al., 2014) and mouse hepatitis virus (MHV) genomic RNAs with short A‐tailed (<25 nt; Gupta et al., 2023), as well as the mRNAs of influenza A virus (IAV) are targeted by TUT4/7 for mono‐/oligo‐uridylation in mammals (Le Pen et al., 2018). The Orsay virus (OrV) RNA genome ends with mono‐U, CDE‐1 (PUP‐1) catalyzes mono‐uridylation of viral genome and resulted viral RNA with di‐U tails is degraded in OrV‐infected C. elegans (Le Pen et al., 2018). Some polyadenylated mRNAs are also uridylated by Cid1 and undergo mRNA decay pathway in S. pombe (Rissland & Norbury, 2009). Taken together, RNA oligo‐U‐tails trigger RNA degradations in many cases (Lubas et al., 2013; Malecki et al., 2013; Rissland & Norbury, 2009).

4.4. An assay TRAID‐seq for defining the TENT processivity and nucleotide preference in vivo

Apart from complicated endogenous mechanisms for substrate specificity, nucleotide preference, and enzyme processivity, a MS2‐based tethering and tailing assay TRAID‐seq (tethered rNTase activity identified by high‐throughput sequencing) performed in yeast can reveal the nucleotide preference and enzyme processivity own by enzyme‐self. In TRAID‐seq, the fusion protein between TENT and MS2 coat protein (MCP) was co‐expressed in yeast with the modified tRNA reporter bearing MS2‐binding sites (MBS) to specify TENT‐own biochemical activity in vivo (Preston et al., 2019). These can provide valuable information for TENT enzyme engineering and technology development, such as sequencing library preparation, molecular interaction detection, and hyper‐active enzyme screening.

In summary, mono‐ and oligo‐uridylation participate in biogenesis, decay, and function of coding and noncoding RNA. In many cases, oligo‐U tags often trigger spatiotemporally‐controlled RNA degradation, as well as abnormal RNA and viral RNA clearance, which plays key roles in embryonic development, cell cycle, circadian rhythm, immunity, and diseases (Beta & Balatsos, 2018; Chang et al., 2018; Le Pen et al., 2018; Liu et al., 2020; Reyes & Ross, 2016; Song et al., 2023).

4.5. RNA tagging identifies RBP target transcripts based on CePUP2 in yeast

A method called RNA tagging was developed, in which RBP was fused with terminal uridylyltransferase CePUP2. The coding sequences of CePUP2 were inserted into downstream of a native RBP gene locus at S. cerevisiae genome thus the expression of fusion protein was under control of the endogenous RBP promoter (Figure 4d; Lapointe et al., 2015). CePUP‐2 carries only a minimal catalytic domain but not any RBD (Figure 2), and can add pretty pure U‐tails with high processivity in vivo when tethered to RNA in TRAID‐seq (Preston et al., 2019). RNA tagging employed CePUP‐2 to leave a covalent mark of U‐tails on RBP's target mRNAs and was applied to profile targets of two RBPs, Puf3 and Bfr1. About 50% sequencing reads in RNA tagging could be uniquely aligned to yeast genome, and hundreds to a thousand of targets were identified with U‐tag lengths ranging from 1 to more than 10 nt, meaning that these uridylated mRNAs were stable sufficient to be observed. These targets were also ranked according to U‐tag count and U‐tag length to quantify the affinity between RBP and target RNAs.

There are still some shortcomings in RNA tagging. At first, U‐tagging might be biased toward short A‐tailed mRNAs. It was reported that different organisms have varying poly(A) tail lengths, approximately 30 nt in yeast, 50 nt in Arabidopsis and Drosophila, and 70 nt in mammals on average according to PAL‐seq results (Subtelny et al., 2014). Also, adding oligo‐U to the end of poly(A) tail possibly induces target RNA degradation in some species (Lim et al., 2014). These will challenge adapting RNA tagging to other organisms. Second, RNA tagging does not provide information about the location of RBP binding sites, and has only been used in S. cerevisiae and not yet tried the RBPs, whose binding sites on mRNA far away from the ends of poly(A) tails to date.

Nevertheless, RNA tagging has avoided some complicated steps, such as cross‐linking, immunoprecipitation, radioactive‐labeling, gel‐running, turning to examination of uridylated RNA terminal. From the application aspect, RNA tagging can be applied for an individual type of cells or tissues, and for proteins directly and indirectly associated with mRNAs. This method is simple, reproducible, sensitive, and high‐throughput, exhibiting high performance in yeasts.

5. CYTIDINE DEAMINASE APOBEC AND ITS DERIVATIVE STAMP

5.1. The cytidine deaminase AID/APOBEC family

The AID/APOBEC polynucleotide cytidine deaminases can lead to C‐to‐T DNA mutation as well as Cytosine‐to‐Uracil (C‐to‐U) RNA editing. AID/APOBEC family includes activation‐induced cytidine deaminase (AID), ApoB mRNA editing enzyme catalytic‐1 (APOBEC1), APOBEC2, APOBEC3, and APOBEC4. Among them, APOBEC3 is specific to mammals, and is composed of a single gene in mice but expanded in primates into seven members, APOBEC3A (A3A), A3B, A3C, A3D, A3F, A3G, and A3H. APOBEC1 is conserved in tetrapods, AID and APOBEC2 are in vertebrates, whereas the orthologues of the mostly conserved APOBEC4 have even been found in invertebrates (Pecori et al., 2022).

According to substrate specificity, these members can be divided into two groups. The first group, C‐to‐U RNA editing enzymes APOBEC1, A3A, and A3G, can deaminate single‐stranded RNA (ssRNA). Peculiarly, they reserved the activity to deaminate single‐stranded DNA (ssDNA), which introduces mutations in genome and viral cDNA. As the first group enzymes can possibly deaminate both RNA and DNA, they were called as generalists. The second group, C‐to‐T DNA mutation enzymes AID and APOBEC2, can only catalyze the deamination of DNA but not RNA, so they are called as specialists. In fact, many members have not been categorized because of limited information about them (Pecori et al., 2022). Surprisingly, APOBEC2 was reported to act on DNA deamination‐independently (Powell et al., 2015). Humans carry the most diverse AID/APOBEC family members, including AID, APOBEC1, APOBEC2, seven APOBEC3 genes, and APOBEC4 (Salter et al., 2016). All members in this family are characterized by the catalytic domain of CMP‐dCMP‐type deaminase domain (Figure 2). Human A3B, A3D, A3F, and A3G have two deaminase domains, while the other members have only one. Some members additionally carry protein–protein interaction region and sub‐cellular localization signal, which are important for their functionality.

The loss of function studies has illustrated the importance of AID/APOBEC family roles. Mouse APOBEC1 functions in monocytic innate immune cells and nervous system by editing hundreds of target mRNAs (Cole et al., 2017; Rayon‐Estrada et al., 2017). APOBEC2 plays conserved roles in animal cardiac and skeletal muscle (Etard et al., 2010; Sato et al., 2010). APOBEC3 restricts viral infection and the transposition of genomic mobile elements through deaminating their cDNA intermediates (Bogerd et al., 2006; Harris & Dudley, 2015; Refsland & Harris, 2013). AID targets the genome regions of immunoglobulin genes in B cells thereby increasing antibody diversification (Muramatsu et al., 2000).

5.2. The substrate specificity of AID/APOBEC enzyme

Rattus norvegicus APOBEC1 (rAPOBEC1), the first member of AID/APOBEC family, was originally discovered in the ApoB mRNA editing event, in which rAPOBEC1 catalyzes C‐to‐U conversion on ssRNA substrates (Figure 2 and 3a; Navaratnam et al., 1993). The editing on ApoB pre‐mRNA generates a truncated protein isoform with distinct functions in lipid transport. Although APOBEC1 alone can catalyze the deamination of single cytosine in ApoB mRNA both in vivo and in vitro, the accessory protein APOBEC1 complementation factor (A1CF) and RNA‐binding motif protein‐47 (RBM47) are required for the efficient editing of ApoB mRNA (Blanc et al., 2019). The minimum required sequence for editing ApoB mRNA is 26 nt long (Figure 3c), which is highly conserved from marsupials to humans. Three cis‐acting elements are required for site specific deamination of ApoB mRNA (Backus & Smith, 1992; Smith et al., 2012; Soleymanjahi et al., 2021): an regulatory sequence upstream of editing C‐site (Figure 3c, black line), which regulates the editing efficiency (Maris et al., 2005); the flanking A/U at the 5′ and 3′ of the C‐site as well as an spacer in ~4 nt long (Figure 3c, blue line); the most important downstream 11 nt mooring sequence (Figure 3c, gray line), recognized by the accessory protein A1CF. In addition, extending the substrate's length to 102 nt enhances the editing efficiency (Wolfe et al., 2019).

Hundreds of transcripts are edited by APOBEC1 in mouse liver, intestine, and so on, and the ±4 nt AU‐rich flanking sequences are found adjacent to the APOBEC1 editing sites (Blanc et al., 2014; Rosenberg et al., 2011). The ~80% editing sites of APOBEC1 are located in 3′ UTR, some of the editing might affect the binding of miRNA to mRNA 3′ UTR so modulate the mRNA translation (Rayon‐Estrada et al., 2017). The substrate selection of APOBEC1 extremely depends on its co‐factor mainly RBM47 but also A1CF in mice, as illustrated by single or double knockout experiments of the co‐factors. RBM47 and A1CF specifically recognize mRNA, and directly recruit APOBEC1 protein for efficient mRNA editing. Each co‐factor has its own targets, they only share part of targets (Blanc et al., 2019; Pecori et al., 2022). Interestingly, the stem‐loop structure and cis‐acting elements similar to ApoB mRNA (Figure 3c) were also observed in these APOBEC1 substrates, and slight distinct features in the elements were expected to be recognized by different co‐factors (Soleymanjahi et al., 2021).

APOBEC1 is distributed in both the nucleus and the cytoplasm, but the nuclear distribution was observed only in cells that were capable of editing ApoB mRNA. It was reported that editing of ApoB mRNA almost occurred in the nucleus. The localization of APOBEC1 is regulated by both an N‐terminal nuclear localization signal (NLS) and a C‐terminal leucine‐rich nuclear export signal (NES), mutations on these regions impair substrate editing (Figure 2; Chester et al., 2003; Teng et al., 1999). Also, its localization may depend on the transport of accessory proteins, so C‐terminal hydrophobic domain of APOBEC1 related to intramolecular interactions further affects its subcellular localization (Wolfe et al., 2020).

The catalytic activity of AID/APOBEC family is provided by one or two conserved core zinc‐dependent cytidine deaminase domains. The catalytic domain comprises five β‐sheets (β1–β5) and is surrounded by six α‐helices (α1–α6), as well as 10 loops (L‐1 to L‐10) connecting them. The active site, a zinc finger motif H‐X‐E‐X23‐28‐P‐C‐X2‐4‐C (His‐Glu and Cys‐Cys, HECC), is highly conserved in the family (Revathidevi et al., 2021). The U‐shaped substrate binding groove on AID/APOBEC, defined by loops 1, 3, 5, and 7 surrounding the active site, interacts with U‐shaped 5–6 nt single‐stranded substrates and largely determines substrate specificity originated from enzyme itself. The amino acid variations in substrate binding groove might explain distinct preference for flanking sequence and DNA/RNA in different members of the family (Bohn et al., 2017; Kouno et al., 2017; Matsuoka et al., 2018; Rathore et al., 2013; Shaban et al., 2018; Shi et al., 2017; Wolfe et al., 2020). In addition, it was reported that the C‐terminal residues of rAPOBEC1 (196–210 aa and 221–229 aa) take part in protein homo‐dimerization (Teng et al., 1999), and C‐terminus‐deleted mAPOBEC1 (N1‐196 aa) neither dimerizes nor edits ApoB mRNA (Oka et al., 1997; Teng et al., 1999), RNA molecules are in turn required for dimerization of APOBEC1 (Ikeda et al., 2016).

5.3. STAMP identifies RBP target transcripts based on rAPOBEC1 in mammals

DART‐seq (Deamination Adjacent to RNA Modification Target; Meyer, 2019) first mingled rAPOBEC1 with the m6A‐binding YTH domain to map m6A sites on mRNAs. Later, Gene W. Yeo group developed STAMP approach to demonstrate the RBP‐binding sites on RNA isoform‐specifically at single‐cell resolution (Figure 3a and 4c; Brannan et al., 2021). STAMP entirely follows the theory of immunopurification‐free identification of RBP targets by the fusion protein of full‐length rAPOBEC1 and a given RBP, but using extremely low or even single‐cell input. As RNA‐seq has thrived on deciphering RNA biology cell‐type specifically and isoform‐specifically with tools of spatial‐omics, single‐cell sequencing (scRNA‐seq), and long‐read RNA‐seq (Figure 5a; Stark et al., 2019), STAMP employed these tools. In a single pooled experiment, STAMP drew support from single‐cell capture and third‐generation long‐read sequencing methods and identified RNA–protein interactions in cell‐type‐specific and RNA isoform‐specific manners (Figure 5a). Also, the editing sites in RBFOX2‐STAMP were found within ±100 bp of expected RBFOX2 binding‐sites.

FIGURE 5.

FIGURE 5

Schematic representation of methods for detecting isoform‐specific mRNA (TGS), poly(A) mRNA (SGS), and nascent RNA (Run‐on method and RNA pol‐II‐directed method). (a) The library preparation procedure for third or second‐generation sequencing (TGS or SGS) is simulated from the kits instruction of PACBIO® and illumina® respectively. Alternatively, rRNA can also be depleted through cleaving rRNA hybridized to DNA probes with RNaseH. (b and c) Run‐on and RNA pol‐II‐directed methods make no mutations on RNA and fit for nascent RNA target detection in RBP–RME method. Amalgamating with TRIBE is described as a model but it is also applicable to STAMP. The same kind of elements are presented for once in the figure, specific nucleotides are enlarged for convenience. BrU, Bromouridine.

Furthermore, ribosome‐subunit STAMP (Ribo‐STAMP) bearing ribosomal subunit and rAPOBEC1 fusion offers ribosome‐interacting transcriptome (Brannan et al., 2021). The research tried two small ribosomal subunits RPS2 and RPS3, and RPS2‐STAMP exhibited better correlation when comparing EPKM (edited reads per kilobase of transcripts per million mapped reads) from Ribo‐STAMP with RPKM from ribosome‐protected fragments (RPFs) of Ribo‐seq. The editing sites in RPS2‐STAMP were mainly found in CDS but also in 3′ UTR, and 3′ UTR edits were confirmed to be effective in the quantification of translation efficiency and might partly arise from association small‐ribosomal subunit with mRNA even after translation termination. It was previously reported that the inhibition of mTOR pathway by its inhibitor like Torin‐1 represses translation preferentially for the mRNAs with 5’ TOP motif, which is enriched in the mRNAs of translational machinery, and further represses translation generally (Yang et al., 2022). RPS2‐STAMP could capture both general and specific translation repression after 72‐h treatment of mTOR inhibitor Torin‐1 (Brannan et al., 2021). While the requirement of time lag (12–24 h) to yield enough detectable edits dampens the capacity to monitor translational responses timely, and also nonsense or frameshift mutations cause artificial observation in Ribo‐STAMP. There is a demand for keeping expression levels of STAMP transgenes to an endogenous level to weaken false‐positive results.

The stem‐loop structure in APOBEC1 endogenous mRNA substrates is more likely needed for specific recognition by APOBEC1 co‐factors, the editing C‐sites are mainly positioned in ssRNA regions of the stem‐loops, such as loops, bulges, and tails (Soleymanjahi et al., 2021). The protein structures of core catalytic pockets of APOBEC1, A3A, and A3G indicate that they prefer binding to U‐shaped ssDNA/RNA (Kouno et al., 2017; Pecori et al., 2022). Therefore, STAMP is expected to edit ssRNA regions and does not require dsRNA structure for editing, even though this point needs further investigation. Anyway, the cytosine editing in STAMP still reserved the preference for 5′ and 3′ A/U rich flanking sequences (Figure 3e). The APOBEC1 alone and its co‐factor‐derived background edits can be removed by control STAMP expressing rAPOBEC1 only. When STAMP is applied in the nucleus, the possibility of genomic DNA editing should be considered. In addition, the subcellular localization of APOBEC1 is tightly regulated by its NLS, NES, and interacting co‐factors, therefore, it is necessary to check whether RBP‐rAPOBEC1 is properly localized prior to STAMP. STAMP has only been used in mammals to date (Brannan et al., 2021; Ghashghaei et al., 2024; Nicholson‐Shaw et al., 2022; Tao et al., 2023), and has shown unworkable in Drosophila (Abruzzi et al., 2023) and plant (Loeser et al., 2024), is still uncertain if it is applicable to other species.

5.4. Combined tribe with stamp

The principles of TRIBE (McMahon et al., 2016) and STAMP (Brannan et al., 2021) methods are similar, they fuse RBP with two different types of deaminase domain. TRIBE performs A‐to‐I (G) edits, and STAMP carries out C‐to‐U edits. Given the similarity in their underlying principles, it is feasible to combine these two methods for using together.

Abruzzi and colleagues compared hyperTRIBE and STAMP in both mammals and flies (Abruzzi et al., 2023). In human cells, both HyperTRIBE (TDP‐43‐hADAR2cd E488Q) and STAMP (TDP‐43‐rAPOBEC1) worked well with comparable performance regarding the numbers of identified editing sites per million reads and editing site reproducibility, and 70% of STAMP target genes were also observed in hyperTRIBE. Average editing percentage was higher in hyperTRIBE and mean number of editing sites per transcript was higher in STAMP (Abruzzi et al., 2023). The flanking sequence preference for editing in two methods was different as reported (Figure 3e). The cross‐checking with two methods would be more successful in identifying high‐confidence targets. Two methods (RBP‐dADARcd E488Q vs. RBP‐rAPOBEC1) were also compared in Drosophila cells, HyperTRIBE worked well as known but STAMP produced only a few editing sites, a similar level to the control editing. The same phenomenon was observed in plant (Loeser et al., 2024). As TRIBE and STAMP can study synergy and/or antagonism of two RBPs, it should put effort into adjusting STAMP in the species other than manmals.

Recently, PIE‐Seq was proposed (Ruan et al., 2023). It is a method of enhancing target fidelity by using double deaminases (Figure 4e). This method integrates an RBP into both hADAR2cd and rAPOBEC1. The PIE‐seq successfully identified target RNAs and binding motifs of 25 human RBPs, and was further applied to mouse brain for cell‐type specific profiling RBP–RNA interaction by single‐cell sequencing. PIE‐Seq reduces biases from enzyme preference for editing sequence context and secondary structure. Compared to a single deaminase, the dual deaminases of PIE‐seq enhance the fidelity of targets.

The TRIBE‐STAMP method (Flamand et al., 2022) applies TRIBE and STAMP simultaneously to two RBPs and observes whether two RBPs work on a single RNA molecule and how they do it (Owens & Liu, 2022). This method involves fusing one RBP with hADAR2cd‐E488Q (RBP1‐hyperTRIBE) and the other RBP with rAPOBEC1 (RBP2‐STAMP) and co‐expresses two constructs in mammalian cells. The authors applied TRIBE‐STAMP to the cytoplasmic m6A readers YTHDF1 (DF1), DF2, and DF3 to answer a long‐standing argument in the field (Flamand et al., 2022): One model had proposed that three paralogs have distinct pools of target genes and control their target RNAs in different ways; another model had suggested that DF1, DF2, and DF3 have redundant function in driving mRNA decay (Flamand et al., 2023). TRIBE‐STAMP was conducted with several combinations of m6A readers, and revealed that DF1, DF2, and DF3 share many target mRNAs and some target molecules could be bound by more than one DF paralogs throughout their lifetime (Flamand et al., 2022). They also observed that DF1 has more unique targets than the others and proposed the model: DF1 binding might not trigger rapid decay of mRNA molecules bound to it, so these mRNAs have chance to bind by DF2 or DF3 later; the mRNA molecules bound by DF2 might undergo rapid degradation so it has less unique targets.

Overall, combining TRIBE and STAMP could enhance detection accuracy, and be feasible to simultaneously study multiple co‐working RBPs regardless of the same complex or functioning redundantly, which has a wide range of applicability.

5.5. SCREENING RNA BASE‐EDITORS FOR DETECTION OF RBP–RNA INTERACTION

With the development of methods based on editing enzymes for detecting RBP–RNA interactions, it is particularly important to understand varying editing enzyme characteristics more thoroughly for application. The experimental and computational framework of PRINTER (Protein–RNA Interaction‐based triaging of enzymes that edit RNA; Medina‐Munoz et al., 2024) was designed in mammalian cells to evaluate the editing efficiency and specificity based on a MS2 stem loop‐MCP tethering reporter system. They screened 31 RNA base‐editors (rBEs) with A‐to‐I and/or C‐to‐U editing activity, and successfully identified 7 rBEs with expanded editing ability. These editors include C‐to‐U editors evoAPOBEC1 and APOBEC3A (Y132G/K30R), A‐to‐I editors TadA‐8e, TadA‐7.10, TadA‐7.10 (V82G), A‐to‐I and C‐to‐U dual editors ADAR2dd (R) and ADAR2dd (R‐S). Among them, APOBEC3A (Y132G/K30R) and TadA‐7.10 (V82G) exhibited the greatest editing activity and signal‐to‐noise ratio in the C‐to‐U and A‐to‐I rBEs, respectively. To identify two RBP targets concurrently, the APOBEC1 (C‐to‐U) and TadA 7.10 (V82G) (A‐to‐I) pair was recommended for tagging two RBP targets respectively in the same cells (Medina‐Munoz et al., 2024). In PRINTER, the activity of Drosophila ADAR E488Q (hyper‐dADAR/dADAR‐E488Q) was also tested and it showed low editing activity in mammalian cells. This result is consistent with the previous report that hyperTRIBE conducted with dADARcd‐E488Q in mammalian cells has low signal‐to‐noise ratio (Jin et al., 2020). Actually, hyperTRIBE in mammalian systems should carry out using catalytic domain of human ADAR2 E488Q (hADAR2cd‐E488Q) to achieve high signal‐to‐noise ratio (Jin et al., 2020). Although PRINTER was only conducted in mammalian system, the provided information will be valuable for adapting TRIBE and STAMP methods in varying species.

When applying rBE‐dependent detection methods, it is necessary to consider factors such as the preference of editing enzymes for flanking sequence and structure, and editing accuracy. In addition, it should be noted that higher enzyme activity may capture more RBP targets but also introduce higher noise, the optimal signal‐to‐noise ratio should be achieved as much as possible.

6. CONCLUSIONS

Despite the utility and strength of staple methodologies of RIP and CLIP, there are limitations in materials and conditions required, antibodies and a bulk of RNA quarried, native conditions of RIP raising nonphysiological interactions in vitro (Hafner, Landthaler, Burger, Khorshid, Hausser, Berninger, Rothballer, Ascano, et al., 2010; Konig et al., 2010; Licatalosi et al., 2008), and UV‐crosslinking desired to be enhanced (Mili & Steitz, 2004; Riley et al., 2012; Zhao et al., 2010), especially for RBP–dsRNA interaction mapping (Chi et al., 2009; Ricci et al., 2014). After overcoming all difficulties, fragmentation strategy of CLIP hides transcript isoform information and the detection of transient interaction in CLIP confounds false positives with biologically meaningful interaction (Figure 4f; Konig et al., 2010). Recent studies have made efforts to create fundamentally different methodology utilizing fusion of RNA modification enzyme to RBP of interest (RBP–RME), named TRIBE, RNA tagging, and STAMP (Jin et al., 2020; Lapointe et al., 2015; McMahon et al., 2016; Medina‐Munoz et al., 2020; Nguyen et al., 2020). In addition, Kleiner team developed RNA‐mediated activity‐based protein profiling (RNABPP; Dai et al., 2021, 2023), which is a metabolic labeling strategy based on reactive modified nucleoside probes to profile RNA modification‐associated proteins in living cells. This kind of approach will widen our knowledge about RME and further advance RBP–RME methodology.

To design an RBP–RME method, several things should be kept in mind. First, available RNA modification enzymes are varying according to organisms: dADARcd E488Q in fly and plant, hADAR2cd E488Q and CePUP‐2 in yeast, hADAR2cd E488Q, rAPOBEC1, APOBEC3A (Y132G/K30R), TadA‐7.10 (V82G), and others in mammals (Figure 3a). Second, the protein localization of RBP–RME should resemble natural RBP localization, not RME. It needs to avoid RME‐mediated mislocalization of RBP–RME. Third, transient and/or endogenous level expression of RBP–RME is recommended for reducing artificial observation. Fourth, when simultaneously exploring several RBPs, whatever they coordinate in one complex or function redundantly, a pair of RMEs are required like in TRIBE‐STAMP method (Medina‐Munoz et al., 2024): one with A‐to‐I editing and the other with C‐to‐U editing activities. The utilization of a TRIBE and STAMP pair can achieve target detection of two RBPs at a single molecular level. Fifth, Ribo‐STAMP advanced translation monitoring methodology. Since it is challenging to set up traditional translation exploration tools, like polysome profiling and Ribo‐seq, Ribo‐STAMP could be an option for the quantitative study of translation and applicable for the experiments of a single cell level (Einstein et al., 2021).

According to the purpose of the study and expected RBP‐binding regions in mRNAs, TRIBE and STAMP should choose different library preparation methods and high‐throughput sequencing. Regular downstream short‐read second‐generation sequencing (SGS) can identify targets at a gene level, long‐read third‐generation sequencing (TGS) can observe at an RNA‐isoform level (Figure 5a), and single‐cell sequencing can be used for cell‐specific monitoring of targets. If an RBP is expected binding to intronic regions of pre‐mRNAs, it should use the library preparation methods, which are able to investigate nascent RNAs. For example, it is possible to use global run‐on sequencing (GRO‐seq; Core et al., 2008), precision nuclear run‐on and sequencing (PRO‐seq) (up to single‐nucleotide resolution compared to the former) (Figure 5b; Kwak et al., 2013), RNA pol II‐directed method NET‐seq (aka native elongating transcript sequencing) (Figure 5c), or chromatin‐associated RNA (caRNA) methods. However, metabolic RNA labeling method is not appropriate because its U‐to‐C conversion will mix up with C‐to‐U edits left by APOBEC1 (Wissink et al., 2019). Many efforts have been made for establishing in yeast (Churchman & Weissman, 2011), fly (Prudencio et al., 2022), and mammalian (Nojima et al., 2015) NET‐seq or run‐on methods (O'Brien et al., 2023), for implementing caRNA in mouse (Menet et al., 2012) and fly (Khodor et al., 2011), which supplies convenience for amalgamating these methods with RBP‐RME methodology.

Since CLIP‐based and RBP–RME methods were designed with completely distinct principles, the methodologies in two different categories will favor revealing real RBP's target profile. Moreover, the following seven points need to be considered in order to choose a more suitable method among CLIP‐based and RBP–RME methods. First, RBP–RME methods can discern cell‐specific interaction between RBP and RNA, while CLIP‐seq cannot reach a single‐cell level. Second, only TRIBE and STAMP can do a survey for isoform‐specific interaction by using long‐read RNA‐seq, while RNA‐tagging is restricted by difficulty in PacBio sequencing library construction, and transcript digesting in CLIP‐seq will lose isoform details. Third, TRIBE and STAMP can detect nascent RNA targets while RNA‐tagging has not fitted for nascent RNA. CLIP‐seq can determine intronic RBP‐binding regions when combined with nuclei purification. Fourth, the complementary target mRNA tied by Argonaute (AGO)‐miRNA can be captured by CLIP (Chi et al., 2009; Rozen‐Gagnon et al., 2021), TRIBE (Sekar et al., 2023), and STAMP, but the object of small RNA (e.g., miRNA) can be only scanned by CLIP. It is believed that RBP–RME method is more effective in editing long RNA (about >100 nt). Fifth, TRIBE can acquire dsRNA targets efficiently while UV‐crosslinking between RBP and dsRNA in CLIP‐seq is likely inefficient. Sixth, CLIP‐seq can get binding data at nucleotide resolution, while TRIBE and STAMP cannot. Seventh, TRIBE‐STAMP can observe two RBPs working on a single RNA molecule (Flamand et al., 2022), whereas multiplexing eCLIP, the ABC method, can explore many RBPs concurrently (Lorenz et al., 2023).

RBPs are key regulatory factors of gene expression, and therefore their dysfunctions can lead to varying diseases. CRISPR‐Cas9 screening was carried out and identified dozens of RBPs, including YTHDF2, working on oncogenic pathways. With eCLIP, m6A sequencing, and single‐cell Ribo‐STAMP (scRibo‐STAMP), they further explored alterations in the translatome at a single‐cell level within YTHDF2‐depleted solid tumors and highlighted the therapeutic potential of YTHDF2 in breast cancers (Einstein et al., 2021). In addition, RNA‐binding protein Ataxin‐2 (Atx2) is associated with the pathogenesis of various neurodegenerative diseases. The extensive data set of Atx2‐target mRNAs were identified by TRIBE in Drosophila (Singh et al., 2021), which contributed to the knowledge of neural translation control mechanism and provided a therapeutic strategy targeting human Atx2. The critical proliferation‐related role of CCDC137 was demonstrated in hepatocellular carcinoma (HCC) patients by STAMP (Tao et al., 2023). Therefore, the continuous development and improvement of RBP–RNA interaction profiling technology will benefit the understanding of human health and disease mechanisms.

The RME‐based strategy has made up a substantial portion of mRNA target identification. Here, we reviewed RNA modification enzymes in ADAR, TENT, and AID/APOBEC family, regarding the biological function, biochemical activity, and substrate specificity of the enzymes and their domains. Also, some protein mutations with engineered activity were discussed. Furthermore, we provided a systematic overview of the basic principles, advantages, disadvantages, and applications of the RNA–protein interaction identification techniques, such as TRIBE, RNA tagging, STAMP, and CLIP‐seq. These distinctive methods will keep contributing to the study of RBP–RNA interactome.

AUTHOR CONTRIBUTIONS

Hua Jin: Funding acquisition (lead); project administration (lead); supervision (equal); validation (equal); writing – review and editing (equal); writing – original draft (lead); . Chong Li: Validation (supporting); visualization (lead); writing – review and editing (equal). Yunxiao Jia: Visualization (supporting); writing – review and editing (supporting). Weilan Piao: Supervision (equal); validation (equal); writing – original draft (lead); writing – review and editing (equal). Yuxuan Qi: Validation (equal); writing – review and editing (equal).

FUNDING INFORMATION

The work was supported by the General Program of the National Natural Science Foundation of China (31970622), and supported by the Fundamental Research Funds for the Central Universities.

CONFLICT OF INTEREST STATEMENT

The authors have declared no conflicts of interest for this article.

RELATED WIREs ARTICLE

To edit or not to edit: Regulation of ADAR editing specificity and efficiency

Jin, H. , Li, C. , Jia, Y. , Qi, Y. , & Piao, W. (2024). Revealing the hidden RBP–RNA interactions with RNA modification enzyme‐based strategies. WIREs RNA, 15(3), e1863. 10.1002/wrna.1863

Hua Jin and Chong Li contributed equally to this study.

Edited by: Jeff Wilusz, Editor‐in‐Chief

DATA AVAILABILITY STATEMENT

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

REFERENCES

  1. Abruzzi, K. C. , Ratner, C. , & Rosbash, M. (2023). Comparison of TRIBE and STAMP for identifying targets of RNA binding proteins in human and drosophila cells. RNA, 29(8), 1230–1242. 10.1261/rna.079608.123 [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. 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. (2019). A cytosine deaminase for programmable single‐base RNA editing. Science, 365(6451), 382–386. 10.1126/science.aax7063 [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Arribas‐Hernandez, L. , Rennie, S. , Koster, T. , Porcelli, C. , Lewinski, M. , Staiger, D. , Andersson, R. , & Brodersen, P. (2021). Principles of mRNA targeting via the Arabidopsis m(6)A‐binding protein ECT2. eLife, 10, e72375. 10.7554/eLife.72375 [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Arribas‐Hernandez, L. , Rennie, S. , Schon, M. , Porcelli, C. , Enugutti, B. , Andersson, R. , Nodine, M. D. , & Brodersen, P. (2021). The YTHDF proteins ECT2 and ECT3 bind largely overlapping target sets and influence target mRNA abundance, not alternative polyadenylation. eLife, 10, e72377. 10.7554/eLife.72377 [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Backus, J. W. , & Smith, H. C. (1992). Three distinct RNA sequence elements are required for efficient apolipoprotein B (apoB) RNA editing in vitro. Nucleic Acids Research, 20(22), 6007–6014. 10.1093/nar/20.22.6007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Barraud, P. , & Allain, F. H. (2012). ADAR proteins: Double‐stranded RNA and Z‐DNA binding domains. Current Topics in Microbiology and Immunology, 353, 35–60. 10.1007/82_2011_145 [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Bazak, L. , Haviv, A. , Barak, M. , Jacob‐Hirsch, J. , Deng, P. , Zhang, R. , Isaacs, F. J. , Rechavi, G. , Li, J. B. , Eisenberg, E. , & Levanon, E. Y. (2014). A‐to‐I RNA editing occurs at over a hundred million genomic sites, located in a majority of human genes. Genome Research, 24(3), 365–376. 10.1101/gr.164749.113 [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Beghini, A. , Ripamonti, C. B. , Peterlongo, P. , Roversi, G. , Cairoli, R. , Morra, E. , & Larizza, L. (2000). RNA hyperediting and alternative splicing of hematopoietic cell phosphatase (PTPN6) gene in acute myeloid leukemia. Human Molecular Genetics, 9(15), 2297–2304. 10.1093/oxfordjournals.hmg.a018921 [DOI] [PubMed] [Google Scholar]
  9. Beta, R. A. A. , & Balatsos, N. A. A. (2018). Tales around the clock: Poly(a) tails in circadian gene expression. WIREs RNA, 9(5), e1484. 10.1002/wrna.1484 [DOI] [PubMed] [Google Scholar]
  10. Biswas, J. , Rahman, R. , Gupta, V. , Rosbash, M. , & Singer, R. H. (2020). MS2‐TRIBE evaluates both protein‐RNA interactions and nuclear Organization of Transcription by RNA editing. iScience, 23(7), 101318. 10.1016/j.isci.2020.101318 [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Blanc, V. , Park, E. , Schaefer, S. , Miller, M. , Lin, Y. , Kennedy, S. , Billing, A. M. , Ben Hamidane, H. , Graumann, J. , Mortazavi, A. , Nadeau, J. H. , & Davidson, N. O. (2014). Genome‐wide identification and functional analysis of Apobec‐1‐mediated C‐to‐U RNA editing in mouse small intestine and liver. Genome Biology, 15(6), R79. 10.1186/gb-2014-15-6-r79 [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Blanc, V. , Xie, Y. , Kennedy, S. , Riordan, J. D. , Rubin, D. C. , Madison, B. B. , Mills, J. C. , Nadeau, J. H. , & Davidson, N. O. (2019). Apobec1 complementation factor (A1CF) and RBM47 interact in tissue‐specific regulation of C to U RNA editing in mouse intestine and liver. RNA, 25(1), 70–81. 10.1261/rna.068395.118 [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Boccaletto, P. , Stefaniak, F. , Ray, A. , Cappannini, A. , Mukherjee, S. , Purta, E. , Kurkowska, M. , Shirvanizadeh, N. , Destefanis, E. , Groza, P. , Avsar, G. , Romitelli, A. , Pir, P. , Dassi, E. , Conticello, S. G. , Aguilo, F. , & Bujnicki, J. M. (2022). MODOMICS: A database of RNA modification pathways. 2021 update. Nucleic Acids Research, 50(D1), D231–D235. 10.1093/nar/gkab1083 [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Bogerd, H. P. , Wiegand, H. L. , Doehle, B. P. , Lueders, K. K. , & Cullen, B. R. (2006). APOBEC3A and APOBEC3B are potent inhibitors of LTR‐retrotransposon function in human cells. Nucleic Acids Research, 34(1), 89–95. 10.1093/nar/gkj416 [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Bohn, J. A. , Thummar, K. , York, A. , Raymond, A. , Brown, W. C. , Bieniasz, P. D. , Hatziioannou, T. , & Smith, J. L. (2017). APOBEC3H structure reveals an unusual mechanism of interaction with duplex RNA. Nature Communications, 8(1), 1021. 10.1038/s41467-017-01309-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Bortolamiol‐Becet, D. , Hu, F. , Jee, D. , Wen, J. , Okamura, K. , Lin, C. J. , Ameres, S. L. , & Lai, E. C. (2015). Selective suppression of the splicing‐mediated MicroRNA pathway by the terminal Uridyltransferase tailor. Molecular Cell, 59(2), 217–228. 10.1016/j.molcel.2015.05.034 [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Bossy‐Wetzel, E. , Schwarzenbacher, R. , & Lipton, S. A. (2004). Molecular pathways to neurodegeneration. Nature Medicine, 10(Suppl), S2–S9. 10.1038/nm1067 [DOI] [PubMed] [Google Scholar]
  18. Brannan, K. W. , Chaim, I. A. , Marina, R. J. , Yee, B. A. , Kofman, E. R. , Lorenz, D. A. , Jagannatha, P. , Dong, K. D. , Madrigal, A. A. , Underwood, J. G. , & Yeo, G. W. (2021). Robust single‐cell discovery of RNA targets of RNA‐binding proteins and ribosomes. Nature Methods, 18(5), 507–519. 10.1038/s41592-021-01128-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Burjoski, V. , & Reddy, A. S. N. (2021). The landscape of RNA‐protein interactions in plants: Approaches and current status. International Journal of Molecular Sciences, 22(6), 2845. 10.3390/ijms22062845 [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Cai, D. , Sun, C. , Murashita, T. , Que, X. , & Chen, S. Y. (2023). ADAR1 non‐editing function in macrophage activation and abdominal aortic aneurysm. Circulation Research, 132(4), e78–e93. 10.1161/CIRCRESAHA.122.321722 [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Chang, H. , Yeo, J. , Kim, J. G. , Kim, H. , Lim, J. , Lee, M. , Kim, H. H. , Ohk, J. , Jeon, H. Y. , Lee, H. , Jung, H. , Kim, K. W. , & Kim, V. N. (2018). Terminal Uridylyltransferases execute programmed clearance of maternal transcriptome in vertebrate embryos. Molecular Cell, 70(1), 72–82.e77. 10.1016/j.molcel.2018.03.004 [DOI] [PubMed] [Google Scholar]
  22. Chen, C. X. , Cho, D. S. , Wang, Q. , Lai, F. , Carter, K. C. , & Nishikura, K. (2000). A third member of the RNA‐specific adenosine deaminase gene family, ADAR3, contains both single‐ and double‐stranded RNA binding domains. RNA, 6(5), 755–767. 10.1017/s1355838200000170 [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Chen, J. , Jin, J. , Jiang, J. , & Wang, Y. (2023). Adenosine deaminase acting on RNA 1 (ADAR1) as crucial regulators in cardiovascular diseases: Structures, pathogenesis, and potential therapeutic approach. Frontiers in Pharmacology, 14, 1194884. 10.3389/fphar.2023.1194884 [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Chen, L. , Li, Y. , Lin, C. H. , Chan, T. H. , Chow, R. K. , Song, Y. , Liu, M. , Yuan, Y. F. , Fu, L. , Kong, K. L. , Qi, L. , Li, Y. , Zhang, N. , Tong, A. H. , Kwong, D. L. , Man, K. , Lo, C. M. , Lok, S. , Tenen, D. G. , & Guan, X. Y. (2013). Recoding RNA editing of AZIN1 predisposes to hepatocellular carcinoma. Nature Medicine, 19(2), 209–216. 10.1038/nm.3043 [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Cheng, L. , Li, F. , Jiang, Y. , Yu, H. , Xie, C. , Shi, Y. , & Gong, Q. (2019). Structural insights into a unique preference for 3′ terminal guanine of mirtron in drosophila TUTase tailor. Nucleic Acids Research, 47(1), 495–508. 10.1093/nar/gky1116 [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Chester, A. , Somasekaram, A. , Tzimina, M. , Jarmuz, A. , Gisbourne, J. , O'Keefe, R. , Scott, J. , & Navaratnam, N. (2003). The apolipoprotein B mRNA editing complex performs a multifunctional cycle and suppresses nonsense‐mediated decay. The EMBO Journal, 22(15), 3971–3982. 10.1093/emboj/cdg369 [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Chi, S. W. , Zang, J. B. , Mele, A. , & Darnell, R. B. (2009). Argonaute HITS‐CLIP decodes microRNA‐mRNA interaction maps. Nature, 460(7254), 479–486. 10.1038/nature08170 [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Chu, C. , Qu, K. , Zhong, F. L. , Artandi, S. E. , & Chang, H. Y. (2011). Genomic maps of long noncoding RNA occupancy reveal principles of RNA‐chromatin interactions. Molecular Cell, 44(4), 667–678. 10.1016/j.molcel.2011.08.027 [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Chu, C. , Zhang, Q. C. , da Rocha, S. T. , Flynn, R. A. , Bharadwaj, M. , Calabrese, J. M. , Magnuson, T. , Heard, E. , & Chang, H. Y. (2015). Systematic discovery of Xist RNA binding proteins. Cell, 161(2), 404–416. 10.1016/j.cell.2015.03.025 [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Churchman, L. S. , & Weissman, J. S. (2011). Nascent transcript sequencing visualizes transcription at nucleotide resolution. Nature, 469(7330), 368–373. 10.1038/nature09652 [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Cole, D. C. , Chung, Y. , Gagnidze, K. , Hajdarovic, K. H. , Rayon‐Estrada, V. , Harjanto, D. , Bigio, B. , Gal‐Toth, J. , Milner, T. A. , McEwen, B. S. , Papavasiliou, F. N. , & Bulloch, K. (2017). Loss of APOBEC1 RNA‐editing function in microglia exacerbates age‐related CNS pathophysiology. Proceedings of the National Academy of Sciences of the United States of America, 114(50), 13272–13277. 10.1073/pnas.1710493114 [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Core, L. J. , Waterfall, J. J. , & Lis, J. T. (2008). Nascent RNA sequencing reveals widespread pausing and divergent initiation at human promoters. Science, 322(5909), 1845–1848. 10.1126/science.1162228 [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Dahm, G. M. , Gubin, M. M. , Magee, J. D. , Techasintana, P. , Calaluce, R. , & Atasoy, U. (2012). Method for the isolation and identification of mRNAs, microRNAs and protein components of ribonucleoprotein complexes from cell extracts using RIP‐Chip. Journal of Visualized Experiments, 67, 3851. 10.3791/3851 [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Dai, W. , Li, A. , Yu, N. J. , Nguyen, T. , Leach, R. W. , Wuhr, M. , & Kleiner, R. E. (2021). Activity‐based RNA‐modifying enzyme probing reveals DUS3L‐mediated dihydrouridylation. Nature Chemical Biology, 17(11), 1178–1187. 10.1038/s41589-021-00874-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Dai, W. , Yu, N. J. , & Kleiner, R. E. (2023). Chemoproteomic approaches to studying RNA modification‐associated proteins. Accounts of Chemical Research, 56(19), 2726–2739. 10.1021/acs.accounts.3c00450 [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Darnell, R. B. (2010). HITS‐CLIP: Panoramic views of protein‐RNA regulation in living cells. WIREs RNA, 1(2), 266–286. 10.1002/wrna.31 [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Deffit, S. N. , & Hundley, H. A. (2016). To edit or not to edit: Regulation of ADAR editing specificity and efficiency. WIREs RNA, 7(1), 113–127. 10.1002/wrna.1319 [DOI] [PubMed] [Google Scholar]
  38. Desterro, J. M. , Keegan, L. P. , Lafarga, M. , Berciano, M. T. , O'Connell, M. , & Carmo‐Fonseca, M. (2003). Dynamic association of RNA‐editing enzymes with the nucleolus. Journal of Cell Science, 116(Pt 9), 1805–1818. 10.1242/jcs.00371 [DOI] [PubMed] [Google Scholar]
  39. Doma, M. K. , & Parker, R. (2007). RNA quality control in eukaryotes. Cell, 131(4), 660–668. 10.1016/j.cell.2007.10.041 [DOI] [PubMed] [Google Scholar]
  40. Eggington, J. M. , Greene, T. , & Bass, B. L. (2011). Predicting sites of ADAR editing in double‐stranded RNA. Nature Communications, 2, 319. 10.1038/ncomms1324 [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Eifler, T. , Pokharel, S. , & Beal, P. A. (2013). RNA‐Seq analysis identifies a novel set of editing substrates for human ADAR2 present in Saccharomyces cerevisiae. Biochemistry, 52(45), 7857–7869. 10.1021/bi4006539 [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Einstein, J. M. , Perelis, M. , Chaim, I. A. , Meena, J. K. , Nussbacher, J. K. , Tankka, A. T. , Yee, B. A. , Li, H. , Madrigal, A. A. , Neill, N. J. , Shankar, A. , Tyagi, S. , Westbrook, T. F. , & Yeo, G. W. (2021). Inhibition of YTHDF2 triggers proteotoxic cell death in MYC‐driven breast cancer. Molecular Cell, 81(15), 3048–3064 e3049. 10.1016/j.molcel.2021.06.014 [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Enstero, M. , Daniel, C. , Wahlstedt, H. , Major, F. , & Ohman, M. (2009). Recognition and coupling of A‐to‐I edited sites are determined by the tertiary structure of the RNA. Nucleic Acids Research, 37(20), 6916–6926. 10.1093/nar/gkp731 [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Etard, C. , Roostalu, U. , & Strahle, U. (2010). Lack of Apobec2‐related proteins causes a dystrophic muscle phenotype in zebrafish embryos. The Journal of Cell Biology, 189(3), 527–539. 10.1083/jcb.200912125 [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Faehnle, C. R. , Walleshauser, J. , & Joshua‐Tor, L. (2017). Multi‐domain utilization by TUT4 and TUT7 in control of let‐7 biogenesis. Nature Structural & Molecular Biology, 24(8), 658–665. 10.1038/nsmb.3428 [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Fecko, C. J. , Munson, K. M. , Saunders, A. , Sun, G. , Begley, T. P. , Lis, J. T. , & Webb, W. W. (2007). Comparison of femtosecond laser and continuous wave UV sources for protein‐nucleic acid crosslinking. Photochemistry and Photobiology, 83(6), 1394–1404. 10.1111/j.1751-1097.2007.00179.x [DOI] [PubMed] [Google Scholar]
  47. Flamand, M. N. , Ke, K. , Tamming, R. , & Meyer, K. D. (2022). Single‐molecule identification of the target RNAs of different RNA binding proteins simultaneously in cells. Genes & Development, 36(17–18), 1002–1015. 10.1101/gad.349983.122 [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Flamand, M. N. , Tegowski, M. , & Meyer, K. D. (2023). The proteins of mRNA modification: Writers, readers, and erasers. Annual Review of Biochemistry, 92, 145–173. 10.1146/annurev-biochem-052521-035330 [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. George, C. X. , & Samuel, C. E. (1999). Human RNA‐specific adenosine deaminase ADAR1 transcripts possess alternative exon 1 structures that initiate from different promoters, one constitutively active and the other interferon inducible. Proceedings of the National Academy of Sciences of the United States of America, 96(8), 4621–4626. 10.1073/pnas.96.8.4621 [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. George, C. X. , Wagner, M. V. , & Samuel, C. E. (2005). Expression of interferon‐inducible RNA adenosine deaminase ADAR1 during pathogen infection and mouse embryo development involves tissue‐selective promoter utilization and alternative splicing. The Journal of Biological Chemistry, 280(15), 15020–15028. 10.1074/jbc.M500476200 [DOI] [PubMed] [Google Scholar]
  51. Gerstberger, S. , Hafner, M. , & Tuschl, T. (2014). A census of human RNA‐binding proteins. Nature Reviews. Genetics, 15(12), 829–845. 10.1038/nrg3813 [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Ghashghaei, M. , Liu, Y. , Ettles, J. , Bombaci, G. , Ramkumar, N. , Liu, Z. , Escano, L. , Miko, S. S. , Kim, Y. , Waldron, J. A. , Do, K. , MacPherson, K. , Yuen, K. A. , Taibi, T. , Yue, M. , Arsalan, A. , Jin, Z. , Edin, G. , Karsan, A. , … Vu, L. P. (2024). Translation efficiency driven by CNOT3 subunit of the CCR4‐NOT complex promotes leukemogenesis. Nature Communications, 15(1), 2340. 10.1038/s41467-024-46665-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Grawe, C. , Stelloo, S. , van Hout, F. A. H. , & Vermeulen, M. (2021). RNA‐centric methods: Toward the Interactome of specific RNA transcripts. Trends in Biotechnology, 39(9), 890–900. 10.1016/j.tibtech.2020.11.011 [DOI] [PubMed] [Google Scholar]
  54. Gupta, A. , Li, Y. , Chen, S. H. , Papas, B. N. , Martin, N. P. , & Morgan, M. (2023). TUT4/7‐mediated uridylation of a coronavirus subgenomic RNAs delays viral replication. Communications Biology, 6(1), 438. 10.1038/s42003-023-04814-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Hafner, M. , Katsantoni, M. , Köster, T. , Marks, J. , Mukherjee, J. , Staiger, D. , Ule, J. , & Zavolan, M. (2021). CLIP and complementary methods. Nature Reviews Methods Primers, 1(1), 20. 10.1038/s43586-021-00018-1 [DOI] [Google Scholar]
  56. Hafner, M. , Landthaler, M. , Burger, L. , Khorshid, M. , Hausser, J. , Berninger, P. , Rothballer, A. , Ascano, M., Jr. , Jungkamp, A. C. , Munschauer, M. , Ulrich, A. , Wardle, G. S. , Dewell, S. , Zavolan, M. , & Tuschl, T. (2010). Transcriptome‐wide identification of RNA‐binding protein and microRNA target sites by PAR‐CLIP. Cell, 141(1), 129–141. 10.1016/j.cell.2010.03.009 [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Hafner, M. , Landthaler, M. , Burger, L. , Khorshid, M. , Hausser, J. , Berninger, P. , Rothballer, A. , Ascano, M. , Jungkamp, A. C. , Munschauer, M. , Ulrich, A. , Wardle, G. S. , Dewell, S. , Zavolan, M. , & Tuschl, T. (2010). PAR‐CliP—A method to identify transcriptome‐wide the binding sites of RNA binding proteins. Journal of Visualized Experiments, (41), 2034. 10.3791/2034 [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Hajji, K. , Sedmik, J. , Cherian, A. , Amoruso, D. , Keegan, L. P. , & O'Connell, M. A. (2022). ADAR2 enzymes: Efficient site‐specific RNA editors with gene therapy aspirations. RNA, 28(10), 1281–1297. 10.1261/rna.079266.122 [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Harris, R. S. , & Dudley, J. P. (2015). APOBECs and virus restriction. Virology, 479–480, 131–145. 10.1016/j.virol.2015.03.012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Hayashi, M. , & Suzuki, T. (2013). Dyschromatosis symmetrica hereditaria. The Journal of Dermatology, 40(5), 336–343. 10.1111/j.1346-8138.2012.01661.x [DOI] [PubMed] [Google Scholar]
  61. Heo, I. , Ha, M. , Lim, J. , Yoon, M. J. , Park, J. E. , Kwon, S. C. , Chang, H. , & Kim, V. N. (2012). Mono‐uridylation of pre‐microRNA as a key step in the biogenesis of group II let‐7 microRNAs. Cell, 151(3), 521–532. 10.1016/j.cell.2012.09.022 [DOI] [PubMed] [Google Scholar]
  62. Heo, I. , Joo, C. , Kim, Y. K. , Ha, M. , Yoon, M. J. , Cho, J. , Yeom, K. H. , Han, J. , & Kim, V. N. (2009). TUT4 in concert with Lin28 suppresses microRNA biogenesis through pre‐microRNA uridylation. Cell, 138(4), 696–708. 10.1016/j.cell.2009.08.002 [DOI] [PubMed] [Google Scholar]
  63. Herbert, A. , & Rich, A. (2001). The role of binding domains for dsRNA and Z‐DNA in the in vivo editing of minimal substrates by ADAR1. Proceedings of the National Academy of Sciences of the United States of America, 98(21), 12132–12137. 10.1073/pnas.211419898 [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Herzog, J. J. , Xu, W. , Deshpande, M. , Rahman, R. , Suib, H. , Rodal, A. A. , Rosbash, M. , & Paradis, S. (2020). TDP‐43 dysfunction restricts dendritic complexity by inhibiting CREB activation and altering gene expression. Proceedings of the National Academy of Sciences of the United States of America, 117(21), 11760–11769. 10.1073/pnas.1917038117 [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Ikeda, T. , Ong, E. B. , Watanabe, N. , Sakaguchi, N. , Maeda, K. , & Koito, A. (2016). Creation of chimeric human/rabbit APOBEC1 with HIV‐1 restriction and DNA mutation activities. Scientific Reports, 6, 19035. 10.1038/srep19035 [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Ivanov, A. , Memczak, S. , Wyler, E. , Torti, F. , Porath, H. T. , Orejuela, M. R. , Piechotta, M. , Levanon, E. Y. , Landthaler, M. , Dieterich, C. , & Rajewsky, N. (2015). Analysis of intron sequences reveals hallmarks of circular RNA biogenesis in animals. Cell Reports, 10(2), 170–177. 10.1016/j.celrep.2014.12.019 [DOI] [PubMed] [Google Scholar]
  67. Jaikaran, D. C. , Collins, C. H. , & MacMillan, A. M. (2002). Adenosine to inosine editing by ADAR2 requires formation of a ternary complex on the GluR‐B R/G site. The Journal of Biological Chemistry, 277(40), 37624–37629. 10.1074/jbc.M204126200 [DOI] [PubMed] [Google Scholar]
  68. Jin, H. , Xu, W. , Rahman, R. , Na, D. , Fieldsend, A. , Song, W. , Liu, S. , Li, C. , & Rosbash, M. (2020). TRIBE editing reveals specific mRNA targets of eIF4E‐BP in Drosophila and in mammals. Science Advances, 6(33), eabb8771. 10.1126/sciadv.abb8771 [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Kallman, A. M. , Sahlin, M. , & Ohman, M. (2003). ADAR2 A→I editing: Site selectivity and editing efficiency are separate events. Nucleic Acids Research, 31(16), 4874–4881. 10.1093/nar/gkg681 [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Kawahara, Y. , Zinshteyn, B. , Sethupathy, P. , Iizasa, H. , Hatzigeorgiou, A. G. , & Nishikura, K. (2007). Redirection of silencing targets by adenosine‐to‐inosine editing of miRNAs. Science, 315(5815), 1137–1140. 10.1126/science.1138050 [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Khodor, Y. L. , Rodriguez, J. , Abruzzi, K. C. , Tang, C. H. , Marr, M. T., 2nd , & Rosbash, M. (2011). Nascent‐seq indicates widespread cotranscriptional pre‐mRNA splicing in drosophila. Genes & Development, 25(23), 2502–2512. 10.1101/gad.178962.111 [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Klattenhoff, C. A. , Scheuermann, J. C. , Surface, L. E. , Bradley, R. K. , Fields, P. A. , Steinhauser, M. L. , Ding, H. , Butty, V. L. , Torrey, L. , Haas, S. , Abo, R. , Tabebordbar, M. , Lee, R. T. , Burge, C. B. , & Boyer, L. A. (2013). Braveheart, a long noncoding RNA required for cardiovascular lineage commitment. Cell, 152(3), 570–583. 10.1016/j.cell.2013.01.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Knebel, U. E. , Peleg, S. , Dai, C. , Cohen‐Fultheim, R. , Jonsson, S. , Poznyak, K. , Israeli, M. , Zamashanski, L. , Glaser, B. , Levanon, E. Y. , Powers, A. C. , Klochendler, A. , & Dor, Y. (2024). Disrupted RNA editing in beta cells mimics early‐stage type 1 diabetes. Cell Metabolism, 36(1), 48–61 e46. 10.1016/j.cmet.2023.11.011 [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Konig, J. , Zarnack, K. , Rot, G. , Curk, T. , Kayikci, M. , Zupan, B. , Turner, D. J. , Luscombe, N. M. , & Ule, J. (2010). iCLIP reveals the function of hnRNP particles in splicing at individual nucleotide resolution. Nature Structural & Molecular Biology, 17(7), 909–915. 10.1038/nsmb.1838 [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Konig, J. , Zarnack, K. , Rot, G. , Curk, T. , Kayikci, M. , Zupan, B. , Turner, D. J. , Luscombe, N. M. , & Ule, J. (2011). iCLIP—Transcriptome‐wide mapping of protein‐RNA interactions with individual nucleotide resolution. Journal of Visualized Experiments, 50, 2638. 10.3791/2638 [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Kouno, T. , Silvas, T. V. , Hilbert, B. J. , Shandilya, S. M. D. , Bohn, M. F. , Kelch, B. A. , Royer, W. E. , Somasundaran, M. , Kurt Yilmaz, N. , Matsuo, H. , & Schiffer, C. A. (2017). Crystal structure of APOBEC3A bound to single‐stranded DNA reveals structural basis for cytidine deamination and specificity. Nature Communications, 8, 15024. 10.1038/ncomms15024 [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Kretz, M. , Siprashvili, Z. , Chu, C. , Webster, D. E. , Zehnder, A. , Qu, K. , Lee, C. S. , Flockhart, R. J. , Groff, A. F. , Chow, J. , Johnston, D. , Kim, G. E. , Spitale, R. C. , Flynn, R. A. , Zheng, G. X. , Aiyer, S. , Raj, A. , Rinn, J. L. , Chang, H. Y. , & Khavari, P. A. (2013). Control of somatic tissue differentiation by the long non‐coding RNA TINCR. Nature, 493(7431), 231–235. 10.1038/nature11661 [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Kroupova, A. , Ivascu, A. , Reimao‐Pinto, M. M. , Ameres, S. L. , & Jinek, M. (2019). Structural basis for acceptor RNA substrate selectivity of the 3′ terminal uridylyl transferase tailor. Nucleic Acids Research, 47(2), 1030–1042. 10.1093/nar/gky1164 [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Kuttan, A. , & Bass, B. L. (2012). Mechanistic insights into editing‐site specificity of ADARs. Proceedings of the National Academy of Sciences of the United States of America, 109(48), E3295–E3304. 10.1073/pnas.1212548109 [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Kwak, H. , Fuda, N. J. , Core, L. J. , & Lis, J. T. (2013). Precise maps of RNA polymerase reveal how promoters direct initiation and pausing. Science, 339(6122), 950–953. 10.1126/science.1229386 [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Lai, F. , Chen, C. X. , Carter, K. C. , & Nishikura, K. (1997). Editing of glutamate receptor B subunit ion channel RNAs by four alternatively spliced DRADA2 double‐stranded RNA adenosine deaminases. Molecular and Cellular Biology, 17(5), 2413–2424. 10.1128/MCB.17.5.2413 [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Lapointe, C. P. , Wilinski, D. , Saunders, H. A. , & Wickens, M. (2015). Protein‐RNA networks revealed through covalent RNA marks. Nature Methods, 12(12), 1163–1170. 10.1038/nmeth.3651 [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. le Pen, J. , Jiang, H. , di Domenico, T. , Kneuss, E. , Kosalka, J. , Leung, C. , Morgan, M. , Much, C. , Rudolph, K. L. M. , Enright, A. J. , O'Carroll, D. , Wang, D. , & Miska, E. A. (2018). Terminal uridylyltransferases target RNA viruses as part of the innate immune system. Nature Structural & Molecular Biology, 25(9), 778–786. 10.1038/s41594-018-0106-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Lee, F. C. Y. , & Ule, J. (2018). Advances in CLIP Technologies for Studies of protein‐RNA interactions. Molecular Cell, 69(3), 354–369. 10.1016/j.molcel.2018.01.005 [DOI] [PubMed] [Google Scholar]
  85. Lee, H. Y. , Haurwitz, R. E. , Apffel, A. , Zhou, K. , Smart, B. , Wenger, C. D. , Laderman, S. , Bruhn, L. , & Doudna, J. A. (2013). RNA‐protein analysis using a conditional CRISPR nuclease. Proceedings of the National Academy of Sciences of the United States of America, 110(14), 5416–5421. 10.1073/pnas.1302807110 [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Lehmann, K. A. , & Bass, B. L. (1999). The importance of internal loops within RNA substrates of ADAR1. Journal of Molecular Biology, 291(1), 1–13. 10.1006/jmbi.1999.2914 [DOI] [PubMed] [Google Scholar]
  87. Lehrbach, N. J. , Armisen, J. , Lightfoot, H. L. , Murfitt, K. J. , Bugaut, A. , Balasubramanian, S. , & Miska, E. A. (2009). LIN‐28 and the poly(U) polymerase PUP‐2 regulate let‐7 microRNA processing in Caenorhabditis elegans. Nature Structural & Molecular Biology, 16(10), 1016–1020. 10.1038/nsmb.1675 [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Leppek, K. , & Stoecklin, G. (2014). An optimized streptavidin‐binding RNA aptamer for purification of ribonucleoprotein complexes identifies novel ARE‐binding proteins. Nucleic Acids Research, 42(2), e13. 10.1093/nar/gkt956 [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Levanon, E. Y. , Eisenberg, E. , Yelin, R. , Nemzer, S. , Hallegger, M. , Shemesh, R. , Fligelman, Z. Y. , Shoshan, A. , Pollock, S. R. , Sztybel, D. , Olshansky, M. , Rechavi, G. , & Jantsch, M. F. (2004). Systematic identification of abundant A‐to‐I editing sites in the human transcriptome. Nature Biotechnology, 22(8), 1001–1005. 10.1038/nbt996 [DOI] [PubMed] [Google Scholar]
  90. Li, Y. , & Maine, E. M. (2018). The balance of poly(U) polymerase activity ensures germline identity, survival and development in Caenorhabditis elegans. Development, 145(19), dev165944. 10.1242/dev.165944 [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Licatalosi, D. D. , Mele, A. , Fak, J. J. , Ule, J. , Kayikci, M. , Chi, S. W. , Clark, T. A. , Schweitzer, A. C. , Blume, J. E. , Wang, X. , Darnell, J. C. , & Darnell, R. B. (2008). HITS‐CLIP yields genome‐wide insights into brain alternative RNA processing. Nature, 456(7221), 464–469. 10.1038/nature07488 [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Lim, J. , Ha, M. , Chang, H. , Kwon, S. C. , Simanshu, D. K. , Patel, D. J. , & Kim, V. N. (2014). Uridylation by TUT4 and TUT7 marks mRNA for degradation. Cell, 159(6), 1365–1376. 10.1016/j.cell.2014.10.055 [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Lim, J. , Kim, D. , Lee, Y. S. , Ha, M. , Lee, M. , Yeo, J. , Chang, H. , Song, J. , Ahn, K. , & Kim, V. N. (2018). Mixed tailing by TENT4A and TENT4B shields mRNA from rapid deadenylation. Science, 361(6403), 701–704. 10.1126/science.aam5794 [DOI] [PubMed] [Google Scholar]
  94. Liu, Y. , Nie, H. , & Lu, F. (2020). Dynamic RNA 3' Uridylation and Guanylation during mitosis. iScience, 23(8), 101402. 10.1016/j.isci.2020.101402 [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Liudkovska, V. , & Dziembowski, A. (2021). Functions and mechanisms of RNA tailing by metazoan terminal nucleotidyltransferases. WIREs RNA, 12(2), e1622. 10.1002/wrna.1622 [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Loeser, J. , Bauer, J. , Janssen, K. , Rockenbach, K. , & Wachter, A. (2024). A transient in planta editing assay identifies specific binding of the splicing regulator PTB as a prerequisite for cassette exon inclusion. Plant Molecular Biology, 114(2), 22. 10.1007/s11103-024-01414-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Lorenz, D. A. , Her, H. L. , Shen, K. A. , Rothamel, K. , Hutt, K. R. , Nojadera, A. C. , Bruns, S. C. , Manakov, S. A. , Yee, B. A. , Chapman, K. B. , & Yeo, G. W. (2023). Multiplexed transcriptome discovery of RNA‐binding protein binding sites by antibody‐barcode eCLIP. Nature Methods, 20(1), 65–69. 10.1038/s41592-022-01708-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Lu, D. , Lu, J. , Liu, Q. , & Zhang, Q. (2023). Emerging role of the RNA‐editing enzyme ADAR1 in stem cell fate and function. Biomarker Research, 11(1), 61. 10.1186/s40364-023-00503-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Lubas, M. , Damgaard, C. K. , Tomecki, R. , Cysewski, D. , Jensen, T. H. , & Dziembowski, A. (2013). Exonuclease hDIS3L2 specifies an exosome‐independent 3′–5′ degradation pathway of human cytoplasmic mRNA. The EMBO Journal, 32(13), 1855–1868. 10.1038/emboj.2013.135 [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Macbeth, M. R. , Lingam, A. T. , & Bass, B. L. (2004). Evidence for auto‐inhibition by the N terminus of hADAR2 and activation by dsRNA binding. RNA, 10(10), 1563–1571. 10.1261/rna.7920904 [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Macbeth, M. R. , Schubert, H. L. , Vandemark, A. P. , Lingam, A. T. , Hill, C. P. , & Bass, B. L. (2005). Inositol hexakisphosphate is bound in the ADAR2 core and required for RNA editing. Science, 309(5740), 1534–1539. 10.1126/science.1113150 [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Malecki, M. , Viegas, S. C. , Carneiro, T. , Golik, P. , Dressaire, C. , Ferreira, M. G. , & Arraiano, C. M. (2013). The exoribonuclease Dis3L2 defines a novel eukaryotic RNA degradation pathway. The EMBO Journal, 32(13), 1842–1854. 10.1038/emboj.2013.63 [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Marcucci, R. , Brindle, J. , Paro, S. , Casadio, A. , Hempel, S. , Morrice, N. , Bisso, A. , Keegan, L. P. , del Sal, G. , & O'Connell, M. A. (2011). Pin1 and WWP2 regulate GluR2 Q/R site RNA editing by ADAR2 with opposing effects. The EMBO Journal, 30(20), 4211–4222. 10.1038/emboj.2011.303 [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Maris, C. , Masse, J. , Chester, A. , Navaratnam, N. , & Allain, F. H. (2005). NMR structure of the apoB mRNA stem‐loop and its interaction with the C to U editing APOBEC1 complementary factor. RNA, 11(2), 173–186. 10.1261/rna.7190705 [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Martin, G. , & Keller, W. (2007). RNA‐specific ribonucleotidyl transferases. RNA, 13(11), 1834–1849. 10.1261/rna.652807 [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Martinez‐Perez, M. , Aparicio, F. , Arribas‐Hernandez, L. , Tankmar, M. D. , Rennie, S. , von Bulow, S. , Lindorff‐Larsen, K. , Brodersen, P. , & Pallas, V. (2023). Plant YTHDF proteins are direct effectors of antiviral immunity against an N6‐methyladenosine‐containing RNA virus. The EMBO Journal, 42(18), e113378. 10.15252/embj.2022113378 [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Matia‐Gonzalez, A. M. , Iadevaia, V. , & Gerber, A. P. (2017). A versatile tandem RNA isolation procedure to capture in vivo formed mRNA‐protein complexes. Methods, 118–119, 93–100. 10.1016/j.ymeth.2016.10.005 [DOI] [PubMed] [Google Scholar]
  108. Matsuoka, T. , Nagae, T. , Ode, H. , Awazu, H. , Kurosawa, T. , Hamano, A. , Matsuoka, K. , Hachiya, A. , Imahashi, M. , Yokomaku, Y. , Watanabe, N. , & Iwatani, Y. (2018). Structural basis of chimpanzee APOBEC3H dimerization stabilized by double‐stranded RNA. Nucleic Acids Research, 46(19), 10368–10379. 10.1093/nar/gky676 [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Matthews, M. M. , Thomas, J. M. , Zheng, Y. , Tran, K. , Phelps, K. J. , Scott, A. I. , Havel, J. , Fisher, A. J. , & Beal, P. A. (2016). Structures of human ADAR2 bound to dsRNA reveal base‐flipping mechanism and basis for site selectivity. Nature Structural & Molecular Biology, 23(5), 426–433. 10.1038/nsmb.3203 [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. McHugh, C. A. , & Guttman, M. (2018). RAP‐MS: A method to identify proteins that interact directly with a specific RNA molecule in cells. Methods in Molecular Biology, 1649, 473–488. 10.1007/978-1-4939-7213-5_31 [DOI] [PubMed] [Google Scholar]
  111. McHugh, C. A. , Russell, P. , & Guttman, M. (2014). Methods for comprehensive experimental identification of RNA‐protein interactions. Genome Biology, 15(1), 203. 10.1186/gb4152 [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. McMahon, A. C. , Rahman, R. , Jin, H. , Shen, J. L. , Fieldsend, A. , Luo, W. , & Rosbash, M. (2016). TRIBE: Hijacking an RNA‐editing enzyme to identify cell‐specific targets of RNA‐binding proteins. Cell, 165(3), 742–753. 10.1016/j.cell.2016.03.007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Medina‐Munoz, H. C. , Kofman, E. , Jagannatha, P. , Boyle, E. A. , Yu, T. , Jones, K. L. , Mueller, J. R. , Lykins, G. D. , Doudna, A. T. , Park, S. S. , Blue, S. M. , Ranzau, B. L. , Kohli, R. M. , Komor, A. C. , & Yeo, G. W. (2024). Expanded palette of RNA base editors for comprehensive RBP‐RNA interactome studies. Nature Communications, 15(1), 875. 10.1038/s41467-024-45009-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Medina‐Munoz, H. C. , Lapointe, C. P. , Porter, D. F. , & Wickens, M. (2020). Records of RNA locations in living yeast revealed through covalent marks. Proceedings of the National Academy of Sciences of the United States of America, 117(38), 23539–23547. 10.1073/pnas.1921408117 [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Melcher, T. , Maas, S. , Herb, A. , Sprengel, R. , Higuchi, M. , & Seeburg, P. H. (1996). RED2, a brain‐specific member of the RNA‐specific adenosine deaminase family. The Journal of Biological Chemistry, 271(50), 31795–31798. 10.1074/jbc.271.50.31795 [DOI] [PubMed] [Google Scholar]
  116. Menet, J. S. , Rodriguez, J. , Abruzzi, K. C. , & Rosbash, M. (2012). Nascent‐Seq reveals novel features of mouse circadian transcriptional regulation. eLife, 1, e00011. 10.7554/eLife.00011 [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Meyer, K. D. (2019). DART‐seq: An antibody‐free method for global m(6)a detection. Nature Methods, 16(12), 1275–1280. 10.1038/s41592-019-0570-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Mili, S. , & Steitz, J. A. (2004). Evidence for reassociation of RNA‐binding proteins after cell lysis: Implications for the interpretation of immunoprecipitation analyses. RNA, 10(11), 1692–1694. 10.1261/rna.7151404 [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Mladenova, D. , Barry, G. , Konen, L. M. , Pineda, S. S. , Guennewig, B. , Avesson, L. , Zinn, R. , Schonrock, N. , Bitar, M. , Jonkhout, N. , Crumlish, L. , Kaczorowski, D. C. , Gong, A. , Pinese, M. , Franco, G. R. , Walkley, C. R. , Vissel, B. , & Mattick, J. S. (2018). Adar3 is involved in learning and memory in mice. Frontiers in Neuroscience, 12, 243. 10.3389/fnins.2018.00243 [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Montiel‐Gonzalez, M. F. , Vallecillo‐Viejo, I. , Yudowski, G. A. , & Rosenthal, J. J. (2013). Correction of mutations within the cystic fibrosis transmembrane conductance regulator by site‐directed RNA editing. Proceedings of the National Academy of Sciences of the United States of America, 110(45), 18285–18290. 10.1073/pnas.1306243110 [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Morgan, M. , Kabayama, Y. , Much, C. , Ivanova, I. , di Giacomo, M. , Auchynnikava, T. , Monahan, J. M. , Vitsios, D. M. , Vasiliauskaite, L. , Comazzetto, S. , Rappsilber, J. , Allshire, R. C. , Porse, B. T. , Enright, A. J. , & O'Carroll, D. (2019). A programmed wave of uridylation‐primed mRNA degradation is essential for meiotic progression and mammalian spermatogenesis. Cell Research, 29(3), 221–232. 10.1038/s41422-018-0128-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Morgan, M. , Much, C. , DiGiacomo, M. , Azzi, C. , Ivanova, I. , Vitsios, D. M. , Pistolic, J. , Collier, P. , Moreira, P. N. , Benes, V. , Enright, A. J. , & O'Carroll, D. (2017). mRNA 3′ uridylation and poly(a) tail length sculpt the mammalian maternal transcriptome. Nature, 548(7667), 347–351. 10.1038/nature23318 [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Mukherjee, N. , Wessels, H. H. , Lebedeva, S. , Sajek, M. , Ghanbari, M. , Garzia, A. , Munteanu, A. , Yusuf, D. , Farazi, T. , Hoell, J. I. , Akat, K. M. , Akalin, A. , Tuschl, T. , & Ohler, U. (2019). Deciphering human ribonucleoprotein regulatory networks. Nucleic Acids Research, 47(2), 570–581. 10.1093/nar/gky1185 [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. Munroe, D. , & Jacobson, A. (1990). mRNA poly(a) tail, a 3′ enhancer of translational initiation. Molecular and Cellular Biology, 10(7), 3441–3455. 10.1128/mcb.10.7.3441-3455.1990 [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Muramatsu, M. , Kinoshita, K. , Fagarasan, S. , Yamada, S. , Shinkai, Y. , & Honjo, T. (2000). Class switch recombination and hypermutation require activation‐induced cytidine deaminase (AID), a potential RNA editing enzyme. Cell, 102(5), 553–563. 10.1016/s0092-8674(00)00078-7 [DOI] [PubMed] [Google Scholar]
  126. Musunuru, K. , & Darnell, R. B. (2001). Paraneoplastic neurologic disease antigens: RNA‐binding proteins and signaling proteins in neuronal degeneration. Annual Review of Neuroscience, 24, 239–262. 10.1146/annurev.neuro.24.1.239 [DOI] [PubMed] [Google Scholar]
  127. Nakahama, T. , Kato, Y. , Shibuya, T. , Inoue, M. , Kim, J. I. , Vongpipatana, T. , Todo, H. , Xing, Y. , & Kawahara, Y. (2021). Mutations in the adenosine deaminase ADAR1 that prevent endogenous Z‐RNA binding induce Aicardi‐Goutieres‐syndrome‐like encephalopathy. Immunity, 54(9), 1976–1988.e1977. 10.1016/j.immuni.2021.08.022 [DOI] [PubMed] [Google Scholar]
  128. Navaratnam, N. , Morrison, J. R. , Bhattacharya, S. , Patel, D. , Funahashi, T. , Giannoni, F. , Teng, B. B. , Davidson, N. O. , & Scott, J. (1993). The p27 catalytic subunit of the apolipoprotein B mRNA editing enzyme is a cytidine deaminase. The Journal of Biological Chemistry, 268(28), 20709–20712. [PubMed] [Google Scholar]
  129. Nguyen, D. T. T. , Lu, Y. , Chu, K. L. , Yang, X. , Park, S. M. , Choo, Z. N. , Chin, C. R. , Prieto, C. , Schurer, A. , Barin, E. , Savino, A. M. , Gourkanti, S. , Patel, P. , Vu, L. P. , Leslie, C. S. , & Kharas, M. G. (2020). HyperTRIBE uncovers increased MUSASHI‐2 RNA binding activity and differential regulation in leukemic stem cells. Nature Communications, 11(1), 2026. 10.1038/s41467-020-15814-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Nicholson‐Shaw, A. L. , Kofman, E. R. , Yeo, G. W. , & Pasquinelli, A. E. (2022). Nuclear and cytoplasmic poly(a) binding proteins (PABPs) favor distinct transcripts and isoforms. Nucleic Acids Research, 50(8), 4685–4702. 10.1093/nar/gkac263 [DOI] [PMC free article] [PubMed] [Google Scholar]
  131. Nishikura, K. (2010). Functions and regulation of RNA editing by ADAR deaminases. Annual Review of Biochemistry, 79, 321–349. 10.1146/annurev-biochem-060208-105251 [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Nojima, T. , Gomes, T. , Grosso, A. R. F. , Kimura, H. , Dye, M. J. , Dhir, S. , Carmo‐Fonseca, M. , & Proudfoot, N. J. (2015). Mammalian NET‐Seq reveals genome‐wide nascent transcription coupled to RNA processing. Cell, 161(3), 526–540. 10.1016/j.cell.2015.03.027 [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Norbury, C. J. (2013). Cytoplasmic RNA: A case of the tail wagging the dog. Nature Reviews. Molecular Cell Biology, 14(10), 643–653. 10.1038/nrm3645 [DOI] [PubMed] [Google Scholar]
  134. O'Brien, M. J. , Gurdziel, K. , & Ansari, A. (2023). Global run‐on sequencing to measure nascent transcription in Saccharomyces cerevisiae . Methods, 217, 18–26. 10.1016/j.ymeth.2023.06.009 [DOI] [PMC free article] [PubMed] [Google Scholar]
  135. Oka, K. , Kobayashi, K. , Sullivan, M. , Martinez, J. , Teng, B. B. , Ishimura‐Oka, K. , & Chan, L. (1997). Tissue‐specific inhibition of apolipoprotein B mRNA editing in the liver by adenovirus‐mediated transfer of a dominant negative mutant APOBEC‐1 leads to increased low density lipoprotein in mice. The Journal of Biological Chemistry, 272(3), 1456–1460. 10.1074/jbc.272.3.1456 [DOI] [PubMed] [Google Scholar]
  136. Owens, M. C. , & Liu, K. F. (2022). TRIBE‐STAMP reveals new insights into the functions of RNA binding proteins. Genes & Development, 36(17–18), 954–955. 10.1101/gad.350207.122 [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Palladino, M. J. , Keegan, L. P. , O'Connell, M. A. , & Reenan, R. A. (2000). A‐to‐I pre‐mRNA editing in drosophila is primarily involved in adult nervous system function and integrity. Cell, 102(4), 437–449. 10.1016/s0092-8674(00)00049-0 [DOI] [PubMed] [Google Scholar]
  138. Patterson, J. B. , & Samuel, C. E. (1995). Expression and regulation by interferon of a double‐stranded‐RNA‐specific adenosine deaminase from human cells: Evidence for two forms of the deaminase. Molecular and Cellular Biology, 15(10), 5376–5388. 10.1128/MCB.15.10.5376 [DOI] [PMC free article] [PubMed] [Google Scholar]
  139. Pecori, R. , di Giorgio, S. , Paulo Lorenzo, J. , & Nina Papavasiliou, F. (2022). Functions and consequences of AID/APOBEC‐mediated DNA and RNA deamination. Nature Reviews. Genetics, 23(8), 505–518. 10.1038/s41576-022-00459-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  140. Phelps, K. J. , Tran, K. , Eifler, T. , Erickson, A. I. , Fisher, A. J. , & Beal, P. A. (2015). Recognition of duplex RNA by the deaminase domain of the RNA editing enzyme ADAR2. Nucleic Acids Research, 43(2), 1123–1132. 10.1093/nar/gku1345 [DOI] [PMC free article] [PubMed] [Google Scholar]
  141. Piao, W. , Li, C. , Sun, P. , Yang, M. , Ding, Y. , Song, W. , Jia, Y. , Yu, L. , Lu, Y. , & Jin, H. (2023). Identification of RNA‐binding protein targets with HyperTRIBE in Saccharomyces cerevisiae. International Journal of Molecular Sciences, 24(10), 9033. 10.3390/ijms24109033 [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Pisacane, P. , & Halic, M. (2017). Tailing and degradation of Argonaute‐bound small RNAs protect the genome from uncontrolled RNAi. Nature Communications, 8, 15332. 10.1038/ncomms15332 [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. Porath, H. T. , Carmi, S. , & Levanon, E. Y. (2014). A genome‐wide map of hyper‐edited RNA reveals numerous new sites. Nature Communications, 5, 4726. 10.1038/ncomms5726 [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Poulsen, H. , Jorgensen, R. , Heding, A. , Nielsen, F. C. , Bonven, B. , & Egebjerg, J. (2006). Dimerization of ADAR2 is mediated by the double‐stranded RNA binding domain. RNA, 12(7), 1350–1360. 10.1261/rna.2314406 [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Poulsen, H. , Nilsson, J. , Damgaard, C. K. , Egebjerg, J. , & Kjems, J. (2001). CRM1 mediates the export of ADAR1 through a nuclear export signal within the Z‐DNA binding domain. Molecular and Cellular Biology, 21(22), 7862–7871. 10.1128/MCB.21.22.7862-7871.2001 [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Powell, C. , Cornblath, E. , & Goldman, D. (2015). Zinc‐binding domain‐dependent, deaminase‐independent actions of apolipoprotein B mRNA‐editing enzyme, catalytic polypeptide 2 (Apobec2), mediate its effect on zebrafish retina regeneration. The Journal of Biological Chemistry, 290(10), 6007. 10.1074/jbc.A114.603043 [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. Preston, M. A. , Porter, D. F. , Chen, F. , Buter, N. , Lapointe, C. P. , Keles, S. , Kimble, J. , & Wickens, M. (2019). Unbiased screen of RNA tailing activities reveals a poly(UG) polymerase. Nature Methods, 16(5), 437–445. 10.1038/s41592-019-0370-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  148. Prudencio, P. , Savisaar, R. , Rebelo, K. , Martinho, R. G. , & Carmo‐Fonseca, M. (2022). Transcription and splicing dynamics during early drosophila development. RNA, 28(2), 139–161. 10.1261/rna.078933.121 [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Quin, J. , Sedmik, J. , Vukic, D. , Khan, A. , Keegan, L. P. , & O'Connell, M. A. (2021). ADAR RNA modifications, the epitranscriptome and innate immunity. Trends in Biochemical Sciences, 46(9), 758–771. 10.1016/j.tibs.2021.02.002 [DOI] [PubMed] [Google Scholar]
  150. Rahman, R. , Xu, W. , Jin, H. , & Rosbash, M. (2018). Identification of RNA‐binding protein targets with HyperTRIBE. Nature Protocols, 13(8), 1829–1849. 10.1038/s41596-018-0020-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  151. Ramanathan, M. , Majzoub, K. , Rao, D. S. , Neela, P. H. , Zarnegar, B. J. , Mondal, S. , Roth, J. G. , Gai, H. , Kovalski, J. R. , Siprashvili, Z. , Palmer, T. D. , Carette, J. E. , & Khavari, P. A. (2018). RNA‐protein interaction detection in living cells. Nature Methods, 15(3), 207–212. 10.1038/nmeth.4601 [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Ramanathan, M. , Porter, D. F. , & Khavari, P. A. (2019). Methods to study RNA‐protein interactions. Nature Methods, 16(3), 225–234. 10.1038/s41592-019-0330-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Rathore, A. , Carpenter, M. A. , Demir, O. , Ikeda, T. , Li, M. , Shaban, N. M. , Law, E. K. , Anokhin, D. , Brown, W. L. , Amaro, R. E. , & Harris, R. S. (2013). The local dinucleotide preference of APOBEC3G can be altered from 5'‐CC to 5'‐TC by a single amino acid substitution. Journal of Molecular Biology, 425(22), 4442–4454. 10.1016/j.jmb.2013.07.040 [DOI] [PMC free article] [PubMed] [Google Scholar]
  154. Rayon‐Estrada, V. , Harjanto, D. , Hamilton, C. E. , Berchiche, Y. A. , Gantman, E. C. , Sakmar, T. P. , Bulloch, K. , Gagnidze, K. , Harroch, S. , McEwen, B. S. , & Papavasiliou, F. N. (2017). Epitranscriptomic profiling across cell types reveals associations between APOBEC1‐mediated RNA editing, gene expression outcomes, and cellular function. Proceedings of the National Academy of Sciences of the United States of America, 114(50), 13296–13301. 10.1073/pnas.1714227114 [DOI] [PMC free article] [PubMed] [Google Scholar]
  155. Refsland, E. W. , & Harris, R. S. (2013). The APOBEC3 family of retroelement restriction factors. Current Topics in Microbiology and Immunology, 371, 1–27. 10.1007/978-3-642-37765-5_1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Reimao‐Pinto, M. M. , Ignatova, V. , Burkard, T. R. , Hung, J. H. , Manzenreither, R. A. , Sowemimo, I. , Herzog, V. A. , Reichholf, B. , Farina‐Lopez, S. , & Ameres, S. L. (2015). Uridylation of RNA hairpins by tailor confines the emergence of MicroRNAs in drosophila. Molecular Cell, 59(2), 203–216. 10.1016/j.molcel.2015.05.033 [DOI] [PMC free article] [PubMed] [Google Scholar]
  157. Revathidevi, S. , Murugan, A. K. , Nakaoka, H. , Inoue, I. , & Munirajan, A. K. (2021). APOBEC: A molecular driver in cervical cancer pathogenesis. Cancer Letters, 496, 104–116. 10.1016/j.canlet.2020.10.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Reyes, J. M. , & Ross, P. J. (2016). Cytoplasmic polyadenylation in mammalian oocyte maturation. WIREs RNA, 7(1), 71–89. 10.1002/wrna.1316 [DOI] [PubMed] [Google Scholar]
  159. Ricci, E. P. , Kucukural, A. , Cenik, C. , Mercier, B. C. , Singh, G. , Heyer, E. E. , Ashar‐Patel, A. , Peng, L. , & Moore, M. J. (2014). Staufen1 senses overall transcript secondary structure to regulate translation. Nature Structural & Molecular Biology, 21(1), 26–35. 10.1038/nsmb.2739 [DOI] [PMC free article] [PubMed] [Google Scholar]
  160. Riley, K. J. , Yario, T. A. , & Steitz, J. A. (2012). Association of Argonaute proteins and microRNAs can occur after cell lysis. RNA, 18(9), 1581–1585. 10.1261/rna.034934.112 [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. Rissland, O. S. , & Norbury, C. J. (2009). Decapping is preceded by 3′ uridylation in a novel pathway of bulk mRNA turnover. Nature Structural & Molecular Biology, 16(6), 616–623. 10.1038/nsmb.1601 [DOI] [PMC free article] [PubMed] [Google Scholar]
  162. Rosenberg, B. R. , Hamilton, C. E. , Mwangi, M. M. , Dewell, S. , & Papavasiliou, F. N. (2011). Transcriptome‐wide sequencing reveals numerous APOBEC1 mRNA‐editing targets in transcript 3' UTRs. Nature Structural & Molecular Biology, 18(2), 230–236. 10.1038/nsmb.1975 [DOI] [PMC free article] [PubMed] [Google Scholar]
  163. Rosenthal, J. J. , & Seeburg, P. H. (2012). A‐to‐I RNA editing: Effects on proteins key to neural excitability. Neuron, 74(3), 432–439. 10.1016/j.neuron.2012.04.010 [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Rozen‐Gagnon, K. , Gu, M. , Luna, J. M. , Luo, J. D. , Yi, S. , Novack, S. , Jacobson, E. , Wang, W. , Paul, M. R. , Scheel, T. K. H. , Carroll, T. , & Rice, C. M. (2021). Argonaute‐CLIP delineates versatile, functional RNAi networks in Aedes aegypti, a major vector of human viruses. Cell Host & Microbe, 29(5), 834–848 e813. 10.1016/j.chom.2021.03.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  165. Ruan, X. , Hu, K. , & Zhang, X. (2023). PIE‐seq: Identifying RNA‐binding protein targets by dual RNA‐deaminase editing and sequencing. Nature Communications, 14(1), 3275. 10.1038/s41467-023-39054-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  166. Ruegger, S. , Miki, T. S. , Hess, D. , & Grosshans, H. (2015). The ribonucleotidyl transferase USIP‐1 acts with SART3 to promote U6 snRNA recycling. Nucleic Acids Research, 43(6), 3344–3357. 10.1093/nar/gkv196 [DOI] [PMC free article] [PubMed] [Google Scholar]
  167. Salter, J. D. , Bennett, R. P. , & Smith, H. C. (2016). The APOBEC protein family: United by structure. Divergent in Function. Trends in Biochemical Sciences, 41(7), 578–594. 10.1016/j.tibs.2016.05.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Sansam, C. L. , Wells, K. S. , & Emeson, R. B. (2003). Modulation of RNA editing by functional nucleolar sequestration of ADAR2. Proceedings of the National Academy of Sciences of the United States of America, 100(24), 14018–14023. 10.1073/pnas.2336131100 [DOI] [PMC free article] [PubMed] [Google Scholar]
  169. Sarkis, S. , Dabo, S. , Lise, M. C. , Neuveut, C. , Meurs, E. F. , Lacoste, V. , & Lavergne, A. (2018). A potential robust antiviral defense state in the common vampire bat: Expression, induction and molecular characterization of the three interferon‐stimulated genes ‐OAS1, ADAR1 and PKR. Developmental and Comparative Immunology, 85, 95–107. 10.1016/j.dci.2018.04.006 [DOI] [PubMed] [Google Scholar]
  170. Sato, Y. , Probst, H. C. , Tatsumi, R. , Ikeuchi, Y. , Neuberger, M. S. , & Rada, C. (2010). Deficiency in APOBEC2 leads to a shift in muscle fiber type, diminished body mass, and myopathy. The Journal of Biological Chemistry, 285(10), 7111–7118. 10.1074/jbc.M109.052977 [DOI] [PMC free article] [PubMed] [Google Scholar]
  171. Savva, Y. A. , Rieder, L. E. , & Reenan, R. A. (2012). The ADAR protein family. Genome Biology, 13(12), 252. 10.1186/gb-2012-13-12-252 [DOI] [PMC free article] [PubMed] [Google Scholar]
  172. Schneider, M. F. , Wettengel, J. , Hoffmann, P. C. , & Stafforst, T. (2014). Optimal guideRNAs for re‐directing deaminase activity of hADAR1 and hADAR2 in trans. Nucleic Acids Research, 42(10), e87. 10.1093/nar/gku272 [DOI] [PMC free article] [PubMed] [Google Scholar]
  173. Sekar, V. , Marmol‐Sanchez, E. , Kalogeropoulos, P. , Stanicek, L. , Sagredo, E. A. , Widmark, A. , Doukoumopoulos, E. , Bonath, F. , Biryukova, I. , & Friedlander, M. R. (2023). Detection of transcriptome‐wide microRNA‐target interactions in single cells with agoTRIBE. Nature Biotechnology. 10.1038/s41587-023-01951-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  174. Seo, K. W. , & Kleiner, R. E. (2023). Profiling dynamic RNA‐protein interactions using small‐molecule‐induced RNA editing. Nature Chemical Biology, 19(11), 1361–1371. 10.1038/s41589-023-01372-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  175. Shaban, N. M. , Shi, K. , Lauer, K. V. , Carpenter, M. A. , Richards, C. M. , Salamango, D. , Wang, J. , Lopresti, M. W. , Banerjee, S. , Levin‐Klein, R. , Brown, W. L. , Aihara, H. , & Harris, R. S. (2018). The antiviral and cancer genomic DNA deaminase APOBEC3H is regulated by an RNA‐mediated dimerization mechanism. Molecular Cell, 69(1), 75–86 e79. 10.1016/j.molcel.2017.12.010 [DOI] [PMC free article] [PubMed] [Google Scholar]
  176. Shi, K. , Carpenter, M. A. , Banerjee, S. , Shaban, N. M. , Kurahashi, K. , Salamango, D. J. , McCann, J. L. , Starrett, G. J. , Duffy, J. V. , Demir, O. , Amaro, R. E. , Harki, D. A. , Harris, R. S. , & Aihara, H. (2017). Structural basis for targeted DNA cytosine deamination and mutagenesis by APOBEC3A and APOBEC3B. Nature Structural & Molecular Biology, 24(2), 131–139. 10.1038/nsmb.3344 [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Simon, M. D. , Wang, C. I. , Kharchenko, P. V. , West, J. A. , Chapman, B. A. , Alekseyenko, A. A. , Borowsky, M. L. , Kuroda, M. I. , & Kingston, R. E. (2011). The genomic binding sites of a noncoding RNA. Proceedings of the National Academy of Sciences of the United States of America, 108(51), 20497–20502. 10.1073/pnas.1113536108 [DOI] [PMC free article] [PubMed] [Google Scholar]
  178. Singh, A. , Hulsmeier, J. , Kandi, A. R. , Pothapragada, S. S. , Hillebrand, J. , Petrauskas, A. , Agrawal, K. , Rt, K. , Thiagarajan, D. , Jayaprakashappa, D. , VijayRaghavan, K. , Ramaswami, M. , & Bakthavachalu, B. (2021). Antagonistic roles for Ataxin‐2 structured and disordered domains in RNP condensation. eLife, 10, e60326. 10.7554/eLife.60326 [DOI] [PMC free article] [PubMed] [Google Scholar]
  179. Siprashvili, Z. , Webster, D. E. , Johnston, D. , Shenoy, R. M. , Ungewickell, A. J. , Bhaduri, A. , Flockhart, R. , Zarnegar, B. J. , Che, Y. , Meschi, F. , Puglisi, J. D. , & Khavari, P. A. (2016). The noncoding RNAs SNORD50A and SNORD50B bind K‐Ras and are recurrently deleted in human cancer. Nature Genetics, 48(1), 53–58. 10.1038/ng.3452 [DOI] [PMC free article] [PubMed] [Google Scholar]
  180. Slotkin, W. , & Nishikura, K. (2013). Adenosine‐to‐inosine RNA editing and human disease. Genome Medicine, 5(11), 105. 10.1186/gm508 [DOI] [PMC free article] [PubMed] [Google Scholar]
  181. Smith, H. C. , Bennett, R. P. , Kizilyer, A. , McDougall, W. M. , & Prohaska, K. M. (2012). Functions and regulation of the APOBEC family of proteins. Seminars in Cell & Developmental Biology, 23(3), 258–268. 10.1016/j.semcdb.2011.10.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  182. Soleymanjahi, S. , Blanc, V. , & Davidson, N. (2021). APOBEC1 mediated C‐to‐U RNA editing: Target sequence and trans‐acting factor contribution to 177 RNA editing events in 119 murine transcripts in‐vivo. RNA, 27(8), 876–890. 10.1261/rna.078678.121 [DOI] [PMC free article] [PubMed] [Google Scholar]
  183. Song, B. , Shiromoto, Y. , Minakuchi, M. , & Nishikura, K. (2022). The role of RNA editing enzyme ADAR1 in human disease. WIREs RNA, 13(1), e1665. 10.1002/wrna.1665 [DOI] [PMC free article] [PubMed] [Google Scholar]
  184. Song, W., Wang, J., Piao, W., Yang, Y., Son, A., Chang, H., Wen, X., Zhang, H., Li, C., Na, D., Lu, Y., Menet, J., Kim, V. N., & Jin, H. (2023). The role of terminal uridyl transferases in the circadian rhythm. New Results, 10.1101/2023.03.13.516290 [DOI]
  185. St Laurent, G. , Tackett, M. R. , Nechkin, S. , Shtokalo, D. , Antonets, D. , Savva, Y. A. , Maloney, R. , Kapranov, P. , Lawrence, C. E. , & Reenan, R. A. (2013). Genome‐wide analysis of A‐to‐I RNA editing by single‐molecule sequencing in drosophila. Nature Structural & Molecular Biology, 20(11), 1333–1339. 10.1038/nsmb.2675 [DOI] [PubMed] [Google Scholar]
  186. Stark, R. , Grzelak, M. , & Hadfield, J. (2019). RNA sequencing: The teenage years. Nature Reviews. Genetics, 20(11), 631–656. 10.1038/s41576-019-0150-2 [DOI] [PubMed] [Google Scholar]
  187. Stefl, R. , Oberstrass, F. C. , Hood, J. L. , Jourdan, M. , Zimmermann, M. , Skrisovska, L. , Maris, C. , Peng, L. , Hofr, C. , Emeson, R. B. , & Allain, F. H. (2010). The solution structure of the ADAR2 dsRBM‐RNA complex reveals a sequence‐specific readout of the minor groove. Cell, 143(2), 225–237. 10.1016/j.cell.2010.09.026 [DOI] [PMC free article] [PubMed] [Google Scholar]
  188. Stefl, R. , Xu, M. , Skrisovska, L. , Emeson, R. B. , & Allain, F. H. (2006). Structure and specific RNA binding of ADAR2 double‐stranded RNA binding motifs. Structure, 14(2), 345–355. 10.1016/j.str.2005.11.013 [DOI] [PubMed] [Google Scholar]
  189. Stephens, O. M. , Haudenschild, B. L. , & Beal, P. A. (2004). The binding selectivity of ADAR2's dsRBMs contributes to RNA‐editing selectivity. Chemistry & Biology, 11(9), 1239–1250. 10.1016/j.chembiol.2004.06.009 [DOI] [PubMed] [Google Scholar]
  190. Stroppel, A. S. , Latifi, N. , Hanswillemenke, A. , Tasakis, R. N. , Papavasiliou, F. N. , & Stafforst, T. (2021). Harnessing self‐labeling enzymes for selective and concurrent A‐to‐I and C‐to‐U RNA base editing. Nucleic Acids Research, 49(16), e95. 10.1093/nar/gkab541 [DOI] [PMC free article] [PubMed] [Google Scholar]
  191. Subtelny, A. O. , Eichhorn, S. W. , Chen, G. R. , Sive, H. , & Bartel, D. P. (2014). Poly(a)‐tail profiling reveals an embryonic switch in translational control. Nature, 508(7494), 66–71. 10.1038/nature13007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  192. Tao, S. , Xie, S. J. , Diao, L. T. , Lv, G. , Hou, Y. R. , Hu, Y. X. , Xu, W. Y. , Du, B. , & Xiao, Z. D. (2023). RNA‐binding protein CCDC137 activates AKT signaling and promotes hepatocellular carcinoma through a novel non‐canonical role of DGCR8 in mRNA localization. Journal of Experimental & Clinical Cancer Research, 42(1), 194. 10.1186/s13046-023-02749-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
  193. Tenenbaum, S. A. , Carson, C. C. , Lager, P. J. , & Keene, J. D. (2000). Identifying mRNA subsets in messenger ribonucleoprotein complexes by using cDNA arrays. Proceedings of the National Academy of Sciences of the United States of America, 97(26), 14085–14090. 10.1073/pnas.97.26.14085 [DOI] [PMC free article] [PubMed] [Google Scholar]
  194. Teng, B. B. , Ochsner, S. , Zhang, Q. , Soman, K. V. , Lau, P. P. , & Chan, L. (1999). Mutational analysis of apolipoprotein B mRNA editing enzyme (APOBEC1). Structure‐function relationships of RNA editing and dimerization. Journal of Lipid Research, 40(4), 623–635. [PubMed] [Google Scholar]
  195. Tonkin, L. A. , Saccomanno, L. , Morse, D. P. , Brodigan, T. , Krause, M. , & Bass, B. L. (2002). RNA editing by ADARs is important for normal behavior in Caenorhabditis elegans. The EMBO Journal, 21(22), 6025–6035. 10.1093/emboj/cdf607 [DOI] [PMC free article] [PubMed] [Google Scholar]
  196. Trippe, R. , Guschina, E. , Hossbach, M. , Urlaub, H. , Luhrmann, R. , & Benecke, B. J. (2006). Identification, cloning, and functional analysis of the human U6 snRNA‐specific terminal uridylyl transferase. RNA, 12(8), 1494–1504. 10.1261/rna.87706 [DOI] [PMC free article] [PubMed] [Google Scholar]
  197. Tsai, B. P. , Wang, X. , Huang, L. , & Waterman, M. L. (2011). Quantitative profiling of in vivo‐assembled RNA‐protein complexes using a novel integrated proteomic approach. Molecular & Cellular Proteomics, 10(4), M110 007385. 10.1074/mcp.M110.007385 [DOI] [PMC free article] [PubMed] [Google Scholar]
  198. Ule, J. , Jensen, K. , Mele, A. , & Darnell, R. B. (2005). CLIP: A method for identifying protein‐RNA interaction sites in living cells. Methods, 37(4), 376–386. 10.1016/j.ymeth.2005.07.018 [DOI] [PubMed] [Google Scholar]
  199. Ule, J. , Jensen, K. B. , Ruggiu, M. , Mele, A. , Ule, A. , & Darnell, R. B. (2003). CLIP identifies Nova‐regulated RNA networks in the brain. Science, 302(5648), 1212–1215. 10.1126/science.1090095 [DOI] [PubMed] [Google Scholar]
  200. Valente, L. , & Nishikura, K. (2007). RNA binding‐independent dimerization of adenosine deaminases acting on RNA and dominant negative effects of nonfunctional subunits on dimer functions. The Journal of Biological Chemistry, 282(22), 16054–16061. 10.1074/jbc.M611392200 [DOI] [PMC free article] [PubMed] [Google Scholar]
  201. van Nostrand, E. L. , Freese, P. , Pratt, G. A. , Wang, X. , Wei, X. , Xiao, R. , Blue, S. M. , Chen, J. Y. , Cody, N. A. L. , Dominguez, D. , Olson, S. , Sundararaman, B. , Zhan, L. , Bazile, C. , Bouvrette, L. P. B. , Bergalet, J. , Duff, M. O. , Garcia, K. E. , Gelboin‐Burkhart, C. , … Yeo, G. W. (2020). A large‐scale binding and functional map of human RNA‐binding proteins. Nature, 583(7818), 711–719. 10.1038/s41586-020-2077-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
  202. van Nostrand, E. L. , Pratt, G. A. , Shishkin, A. A. , Gelboin‐Burkhart, C. , Fang, M. Y. , Sundararaman, B. , Blue, S. M. , Nguyen, T. B. , Surka, C. , Elkins, K. , Stanton, R. , Rigo, F. , Guttman, M. , & Yeo, G. W. (2016). Robust transcriptome‐wide discovery of RNA‐binding protein binding sites with enhanced CLIP (eCLIP). Nature Methods, 13(6), 508–514. 10.1038/nmeth.3810 [DOI] [PMC free article] [PubMed] [Google Scholar]
  203. van Nostrand, E. L. , Pratt, G. A. , Yee, B. A. , Wheeler, E. C. , Blue, S. M. , Mueller, J. , Park, S. S. , Garcia, K. E. , Gelboin‐Burkhart, C. , Nguyen, T. B. , Rabano, I. , Stanton, R. , Sundararaman, B. , Wang, R. , Fu, X. D. , Graveley, B. R. , & Yeo, G. W. (2020). Principles of RNA processing from analysis of enhanced CLIP maps for 150 RNA binding proteins. Genome Biology, 21(1), 90. 10.1186/s13059-020-01982-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  204. Vogel, P. , Schneider, M. F. , Wettengel, J. , & Stafforst, T. (2014). Improving site‐directed RNA editing in vitro and in cell culture by chemical modification of the guideRNA. Angewandte Chemie (International Ed. in English), 53(24), 6267–6271. 10.1002/anie.201402634 [DOI] [PubMed] [Google Scholar]
  205. Vogel, P. , & Stafforst, T. (2014). Site‐directed RNA editing with antagomir deaminases—A tool to study protein and RNA function. ChemMedChem, 9(9), 2021–2025. 10.1002/cmdc.201402139 [DOI] [PubMed] [Google Scholar]
  206. Wang, F. , Tidei, J. J. , Polich, E. D. , Gao, Y. , Zhao, H. , Perrone‐Bizzozero, N. I. , Guo, W. , & Zhao, X. (2015). Positive feedback between RNA‐binding protein HuD and transcription factor SATB1 promotes neurogenesis. Proceedings of the National Academy of Sciences of the United States of America, 112(36), E4995–E5004. 10.1073/pnas.1513780112 [DOI] [PMC free article] [PubMed] [Google Scholar]
  207. Wang, Q. , Hui, H. , Guo, Z. , Zhang, W. , Hu, Y. , He, T. , Tai, Y. , Peng, P. , & Wang, L. (2013). ADAR1 regulates ARHGAP26 gene expression through RNA editing by disrupting miR‐30b‐3p and miR‐573 binding. RNA, 19(11), 1525–1536. 10.1261/rna.041533.113 [DOI] [PMC free article] [PubMed] [Google Scholar]
  208. Wang, Y. , & Beal, P. A. (2016). Probing RNA recognition by human ADAR2 using a high‐throughput mutagenesis method. Nucleic Acids Research, 44(20), 9872–9880. 10.1093/nar/gkw799 [DOI] [PMC free article] [PubMed] [Google Scholar]
  209. Wang, Y. , Chung, D. H. , Monteleone, L. R. , Li, J. , Chiang, Y. , Toney, M. D. , & Beal, P. A. (2019). RNA binding candidates for human ADAR3 from substrates of a gain of function mutant expressed in neuronal cells. Nucleic Acids Research, 47(20), 10801–10814. 10.1093/nar/gkz815 [DOI] [PMC free article] [PubMed] [Google Scholar]
  210. Wang, Y. , Havel, J. , & Beal, P. A. (2015). A phenotypic screen for functional mutants of human adenosine deaminase acting on RNA 1. ACS Chemical Biology, 10(11), 2512–2519. 10.1021/acschembio.5b00711 [DOI] [PMC free article] [PubMed] [Google Scholar]
  211. Wang, Y. , Park, S. , & Beal, P. A. (2018). Selective recognition of RNA substrates by ADAR deaminase domains. Biochemistry, 57(10), 1640–1651. 10.1021/acs.biochem.7b01100 [DOI] [PMC free article] [PubMed] [Google Scholar]
  212. Wang, Y. , Zheng, Y. , & Beal, P. A. (2017). Adenosine deaminases that act on RNA (ADARs). Enzyme, 41, 215–268. 10.1016/bs.enz.2017.03.006 [DOI] [PubMed] [Google Scholar]
  213. Warkocki, Z. , Liudkovska, V. , Gewartowska, O. , Mroczek, S. , & Dziembowski, A. (2018). Terminal nucleotidyl transferases (TENTs) in mammalian RNA metabolism. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 373(1762), 20180162. 10.1098/rstb.2018.0162 [DOI] [PMC free article] [PubMed] [Google Scholar]
  214. Wen, X. , Irshad, A. , & Jin, H. (2023). The Battle for survival: The role of RNA non‐canonical tails in the virus‐host interaction. Metabolites, 13(9), 1009 . 10.3390/metabo13091009 [DOI] [PMC free article] [PubMed] [Google Scholar]
  215. West, J. A. , Davis, C. P. , Sunwoo, H. , Simon, M. D. , Sadreyev, R. I. , Wang, P. I. , Tolstorukov, M. Y. , & Kingston, R. E. (2014). The long noncoding RNAs NEAT1 and MALAT1 bind active chromatin sites. Molecular Cell, 55(5), 791–802. 10.1016/j.molcel.2014.07.012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  216. West, S. , Gromak, N. , Norbury, C. J. , & Proudfoot, N. J. (2006). Adenylation and exosome‐mediated degradation of cotranscriptionally cleaved pre‐messenger RNA in human cells. Molecular Cell, 21(3), 437–443. 10.1016/j.molcel.2005.12.008 [DOI] [PubMed] [Google Scholar]
  217. Wheeler, E. C. , Washburn, M. C. , Major, F. , Rusch, D. B. , & Hundley, H. A. (2015). Noncoding regions of C. Elegans mRNA undergo selective adenosine to inosine deamination and contain a small number of editing sites per transcript. RNA Biology, 12(2), 162–174. 10.1080/15476286.2015.1017220 [DOI] [PMC free article] [PubMed] [Google Scholar]
  218. Wissink, E. M. , Vihervaara, A. , Tippens, N. D. , & Lis, J. T. (2019). Nascent RNA analyses: Tracking transcription and its regulation. Nature Reviews. Genetics, 20(12), 705–723. 10.1038/s41576-019-0159-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  219. Wolfe, A. D. , Arnold, D. B. , & Chen, X. S. (2019). Comparison of RNA editing activity of APOBEC1‐A1CF and APOBEC1‐RBM47 complexes reconstituted in HEK293T cells. Journal of Molecular Biology, 431(7), 1506–1517. 10.1016/j.jmb.2019.02.025 [DOI] [PMC free article] [PubMed] [Google Scholar]
  220. Wolfe, A. D. , Li, S. , Goedderz, C. , & Chen, X. S. (2020). The structure of APOBEC1 and insights into its RNA and DNA substrate selectivity. NAR Cancer, 2(4), zcaa027. 10.1093/narcan/zcaa027 [DOI] [PMC free article] [PubMed] [Google Scholar]
  221. Wong, S. K. , Sato, S. , & Lazinski, D. W. (2001). Substrate recognition by ADAR1 and ADAR2. RNA, 7(6), 846–858. 10.1017/s135583820101007x [DOI] [PMC free article] [PubMed] [Google Scholar]
  222. Xing, Y. , Nakahama, T. , Wu, Y. , Inoue, M. , Kim, J. I. , Todo, H. , Shibuya, T. , Kato, Y. , & Kawahara, Y. (2023). RNA editing of AZIN1 coding sites is catalyzed by ADAR1 p150 after splicing. The Journal of Biological Chemistry, 299(7), 104840. 10.1016/j.jbc.2023.104840 [DOI] [PMC free article] [PubMed] [Google Scholar]
  223. Xu, W. , Rahman, R. , & Rosbash, M. (2018). Mechanistic implications of enhanced editing by a HyperTRIBE RNA‐binding protein. RNA, 24(2), 173–182. 10.1261/rna.064691.117 [DOI] [PMC free article] [PubMed] [Google Scholar]
  224. Yang, J. H. , Nie, Y. , Zhao, Q. , Su, Y. , Pypaert, M. , Su, H. , & Rabinovici, R. (2003). Intracellular localization of differentially regulated RNA‐specific adenosine deaminase isoforms in inflammation. The Journal of Biological Chemistry, 278(46), 45833–45842. 10.1074/jbc.M308612200 [DOI] [PubMed] [Google Scholar]
  225. Yang, M. , Lu, Y. , Piao, W. , & Jin, H. (2022). The translational regulation in mTOR pathway. Biomolecules, 12(6), 802. 10.3390/biom12060802 [DOI] [PMC free article] [PubMed] [Google Scholar]
  226. Yang, Y. , Okada, S. , & Sakurai, M. (2021). Adenosine‐to‐inosine RNA editing in neurological development and disease. RNA Biology, 18(7), 999–1013. 10.1080/15476286.2020.1867797 [DOI] [PMC free article] [PubMed] [Google Scholar]
  227. Yashiro, Y. , & Tomita, K. (2018). Function and regulation of human terminal Uridylyltransferases. Frontiers in Genetics, 9, 538. 10.3389/fgene.2018.00538 [DOI] [PMC free article] [PubMed] [Google Scholar]
  228. Yu, S. , & Kim, V. N. (2020). A tale of non‐canonical tails: Gene regulation by post‐transcriptional RNA tailing. Nature Reviews. Molecular Cell Biology, 21(9), 542–556. 10.1038/s41580-020-0246-8 [DOI] [PubMed] [Google Scholar]
  229. Zalfa, F. , Giorgi, M. , Primerano, B. , Moro, A. , di Penta, A. , Reis, S. , Oostra, B. , & Bagni, C. (2003). The fragile X syndrome protein FMRP associates with BC1 RNA and regulates the translation of specific mRNAs at synapses. Cell, 112(3), 317–327. 10.1016/s0092-8674(03)00079-5 [DOI] [PubMed] [Google Scholar]
  230. Zarnegar, B. J. , Flynn, R. A. , Shen, Y. , Do, B. T. , Chang, H. Y. , & Khavari, P. A. (2016). irCLIP platform for efficient characterization of protein‐RNA interactions. Nature Methods, 13(6), 489–492. 10.1038/nmeth.3840 [DOI] [PMC free article] [PubMed] [Google Scholar]
  231. Zeng, F. , Peritz, T. , Kannanayakal, T. J. , Kilk, K. , Eiriksdottir, E. , Langel, U. , & Eberwine, J. (2006). A protocol for PAIR: PNA‐assisted identification of RNA binding proteins in living cells. Nature Protocols, 1(2), 920–927. 10.1038/nprot.2006.81 [DOI] [PubMed] [Google Scholar]
  232. Zhang, J. , Zhang, G. , Zhang, W. , Bai, L. , Wang, L. , Li, T. , Yan, L. , Xu, Y. , Chen, D. , Gao, W. , Gao, C. , Chen, C. , Ren, M. , Jiao, Y. , Qin, H. , Sun, Y. , Zhi, L. , Qi, Y. , Zhao, J. , … Wang, Y. (2022). Loss of RBMS1 promotes anti‐tumor immunity through enabling PD‐L1 checkpoint blockade in triple‐negative breast cancer. Cell Death and Differentiation, 29(11), 2247–2261. 10.1038/s41418-022-01012-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  233. Zhao, J. , Ohsumi, T. K. , Kung, J. T. , Ogawa, Y. , Grau, D. J. , Sarma, K. , Song, J. J. , Kingston, R. E. , Borowsky, M. , & Lee, J. T. (2010). Genome‐wide identification of polycomb‐associated RNAs by RIP‐seq. Molecular Cell, 40(6), 939–953. 10.1016/j.molcel.2010.12.011 [DOI] [PMC free article] [PubMed] [Google Scholar]
  234. Zheng, X. , Cho, S. , Moon, H. , Loh, T. J. , Jang, H. N. , & Shen, H. (2016). Detecting RNA‐protein interaction using end‐labeled biotinylated RNA oligonucleotides and immunoblotting. Methods in Molecular Biology, 1421, 35–44. 10.1007/978-1-4939-3591-8_4 [DOI] [PubMed] [Google Scholar]
  235. Zhou, G. , Niu, R. , Zhou, Y. , Luo, M. , Peng, Y. , Wang, H. , Wang, Z. , & Xu, G. (2021). Proximity editing to identify RNAs in phase‐separated RNA binding protein condensates. Cell Discovery, 7(1), 72. 10.1038/s41421-021-00288-9 [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

Data sharing is not applicable to this article as no new data were created or analyzed in this study.


Articles from Wiley Interdisciplinary Reviews. RNA are provided here courtesy of Wiley

RESOURCES