Highlights
-
•
Methodological progresses and piling evidence prove the rG4 biology in vivo.
-
•
rG4s step in virtually every aspect of RNA biology.
-
•
Helicases unwinding of rG4s is a fine regulatory layer to the downstream processes and general cell homeostasis.
-
•
The current knowledge is however limited to a few cell lines.
-
•
The regulation of helicases themselves is delineating as a important question.
-
•
Non-helicase rG4-processing proteins likely play a role.
Keywords: RNA, Quadruplex, Helicase
Abstract
The nucleic acid structure called G-quadruplex (G4) is currently discussed to function in nucleic acid-based mechanisms that influence several cellular processes. They can modulate the cellular machinery either positively or negatively, both at the DNA and RNA level. The majority of what we know about G4 biology comes from DNA G4 (dG4) research. RNA G4s (rG4), on the other hand, are gaining interest as researchers become more aware of their role in several aspects of cellular homeostasis. In either case, the correct regulation of G4 structures within cells is essential and demands specialized proteins able to resolve them. Small changes in the formation and unfolding of G4 structures can have severe consequences for the cells that could even stimulate genome instability, apoptosis or proliferation. Helicases are the most relevant negative G4 regulators, which prevent and unfold G4 formation within cells during different pathways. Yet, and despite their importance only a handful of rG4 unwinding helicases have been identified and characterized thus far. This review addresses the current knowledge on rG4s-processing helicases with a focus on methodological approaches. An example of a non-helicase rG4s-unwinding protein is also briefly described.
1. Introduction
Nucleic acid secondary structures provide additional biological information beyond the primary sequence. The formation of nucleic acid structures is widely recognized for their importance during several cellular processes adding up layers of complexity to the web of cellular regulatory mechanisms. Among these are G-quadruplex structures (G4s), stable DNA or RNA structures formed within guanine-rich sequences, which can self-assemble via Hoogsteen hydrogen bonding into (at least two, most frequently three) stacking guanine tetrads stabilized by monovalent cations [1], [2], [3]. The relative orientation of each guanine in the tetrads defines the overall quadruplex topology, which can be summarized in the conventional parallel, antiparallel and hybrid scheme [4]. The inherent biophysical differences between DNA and RNA reflect onto the respective G4 structures [5], [6], because the extra 2′-OH moiety in RNA enables additional hydrogen bonding and affects molecular hydration thus making RNA G4s (rG4s) more stable than the DNA counterpart [1], [7], [8]. In contrast to DNA G4s (dG4s) that can explore a broader conformational space, the 2′-OH moiety in rG4s also introduces steric constriction that forces rG4s into the parallel topology [9], [10] with some remarkable exceptions [11], [12], [13]. In addition to the biophysical aspects the biology of the two classes of G4 diverges because of the diverse processes they operate in although through similar underlying principles.
Roughly 700.000 putative G4-forming sequences are spread throughout the human genome. They are enriched at functional loci, hinting at the role that G4s have in a vast network of cellular processes [14], [15], [16], [17]. Notwithstanding early skepticism there is cogent evidence for the in vivo formation and involvement of dG4s in transcription regulation [18], [19], telomeres stability [20], [21], epigenetics [22], [23], [24], DNA repair mechanisms [25], [26], replication [27], [28], [29], [30], genome stability [30], [31], [32] and recombination [33], [34], [35], which are all reviewed in detail elsewhere [36], [37], [38]. In particular the finding of dG4 formation within promoters of prominent oncogene e.g. KRAS [39], [40], [41], [42], VEGF [43], KIT [44], [45], BCL2 [46] and MYC [47], [48], [49], [50] and the potential to use these structure to control the gene expression, telomerase action and genome stability opened the possibility to use G4s as potential druggable anticancer targets [37], [51], [52], [53], [54]. To improve current therapeutic strategies, it is essential to understand the relevance of DNA and RNA G4s within cells and how they are regulated within cells by helicases.
Likewise dG4s, the actual folding of rG4s in living cells has been long debated given the emergence of conflicting findings [55]. Most rG4s-forming sequences are enriched at ncRNAs and 5′- and 3′-UTR non-coding regions [14] and were first visualized in human cells by ligand-based rG4-selective probing [56], [57] with roughly 13.000 putative rG4s mapped to the transcriptome by reverse transcriptase stop assay combined with next-generation sequencing [58]. Controversy arose upon the report of the coupled DMS/RT-stop analysis from Guo and Bartel [59] arguing for mostly unfolded rG4s in eukaryotes, at equilibrium. This finding is in contrast to other publications showing that although DMS can be used to measure G4-folding kinetics in vitro, DMS-seq can not be used to map G4 formation in living cells. Using a single molecular live cell imaging approach, a recent publication showed that G4 structure formation dynamically changes in the cell, which makes DMS ineffective, because only the unfolded state will be trapped by this method [60]. Furthermore, different methods in different model systems and cell types support the existence of rG4s. For example, with the fluorescent probe QUMA-1 [61] as well as with the biotinylated template-assembled synthetic G-quartets (Bio-TASQ) [62], [63] rG4 visualization, pull-down analysis and global mappings are possible. These and other experiments demonstrate the relevance and function of rG4s in a dense grid of cellular processes [55], [64], [65]. In detail, rG4 have been shown to modulate post-transcriptional events, particularly translation [66], [67], [68], [69], [70], [71], [72], localization [73], [74] and splicing [75], [76], [77], [78], [79], [80], [81], [82], [83], [84]. These findings advocate for a dynamic protein-regulated rG4 scenario with profound implications in living cells. The apparent incoherence between the work of Guo and Bartel with other experimental findings about rG4 is summarized and discussed in recent reviews [64], [85].
rG4s are now accepted to form dynamically in vivo as transient events in an ever-changing rG4 landscape influenced by exogenous and endogenous stimuli and under control of the cellular regulatory machinery. This model can consistently explain the seemingly contradictory findings from the aforementioned experimental approaches. Accordingly, only specific rG4 subsets exist at any given time as a function of cell type, cell-cycle progression [86], environmental stress [87], etc. The higher turnover of the rG4 pool compared to dG4 is also conceivable considering the inherent nature of RNA and its very dynamic processes which, unlike given for DNA, are intrinsically transient and relatively short-lived to ensure readily responsive regulation of cellular functions [88].
Regardless of the polymer nature highly stable G4 structures can act as kinetic traps and must ultimately be unfolded to ensure plasticity to the regulatory system and prevent detrimental effects on downstream processes. Several works probing the misregulation of G4s using stabilizing ligands [22], [89], [90] or targeted knockouts (KO) [91] have been conducted. Whilst the folding process is thermodynamically driven and can occur regardless of supporting proteins [26], [92], [93], [94], G4 unfolding requires proteins that either compete with the folding of the quadruplex (i.e. binding and ‘sequestering’ the unfolded or alternative species) [68], [95] or specialized helicases actively unfolding the G4 [19], [33], [96], [97], [98], [99], [100]. Although the information gathered from dG4s largely dominates the field the pervasive impact on the cell homeostasis of rG4-mediated processes and the identification of rG4-unwinding helicases and proteins increasingly shifted the interest to also address how rG4s interface with the cellular regulatory machinery [55], [64], [96], [62], [101], [102], [103], [104], [105], [106], [107], [108], [109], [110], [111], [112], [113].
Several recent reviews exhaustively document the generalities the biology of processes mediated by dG4-unwinding helicases [55], [64], [114], [115]. Here, we focus on rG4-unwinding helicases and the biological processes they serve (Fig. 1) with an emphasis on experimental approaches that were performed to investigate rG4-helicase action.
Fig. 1.
Helicases intervene in a multitude of RNA processes through the unwinding of rG4s. Genome homeostasis is kept through the processing of telomeric rG4. The resolution of rG4 in 5′-UTRs of mRNAs is a regulatory layer in translation, whereas in 3′-UTRs rG4 processing is linked to stress response and transcripts stability. Helicases modulate the splicing pattern of transcripts with rG4s within CDSs. Also, the miRNA metabolism is subject to helicase modulation of rG4 that would otherwise outcompete the cleavage targets of DICER. At last, rG4-processing by helicase impacts the IgH class recombination.
1.1. Excurse: Methods to demonstrate G4 formation in vitro and in vivo
As described above G4 DNA and RNA structures can form within specific guanine-rich regions harboring a G4 motif. Originally this motif harbored four G-tracts with at least three guanines next to each other that were separated by variable regions (loop regions) with a length of 1–7 nucleotides (nt) (GGGN1-7GGGN1-7GGGN1-7GGG) [116], [117]. Recent studies demonstrated that dG4 as well as rG4 can also form with two guanines per G-tract as well as with loops that are longer [118], [119]. Note, the stability of the G4 structures positively correlates with the amounts of guanines per G-tract and decreases with increasing loop length [120]. Further, G4s are stabilized by cations that are centrally coordinated to the O6 of the guanines [117]. The dependency of G4 stability on salt might be of interest for studies using helicases, because also helicase action is salt dependent.
Studying G4 formation in vitro and in vivo is challenging due to different topologies and stabilities of the G4 (reviewed in [121]). G4 formation in vitro can be observed using X-ray crystallography, NMR spectroscopy, circular dichroism, ultraviolet melting and differences in electromobility behavior [122], [123]. Also different methods have been developed to map G4 in cells in vivo [124]. For example, G4 visualization using immunofluorescence can be achieved using either G4-specific antibodies (e.g. Sty49, 1H6, BG4) [125], [126], [127], [86], [128] or synthetic small molecules that recognize and stabilize G4s and are coupled to a fluorescence probe (e.g. PDS, PhenDC3, DAOTA_M2, QUMA) [129], [130], [131], [61]. Recently, also G4 quantification by flow cytometry has been demonstrated [132]. Different genome-wide sequencing approaches are established using G4-specific antibodies (BG4) [133], G4 forming peptides [134], G4-binding molecules [91], [135], [136], mapping of G4-binding proteins or by using the ability of G4s to stop polymerases [15].
1.2. RNA G4-unwinding helicases
Most RNA molecules exhibit a precisely defined three-dimensional arrangement. Still, the RNA folding landscape is jagged by many thermodynamically stable intermediates that kinetically trap long-lived misfolded species, particularly in vitro [137], [138]. The efficient RNA folding in vivo is secured by specialized proteins that guide RNAs to their native state and resolve kinetic folding traps [139], [140], [141]. Helicases represent the largest group of multifunctional, ubiquitous and pleiotropic DNA- and RNA-remodeling enzymes [142], [143]. They are classified as ATP-dependent enzymes and are sorted in six superfamilies (SF1-6) according to conserved sequence motifs and include dual-purpose DNA/RNA processing enzymes [144]. All the RNA helicases identified so far to intervene in rG4 metabolism (Fig. 1), belong to the superfamilies (SF) 1 and 2. SF1 and SF2 share two highly similar helicase domains resembling the recombination protein RecA [145] and include protein families with distinct sequence, structural and mechanistic features, such as the Pif1-like family in SF1 [146] or the DExD/H-box in SF2 [147], [148].
The SF1 helicase family is marked by the GDxxQ motif, although several variations are found in the group and shared with SF2. Based upon the primary structure the SF1 branches into three distinct families; UvrD/Rep-like, the Upf1-like and Pif1-like enzymes [149]. UvrD-like enzymes are predominantly SF1Aα 3′-5′ DNA helicases, whereas Upf1-like and Pif1 helicases are SF1Bα enzymes (with 5′-3′ processivity) and include examples of proteins that process both DNA and RNA. Slight variations are observed among the helicase motifs throughout the three families presumably correlating with their different motor properties. The founding member of the family, S. cerevisiae Pif1 (P07271), is a ubiquitous eukaryotic protein fulfilling an assortment of tasks in DNA maintenance including dG4 unwinding [98], telomerase inhibition at DNA breaks and Okazaki fragment maturation [150]. To date, the human helicase MOV10 and its testis paralog MOV10L are the sole SF1 members identified to process rG4s.
SF2 is the largest group and comprises most G4-processing helicases identified thus far. Its members share high structural similarities in the catalytic core and intervene in every aspect of RNA [142] and DNA metabolism [145], [149]. The catalytic core is characterized by a flexible linker that allows the correct arrangement of two RecA-like domains reported to be required for sequence-unspecific ssRNA binding and ATP loading [145]. Both RecA-like domains unspecifically contact ssRNAs at the phosphate backbone, whereas auxiliary flanking domains provide specific functions or interaction and recruitment of partner proteins [151], [152], [153]. A remarkable exception is the eukaryotic Initiation Factor 4A (eIF4A) that represents the archetype of a minimal core helicase [154]. The most prominent SF2 are the DEAD-box (from the Asp-Glu-Ala-Asp motif, also referred to as DExD-box due to motif variations, alias DDX with 40 human members) and the DEAH-box family (alias DHX, 15 human members) [142]. Only a limited set of DExD/H helicases has directional processivity with most exhibiting local, chaperone-like RNA unwinding. They are still subject to substrate secondary structure features or availability of flanking overhangs [155], [156].
1.2.1. DHX36
The ATP-dependent 3′-5′ DEAH-box helicase DHX36 (Q9H2U1), also known as RHAU (RNA helicase associated with AU-rich element) due to its regulation of mRNA stability through the binding of AU-rich elements (AREs) [157], is a ∼ 114 kDa SF2 helicase. It has two isoforms, which at least in HEK293T cells operate in different cellular compartments, the nucleus (isoform 1) or the cytoplasm (isoform 2) [158]. DHX36 is a multi-functional helicase that can act on different substrates [157], [159], [160] by efficiently recognizing and processing both dG4s (Kd ∼77 pM) and rG4s (Kd ∼39 pM) [161], [162], [163]. DHX36 function is linked to human health e.g. heart development [164], hematopoiesis [165] and spermatogonia differentiation [166]. Among other functions, at the DNA level DHX36 intervenes to resolve dG4s in gene promoters and thereby activating transcription [159], [161], [167], [168], [169], [170]. The detailed discussion of DHX36 function within cells is reviewed elsewhere [171]. Like other DExD/H members [172], [173], DHX36 bears two tandem RecA-like domains making up the core helicase and ATPase functions, a winged-helix DNA/RNA-binding domain, a “ratchet” like domain (RL) and the DExD/H-specific β-barrel oligonucleotide- oligosaccharide-binding (OB) domain [174] (Fig. 2). The N-terminal, DHX36 specific motif (DSM) confers the high affinity for G4s [175], [176] by contacting the outer G4-tetrad (Fig. 2) through hydrophobic interactions due to its amphipathic α-helix, as seen by solution NMR (PDB 2N21) [177], X-ray crystallography (PDB 6Q6R) [178] and molecular dynamics [179]. The crystallographic structures of the highly-conserved orthologues from B. taurus in complex with the dG4 from the MYC promoter (Q05B79 - PDB 5VHE) [180] and D. melanogaster with ssRNAs (Q8SWT2 - PDB 5N98) [181] elucidate the DSM-dG4 recognition in the context of the full-length protein and the role of the helicase core, the OB and ratchet domains in creating a positively-charged cage to complement the interaction, whereas the helicase core binds to 3’single-stranded substrates longer than 5 nt. Single-molecule fluorescence resonance energy transfer (FRET) experiments [180], [182] support the current model of a partial, ATP-independent G4 destabilization by DHX36 consequent to the binding, followed by the complete G4 resolution by an ATP-driven, translocation-based mechanism threading the single-stranded nucleotides through the helicase core channel [183], similar to double-stranded substrate processing seen in other DExD/H helicases [184], [185], [186]. The functions of the DHX36 auxiliary domains in processing G4s were further clarified for the M. musculus orthologue (Q8VHK9 – PDB 6UP4) [187]. Pre-steady-state reaction RNA remodeling experiments, i.e. under conditions allowing repeated cycles of enzyme-substrate interaction and easier readout as compared to steady-state reactions [188], allowed to assess the stringent ATP-dependence of mDHX36 for the complete remodeling of both dsRNA and rG4 in the presence of a complementary trap sequence preventing the re-annealing [187]. Substrate affinity and ATPase assays clarified the essential function of the OB domain in coupling the ATPase moiety to nucleic acid binding and thus for the ATP-driven remodeling of quadruplex and duplex structures alike. mDHX36ΔOB mutants in fact exhibit a reduced binding affinity for all substrates and basal ATPase activity comparable to mDHX36. Unlike the wild-type, the mDHX36ΔOB ATPase activity was not stimulated in the presence of the substrate pointing to the loss of coupling function from the OB domain. Similarly, the conserved N-terminal β-barrel domain interacting with the OB domain proved to be essential for contacting and remodeling both the duplex and the G-quadruplex [187].
Fig. 2.
Structure and domains of Bos taurus Dhx36 in complex with the parallel quadruplex from the MYC promoter (PDB 5VHE). The DSM domain (dark red) is primarily responsible for the G4 (orange) recognition by contacting the exposed G4-tetrad through its hydrophobic surface. The interaction is complemented by the OB domain (purple) and RecA1 (light blue) engaging the G4 backbone by creating a positively-charge “cage”, which also contributes to the proposed unwinding mechanism. The N-terminal linker (NL) is depicted in white. Rotations refer to the upper-left model. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
The rG4 biology of DHX36 embraces the whole of post-transcriptional processing, including miRNA and pre-mRNA maturation, RNA degradation, telomeres maintenance and translational regulation. DHX36 intervenes in the rG4-mediated translational regulation through diverse mechanisms, yet not completely rationalized. There appears to be a twofold regulation fate for mRNAs based on the localization of rG4s and DHX36-binding at either the 5′- or 3′-UTR. A system-wide, nucleotide-resolution photoactivatable ribonucleoside enhanced crosslinking and immunoprecipitation (PAR-CLIP) [189] analysis in human cells mapped the preferential binding of DHX36 to rG4-prone mRNA sequences [158]. RNA-seq showed that a CRISP/Cas9 DHX36-KO resulted in a significant increase of target mRNAs, yet ribosome foot-printing and stable isotope labeling in cells (SILAC) coupled with mass spectrometry proved the concomitant reduction in ribosome occupancy and protein translation, respectively. The KO also induced the formation of stress granules and protein kinase R (PKR/EIF2AK2) phosphorylation which, along with the finding that mRNAs targets of DHX36 that harbor rG4s preferentially localize in stress granules, suggests that DHX36 functions in resolving stress responses to re-establish cellular homeostasis [158]. This regulatory dichotomy can be explained in the light of the regulation of NKX2-5 by DHX36 in M. musculus [164]. DHX36 -/- cells had significantly higher mRNA levels as measured by microarray and RT-qPCR, yet reduced protein at western blot. RNA pull-down showed a strong DHX36 association at 3′-ARE (AU-rich Elements), which was abolished by using mutant with depleted 3′-ARE, proving that the binding of DHX36 to ARE in the 3′-UTR labels the mRNA for degradation. This is supported by a similar experimental design that proved the negative regulation operated by DHX36 on the translation of the transcription factor PITX1 through processing 3′-UTR rG4s in its mRNA [190]. Conversely, either mutating the NKX2 rG4 upstream of a GFP expression vector or restoring DHX36 expression significantly increased the protein levels, proving the role of DHX36 in regulating translation by resolving the 5′-UTR rG4s [164].
A label-free, liquid chromatography-tandem mass spectrometry (LC-ESI-MS/MS) global proteomic analysis upon DHX36 RNAi silencing identified a small set of significantly downregulated proteins enriched in 5′-UTR rG4s [191]. BCL2 and NRAS bear well-characterized examples of 5-UTR rG4-mediated translation regulation. Nonetheless, neither BCL2 [192] nor NRAS [193] are altered upon DHX36 silencing [191], perhaps due to functional redundancy of other helicases, e.g. of the paralog DHX9 [194]. The two paralogs indeed appear to regulate the translation efficiency, measured by transcriptome-wide ribosomal profiling, on a common subset of transcripts in HeLa cells [194]. Noteworthily, the multifunctional ATP-dependent RNA helicase DDX3X (O00571), whose rG4-unwinding activity still lacks evidence, was identified to interact with the NRAS 5′-UTR rG4 along with two known rG4 helicases, DDX5 and DDX17 [101].
The rG4 biology of DHX36 spans over the modulation of genome stability by regulating telomere maintenance and the expression of the gatekeeper p53. Telomeres, the terminal structures protecting most eukaryotic chromosomes from degradation, end-to-end joining or recognition as double-strand breaks (DSBs) [195], [196], are maintained by telomerase, a reverse transcriptase harboring an internal RNA template for the synthesis of telomeric repeats. The template RNA 5′-end controls the transcription termination through a stem-loop structure (P1) [197], [198] which, due to its high guanine content, is challenged by an rG4 shown to disrupt the P1 helix [199], [200]. DHX36 directly binds this rG4 as demonstrated by independent non-crosslink co-immunoprecipitation/RT-qPCR [201], [202] and non-crosslink RNA immunoprecipitation microarray analyses [203]. Knockdown (KD) of DHX36 by siRNAs leads to decreased telomerase efficiency and shorter telomeres [201], as assessed by monochrome multiplex quantitative PCR [204].
DHX36 also indirectly controls genomic stability by processing a 3′-rG4 in the pre-mRNA of p53 [205], the “guardian of genome stability” [206], through a seemingly different mechanism. The direct interaction between DHX36 and the p53 rG4, proved by RNA immunoprecipitation, is necessary to upkeep the p53 pre-mRNA 3′-end processing upon UV-induced damage. The pre-mRNA 3’-end processing efficiency, as assessed by the ratio between unprocessed RNA and total nuclear RNA by RT-qPCR, was significantly diminished upon DHX36 siRNA KD, resulting in reduced p53 response to the UV-induced damage. Similar effects were obtained using G4-stabilizing ligands [207].
The DHX36-mediated resolution of rG4s is also required for correct miRNA maturation, a stream of coordinated processes pervading many cellular functions [208]. Interestingly, potential rG4s map to miRNA regions [187] and several rG4s are proved to compete with the hairpin structures necessary for pre-miRNA cleavage by Dicer [209], [210], [211], thus prospecting vast implications of rG4s for miRNA-controlled pathways. DHX36 was recently proved to intervene with the maturation of miR26-a [212], a miRNA involved in multiple processes and regarded as a tumor suppressor that is found downregulated in cancers [213], [214], [215]. The miR26-a precursor harbors a stable rG4, as seen by CD and EMSA, which impairs the post-transcriptional maturation with detrimental effects both in vitro and in vivo [212]. MiR26-a expression was higher in cells transfected with a plasmid bearing a mutant, not rG4-forming sequence than with the wild-type sequence, which was also sensitive to the G4-stabilizer PDS. Also, DHX36 overexpression significantly enhanced miR26-a levels only in the wild-type, as measured by RT-qPCR [212].
1.2.2. DHX9 (DDX9)
The 3′-5′ ATP-dependent DEAH-box helicase DHX9 (Q08211), also known as nuclear DNA helicase II (NDH II) or RNA helicase A (RHA), is a paralog to DHX36 and also a member of SF2. DHX9 bears an RGG-box but lacks the N-terminal G4-binding domain present in DHX36 and intervenes in several cellular processes, both at DNA and RNA level. These include replication [216], transcription [217], [218], translation [219] and maintenance of genomic stability [220]. The complex biology of DHX9 was recently reviewed [221].
Helicase and competition assays proved DHX9 to bind and resolve G4s with stronger kinetics toward rG4 and strict ATP-dependence in vitro, as confirmed by the complete loss of G4-unwinding function upon ATPase inhibition by N-ethylmaleimide poisoning [222]. More recently, DHX9 was confirmed to directly intervene in rG4 biology in vivo as a translational modulator of mRNAs harboring rG4s in their 5′-UTR [194]. Polysome fractioning coupled with mass spectrometry demonstrated that DHX9 is bound to fully-assembled polysome, thus actively translating mRNAs. A further indication of an rG4-specific translational modulation role was provided by cell proliferation and eIF2 phosphorylation (a marker of the global translation efficiency) assays upon DHX9 siRNA KD: neither DHX9-KD nor DHX36-KO had a relevant impact on the overall translation, advocating for functional redundancy. However, the translational efficiency of a subset of mRNA with 5′-UTR rG4s was significantly impacted. Individual-nucleotide resolution UV crosslinking and immunoprecipitation (iCLIP) [223] followed by deep sequencing experiments proved the correlation with 5′-UTR rG4s sequences. DHX9 peaks and rG4s centered ∼ 40 nt apart on average possibly reflecting the helicase loading downstream of the rG4 in a pre-processing stage [194] in agreement with the reported need for a 3′single-stranded tail for the DHX9 helicase function [224]. The same work confirms the DHX9-dependent (and DHX36 alike) translational modulation by GFP-reporter assay [225].
1.2.3. DDX5/DDX17/Dbp2 subfamily
The human DEAD-box paralogs DDX5 (P17844) and DDX17 (Q92841) and the S. cerevisiae orthologue Dbp2 (P24783) [226] belong to the SF2 and share high sequence similarity among each other [227]. The two human paralogs mostly differ in the N-terminal domain likely responsible for yet uncharacterized specialized functions. The C-terminal proline-rich element likely reported to protein–protein interactions [228] distinguishes both from non-mammalian orthologs. All three helicases are predominantly nuclear [229], [230]. There is evidence that DDX5 is a shuttling protein [231] whose localization varies in different cell lines [232] and as a function of the cell cycle [233], [234], [235]. This possibly reflects the changing G4 landscape throughout the cell cycle. DDX5 lacks sequence stringency in vitro with conflicting evidence regarding the helicase directionality in processing dsRNA [126], [134], [135], [136]. Accordingly, the DDX5/Dbp2 subfamily functions in virtually every step of the post‐transcriptional gene regulation [236], [237], [238], [239], [240], [241] and correlates with cancer onset and progression [242], [243], [244], [245], [246], HIV [247], RVFV (Rift Valley Fever Virus) [248] and general innate antiviral response [249]. KD experiments also prove the functional redundancy of the human paralogs [250]. The vast non-G4 biology of the DDX5/Dbp2 subfamily is detailed in a recent review [227].
The first indirect evidence for the involvement of DDX5 and DDX17 in rG4 biology came from investigating the co-dependent functionality with hnRNP H/F, a splicing factor previously reported to bind G4-prone sequences [251] and recently shown to intervene in rG4-mediated translational regulation [252]. The silencing of either helicase by siRNA significantly impacts splicing patterns prompting the authors to postulate that DDX5/17 could regulate mRNA splicing by binding of hnRNP H/F to G-rich target motifs otherwise embedded within folded G4s. Further, the transient expression of either DDX5/17, but not helicase-impaired mutants, can rescue the phenotype, bolstering the functional redundancy and backing the central role of the helicase function. This model is supported by the observation that the treatment with the G4-stabilizer TMPyP4 [253] had the same effects on splicing as seen with siRNA silencing. It has to be noted that the ligand is unspecific and the effects cannot be univocally ascribed to the stabilization or rG4s. Proximity ligation assays [254] confirmed the close interaction of the two helicases with hnRNP H/F in vivo. However, these were not definitive evidence that the function was actually due to the unwinding of rG4.
Direct observation of the DDX5 and DDX17 interaction with rG4s has been obtained using an rG4-bait pulldown coupled with LC-mass spectrometry which, among other proteins, also identified hnRNP H/F [101]. The same experimental approach in S. cerevisiae [255] identified the homologous Dbp2 and three other DEAD-box RNA helicases, Ded1 (P06634) [256], Dbp1 (P24784) [257] and Mss116 (P15424) [258]. Surprisingly, fluorescence anisotropy and experiments using a trap single-strand complementary sequence sequestering the rG4-forming sequence upon unfolding [259] showed that Ddp2 destabilizes rG4s regardless of ATP antipodal to previous findings with dsRNA [226] and jarring with in vivo findings using the ATPase-deficient mutants [251] whilst confirming a threefold binding affinity enhancement by 3′-overhangs.
Finally, Wu and colleagues demonstrated DDX5-binding and unwinding activity on both dG4 and rG4 using biotin-dG4s immobilized on streptavidin-coated plates and ELISA readout [97]. They also provided proof of DDX5-dG4-mediated modulation of the oncogene MYC transcription in vivo. The unwinding activity was tracked by FRET using a dG4 from the MYC promoter labeled with 6-fluorescein (6-FAM) at the 3′-end and black hole-1 quencher (BHQ-1) at the 5′-end. FRET with either shorter or no flanking regions returned similar results indicating that DDX5 does not need 3′-overhangs for its activity. Competition assays on the other hand support the influence of siding regions on binding affinity, at least with dG4s, because the dG4 “MYCpu28” binding appears to be an order of magnitude higher. Surprisingly, both K144N (unable to bind ATP) and D248N (unable to hydrolyze ATP) mutants [260] were able to destabilize dG4s, as also confirmed in glucose/hexokinase ATP-free experiments (ATP-dependent phosphorylation to 6-p-glucose) [97]. Nonetheless, whether ATP is required to resolve the more stable rG4s remains unclear, because only data from dG4 was reported. Whilst not producing effects on the unwinding activity, ATP significantly reduced the binding of DDX5 to immobilized dG4, suggesting it might be required for unloading the unfolded oligonucleotide from the protein, as also backed by DMS protection assays [97].
1.2.4. DDX21
The DEAD-box RNA helicase DDX21 (Q9NR30), alias nucleolar RNA helicase 2, was initially proposed to possess a 5′-3′ ATP-dependent core helicase domain [261] and an ATP-independent C-terminal “foldase” function [262], [263], [264], i.e. able to remodel secondary structure in a seemingly single-stranded RNA. However, a recent FRET continuous helicase assay [265] proved its dsRNA helicase activity to be stimulated by, but not to depend on, ATP [266]. DDX21 intervenes in fundamental aspects of RNA metabolism. It is essential for ribosomal RNA biogenesis [267], [268], [269] in tandem with RNA Pol I-mediated transcription [270] and for pre-ribosomal complexes assembly [271]. Functions of DDX21 have also been reported in the immune response by double-stranded RNA sensing [272], regulation of viral RNA synthesis [273], HIV replication [274], HCMV (human β-herpesvirus 5) replication [275] and activation of the innate antiviral response [276], [277]. DDX21 acts as an epigenetic modificatory of gene expression [278] and correlates with cancer [279]. DDX21 is abnormally expressed in colorectal [280], [281], [282] and breast cancers [283]. It intervenes in R-loops regulation [275], [284], [285] and acts as a sensor and mediator of transcription during the nucleotide stress response [286].
DDX21 was identified to interact with rG4s by a streptavidin pull-down coupled with LC-mass spectrometry from HEK293T cellular lysates using the rG4 from the PITX1 3′-UTR as a bait [287]. Transient expression and pull-down of different DDX21 truncations indicated that the C-terminal DDX21(574–783) domain is necessary and sufficient for binding rG4s. However, the binding affinities determined by microscale thermophoresis (MST) [288], [289] demonstrated that the recognition of rG4s also involves other domains of the protein, as seen with DHX36. MST [289] and isothermal titration calorimetry [290] confirmed the direct interaction of DDX21(574–783) with the human telomeric rG4 TERRA [291] in the nanomolar range with a remarkable higher affinity compared to dG4 [292]. The bulk interaction is driven by electrostatic interactions as observed by a reduction in affinity at higher ionic strength. Bidimensional saturation transfer difference NMR allowed mapping the contact with the 2′-OH of loop nucleotides, which might explain the rG4 selectivity [292].
CD measurements and thioflavin T-based assays proved the G-rich RNA sequence used to define the foldase activity [261], [262], [263], [264] to fold into an rG4 [287] and led to question whether such observations could rather be the result of rG4 unwinding activity, rather than a dsRNA re-modeling. In facts, digestion with either RNase A or T1, enzymes that cut at specific single-stranded regions (T1 cuts after guanines, RNase A after cytosines and uracils [293]) showed significant differences only for the T1 cutting pattern in the presence of DDX21, indicating a change in the accessibility of guanine residues but not in cytosine or uracil as a result of DDX21 rG4 unwinding [287]. Altogether, DDX21 appears to bind and process rG4 primarily through its conserved (F/PRGQR) C-terminal repeat, yet, with the determinant support from other domains [292]. This was recently confirmed by RNase T1-based helicase assays using different DDX21 mutants [294].
Gene reporter assays upon transient DDX21 silencing by siRNA provided in vivo evidence of DDX21 to suppress genes with 3′-UTR rG4 as a consequence of its rG4-unwinding activity, regardless of ATP [287]. DDX21 was also shown to regulate the translation of MAGED2 (Q9UNF1), a transcriptional repressor of tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) death receptor 2 (TRAIL-R2 - O14763) mRNA [295], [296] through the processing of its 5′-UTR rG4 [297]. MAGED2 was identified as a candidate DDX21-regulated protein by differential proteomic on HEK293T cells expressing either the wild-type or a helicase impaired DDX21 mutant. This latter was unable to express MAGED2 at a luciferase assay [297].
1.2.5. DDX2 (eIF4)
The ATP-dependent DEAD-box DDX2 helicase family [298], commonly known as the eukaryotic initiation factor 4 (eIF4), comprises three paralogs: DDX2A (P60842) and DDX2B (Q14240) (eIF4A1 and A2, respectively) are highly similar cytosolic helicases that as part of the eIF4F complex exhibit the critical ATP-dependent [299], [300], [301] unwinding of 5′-UTR secondary structures for the recruitment of the 40S complex in cap-dependent initiation of translation [302], [303], [304], [305]. The two are largely redundant and will be interchangeably referred to as DDX2, although there is emerging evidence of different functions in vivo [306], [307]. A specific role of DDX2B in esophageal squamous cell carcinoma was recently reported [308]. DDX2 upregulation drives translation reprogramming in pancreatic [309], [310], breast [311], [312] and other cancers [302], [313], [314], [315]. It is sequestered in trophoblasts to re-modulate the global protein production through a recently discovered mechanism [316]. DDX2 is also hijacked to support viral infections through different mechanisms [317]. The third paralog, DDX48 (P38919 - eIF4A3), is localized in the nucleus and is part of the exon junction complex [318], [319], [320].
Although it is evident that DDX2 is necessary for modulating the translational efficiency of mRNAs by processing secondary structures in their UTRs the extent this can be ascribed to G-quadruplexes is subject to conflicting findings. The first evidence for the intervention of DDX2 in rG4-mediated processes was reported by Wolfe et al. [321]. Ribosomal foot-printing and deep sequencing with and without the specific DDX2 inhibitor silvesterol showed reduced translational efficiency for a set of transcripts with enriched putative rG4-forming motifs in their 5′-UTR. Notably, many were transcription factors and oncogenes, in agreement with other findings [312], [322]. In firefly luciferase assays the translational efficiency of constructs with 5′-UTR rG4s was significantly reduced over scrambled-sequence controls, in the presence of silvesterol. This could be reversed by overexpressing DDX2 or using the silvesterol-insensitive P159Q DDX2 mutant [323] leading to conclude that DDX2 helicase activity is a limiting factor for the translation of mRNAs harboring 5′-UTR rG4s, effectively adding up an extra regulatory layer to such set of transcripts.
This generalized model is, however, questioned by later reports from Waldorn and co-workers [324], [325]. Reverse transcriptase stop and in vitro translation assays reported only three-layered rG4s (rG4(3)) to form in full transcripts and to exhibit sensitivity to the DDX2 inhibitor hippuristanol [324], [326]. Two-layered rG4s (rG4(2)) could not form in full transcripts possibly due to competing hairpin structures. Only under G4-stabilizing conditions, i.e., in the presence of the G4-stabilizing ligand PDS, rG4(2) could reduce the translation efficiency. This stands in contrast to finding of Wolfe et al. [321]. Experiments with constructs bearing 7-deazaguanine (c7G), which impairs Hoogsteen but not Watson-Crick pairing [327], confirmed a greater translational gain for c7rG4(3) over the controls. However, c7rG4s(3) were still significantly sensitive to hippuristanol, suggesting the intervention of alternative, non-rG4 secondary structures. Two constructs bearing either a GGAx or GGCx motif (the former likely forming rG4s but less enriched in the foot-printing analysis, the latter able to form stable non-G4 secondary structures) were compared with and without either PDS or hippuristanol. PDS had a significant effect only on the rG4(2) sensitivity to hippuristanol suggesting that the DDX2 translational regulation is more reliant on canonical hairpin structures than on rG4s, notwithstanding translational repression operated by rG4(3) [324].
A combination of polysome profiling, the DMS-based Structure-seq2 method [328] and DDX2 inhibition by hippuristanol was later used to investigate the structural composition of translationally DDX2-dependent 5′-UTRs [325]. A set of DDX2-dependent transcripts with longer than average 5′-UTRs was enriched in Cs, but not Gs. Computational prediction of rG4 using G4RNA [329] did not show any significant prevalence compared to a set of DDX2-independent mRNAs. Similarly, DMS Structure-seq2 under DDX2-inhibition did not point out any significant difference between the two sets of transcripts, further advocating that the enrichment in (GGC)4 motif in the DDX2-dependent set is due to canonical secondary structures outcompeting the more common rG4s(2) [325].
However, indirect evidence for the rG4 role comes from the oligonucleotide inhibitors of the 48S-PIC complex ES6S region, which constitutes the extended binding channel for the DDX2-mediated mRNA unwinding and scanning [330]. The ES6S region consists of four RNA helices at the bottom of the 40S ribosomal subunit. Polysome mapping upon ES6S inhibition showed significant enrichment in rG4-forming sequences in the set of transcripts with reduced translational efficiency. To note, the authors used the rG4-search algorithm QuadBase2 [331]. The translational efficiency dropped consistently with the 5′-UTR rG4s stability by comparison with constructs bearing more stable rG4 motifs from BCL2. Furthermore, ribosome profiling after treatment with the novel DDX2 inhibitor eFT226 also yielded enriched rG4-forming sequences (the applied search criteria are unclear) [332]. The translational dependence on the original 5′-UTR was proved by comparison with constructs unable to form rG4s. Neither work offers definite evidence for the degree to which the DDX2-dependent translational control relies on rG4s or outcompeting canonical secondary structures, because there is no direct proof for the rG4 folding. The KRAS-controlled, DDX2-mediated translation of the 5′-UTR rG4-containing ARF6 led to similar conclusions [333].
1.2.6. DDX1
The DEAD-box RNA helicase DDX1 (Q92499) has three human splicing isoforms that localize both in the nucleus and the cytoplasm and exhibit a 5′single-stranded RNase activity and participate in several processes, including translation activation [334], splicing regulation [335], fatty-acid dependent insulin regulation [336], double-strand break repair [337] and microRNA maturation [338]. Loss of DDX1 results in embryonic lethality [339] and severe gametogenesis defects [340]. It also has a multifaced role in either inhibiting or facilitating viral infections, including HIV [341], tumorigenesis [342] and cancer progression [343].
DDX1 was shown to bind and process lncRNA rG4s, facilitating R-loop formation and thus assisting in the recruitment of the activation-induced cytidine deaminase (Q9GZX7) to initiate IgH class recombination [344]. DDX1 was pulled down using the S-region containing rG4-forming sequences as a bait, but not with scrambled controls or under Li+ rG4-destabilizing conditions. DDX1 was also extracted from CRISPR/Cas9 AID-KO cells demonstrating that its interaction is AID-independent, as verified by native EMSA assays. Furthermore, DDX1-K52A mutants with impaired ATP-binding capability tended to form a more stable complex but it is uncertain whether this is due to decreased ATP-dependent unwinding or decreased ATP-dependent substrate turnover, i.e., substrate release. Interestingly, the interaction is RNA-specific as no interaction was detected with dG4s. Mouse lymphoma CH12 cells (CVCL_6818) [345] transiently transfected with DDX1-K52A showed reduced class switch recombination and R-loops. In vitro assay with 32P-rG4 and complementary trap DNA strands proved DDX1 to directly convert rG4s into hybrid RNA:DNA R-loops. This depends on ATP because DDX1-K52A was significantly less efficient in the conversion. To date, this is the only direct example of the intervention of DDX1 in rG4 biology [344].
1.2.7. MOV10/L
The mammalian ATP-dependent 5′-3′ SF1 RNA-helicases MOV10 (Q9HCE1) unspecifically loads on single-stranded RNA proximal to the stop codon and translocates along 3′-UTRs intervening in nonsense-mediated decay [346] and suppression of retrotransposition [347], [348]. Its testis-specific paralog MOV10L1 (Q9BXT6) is critical for the control of PIWI-interacting RNA (piRNA) biogenesis. piRNAs are important for retrotransposon silencing and protect the genome integrity of post-meiotic germ cells [349]. In fact, impaired ATP-binding and ATP-hydrolysis motifs cause aberrant piRNA biogenesis and male sterility [350], [351]. Both paralogs preferentially bind G-rich sequences, as proved by CLIP-seq experiments [351], [352].
In vitro helicase assays using complementary nucleotide trap sequences (to prevent rG4 refolding) showed both paralogs of MOV10 unwind rG4 depending on ATP. Neither helicase could process dG4s. The ATP dependence was confirmed using the non-functional K778A and DE888AA mutants and the non-hydrolyzable ATP-analogues AMP-PNP or ATPγS [353]. MOV10L, which requires the N- and C-terminal domains to unwind rG4s also proved higher unwinding efficiency than MOV10 [353]. The MOV10 N-terminal domain is necessary and sufficient for the rG4-unwinding function, as seen by mutagenesis and luciferase assays [354].
MOV10 was showed to intervene in the regulation of miRNAs by exposing the 3′-UTR miRNA recognition elements (MRE) to FMRP (Q06787), the fragile X mental retardation RNA binding protein, possibly through unwinding rG4s [354].
1.3. CNBP, a non-helicase rG4-resolving protein
Helicases are defined as motor proteins that move directionally along the phosphodiester backbone resolving secondary or tertiary nucleic acid structures using energy from ATP hydrolysis [355]. Although largely responsible for the G4 processing in the cell such classification is increasingly failing to embrace proteins still able to untie and impact the biology of G4s.
One example is the cellular nucleic acid binding protein (CNBP – P62633), a conserved eukaryotic zinc finger protein featuring seven CCHC-type ZnF domains and an arginine-glycine (RG/RGG)-rich motif. CNBP is ubiquitous and highly expressed in adult tissues, essential for embryonal development in mice and was reported to relate with myotonic dystrophy type 2 [356]. It binds to both G-rich single-stranded DNA [357] and RNA [358] in vitro and is demonstrated to promote the translation of specific mRNAs by associating to their 5′ oligopyrimidine sequences [359], [360]. PAR-CLIP analysis [189] mapped CNBP onto mRNA G-rich regions proximal to start codons previously reported to form rG4s or other secondary structures [59]. The translational regulatory role of CNBP could be defined by ruling out a significant effect on transcription given only minor, yet statistically significant, changes in target mRNA abundance in CRISPR/Cas9 CNBP-KO cells, as determined by transcriptome-wide RNA-seq. Noteworthy, this sufficed in reducing long-term cell viability (15 days), suggesting a specialized function of CNBP in supporting cell survival. Conversely, Ribo-seq and mass spectrometry highlighted a sizeable reduction of CNBP targets at the protein level pointing out the regulatory role of CNBP onto the translation of such mRNAs by de-structuring rG4s within the CDS, as supported by in vitro SHAPE and CD experiments [95] and consistent with previous findings [59]. Substantial biophysical and biochemical evidence for the ATP-independent CNBP unfolding of G4s, although limited to dG4s, was independently gathered by a combination of CD, NMR, polymerase stop assay, EMSA and unwinding PAGE assays [19]. Accordingly, CNBP might disrupt the G4 folding-unfolding equilibrium by preferentially binding to the unfolded species.
2. Concluding remarks
G-quadruplexes, which are somehow still considered non-canonical, are becoming more accepted as a regulating tool within cells that can alter gene expression and/or maintenance of genome stability. Accumulating evidence about their complex biology moved G4s from being considered “mere” targets for small molecules in the anticancer perspective toward a deeper understanding of their functions in a variety of cellular processes, most often at the DNA level. The knowledge about the RNA counterpart has long lagged owing to peculiar experimental difficulties in assessing their existence in vivo because of the inherent ephemeral RNA nature and has eventually been called into question. The major claim against rG4s is now clarified as resulting from an inherent weakness in the DMS/RT-stop profiling assay, which is better suited to determining G4 equilibrium shifts rather than capturing snapshots of folded rG4s [60]. Besides, rG4s are also supported by interesting biological evidence gathered by independent novel approaches, all of which point to a complex protein-regulated rG4 scenario with far-reaching implications in living cells. In vivo, rG4s can now be thought of as fast-paced, transient events shaping a swinging rG4 environment determined by exogenous or endogenous stimuli and regulated by the cellular regulatory machinery, primarily helicases. Anything from splicing, mRNA stability, miRNA and piRNA maturation, telomeres maintenance, IgH class recombination and transcriptional regulation is now clarified to be controlled by the helicase intervention on rG4s. However, despite recent advances in the structural and biochemical understanding of helicase-mediated rG4s processing a consistent and conclusive mechanistic model appears elusive.
The rG4s biology increasingly appears to reach far beyond individual cellular processes. It rather embraces systemic aspects of the cellular functioning with specific rG4s subset coordinated through the cell cycle, possibly idiosyncratic across cell types and tissues. However, the knowledge of the rG4 biology and more so its helicase/non-helicase regulation is limited to a fairly small set of cell types. Besides, along with several examples in the context of dG4s the paradigm of helicases as the sole source of negative G4 control in the cell is further called into question by the identification of CNBP as the only example (as yet) of non-helicase rG4-destabilizer. It is then utterly relevant to clarify the regulatory network beyond helicases with likely many more direct or indirect rG4 modulators to be identified. Indeed, the network of processes controlled by rG4 appears far too extended and sensitive to the cell to be sufficiently fine regulated by the relatively small number of proteins identified so far. These considerations lead to many more pending questions: how are the regulators regulated themselves? What are the cellular signals and intermediate modulators driving the helicase-mediated and non-helicase response to rG4 after a given stimulus? At last, it is worth noting the knowledge of the rG4 biology and more so its helicase/non-helicase regulation is limited to a fairly small set of mostly transformed cell lines.
CRediT authorship contribution statement
Marco Caterino: Data curation, Investigation, Writing – original draft. Katrin Paeschke: Supervision, Funding acquisition, Conceptualization, Writing – original draft, Writing – review & editing.
Declaration of Competing Interest
The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
Acknowledgments
Acknowledgement
We thank Daniel Hilbig and Stefan Juranek for careful reading of the manuscript.
Research in the Paeschke laboratory is fund by an ERC Stg Grant (638988-G4DSB) and by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) – Project-ID 369799452 – TRR237” as well as by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy – EXC2151 – 390873048.
References
- 1.Lightfoot H.L., Hagen T., Tatum N.J., Hall J. The diverse structural landscape of quadruplexes. FEBS Lett. 2019;593:2083–2102. doi: 10.1002/1873-3468.13547. [DOI] [PubMed] [Google Scholar]
- 2.Takahashi S., Sugimoto N. Stability prediction of canonical and non-canonical structures of nucleic acids in various molecular environments and cells. Chem. Soc. Rev. 2020;49:8439–8468. doi: 10.1039/D0CS00594K. [DOI] [PubMed] [Google Scholar]
- 3.Bhattacharyya D., Mirihana Arachchilage G., Basu S. Metal Cations in G-Quadruplex Folding and Stability. Front. Chem. 2016;4:38. doi: 10.3389/fchem.2016.00038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Webba da Silva M. Geometric formalism for DNA quadruplex folding. Chem. Weinh. Bergstr. Ger. 2007;13:9738–9745. doi: 10.1002/chem.200701255. [DOI] [PubMed] [Google Scholar]
- 5.Joachimi A., Benz A., Hartig J.S. A comparison of DNA and RNA quadruplex structures and stabilities. Bioorg. Med. Chem. 2009;17:6811–6815. doi: 10.1016/j.bmc.2009.08.043. [DOI] [PubMed] [Google Scholar]
- 6.Arora A., Maiti S. Differential biophysical behavior of human telomeric RNA and DNA quadruplex. J. Phys. Chem. B. 2009;113:10515–10520. doi: 10.1021/jp810638n. [DOI] [PubMed] [Google Scholar]
- 7.Agarwal T., Jayaraj G., Pandey S.P., Agarwala P., Maiti S. RNA G-quadruplexes: G-quadruplexes with “U” turns. Curr. Pharm. Des. 2012;18:2102–2111. doi: 10.2174/138161212799958468. [DOI] [PubMed] [Google Scholar]
- 8.Malgowska M., Czajczynska K., Gudanis D., Tworak A., Gdaniec Z. Overview of the RNA G-quadruplex structures. Acta Biochim. Pol. 2016;63:609–621. doi: 10.18388/abp.2016_1335. [DOI] [PubMed] [Google Scholar]
- 9.Tang C.-F., Shafer R.H. Engineering the Quadruplex Fold: Nucleoside Conformation Determines Both Folding Topology and Molecularity in Guanine Quadruplexes. J. Am. Chem. Soc. 2006;128:5966–5973. doi: 10.1021/ja0603958. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Banco M., Ferre-D’Amare A. The emerging structural complexity of G-quadruplex RNAs. RNA N. Y. N. 2021 doi: 10.1261/rna.078238.120. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Xiao C.-D., Ishizuka T., Xu Y. Antiparallel RNA G-quadruplex Formed by Human Telomere RNA Containing 8-Bromoguanosine. Sci. Rep. 2017;7:6695. doi: 10.1038/s41598-017-07050-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Xiao C.-D., Shibata T., Yamamoto Y., Xu Y. An intramolecular antiparallel G-quadruplex formed by human telomere RNA. Chem. Commun. 2018;54:3944–3946. doi: 10.1039/C8CC01427B. [DOI] [PubMed] [Google Scholar]
- 13.Bao H.-L., Liu H., Xu Y. Hybrid-type and two-tetrad antiparallel telomere DNA G-quadruplex structures in living human cells. Nucleic Acids Res. 2019;47:4940–4947. doi: 10.1093/nar/gkz276. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Marsico G., Chambers V.S., Sahakyan A.B., McCauley P., Boutell J.M., Antonio M.D., Balasubramanian S. Whole genome experimental maps of DNA G-quadruplexes in multiple species. Nucleic Acids Res. 2019;47:3862–3874. doi: 10.1093/nar/gkz179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Chambers V.S., Marsico G., Boutell J.M., Di Antonio M., Smith G.P., Balasubramanian S. High-throughput sequencing of DNA G-quadruplex structures in the human genome. Nat. Biotechnol. 2015;33:877–881. doi: 10.1038/nbt.3295. [DOI] [PubMed] [Google Scholar]
- 16.Capra J.A., Paeschke K., Singh M., Zakian V.A. G-Quadruplex DNA Sequences Are Evolutionarily Conserved and Associated with Distinct Genomic Features in Saccharomyces cerevisiae. PLOS Comput. Biol. 2010;6 doi: 10.1371/journal.pcbi.1000861. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Frees S., Menendez C., Crum M., Bagga P.S. QGRS-Conserve: a computational method for discovering evolutionarily conserved G-quadruplex motifs. Hum. Genomics. 2014;8:8. doi: 10.1186/1479-7364-8-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Kim N. The Interplay between G-quadruplex and Transcription. Curr. Med. Chem. 2019;26:2898–2917. doi: 10.2174/0929867325666171229132619. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.David A.P., Pipier A., Pascutti F., Binolfi A., Weiner A.M.J., Challier E., Heckel S., Calsou P., Gomez D., Calcaterra N.B., Armas P. CNBP controls transcription by unfolding DNA G-quadruplex structures. Nucleic Acids Res. 2019;47:7901–7913. doi: 10.1093/nar/gkz527. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Tan J., Lan L. The DNA secondary structures at telomeres and genome instability. Cell Biosci. 2020;10:47. doi: 10.1186/s13578-020-00409-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Bryan T.M. G-Quadruplexes at Telomeres: Friend or Foe? Molecules. 2020;25:3686. doi: 10.3390/molecules25163686. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Guilbaud G., Murat P., Recolin B., Campbell B.C., Maiter A., Sale J.E., Balasubramanian S. Local epigenetic reprogramming induced by G-quadruplex ligands. Nat. Chem. 2017;9:1110–1117. doi: 10.1038/nchem.2828. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Mukherjee A.K., Sharma S., Chowdhury S. Non-duplex G-Quadruplex Structures Emerge as Mediators of Epigenetic Modifications. Trends Genet. 2019;35:129–144. doi: 10.1016/j.tig.2018.11.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Reina C., Cavalieri V. Epigenetic Modulation of Chromatin States and Gene Expression by G-Quadruplex Structures. Int. J. Mol. Sci. 2020;21 doi: 10.3390/ijms21114172. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Fleming A.M., Zhu J., Ding Y., Visser J.A., Zhu J., Burrows C.J. Human DNA repair genes possess potential G-quadruplex sequences in their promoters and 5‘-untranslated regions. Biochemistry. 2018;57:991–1002. doi: 10.1021/acs.biochem.7b01172. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.De Magis A., Götz S., Hajikazemi M., Fekete-Szücs E., Caterino M., Juranek S., Paeschke K. Zuo1 supports G4 structure formation and directs repair toward nucleotide excision repair. Nat. Commun. 2020;11:3907. doi: 10.1038/s41467-020-17701-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Valton A.-L., Prioleau M.-N. G-Quadruplexes in DNA Replication: A Problem or a Necessity? Trends Genet. 2016;32:697–706. doi: 10.1016/j.tig.2016.09.004. [DOI] [PubMed] [Google Scholar]
- 28.Prorok P., Artufel M., Aze A., Coulombe P., Peiffer I., Lacroix L., Guédin A., Mergny J.-L., Damaschke J., Schepers A., Cayrou C., Teulade-Fichou M.-P., Ballester B., Méchali M. Involvement of G-quadruplex regions in mammalian replication origin activity. Nat. Commun. 2019;10:3274. doi: 10.1038/s41467-019-11104-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Bryan T.M. Mechanisms of DNA Replication and Repair: Insights from the Study of G-Quadruplexes. Molecules. 2019;24:3439. doi: 10.3390/molecules24193439. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.K. Paeschke, P. Burkovics, Mgs1 function at G-quadruplex structures during DNA replication, Curr. Genet. (n.d.). doi: 10.1007/s00294-020-01128-1. [DOI] [PMC free article] [PubMed]
- 31.Wang Y., Yang J., Wild A.T., Wu W.H., Shah R., Danussi C., Riggins G.J., Kannan K., Sulman E.P., Chan T.A., Huse J.T. G-quadruplex DNA drives genomic instability and represents a targetable molecular abnormality in ATRX-deficient malignant glioma. Nat. Commun. 2019;10:943. doi: 10.1038/s41467-019-08905-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.De Magis A., Manzo S., Russo M., Marinello J., Morigi R., Sordet O., Capranico G. DNA damage and genome instability by G-quadruplex ligands are mediated by R loops in human cancer cells. Proc. Natl. Acad. Sci. 2018;116:201810409. doi: 10.1073/pnas.1810409116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.van Wietmarschen N., Merzouk S., Halsema N., Spierings D.C.J., Guryev V., Lansdorp P.M. BLM helicase suppresses recombination at G-quadruplex motifs in transcribed genes. Nat. Commun. 2018;9:271. doi: 10.1038/s41467-017-02760-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Saranathan N., Biswas B., Patra A., Vivekanandan P. G-quadruplexes may determine the landscape of recombination in HSV-1. BMC Genomics. 2019;20:382. doi: 10.1186/s12864-019-5731-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Dalloul Z., Chenuet P., Dalloul I., Boyer F., Aldigier J.-C., Laffleur B., El Makhour Y., Ryffel B., Quesniaux V.F.J., Togbé D., Mergny J.-L., Cook-Moreau J., Cogné M. G-quadruplex DNA targeting alters class-switch recombination in B cells and attenuates allergic inflammation. J. Allergy Clin. Immunol. 2018;142:1352–1355. doi: 10.1016/j.jaci.2018.06.011. [DOI] [PubMed] [Google Scholar]
- 36.Bochman M.L., Paeschke K., Zakian V.A. DNA secondary structures: stability and function of G-quadruplex structures. Nat. Rev. Genet. 2012;13:770–780. doi: 10.1038/nrg3296. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Varshney D., Spiegel J., Zyner K., Tannahill D., Balasubramanian S. The regulation and functions of DNA and RNA G-quadruplexes. Nat. Rev. Mol. Cell Biol. 2020;21:459–474. doi: 10.1038/s41580-020-0236-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Arthanari H., Bolton P.H. Functional and dysfunctional roles of quadruplex DNA in cells. Chem. Biol. 2001;8:221–230. doi: 10.1016/S1074-5521(01)00007-2. [DOI] [PubMed] [Google Scholar]
- 39.Cogoi S., Xodo L.E. G-quadruplex formation within the promoter of the KRAS proto-oncogene and its effect on transcription. Nucleic Acids Res. 2006;34:2536–2549. doi: 10.1093/nar/gkl286. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Ou A., Schmidberger J.W., Wilson K.A., Evans C.W., Hargreaves J.A., Grigg M., O’Mara M.L., Iyer K.S., Bond C.S., Smith N.M. High resolution crystal structure of a KRAS promoter G-quadruplex reveals a dimer with extensive poly-A π-stacking interactions for small-molecule recognition. Nucleic Acids Res. 2020;48:5766–5776. doi: 10.1093/nar/gkaa262. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Caterino M., D’Aria F., Kustov A., Belykh D., Khudyaeva I., Starseva O., Berezin D., Pylina Y., Usacheva T., Amato J., Giancola C. Selective binding of a bioactive porphyrin-based photosensitizer to the G-quadruplex from the KRAS oncogene promoter. Int. J. Biol. Macromol. 2019;145 doi: 10.1016/j.ijbiomac.2019.12.152. [DOI] [PubMed] [Google Scholar]
- 42.D’Aria F., D’Amore V.M., Di Leva F.S., Amato J., Caterino M., Russomanno P., Salerno S., Barresi E., De Leo M., Marini A.M., Taliani S., Da Settimo F., Salgado G.F., Pompili L., Zizza P., Shirasawa S., Novellino E., Biroccio A., Marinelli L., Giancola C. Targeting the KRAS oncogene: Synthesis, physicochemical and biological evaluation of novel G-Quadruplex DNA binders. Eur. J. Pharm. Sci. 2020;149 doi: 10.1016/j.ejps.2020.105337. [DOI] [PubMed] [Google Scholar]
- 43.Sun D., Liu W.-J., Guo K., Rusche J.J., Ebbinghaus S., Gokhale V., Hurley L.H. Proximal promoter region of the human vascular endothelial growth factor gene has a G-quadruplex structure which can be targeted by G-quadruplex-interactive agents. Mol. Cancer Ther. 2008;7:880–889. doi: 10.1158/1535-7163.MCT-07-2119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Ducani C., Bernardinelli G., Högberg B., Keppler B.K., Terenzi A. Interplay of Three G-Quadruplex Units in the KIT Promoter. J. Am. Chem. Soc. 2019;141:10205–10213. doi: 10.1021/jacs.8b12753. [DOI] [PubMed] [Google Scholar]
- 45.Fernando H., Reszka A.P., Huppert J., Ladame S., Rankin S., Venkitaraman A.R., Neidle S., Balasubramanian S. A Conserved Quadruplex Motif Located in a Transcription Activation Site of the Human c-kit Oncogene. Biochemistry. 2006;45:7854–7860. doi: 10.1021/bi0601510. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Agrawal P., Lin C., Mathad R.I., Carver M., Yang D. The major G-quadruplex formed in the human BCL-2 proximal promoter adopts a parallel structure with a 13-nt loop in K+ solution. J. Am. Chem. Soc. 2014;136:1750–1753. doi: 10.1021/ja4118945. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Brooks T.A., Hurley L.H. Targeting MYC Expression through G-Quadruplexes. Genes Cancer. 2010;1:641–649. doi: 10.1177/1947601910377493. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Siddiqui-Jain A., Grand C.L., Bearss D.J., Hurley L.H. Direct evidence for a G-quadruplex in a promoter region and its targeting with a small molecule to repress c-MYC transcription. Proc. Natl. Acad. Sci. U. S. A. 2002;99:11593–11598. doi: 10.1073/pnas.182256799. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Stump S., Mou T.-C., Sprang S.R., Natale N.R., Beall H.D. Crystal structure of the major quadruplex formed in the promoter region of the human c-MYC oncogene. PLoS ONE. 2018;13 doi: 10.1371/journal.pone.0205584. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Wang W., Hu S., Gu Y., Yan Y., Stovall D.B., Li D., Sui G. Human MYC G-quadruplex: From discovery to a cancer therapeutic target. Biochim. Biophys. Acta-Rev. Cancer. 2020;1874 doi: 10.1016/j.bbcan.2020.188410. [DOI] [PubMed] [Google Scholar]
- 51.Kosiol N., Juranek S., Brossart P., Heine A., Paeschke K. G-quadruplexes: a promising target for cancer therapy. Mol. Cancer. 2021;20:40. doi: 10.1186/s12943-021-01328-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Neidle S. Quadruplex nucleic acids as targets for anticancer therapeutics. Nat. Rev. Chem. 2017;1:1–10. doi: 10.1038/s41570-017-0041. [DOI] [Google Scholar]
- 53.Carvalho J., Mergny J.-L., Salgado G.F., Queiroz J.A., Cruz C. G-quadruplex, Friend or Foe: The Role of the G-quartet in Anticancer Strategies. Trends Mol. Med. 2020;26:848–861. doi: 10.1016/j.molmed.2020.05.002. [DOI] [PubMed] [Google Scholar]
- 54.Sanchez-Martin V., Lopez-Pujante C., Soriano-Rodriguez M., Garcia-Salcedo J.A. An Updated Focus on Quadruplex Structures as Potential Therapeutic Targets in Cancer. Int. J. Mol. Sci. 2020;21:8900. doi: 10.3390/ijms21238900. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Fay M.M., Lyons S.M., Ivanov P. RNA G-quadruplexes in biology: principles and molecular mechanisms. J. Mol. Biol. 2017;429:2127–2147. doi: 10.1016/j.jmb.2017.05.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Di Antonio M., Biffi G., Mariani A., Raiber E.-A., Rodriguez R., Balasubramanian S. Selective RNA versus DNA G-quadruplex targeting by in situ click chemistry. Angew. Chem. Int. Ed Engl. 2012;51:11073–11078. doi: 10.1002/anie.201206281. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Biffi G., Di Antonio M., Tannahill D., Balasubramanian S. Visualization and selective chemical targeting of RNA G-quadruplex structures in the cytoplasm of human cells. Nat. Chem. 2014;6:75–80. doi: 10.1038/nchem.1805. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Kwok C.K., Marsico G., Sahakyan A.B., Chambers V.S., Balasubramanian S. rG4-seq reveals widespread formation of G-quadruplex structures in the human transcriptome. Nat. Methods. 2016;13:841–844. doi: 10.1038/nmeth.3965. [DOI] [PubMed] [Google Scholar]
- 59.Guo J.U., Bartel D.P. RNA G-quadruplexes are globally unfolded in eukaryotic cells and depleted in bacteria. Science. 2016;353 doi: 10.1126/science.aaf5371. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Di Antonio M., Ponjavic A., Radzevičius A., Ranasinghe R.T., Catalano M., Zhang X., Shen J., Needham L.-M., Lee S.F., Klenerman D., Balasubramanian S. Single-molecule visualization of DNA G-quadruplex formation in live cells. Nat. Chem. 2020;12:832–837. doi: 10.1038/s41557-020-0506-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Chen X.-C., Chen S.-B., Dai J., Yuan J.-H., Ou T.-M., Huang Z.-S., Tan J.-H. Tracking the Dynamic Folding and Unfolding of RNA G-Quadruplexes in Live Cells. Angew. Chem. Int. Ed Engl. 2018;57:4702–4706. doi: 10.1002/anie.201801999. [DOI] [PubMed] [Google Scholar]
- 62.Yang S.Y., Lejault P., Chevrier S., Boidot R., Robertson A.G., Wong J.M.Y., Monchaud D. Transcriptome-wide identification of transient RNA G-quadruplexes in human cells. Nat. Commun. 2018;9:4730. doi: 10.1038/s41467-018-07224-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Renard I., Grandmougin M., Roux A., Yang S.Y., Lejault P., Pirrotta M., Wong J.M.Y., Monchaud D. Small-molecule affinity capture of DNA/RNA quadruplexes and their identification in vitro and in vivo through the G4RP protocol. Nucleic Acids Res. 2019;47:5502–5510. doi: 10.1093/nar/gkz215. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Kharel P., Becker G., Tsvetkov V., Ivanov P. Properties and biological impact of RNA G-quadruplexes: from order to turmoil and back. Nucleic Acids Res. 2020;48:12534–12555. doi: 10.1093/nar/gkaa1126. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Agarwala P., Pandey S., Maiti S. The tale of RNA G-quadruplex. Org. Biomol. Chem. 2015;13:5570–5585. doi: 10.1039/C4OB02681K. [DOI] [PubMed] [Google Scholar]
- 66.Song J., Perreault J.-P., Topisirovic I., Richard S. RNA G-quadruplexes and their potential regulatory roles in translation. Translation. 2016;4 doi: 10.1080/21690731.2016.1244031. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Endoh T., Sugimoto N. Conformational Dynamics of the RNA G-Quadruplex and its Effect on Translation Efficiency. Molecules. 2019;24 doi: 10.3390/molecules24081613. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Herviou P., Le Bras M., Dumas L., Hieblot C., Gilhodes J., Cioci G., Hugnot J.-P., Ameadan A., Guillonneau F., Dassi E., Cammas A., Millevoi S. hnRNP H/F drive RNA G-quadruplex-mediated translation linked to genomic instability and therapy resistance in glioblastoma. Nat. Commun. 2020;11:2661. doi: 10.1038/s41467-020-16168-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Endoh T., Sugimoto N. Mechanical insights into ribosomal progression overcoming RNA G-quadruplex from periodical translation suppression in cells. Sci. Rep. 2016;6:22719. doi: 10.1038/srep22719. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Jodoin R., Carrier J.C., Rivard N., Bisaillon M., Perreault J.-P. G-quadruplex located in the 5′UTR of the BAG-1 mRNA affects both its cap-dependent and cap-independent translation through global secondary structure maintenance. Nucleic Acids Res. 2019;47:10247–10266. doi: 10.1093/nar/gkz777. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Arora A., Suess B. An RNA G-quadruplex in the 3’ UTR of the proto-oncogene PIM1 represses translation. RNA Biol. 2011;8:802–805. doi: 10.4161/rna.8.5.16038. [DOI] [PubMed] [Google Scholar]
- 72.Dabral P., Babu J., Zareie A., Verma S.C. LANA and hnRNP A1 Regulate the Translation of LANA mRNA through G-Quadruplexes. J. Virol. 2020;94 doi: 10.1128/JVI.01508-19. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Kanai Y., Dohmae N., Hirokawa N. Kinesin transports RNA: isolation and characterization of an RNA-transporting granule. Neuron. 2004;43:513–525. doi: 10.1016/j.neuron.2004.07.022. [DOI] [PubMed] [Google Scholar]
- 74.Subramanian M., Rage F., Tabet R., Flatter E., Mandel J.-L., Moine H. G–quadruplex RNA structure as a signal for neurite mRNA targeting. EMBO Rep. 2011;12:697–704. doi: 10.1038/embor.2011.76. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Huppert J.L., Bugaut A., Kumari S., Balasubramanian S. G-quadruplexes: the beginning and end of UTRs. Nucleic Acids Res. 2008;36:6260–6268. doi: 10.1093/nar/gkn511. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Huang H., Zhang J., Harvey S.E., Hu X., Cheng C. RNA G-quadruplex secondary structure promotes alternative splicing via the RNA-binding protein hnRNPF. Genes Dev. 2017;31:2296–2309. doi: 10.1101/gad.305862.117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Kwok C.K., Marsico G., Balasubramanian S. Detecting RNA G-Quadruplexes (rG4s) in the Transcriptome. Cold Spring Harb. Perspect. Biol. 2018;10 doi: 10.1101/cshperspect.a032284. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.M.-A. Turcotte, Guanine Nucleotide-Binding Protein-Like 1 (GNL1) binds RNA G-quadruplex structures in genes associat, (n.d.) 16. [DOI] [PMC free article] [PubMed]
- 79.Marcel V., Tran P.L.T., Sagne C., Martel-Planche G., Vaslin L., Teulade-Fichou M.-P., Hall J., Mergny J.-L., Hainaut P., Van Dyck E. G-quadruplex structures in TP53 intron 3: role in alternative splicing and in production of p53 mRNA isoforms. Carcinogenesis. 2011;32:271–278. doi: 10.1093/carcin/bgq253. [DOI] [PubMed] [Google Scholar]
- 80.Millevoi S., Moine H., Vagner S. G-quadruplexes in RNA biology. WIREs RNA. 2012;3:495–507. doi: 10.1002/wrna.1113. [DOI] [PubMed] [Google Scholar]
- 81.Weldon C., Dacanay J.G., Gokhale V., Boddupally P.V.L., Behm-Ansmant I., Burley G.A., Branlant C., Hurley L.H., Dominguez C., Eperon I.C. Specific G-quadruplex ligands modulate the alternative splicing of Bcl-X. Nucleic Acids Res. 2018;46:886–896. doi: 10.1093/nar/gkx1122. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Jara-Espejo M., Fleming A.M., Burrows C.J. Potential G-Quadruplex Forming Sequences and N6-Methyladenosine Colocalize at Human Pre-mRNA Intron Splice Sites. ACS Chem. Biol. 2020;15:1292–1300. doi: 10.1021/acschembio.0c00260. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Xu B., Meng Y., Jin Y. RNA structures in alternative splicing and back-splicing. WIREs RNA. 2021;12 doi: 10.1002/wrna.1626. [DOI] [PubMed] [Google Scholar]
- 84.Taylor K., Sobczak K. Intrinsic Regulatory Role of RNA Structural Arrangement in Alternative Splicing Control. Int. J. Mol. Sci. 2020;21:5161. doi: 10.3390/ijms21145161. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Raguseo F., Chowdhury S., Minard A., Antonio M.D. Chemical-biology approaches to probe DNA and RNA G-quadruplex structures in the genome. Chem. Commun. 2020;56:1317–1324. doi: 10.1039/C9CC09107F. [DOI] [PubMed] [Google Scholar]
- 86.Biffi G., Tannahill D., McCafferty J., Balasubramanian S. Quantitative visualization of DNA G-quadruplex structures in human cells. Nat. Chem. 2013;5:182–186. doi: 10.1038/nchem.1548. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Byrd A.K., Zybailov B.L., Maddukuri L., Gao J., Marecki J.C., Jaiswal M., Bell M.R., Griffin W.C., Reed M.R., Chib S., Mackintosh S.G., MacNicol A.M., Baldini G., Eoff R.L., Raney K.D. Evidence That G-quadruplex DNA Accumulates in the Cytoplasm and Participates in Stress Granule Assembly in Response to Oxidative Stress. J. Biol. Chem. 2016;291:18041–18057. doi: 10.1074/jbc.M116.718478. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Baudrimont A., Voegeli S., Viloria E.C., Stritt F., Lenon M., Wada T., Jaquet V., Becskei A. Multiplexed gene control reveals rapid mRNA turnover. Sci. Adv. 2017;3 doi: 10.1126/sciadv.1700006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Halder R., Riou J.-F., Teulade-Fichou M.-P., Frickey T., Hartig J.S. Bisquinolinium compounds induce quadruplex-specific transcriptome changes in HeLa S3 cell lines. BMC Res. Notes. 2012;5:138. doi: 10.1186/1756-0500-5-138. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Xu H., Di Antonio M., McKinney S., Mathew V., Ho B., O’Neil N.J., Santos N.D., Silvester J., Wei V., Garcia J., Kabeer F., Lai D., Soriano P., Banáth J., Chiu D.S., Yap D., Le D.D., Ye F.B., Zhang A., Thu K., Soong J., Lin S., Tsai A.H.C., Osako T., Algara T., Saunders D.N., Wong J., Xian J., Bally M.B., Brenton J.D., Brown G.W., Shah S.P., Cescon D., Mak T.W., Caldas C., Stirling P.C., Hieter P., Balasubramanian S., Aparicio S. CX-5461 is a DNA G-quadruplex stabilizer with selective lethality in BRCA1/2 deficient tumours. Nat. Commun. 2017;8:14432. doi: 10.1038/ncomms14432. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Gray L.T., Vallur A.C., Eddy J., Maizels N. G-quadruplexes are genomewide targets of transcriptional helicases XPB and XPD. Nat. Chem. Biol. 2014;10:313–318. doi: 10.1038/nchembio.1475. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Lane A.N., Chaires J.B., Gray R.D., Trent J.O. Stability and kinetics of G-quadruplex structures. Nucleic Acids Res. 2008;36:5482–5515. doi: 10.1093/nar/gkn517. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Borgognone M., Armas P., Calcaterra N.B. Cellular nucleic-acid-binding protein, a transcriptional enhancer of c-Myc, promotes the formation of parallel G-quadruplexes. Biochem. J. 2010;428:491–498. doi: 10.1042/BJ20100038. [DOI] [PubMed] [Google Scholar]
- 94.Zanotti K.J., Lackey P.E., Evans G.L., Mihailescu M.-R. Thermodynamics of the Fragile X Mental Retardation Protein RGG Box Interactions with G Quartet Forming RNA. Biochemistry. 2006;45:8319–8330. doi: 10.1021/bi060209a. [DOI] [PubMed] [Google Scholar]
- 95.Benhalevy D., Gupta S.K., Danan C.H., Ghosal S., Sun H.-W., Kazemier H.G., Paeschke K., Hafner M., Juranek S.A. The Human CCHC-type Zinc Finger Nucleic Acid-Binding Protein Binds G-Rich Elements in Target mRNA Coding Sequences and Promotes Translation. Cell Rep. 2017;18:2979–2990. doi: 10.1016/j.celrep.2017.02.080. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Meier M., Patel T.R., Booy E.P., Marushchak O., Okun N., Deo S., Howard R., McEleney K., Harding S.E., Stetefeld J., McKenna S.A. Binding of G-quadruplexes to the N-terminal recognition domain of the RNA helicase associated with AU-rich element (RHAU) J. Biol. Chem. 2013;288:35014–35027. doi: 10.1074/jbc.M113.512970. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Wu G., Xing Z., Tran E.J., Yang D. DDX5 helicase resolves G-quadruplex and is involved in MYC gene transcriptional activation. Proc. Natl. Acad. Sci. 2019;116:20453–20461. doi: 10.1073/pnas.1909047116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Paeschke K., Bochman M.L., Garcia P.D., Cejka P., Friedman K.L., Kowalczykowski S.C., Zakian V.A. Pif1 family helicases suppress genome instability at G-quadruplex motifs. Nature. 2013;497:458–462. doi: 10.1038/nature12149. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Wu Y., Shin-ya K., Brosh R.M. FANCJ Helicase Defective in Fanconia Anemia and Breast Cancer Unwinds G-Quadruplex DNA To Defend Genomic Stability. Mol. Cell. Biol. 2008;28:4116–4128. doi: 10.1128/MCB.02210-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Brosh R.M. DNA helicases involved in DNA repair and their roles in cancer. Nat. Rev. Cancer. 2013;13:542–558. doi: 10.1038/nrc3560. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Herdy B., Mayer C., Varshney D., Marsico G., Murat P., Taylor C., D’Santos C., Tannahill D., Balasubramanian S. Analysis of NRAS RNA G-quadruplex binding proteins reveals DDX3X as a novel interactor of cellular G-quadruplex containing transcripts. Nucleic Acids Res. 2018;46:11592–11604. doi: 10.1093/nar/gky861. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Mishra S.K., Tawani A., Mishra A., Kumar A. G4IPDB: A database for G-quadruplex structure forming nucleic acid interacting proteins. Sci. Rep. 2016;6:38144. doi: 10.1038/srep38144. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.von Hacht A., Seifert O., Menger M., Schütze T., Arora A., Konthur Z., Neubauer P., Wagner A., Weise C., Kurreck J. Identification and characterization of RNA guanine-quadruplex binding proteins. Nucleic Acids Res. 2014;42:6630–6644. doi: 10.1093/nar/gku290. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Mori K., Lammich S., Mackenzie I.R.A., Forné I., Zilow S., Kretzschmar H., Edbauer D., Janssens J., Kleinberger G., Cruts M., Herms J., Neumann M., Van Broeckhoven C., Arzberger T., Haass C. hnRNP A3 binds to GGGGCC repeats and is a constituent of p62-positive/TDP43-negative inclusions in the hippocampus of patients with C9orf72 mutations. Acta Neuropathol. (Berl.) 2013;125:413–423. doi: 10.1007/s00401-013-1088-7. [DOI] [PubMed] [Google Scholar]
- 105.Dai J., Liu Z.-Q., Wang X.-Q., Lin J., Yao P.-F., Huang S.-L., Ou T.-M., Tan J.-H., Li D., Gu L.-Q., Huang Z.-S. Discovery of Small Molecules for Up-Regulating the Translation of Antiamyloidogenic Secretase, a Disintegrin and Metalloproteinase 10 (ADAM10), by Binding to the G-Quadruplex-Forming Sequence in the 5’ Untranslated Region (UTR) of Its mRNA. J. Med. Chem. 2015;58:3875–3891. doi: 10.1021/acs.jmedchem.5b00139. [DOI] [PubMed] [Google Scholar]
- 106.Nagata T., Takada Y., Ono A., Nagata K., Konishi Y., Nukina T., Ono M., Matsugami A., Furukawa A., Fujimoto N., Fukuda H., Nakagama H., Katahira M. Elucidation of the mode of interaction in the UP1-telomerase RNA-telomeric DNA ternary complex which serves to recruit telomerase to telomeric DNA and to enhance the telomerase activity. Nucleic Acids Res. 2008;36:6816–6824. doi: 10.1093/nar/gkn767. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Takahama K., Takada A., Tada S., Shimizu M., Sayama K., Kurokawa R., Oyoshi T. Regulation of telomere length by G-quadruplex telomere DNA- and TERRA-binding protein TLS/FUS. Chem. Biol. 2013;20:341–350. doi: 10.1016/j.chembiol.2013.02.013. [DOI] [PubMed] [Google Scholar]
- 108.Zhang Y., Gaetano C.M., Williams K.R., Bassell G.J., Mihailescu M.R. FMRP interacts with G-quadruplex structures in the 3’-UTR of its dendritic target Shank1 mRNA. RNA Biol. 2014;11:1364–1374. doi: 10.1080/15476286.2014.996464. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Biffi G., Tannahill D., Balasubramanian S. An intramolecular G-quadruplex structure is required for binding of telomeric repeat-containing RNA to the telomeric protein TRF2. J. Am. Chem. Soc. 2012;134:11974–11976. doi: 10.1021/ja305734x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Al-Furoukh N., Goffart S., Szibor M., Wanrooij S., Braun T. Binding to G-quadruplex RNA activates the mitochondrial GTPase NOA1. Biochim. Biophys. Acta. 1833;2013:2933–2942. doi: 10.1016/j.bbamcr.2013.07.022. [DOI] [PubMed] [Google Scholar]
- 111.Bolduc F., Turcotte M.-A., Perreault J.-P. The Small Nuclear Ribonucleoprotein Polypeptide A (SNRPA) binds to the G-quadruplex of the BAG-1 5’UTR. Biochimie. 2020;176:122–127. doi: 10.1016/j.biochi.2020.06.013. [DOI] [PubMed] [Google Scholar]
- 112.Simko E.A.J., Liu H., Zhang T., Velasquez A., Teli S., Haeusler A.R., Wang J. G-quadruplexes offer a conserved structural motif for NONO recruitment to NEAT1 architectural lncRNA. Nucleic Acids Res. 2020;48:7421–7438. doi: 10.1093/nar/gkaa475. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Brázda V., Hároníková L., Liao J.C.C., Fojta M. DNA and RNA Quadruplex-Binding Proteins. Int. J. Mol. Sci. 2014;15:17493–17517. doi: 10.3390/ijms151017493. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Sauer M., Paeschke K. G-quadruplex unwinding helicases and their function in vivo. Biochem. Soc. Trans. 2017;45:1173–1182. doi: 10.1042/BST20170097. [DOI] [PubMed] [Google Scholar]
- 115.Mendoza O., Bourdoncle A., Boulé J.-B., Brosh R.M., Mergny J.-L. G-quadruplexes and helicases. Nucleic Acids Res. 2016;44:1989–2006. doi: 10.1093/nar/gkw079. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Gellert M., Lipsett M.N., Davies D.R. Helix formation by guanylic acid. Proc. Natl. Acad. Sci. U. S. A. 1962;48:2013–2018. doi: 10.1073/pnas.48.12.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Sen D., Gilbert W. A sodium-potassium switch in the formation of four-stranded G4-DNA. Nature. 1990;344:410–414. doi: 10.1038/344410a0. [DOI] [PubMed] [Google Scholar]
- 118.Mukundan V.T., Phan A.T. Bulges in G-quadruplexes: broadening the definition of G-quadruplex-forming sequences. J. Am. Chem. Soc. 2013;135:5017–5028. doi: 10.1021/ja310251r. [DOI] [PubMed] [Google Scholar]
- 119.Sengar A., Vandana J.J., Chambers V.S., Di Antonio M., Winnerdy F.R., Balasubramanian S., Phan A.T. Structure of a (3+1) hybrid G-quadruplex in the PARP1 promoter. Nucleic Acids Res. 2019;47:1564–1572. doi: 10.1093/nar/gky1179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Cheng M., Cheng Y., Hao J., Jia G., Zhou J., Mergny J.-L., Li C. Loop permutation affects the topology and stability of G-quadruplexes. Nucleic Acids Res. 2018;46:9264–9275. doi: 10.1093/nar/gky757. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Hänsel-Hertsch R., Di Antonio M., Balasubramanian S. DNA G-quadruplexes in the human genome: detection, functions and therapeutic potential. Nat. Rev. Mol. Cell Biol. 2017;18:279–284. doi: 10.1038/nrm.2017.3. [DOI] [PubMed] [Google Scholar]
- 122.Mergny J., Lacroix L. UV Melting of G-Quadruplexes. Curr. Protoc. Nucleic Acid Chem. 2009;37 doi: 10.1002/0471142700.nc1701s37. [DOI] [PubMed] [Google Scholar]
- 123.Burge S., Parkinson G.N., Hazel P., Todd A.K., Neidle S. Quadruplex DNA: sequence, topology and structure. Nucleic Acids Res. 2006;34:5402–5415. doi: 10.1093/nar/gkl655. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.D. Monchaud, Quadruplex detection in human cells, in: Annu. Rep. Med. Chem., Elsevier, 2020: pp. 133–160. 10.1016/bs.armc.2020.04.007.
- 125.Paeschke K., Simonsson T., Postberg J., Rhodes D., Lipps H.J. Telomere end-binding proteins control the formation of G-quadruplex DNA structures in vivo. Nat. Struct. Mol. Biol. 2005;12:847–854. doi: 10.1038/nsmb982. [DOI] [PubMed] [Google Scholar]
- 126.Paeschke K., Juranek S., Simonsson T., Hempel A., Rhodes D., Lipps H.J. Telomerase recruitment by the telomere end binding protein-β facilitates G-quadruplex DNA unfolding in ciliates. Nat. Struct. Mol. Biol. 2008;15:598–604. doi: 10.1038/nsmb.1422. [DOI] [PubMed] [Google Scholar]
- 127.Henderson A., Wu Y., Huang Y.C., Chavez E.A., Platt J., Johnson F.B., Brosh R.M., Sen D., Lansdorp P.M. Detection of G-quadruplex DNA in mammalian cells. Nucleic Acids Res. 2014;42:860–869. doi: 10.1093/nar/gkt957. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128.Kazemier H.G., Paeschke K., Lansdorp P.M. Guanine quadruplex monoclonal antibody 1H6 cross-reacts with restrained thymidine-rich single stranded DNA. Nucleic Acids Res. 2017;45:5913–5919. doi: 10.1093/nar/gkx245. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Rodriguez R., Miller K.M., Forment J.V., Bradshaw C.R., Nikan M., Britton S., Oelschlaegel T., Xhemalce B., Balasubramanian S., Jackson S.P. Small-molecule–induced DNA damage identifies alternative DNA structures in human genes. Nat. Chem. Biol. 2012;8:301–310. doi: 10.1038/nchembio.780. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Lefebvre J., Guetta C., Poyer F., Mahuteau-Betzer F., Teulade-Fichou M.-P. Copper-Alkyne Complexation Responsible for the Nucleolar Localization of Quadruplex Nucleic Acid Drugs Labeled by Click Reactions. Angew. Chem. Int. Ed. 2017;56:11365–11369. doi: 10.1002/anie.201703783. [DOI] [PubMed] [Google Scholar]
- 131.Shivalingam A., Izquierdo M.A., Marois A.L., Vyšniauskas A., Suhling K., Kuimova M.K., Vilar R. The interactions between a small molecule and G-quadruplexes are visualized by fluorescence lifetime imaging microscopy. Nat. Commun. 2015;6:8178. doi: 10.1038/ncomms9178. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.De Magis A., Kastl M., Brossart P., Heine A., Paeschke K. BG-flow, a new flow cytometry tool for G-quadruplex quantification in fixed cells. BMC Biol. 2021;19:45. doi: 10.1186/s12915-021-00986-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Hänsel-Hertsch R., Beraldi D., Lensing S.V., Marsico G., Zyner K., Parry A., Di Antonio M., Pike J., Kimura H., Narita M., Tannahill D., Balasubramanian S. G-quadruplex structures mark human regulatory chromatin. Nat. Genet. 2016;48:1267–1272. doi: 10.1038/ng.3662. [DOI] [PubMed] [Google Scholar]
- 134.Zheng K., Zhang J., He Y., Gong J., Wen C., Chen J., Hao Y., Zhao Y., Tan Z. Detection of genomic G-quadruplexes in living cells using a small artificial protein. Nucleic Acids Res. 2020;48:11706–11720. doi: 10.1093/nar/gkaa841. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.Law M.J., Lower K.M., Voon H.P.J., Hughes J.R., Garrick D., Viprakasit V., Mitson M., De Gobbi M., Marra M., Morris A., Abbott A., Wilder S.P., Taylor S., Santos G.M., Cross J., Ayyub H., Jones S., Ragoussis J., Rhodes D., Dunham I., Higgs D.R., Gibbons R.J. ATR-X Syndrome Protein Targets Tandem Repeats and Influences Allele-Specific Expression in a Size-Dependent Manner. Cell. 2010;143:367–378. doi: 10.1016/j.cell.2010.09.023. [DOI] [PubMed] [Google Scholar]
- 136.Paeschke K., Capra J.A., Zakian V.A. DNA Replication through G-Quadruplex Motifs Is Promoted by the Saccharomyces cerevisiae Pif1 DNA Helicase. Cell. 2011;145:678–691. doi: 10.1016/j.cell.2011.04.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Treiber D.K., Williamson J.R. Exposing the kinetic traps in RNA folding. Curr. Opin. Struct. Biol. 1999;9:339–345. doi: 10.1016/S0959-440X(99)80045-1. [DOI] [PubMed] [Google Scholar]
- 138.Thirumalai D., Hyeon C. RNA and protein folding: common themes and variations. Biochemistry. 2005;44:4957–4970. doi: 10.1021/bi047314+. [DOI] [PubMed] [Google Scholar]
- 139.Weeks K.M. Protein-facilitated RNA folding. Curr. Opin. Struct. Biol. 1997;7:336–342. doi: 10.1016/s0959-440x(97)80048-6. [DOI] [PubMed] [Google Scholar]
- 140.Woodson S.A. Taming free energy landscapes with RNA chaperones. RNA Biol. 2010;7:677–686. doi: 10.4161/rna.7.6.13615. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141.Zemora G., Waldsich C. RNA folding in living cells. RNA Biol. 2010;7:634–641. doi: 10.4161/rna.7.6.13554. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Bourgeois C.F., Mortreux F., Auboeuf D. The multiple functions of RNA helicases as drivers and regulators of gene expression. Nat. Rev. Mol. Cell Biol. 2016;17:426–438. doi: 10.1038/nrm.2016.50. [DOI] [PubMed] [Google Scholar]
- 143.Rocak S., Linder P. DEAD-box proteins: the driving forces behind RNA metabolism. Nat. Rev. Mol. Cell Biol. 2004;5:232–241. doi: 10.1038/nrm1335. [DOI] [PubMed] [Google Scholar]
- 144.Gorbalenya A.E., Koonin E.V. Helicases: amino acid sequence comparisons and structure-function relationships. Curr. Opin. Struct. Biol. 1993;3:419–429. doi: 10.1016/S0959-440X(05)80116-2. [DOI] [Google Scholar]
- 145.Singleton M.R., Dillingham M.S., Wigley D.B. Structure and mechanism of helicases and nucleic acid translocases. Annu. Rev. Biochem. 2007;76:23–50. doi: 10.1146/annurev.biochem.76.052305.115300. [DOI] [PubMed] [Google Scholar]
- 146.Bessler J.B., Torres J.Z., Zakian V.A. The Pif1p subfamily of helicases: region-specific DNA helicases? Trends Cell Biol. 2001;11:60–65. doi: 10.1016/s0962-8924(00)01877-8. [DOI] [PubMed] [Google Scholar]
- 147.Dürr H., Flaus A., Owen-Hughes T., Hopfner K.-P. Snf2 family ATPases and DExx box helicases: differences and unifying concepts from high-resolution crystal structures. Nucleic Acids Res. 2006;34:4160–4167. doi: 10.1093/nar/gkl540. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.Linder P. Dead-box proteins: a family affair—active and passive players in RNP-remodeling. Nucleic Acids Res. 2006;34:4168–4180. doi: 10.1093/nar/gkl468. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149.Fairman-Williams M.E., Guenther U.-P., Jankowsky E. SF1 and SF2 helicases: family matters. Curr. Opin. Struct. Biol. 2010;20:313–324. doi: 10.1016/j.sbi.2010.03.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150.Bochman M.L., Sabouri N., Zakian V.A. Unwinding the functions of the Pif1 family helicases. DNA Repair. 2010;9:237–249. doi: 10.1016/j.dnarep.2010.01.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.Putnam A.A., Jankowsky E. DEAD-box helicases as integrators of RNA, nucleotide and protein binding. Biochim. Biophys. Acta. 1829;2013:884–893. doi: 10.1016/j.bbagrm.2013.02.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152.Jankowsky E. RNA helicases at work: binding and rearranging. Trends Biochem. Sci. 2011;36:19–29. doi: 10.1016/j.tibs.2010.07.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153.Ozgur S., Buchwald G., Falk S., Chakrabarti S., Prabu J.R., Conti E. The conformational plasticity of eukaryotic RNA-dependent ATPases. FEBS J. 2015;282:850–863. doi: 10.1111/febs.13198. [DOI] [PubMed] [Google Scholar]
- 154.Andreou A.Z., Klostermeier D. The DEAD-box helicase eIF4A. RNA Biol. 2013;10:19–32. doi: 10.4161/rna.21966. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Awano N., Xu C., Ke H., Inoue K., Inouye M., Phadtare S. Complementation Analysis of the Cold-Sensitive Phenotype of the Escherichia coli csdA Deletion Strain. J. Bacteriol. 2007;189:5808–5815. doi: 10.1128/JB.00655-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Rajkowitsch L., Chen D., Stampfl S., Semrad K., Waldsich C., Mayer O., Jantsch M.F., Konrat R., Bläsi U., Schroeder R. RNA Chaperones, RNA Annealers and RNA Helicases. RNA Biol. 2007;4:118–130. doi: 10.4161/rna.4.3.5445. [DOI] [PubMed] [Google Scholar]
- 157.Tran H., Schilling M., Wirbelauer C., Hess D., Nagamine Y. Facilitation of mRNA Deadenylation and Decay by the Exosome-Bound, DExH Protein RHAU. Mol. Cell. 2004;13:101–111. doi: 10.1016/S1097-2765(03)00481-7. [DOI] [PubMed] [Google Scholar]
- 158.Sauer M., Juranek S.A., Marks J., De Magis A., Kazemier H.G., Hilbig D., Benhalevy D., Wang X., Hafner M., Paeschke K. DHX36 prevents the accumulation of translationally inactive mRNAs with G4-structures in untranslated regions. Nat. Commun. 2019;10:2421. doi: 10.1038/s41467-019-10432-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159.Iwamoto F., Stadler M., Chalupníková K., Oakeley E., Nagamine Y. Transcription-dependent nucleolar cap localization and possible nuclear function of DExH RNA helicase RHAU. Exp. Cell Res. 2008;314:1378–1391. doi: 10.1016/j.yexcr.2008.01.006. [DOI] [PubMed] [Google Scholar]
- 160.Booy E.P., McRae E.K.S., Howard R., Deo S.R., Ariyo E.O., Dzananovic E., Meier M., Stetefeld J., McKenna S.A. RNA Helicase Associated with AU-rich Element (RHAU/DHX36) Interacts with the 3′-Tail of the Long Non-coding RNA BC200 (BCYRN1)*. J. Biol. Chem. 2016;291:5355–5372. doi: 10.1074/jbc.M115.711499. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 161.Giri B., Smaldino P.J., Thys R.G., Creacy S.D., Routh E.D., Hantgan R.R., Lattmann S., Nagamine Y., Akman S.A., Vaughn J.P. G4 resolvase 1 tightly binds and unwinds unimolecular G4-DNA. Nucleic Acids Res. 2011;39:7161–7178. doi: 10.1093/nar/gkr234. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 162.Creacy S.D., Routh E.D., Iwamoto F., Nagamine Y., Akman S.A., Vaughn J.P. G4 Resolvase 1 Binds Both DNA and RNA Tetramolecular Quadruplex with High Affinity and Is the Major Source of Tetramolecular Quadruplex G4-DNA and G4-RNA Resolving Activity in HeLa Cell Lysates*. J. Biol. Chem. 2008;283:34626–34634. doi: 10.1074/jbc.M806277200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 163.Vaughn J.P., Creacy S.D., Routh E.D., Joyner-Butt C., Jenkins G.S., Pauli S., Nagamine Y., Akman S.A. The DEXH protein product of the DHX36 gene is the major source of tetramolecular quadruplex G4-DNA resolving activity in HeLa cell lysates. J. Biol. Chem. 2005;280:38117–38120. doi: 10.1074/jbc.C500348200. [DOI] [PubMed] [Google Scholar]
- 164.Nie J., Jiang M., Zhang X., Tang H., Jin H., Huang X., Yuan B., Zhang C., Lai J.C., Nagamine Y., Pan D., Wang W., Yang Z. Post-transcriptional Regulation of Nkx2-5 by RHAU in Heart Development. Cell Rep. 2015;13:723–732. doi: 10.1016/j.celrep.2015.09.043. [DOI] [PubMed] [Google Scholar]
- 165.Lai J.C., Ponti S., Pan D., Kohler H., Skoda R.C., Matthias P., Nagamine Y. The DEAH-box helicase RHAU is an essential gene and critical for mouse hematopoiesis. Blood. 2012;119:4291–4300. doi: 10.1182/blood-2011-08-362954. [DOI] [PubMed] [Google Scholar]
- 166.Gao X., Ma W., Nie J., Zhang C., Zhang J., Yao G., Han J., Xu J., Hu B., Du Y., Shi Q., Yang Z., Huang X., Zhang Y. A G-quadruplex DNA structure resolvase, RHAU, is essential for spermatogonia differentiation. Cell Death Dis. 2015;6 doi: 10.1038/cddis.2014.571. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 167.Chashchina G.V., Beniaminov A.D., Kaluzhny D.N. Stable G-Quadruplex Structures of Oncogene Promoters Induce Potassium-Dependent Stops of Thermostable DNA Polymerase. Biochem. Mosc. 2019;84:562–569. doi: 10.1134/S0006297919050109. [DOI] [PubMed] [Google Scholar]
- 168.Huang W., Smaldino P.J., Zhang Q., Miller L.D., Cao P., Stadelman K., Wan M., Giri B., Lei M., Nagamine Y., Vaughn J.P., Akman S.A., Sui G. Yin Yang 1 contains G-quadruplex structures in its promoter and 5′-UTR and its expression is modulated by G4 resolvase 1. Nucleic Acids Res. 2012;40:1033–1049. doi: 10.1093/nar/gkr849. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 169.Kim H.-N., Lee J.-H., Bae S.-C., Ryoo H.-M., Kim H.-H., Ha H., Lee Z.H. Histone deacetylase inhibitor MS-275 stimulates bone formation in part by enhancing Dhx36-mediated TNAP transcription. J. Bone Miner. Res. 2011;26:2161–2173. doi: 10.1002/jbmr.426. [DOI] [PubMed] [Google Scholar]
- 170.Schlag K., Steinhilber D., Karas M., Sorg B.L. Analysis of proximal ALOX5 promoter binding proteins by quantitative proteomics. FEBS J. 2020;287:4481–4499. doi: 10.1111/febs.15259. [DOI] [PubMed] [Google Scholar]
- 171.Schult P., Paeschke K. The DEAH helicase DHX36 and its role in G-quadruplex-dependent processes. Biol. Chem. 2020 doi: 10.1515/hsz-2020-0292. [DOI] [PubMed] [Google Scholar]
- 172.Dhote V., Sweeney T.R., Kim N., Hellen C.U.T., Pestova T.V. Roles of individual domains in the function of DHX29, an essential factor required for translation of structured mammalian mRNAs. Proc. Natl. Acad. Sci. 2012;109:E3150–E3159. doi: 10.1073/pnas.1208014109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173.Tanner N.K., Linder P. DExD/H Box RNA Helicases: From Generic Motors to Specific Dissociation Functions. Mol. Cell. 2001;8:251–262. doi: 10.1016/S1097-2765(01)00329-X. [DOI] [PubMed] [Google Scholar]
- 174.He Y., Andersen G.R., Nielsen K.H. Structural basis for the function of DEAH helicases. EMBO Rep. 2010;11:180–186. doi: 10.1038/embor.2010.11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175.Lattmann S., Giri B., Vaughn J.P., Akman S.A., Nagamine Y. Role of the amino terminal RHAU-specific motif in the recognition and resolution of guanine quadruplex-RNA by the DEAH-box RNA helicase RHAU. Nucleic Acids Res. 2010;38:6219–6233. doi: 10.1093/nar/gkq372. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 176.Chalupníková K., Lattmann S., Selak N., Iwamoto F., Fujiki Y., Nagamine Y. Recruitment of the RNA helicase RHAU to stress granules via a unique RNA-binding domain. J. Biol. Chem. 2008;283:35186–35198. doi: 10.1074/jbc.M804857200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177.Heddi B., Cheong V.V., Martadinata H., Phan A.T. Insights into G-quadruplex specific recognition by the DEAH-box helicase RHAU: Solution structure of a peptide-quadruplex complex. Proc. Natl. Acad. Sci. U. S. A. 2015;112:9608–9613. doi: 10.1073/pnas.1422605112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 178.Heddi B., Cheong V.V., Schmitt E., Mechulam Y., Phan A.T. Recognition of different base tetrads by RHAU (DHX36): X-ray crystal structure of the G4 recognition motif bound to the 3’-end tetrad of a DNA G-quadruplex. J. Struct. Biol. 2020;209 doi: 10.1016/j.jsb.2019.10.001. [DOI] [PubMed] [Google Scholar]
- 179.Hossain K.A., Jurkowski M., Czub J., Kogut M. Mechanism of recognition of parallel G-quadruplexes by DEAH/RHAU helicase DHX36 explored by molecular dynamics simulations. Comput. Struct. Biotechnol. J. 2021;19:2526–2536. doi: 10.1016/j.csbj.2021.04.039. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 180.Chen M.C., Tippana R., Demeshkina N.A., Murat P., Balasubramanian S., Myong S., Ferré-D’Amaré A.R. Structural basis of G-quadruplex unfolding by the DEAH/RHA helicase DHX36. Nature. 2018;558:465–469. doi: 10.1038/s41586-018-0209-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 181.Chen W.-F., Rety S., Guo H.-L., Dai Y.-X., Wu W.-Q., Liu N.-N., Auguin D., Liu Q.-W., Hou X.-M., Dou S.-X., Xi X.-G. Molecular Mechanistic Insights into Drosophila DHX36-Mediated G-Quadruplex Unfolding: A Structure-Based Model. Struct. Lond. Engl. 2018;1993(26):403–415.e4. doi: 10.1016/j.str.2018.01.008. [DOI] [PubMed] [Google Scholar]
- 182.Tippana R., Chen M.C., Demeshkina N.A., Ferré-D’Amaré A.R., Myong S. RNA G-quadruplex is resolved by repetitive and ATP-dependent mechanism of DHX36. Nat. Commun. 2019;10:1855. doi: 10.1038/s41467-019-09802-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 183.Yangyuoru P.M., Bradburn D.A., Liu Z., Xiao T.S., Russell R. The G-quadruplex (G4) resolvase DHX36 efficiently and specifically disrupts DNA G4s via a translocation-based helicase mechanism. J. Biol. Chem. 2018;293:1924–1932. doi: 10.1074/jbc.M117.815076. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 184.He Y., Staley J.P., Andersen G.R., Nielsen K.H. Structure of the DEAH/RHA ATPase Prp43p bound to RNA implicates a pair of hairpins and motif Va in translocation along RNA. RNA. 2017;23:1110–1124. doi: 10.1261/rna.060954.117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 185.Prabu J.R., Müller M., Thomae A.W., Schüssler S., Bonneau F., Becker P.B., Conti E. Structure of the RNA Helicase MLE Reveals the Molecular Mechanisms for Uridine Specificity and RNA-ATP Coupling. Mol. Cell. 2015;60:487–499. doi: 10.1016/j.molcel.2015.10.011. [DOI] [PubMed] [Google Scholar]
- 186.Hamann F., Enders M., Ficner R. Structural basis for RNA translocation by DEAH-box ATPases. Nucleic Acids Res. 2019;47:4349–4362. doi: 10.1093/nar/gkz150. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 187.Srinivasan S., Liu Z., Chuenchor W., Xiao T.S., Jankowsky E. Function of Auxiliary Domains of the DEAH/RHA Helicase DHX36 in RNA Remodeling. J. Mol. Biol. 2020;432:2217–2231. doi: 10.1016/j.jmb.2020.02.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188.Jankowsky E., Putnam A. Duplex unwinding with DEAD-box proteins. Methods Mol. Biol. Clifton NJ. 2010;587:245–264. doi: 10.1007/978-1-60327-355-8_18. [DOI] [PubMed] [Google Scholar]
- 189.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. Transcriptome-wide Identification of RNA-Binding Protein and MicroRNA Target Sites by PAR-CLIP. Cell. 2010;141:129–141. doi: 10.1016/j.cell.2010.03.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 190.Booy E.P., Howard R., Marushchak O., Ariyo E.O., Meier M., Novakowski S.K., Deo S.R., Dzananovic E., Stetefeld J., McKenna S.A. The RNA helicase RHAU (DHX36) suppresses expression of the transcription factor PITX1. Nucleic Acids Res. 2014;42:3346–3361. doi: 10.1093/nar/gkt1340. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 191.Vester K., Eravci M., Serikawa T., Schütze T., Weise C., Kurreck J. RNAi-mediated knockdown of the Rhau helicase preferentially depletes proteins with a Guanine-quadruplex motif in the 5’-UTR of their mRNA. Biochem. Biophys. Res. Commun. 2019;508:756–761. doi: 10.1016/j.bbrc.2018.11.186. [DOI] [PubMed] [Google Scholar]
- 192.Shahid R., Bugaut A., Balasubramanian S. The BCL-2 5′ Untranslated Region Contains an RNA G-Quadruplex-Forming Motif That Modulates Protein Expression. Biochemistry. 2010;49:8300–8306. doi: 10.1021/bi100957h. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193.Kumari S., Bugaut A., Huppert J.L., Balasubramanian S. An RNA G-quadruplex in the 5’ UTR of the NRAS proto-oncogene modulates translation. Nat. Chem. Biol. 2007;3:218–221. doi: 10.1038/nchembio864. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 194.Murat P., Marsico G., Herdy B., Ghanbarian A., Portella G., Balasubramanian S. RNA G-quadruplexes at upstream open reading frames cause DHX36-and DHX9-dependent translation of human mRNAs. Genome Biol. 2018;19:229. doi: 10.1186/s13059-018-1602-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 195.Blackburn E.H., Epel E.S., Lin J. Human telomere biology: A contributory and interactive factor in aging, disease risks, and protection. Science. 2015;350:1193–1198. doi: 10.1126/science.aab3389. [DOI] [PubMed] [Google Scholar]
- 196.de Lange T. How Telomeres Solve the End-Protection Problem. Science. 2009;326:948–952. doi: 10.1126/science.1170633. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 197.Chen J.-L., Blasco M.A., Greider C.W. Secondary Structure of Vertebrate Telomerase RNA. Cell. 2000;100:503–514. doi: 10.1016/S0092-8674(00)80687-X. [DOI] [PubMed] [Google Scholar]
- 198.Webb C.J., Zakian V.A. Telomerase RNA stem terminus element affects template boundary element function, telomere sequence, and shelterin binding. Proc. Natl. Acad. Sci. 2015;112:11312–11317. doi: 10.1073/pnas.1503157112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 199.Gros J., Guédin A., Mergny J.-L., Lacroix L. G-Quadruplex Formation Interferes with P1 Helix Formation in the RNA Component of Telomerase hTERC. ChemBioChem. 2008;9:2075–2079. doi: 10.1002/cbic.200800300. [DOI] [PubMed] [Google Scholar]
- 200.Li X., Nishizuka H., Tsutsumi K., Imai Y., Kurihara Y., Uesugi S. Structure, Interactions and Effects on Activity of the 5′-terminal Region of Human telomerase RNA. J. Biochem. (Tokyo) 2007;141:755–765. doi: 10.1093/jb/mvm081. [DOI] [PubMed] [Google Scholar]
- 201.Booy E.P., Meier M., Okun N., Novakowski S.K., Xiong S., Stetefeld J., McKenna S.A. The RNA helicase RHAU (DHX36) unwinds a G4-quadruplex in human telomerase RNA and promotes the formation of the P1 helix template boundary. Nucleic Acids Res. 2012;40:4110–4124. doi: 10.1093/nar/gkr1306. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 202.Sexton A.N., Collins K. The 5′ Guanosine Tracts of Human Telomerase RNA Are Recognized by the G-Quadruplex Binding Domain of the RNA Helicase DHX36 and Function To Increase RNA Accumulation. Mol. Cell. Biol. 2011;31:736–743. doi: 10.1128/MCB.01033-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 203.Lattmann S., Stadler M.B., Vaughn J.P., Akman S.A., Nagamine Y. The DEAH-box RNA helicase RHAU binds an intramolecular RNA G-quadruplex in TERC and associates with telomerase holoenzyme. Nucleic Acids Res. 2011;39:9390–9404. doi: 10.1093/nar/gkr630. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 204.Cawthon R.M. Telomere length measurement by a novel monochrome multiplex quantitative PCR method. Nucleic Acids Res. 2009;37 doi: 10.1093/nar/gkn1027. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 205.Decorsière A., Cayrel A., Vagner S., Millevoi S. Essential role for the interaction between hnRNP H/F and a G quadruplex in maintaining p53 pre-mRNA 3’-end processing and function during DNA damage. Genes Dev. 2011;25:220–225. doi: 10.1101/gad.607011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 206.Mantovani F., Collavin L., Del Sal G. Mutant p53 as a guardian of the cancer cell. Cell Death Differ. 2019;26:199–212. doi: 10.1038/s41418-018-0246-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 207.Newman M., Sfaxi R., Saha A., Monchaud D., Teulade-Fichou M.-P., Vagner S. The G-Quadruplex-Specific RNA Helicase DHX36 Regulates p53 Pre-mRNA 3′-End Processing Following UV-Induced DNA Damage. J. Mol. Biol. 2017;429:3121–3131. doi: 10.1016/j.jmb.2016.11.033. [DOI] [PubMed] [Google Scholar]
- 208.O’Brien J., Hayder H., Zayed Y., Peng C. Overview of MicroRNA Biogenesis, Mechanisms of Actions, and Circulation. Front. Endocrinol. 2018;9 doi: 10.3389/fendo.2018.00402. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 209.O’Day E., Le M.T.N., Imai S., Tan S.M., Kirchner R., Arthanari H., Hofmann O., Wagner G., Lieberman J. An RNA-binding Protein, Lin28, Recognizes and Remodels G-quartets in the MicroRNAs (miRNAs) and mRNAs It Regulates. J. Biol. Chem. 2015;290:17909–17922. doi: 10.1074/jbc.M115.665521. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 210.Mirihana Arachchilage G., Dassanayake A.C., Basu S. A potassium ion-dependent RNA structural switch regulates human pre-miRNA 92b maturation. Chem. Biol. 2015;22:262–272. doi: 10.1016/j.chembiol.2014.12.013. [DOI] [PubMed] [Google Scholar]
- 211.Pandey S., Agarwala P., Jayaraj G.G., Gargallo R., Maiti S. The RNA Stem-Loop to G-Quadruplex Equilibrium Controls Mature MicroRNA Production inside the Cell. Biochemistry. 2015;54:7067–7078. doi: 10.1021/acs.biochem.5b00574. [DOI] [PubMed] [Google Scholar]
- 212.Liu G., Du W., Xu H., Sun Q., Tang D., Zou S., Zhang Y., Ma M., Zhang G., Du X., Ju S., Cheng W., Tian Y., Fu X. RNA G-quadruplex regulates microRNA-26a biogenesis and function. J. Hepatol. 2020;73:371–382. doi: 10.1016/j.jhep.2020.02.032. [DOI] [PubMed] [Google Scholar]
- 213.Fu X., Meng Z., Liang W., Tian Y., Wang X., Han W., Lou G., Wang X., Lou F., Yen Y., Yu H., Jove R., Huang W. miR-26a enhances miRNA biogenesis by targeting Lin28B and Zcchc11 to suppress tumor growth and metastasis. Oncogene. 2014;33:4296–4306. doi: 10.1038/onc.2013.385. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 214.Kota J., Chivukula R.R., O’Donnell K.A., Wentzel E.A., Montgomery C.L., Hwang H.-W., Chang T.-C., Vivekanandan P., Torbenson M., Clark K.R., Mendell J.R., Mendell J.T. Therapeutic microRNA delivery suppresses tumorigenesis in a murine liver cancer model. Cell. 2009;137:1005–1017. doi: 10.1016/j.cell.2009.04.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 215.Ji J., Shi J., Budhu A., Yu Z., Forgues M., Roessler S., Ambs S., Chen Y., Meltzer P.S., Croce C.M., Qin L.-X., Man K., Lo C.-M., Lee J., Ng I.O.L., Fan J., Tang Z.-Y., Sun H.-C., Wang X.W. MicroRNA expression, survival, and response to interferon in liver cancer. N. Engl. J. Med. 2009;361:1437–1447. doi: 10.1056/NEJMoa0901282. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 216.Thacker U., Pauzaite T., Tollitt J., Twardowska M., Harrison C., Dowle A., Coverley D., Copeland N.A. Identification of DHX9 as a cell cycle regulated nucleolar recruitment factor for CIZ1. Sci. Rep. 2020;10:18103. doi: 10.1038/s41598-020-75160-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 217.Aratani S., Fujii R., Oishi T., Fujita H., Amano T., Ohshima T., Hagiwara M., Fukamizu A., Nakajima T. Dual roles of RNA helicase A in CREB-dependent transcription. Mol. Cell. Biol. 2001;21:4460–4469. doi: 10.1128/MCB.21.14.4460-4469.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 218.Tetsuka T., Uranishi H., Sanda T., Asamitsu K., Yang J.-P., Wong-Staal F., Okamoto T. RNA helicase A interacts with nuclear factor kappaB p65 and functions as a transcriptional coactivator. Eur. J. Biochem. 2004;271:3741–3751. doi: 10.1111/j.1432-1033.2004.04314.x. [DOI] [PubMed] [Google Scholar]
- 219.Hartman T.R., Qian S., Bolinger C., Fernandez S., Schoenberg D.R., Boris-Lawrie K. RNA helicase A is necessary for translation of selected messenger RNAs. Nat. Struct. Mol. Biol. 2006;13:509–516. doi: 10.1038/nsmb1092. [DOI] [PubMed] [Google Scholar]
- 220.Jain A., Bacolla A., Del Mundo I.M., Zhao J., Wang G., Vasquez K.M. DHX9 helicase is involved in preventing genomic instability induced by alternatively structured DNA in human cells. Nucleic Acids Res. 2013;41:10345–10357. doi: 10.1093/nar/gkt804. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 221.Lee T., Pelletier J. The biology of DHX9 and its potential as a therapeutic target. Oncotarget. 2016;7:42716–42739. doi: 10.18632/oncotarget.8446. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 222.Chakraborty P., Grosse F. Human DHX9 helicase preferentially unwinds RNA-containing displacement loops (R-loops) and G-quadruplexes. DNA Repair. 2011;10:654–665. doi: 10.1016/j.dnarep.2011.04.013. [DOI] [PubMed] [Google Scholar]
- 223.Huppertz I., Attig J., D’Ambrogio A., Easton L.E., Sibley C.R., Sugimoto Y., Tajnik M., König J., Ule J. iCLIP: protein-RNA interactions at nucleotide resolution. Methods San Diego Calif. 2014;65:274–287. doi: 10.1016/j.ymeth.2013.10.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 224.Lee C.G., Hurwitz J. A new RNA helicase isolated from HeLa cells that catalytically translocates in the 3’ to 5’ direction. J. Biol. Chem. 1992;267:4398–4407. [PubMed] [Google Scholar]
- 225.Soboleski M.R., Oaks J., Halford W.P. Green fluorescent protein is a quantitative reporter of gene expression in individual eukaryotic cells. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2005;19:440–442. doi: 10.1096/fj.04-3180fje. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 226.Xing Z., Wang S., Tran E.J. Characterization of the mammalian DEAD-box protein DDX5 reveals functional conservation with S. cerevisiae ortholog Dbp2 in transcriptional control and glucose metabolism. RNA N. Y. N. 2017;23:1125–1138. doi: 10.1261/rna.060335.116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 227.Xing Z., Ma W.K., Tran E.J. The DDX5/Dbp2 subfamily of DEAD-box RNA helicases. WIREs RNA. 2019;10 doi: 10.1002/wrna.1519. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 228.Kay B.K., Williamson M.P., Sudol M. The importance of being proline: the interaction of proline-rich motifs in signaling proteins with their cognate domains. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2000;14:231–241. [PubMed] [Google Scholar]
- 229.Cloutier S.C., Ma W.K., Nguyen L.T., Tran E.J. The DEAD-box RNA helicase Dbp2 connects RNA quality control with repression of aberrant transcription. J. Biol. Chem. 2012;287:26155–26166. doi: 10.1074/jbc.M112.383075. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 230.Kahlina K., Goren I., Pfeilschifter J., Frank S. p68 DEAD box RNA helicase expression in keratinocytes. Regulation, nucleolar localization, and functional connection to proliferation and vascular endothelial growth factor gene expression. J. Biol. Chem. 2004;279:44872–44882. doi: 10.1074/jbc.M402467200. [DOI] [PubMed] [Google Scholar]
- 231.Wang H., Gao X., Huang Y., Yang J., Liu Z.-R. P68 RNA helicase is a nucleocytoplasmic shuttling protein. Cell Res. 2009;19:1388–1400. doi: 10.1038/cr.2009.113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 232.Rossow K.L., Janknecht R. Synergism between p68 RNA helicase and the transcriptional coactivators CBP and p300. Oncogene. 2003;22:151–156. doi: 10.1038/sj.onc.1206067. [DOI] [PubMed] [Google Scholar]
- 233.Iggo R.D., Jamieson D.J., MacNeill S.A., Southgate J., McPheat J., Lane D.P. p68 RNA helicase: identification of a nucleolar form and cloning of related genes containing a conserved intron in yeasts. Mol. Cell. Biol. 1991;11:1326–1333. doi: 10.1128/mcb.11.3.1326. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 234.Nicol S.M., Causevic M., Prescott A.R., Fuller-Pace F.V. The nuclear DEAD box RNA helicase p68 interacts with the nucleolar protein fibrillarin and colocalizes specifically in nascent nucleoli during telophase. Exp. Cell Res. 2000;257:272–280. doi: 10.1006/excr.2000.4886. [DOI] [PubMed] [Google Scholar]
- 235.Choi Y.-J., Lee S.-G. The DEAD-box RNA helicase DDX3 interacts with DDX5, co-localizes with it in the cytoplasm during the G2/M phase of the cycle, and affects its shuttling during mRNP export. J. Cell. Biochem. 2012;113:985–996. doi: 10.1002/jcb.23428. [DOI] [PubMed] [Google Scholar]
- 236.Ngo T.D., Partin A.C., Nam Y. RNA Specificity and Autoregulation of DDX17, a Modulator of MicroRNA Biogenesis. Cell Rep. 2019;29:4024-+. doi: 10.1016/j.celrep.2019.11.059. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 237.Clark E.L., Hadjimichael C., Temperley R., Barnard A., Fuller-Pace F.V., Robson C.N. p68/DdX5 Supports beta-Catenin & RNAP II during Androgen Receptor Mediated Transcription in Prostate Cancer. PLoS ONE. 2013;8 doi: 10.1371/journal.pone.0054150. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 238.Zhao X., Li D., Yang F., Lian H., Wang J., Wang X., Fang E., Song H., Hu A., Guo Y., Liu Y., Li H., Chen Y., Huang K., Zheng L., Tong Q. Long Noncoding RNA NHEG1 Drives beta-Catenin Transactivation and Neuroblastoma Progression through Interacting with DDX5. Mol. Ther. 2020;28:946–962. doi: 10.1016/j.ymthe.2019.12.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 239.You Z., Liu C., Wang C., Ling Z., Wang Y., Wang Y., Zhang M., Chen S., Xu B., Guan H., Chen M. LncRNA CCAT1 Promotes Prostate Cancer Cell Proliferation by Interacting with DDX5 and MIR-28-5P. Mol. Cancer Ther. 2019;18:2469–2479. doi: 10.1158/1535-7163.MCT-19-0095. [DOI] [PubMed] [Google Scholar]
- 240.Tedeschi F.A., Cloutier S.C., Tran E.J., Jankowsky E. The DEAD-box protein Dbp2p is linked to noncoding RNAs, the helicase Sen1p, and R-loops. RNA. 2018;24:1693–1705. doi: 10.1261/rna.067249.118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 241.Das M., Renganathan A., Dighe S.N., Bhaduri U., Shettar A., Mukherjee G., Kondaiah P., Rao M.R.S. DDX5/p68 associated lncRNA LOC284454 is differentially expressed in human cancers and modulates gene expression. Rna Biol. 2018;15:214–230. doi: 10.1080/15476286.2017.1397261. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 242.Abbasi N., Long T., Li Y., Yee B.A., Cho B.S., Hernandez J.E., Ma E., Patel P.R., Sahoo D., Sayed I.M., Varki N., Das S., Ghosh P., Yeo G.W., Huang W.J.M. DDX5 promotes oncogene C3 and FABP1 expressions and drives intestinal inflammation and tumorigenesis. Life Sci. Alliance. 2020;3:e202000772. doi: 10.26508/lsa.202000772. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 243.Hashemi V., Masjedi A., Hazhir-karzar B., Tanomand A., Shotorbani S.S., Hojjat-Farsangi M., Ghalamfarsa G., Azizi G., Anvari E., Baradaran B., Jadidi-Niaragh F. The role of DEAD-box RNA helicase p68 (DDX5) in the development and treatment of breast cancer. J. Cell. Physiol. 2019;234:5478–5487. doi: 10.1002/jcp.26912. [DOI] [PubMed] [Google Scholar]
- 244.Nyamao R.M., Wu J., Yu L., Xiao X., Zhang F.-M. Roles of DDX5 in the tumorigenesis, proliferation, differentiation, metastasis and pathway regulation of human malignancies. Biochim. Biophys. Acta-Rev. Cancer. 1871;2019:85–98. doi: 10.1016/j.bbcan.2018.11.003. [DOI] [PubMed] [Google Scholar]
- 245.Z. Quan, B. Zhang, F. Yin, J. Du, Y. Zhi, J. Xu, N. Song, DDX5 Silencing Suppresses the Migration of Basal cell Carcinoma Cells by Downregulating JAK2/STAT3 Pathway, Technol. Cancer Res. Treat. 18 (2019) 1533033819892258. doi: 10.1177/1533033819892258. [DOI] [PMC free article] [PubMed]
- 246.Zhang T., Yang X., Zhou W., Xu W., Wang J., Wu D., Hong Z., Yuan S., Zeng Z., Jia X., Luz S., Safadi R., Han S., Yang Z., Liangpunsakul S., Lu Y. Heat Shock Protein 90 Promotes Rna Helicase Ddx5 Accumulation and Exacerbates Liver Cancer by Inhibiting Autophagy. Hepatology. 2020;72:664A–665A. doi: 10.20892/j.issn.2095-3941.2020.0262. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 247.Sithole N., Williams C.A., Abbink T.E.M., Lever A.M.L. DDX5 potentiates HIV-1 transcription as a co-factor of Tat. Retrovirology. 2020;17:6. doi: 10.1186/s12977-020-00514-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 248.Nelson C.R., Mrozowich T., Park S.M., D’souza S., Henrickson A., Vigar J.R.J., Wieden H.-J., Owens R.J., Demeler B., Patel T.R. Human DDX17 Unwinds Rift Valley Fever Virus Non-Coding RNAs. Int. J. Mol. Sci. 2020;22 doi: 10.3390/ijms22010054. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 249.Zan J., Xu R., Tang X., Lu M., Xie S., Cai J., Huang Z., Zhang J. RNA helicase DDX5 suppresses IFN-I antiviral innate immune response by interacting with PP2A-C beta to deactivate IRF3. Exp. Cell Res. 2020;396 doi: 10.1016/j.yexcr.2020.112332. [DOI] [PubMed] [Google Scholar]
- 250.Jalal C., Uhlmann-Schiffler H., Stahl H. Redundant role of DEAD box proteins p68 (Ddx5) and p72/p82 (Ddx17) in ribosome biogenesis and cell proliferation. Nucleic Acids Res. 2007;35:3590–3601. doi: 10.1093/nar/gkm058. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 251.Dardenne E., Polay Espinoza M., Fattet L., Germann S., Lambert M.-P., Neil H., Zonta E., Mortada H., Gratadou L., Deygas M., Chakrama F.Z., Samaan S., Desmet F.-O., Tranchevent L.-C., Dutertre M., Rimokh R., Bourgeois C.F., Auboeuf D. RNA helicases DDX5 and DDX17 dynamically orchestrate transcription, miRNA, and splicing programs in cell differentiation. Cell Rep. 2014;7:1900–1913. doi: 10.1016/j.celrep.2014.05.010. [DOI] [PubMed] [Google Scholar]
- 252.Herviou P., Le Bras M., Dumas L., Hieblot C., Gilhodes J., Cioci G., Hugnot J.-P., Ameadan A., Guillonneau F., Dassi E., Cammas A., S. Millevoi, hnRNP H/F drive RNA G-quadruplex-mediated translation linked to genomic instability and therapy resistance in glioblastoma. Nat. Commun. 2020;11:2661. doi: 10.1038/s41467-020-16168-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 253.Wheelhouse R.T., Sun D., Han H., Han F.X., Hurley L.H. Cationic Porphyrins as Telomerase Inhibitors: the Interaction of Tetra-(N-methyl-4-pyridyl)porphine with Quadruplex DNA. J. Am. Chem. Soc. 1998;120:3261–3262. doi: 10.1021/ja973792e. [DOI] [Google Scholar]
- 254.Gullberg M., Gústafsdóttir S.M., Schallmeiner E., Jarvius J., Bjarnegård M., Betsholtz C., Landegren U., Fredriksson S. Cytokine detection by antibody-based proximity ligation. Proc. Natl. Acad. Sci. U. S. A. 2004;101:8420–8424. doi: 10.1073/pnas.0400552101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 255.Gao J., Byrd A.K., Zybailov B.L., Marecki J.C., Guderyon M.J., Edwards A.D., Chib S., West K.L., Waldrip Z.J., Mackintosh S.G., Gao Z., Putnam A.A., Jankowsky E., Raney K.D. DEAD-box RNA helicases Dbp2, Ded1 and Mss116 bind to G-quadruplex nucleic acids and destabilize G-quadruplex RNA. Chem. Commun. Camb. Engl. 2019;55:4467–4470. doi: 10.1039/c8cc10091h. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 256.Iost I., Dreyfus M., Linder P. Ded1p, a DEAD-box protein required for translation initiation in Saccharomyces cerevisiae, is an RNA helicase. J. Biol. Chem. 1999;274:17677–17683. doi: 10.1074/jbc.274.25.17677. [DOI] [PubMed] [Google Scholar]
- 257.Sen N.D., Gupta N., Archer S.K., Preiss T., Lorsch J.R., Hinnebusch A.G. Functional interplay between DEAD-box RNA helicases Ded1 and Dbp1 in preinitiation complex attachment and scanning on structured mRNAs in vivo. Nucleic Acids Res. 2019;47:8785–8806. doi: 10.1093/nar/gkz595. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 258.Markov D.A., Wojtas I.D., Tessitore K., Henderson S., McAllister W.T. Yeast DEAD box protein Mss116p is a transcription elongation factor that modulates the activity of mitochondrial RNA polymerase. Mol. Cell. Biol. 2014;34:2360–2369. doi: 10.1128/MCB.00160-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 259.Byrd A.K., Raney K.D. A Parallel Quadruplex DNA Is Bound Tightly but Unfolded Slowly by Pif1 Helicase*. J. Biol. Chem. 2015;290:6482–6494. doi: 10.1074/jbc.M114.630749. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 260.Zhang H., Xing Z., Mani S.K.K., Bancel B., Durantel D., Zoulim F., Tran E.J., Merle P., Andrisani O. RNA helicase DEAD box protein 5 regulates Polycomb repressive complex 2/Hox transcript antisense intergenic RNA function in hepatitis B virus infection and hepatocarcinogenesis. Hepatology. 2016;64:1033–1048. doi: 10.1002/hep.28698. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 261.Flores-Rozas H., Hurwitz J. Characterization of a new RNA helicase from nuclear extracts of HeLa cells which translocates in the 5’ to 3’ direction. J. Biol. Chem. 1993;268:21372–21383. [PubMed] [Google Scholar]
- 262.Valdez B.C., Henning D., Perumal K., Busch H. RNA-unwinding and RNA-folding activities of RNA helicase II/Gu–two activities in separate domains of the same protein. Eur. J. Biochem. 1997;250:800–807. doi: 10.1111/j.1432-1033.1997.00800.x. [DOI] [PubMed] [Google Scholar]
- 263.Ou Y., Fritzler M.J., Valdez B.C., Rattner J.B. Mapping and characterization of the functional domains of the nucleolar protein RNA helicase II/Gu. Exp. Cell Res. 1999;247:389–398. doi: 10.1006/excr.1998.4365. [DOI] [PubMed] [Google Scholar]
- 264.Valdez B.C. Structural domains involved in the RNA folding activity of RNA helicase II/Gu protein. Eur. J. Biochem. 2000;267:6395–6402. doi: 10.1046/j.1432-1327.2000.01727.x. [DOI] [PubMed] [Google Scholar]
- 265.Bizebard T., Dreyfus M. In: RNA Remodel. Proteins Methods Protoc. Boudvillain M., editor. Springer; New York, NY: 2015. A FRET-Based, Continuous Assay for the Helicase Activity of DEAD-Box Proteins; pp. 199–209. [DOI] [PubMed] [Google Scholar]
- 266.Hammond J.A., Zhou L., Lamichhane R., Chu H.-Y., Millar D.P., Gerace L., Williamson J.R. A Survey of DDX21 Activity During Rev/RRE Complex Formation. J. Mol. Biol. 2018;430:537–553. doi: 10.1016/j.jmb.2017.06.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 267.Kim D.-S., Camacho C.V., Nagari A., Malladi V.S., Challa S., Kraus W.L. Activation of PARP-1 by snoRNAs Controls Ribosome Biogenesis and Cell Growth via the RNA Helicase DDX21. Mol. Cell. 2019;75:1270-+. doi: 10.1016/j.molcel.2019.06.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 268.Henning D., So R.B., Jin R., Lau L.F., Valdez B.C. Silencing of RNA helicase II/Gualpha inhibits mammalian ribosomal RNA production. J. Biol. Chem. 2003;278:52307–52314. doi: 10.1074/jbc.M310846200. [DOI] [PubMed] [Google Scholar]
- 269.Yang H., Zhou J., Ochs R.L., Henning D., Jin R., Valdez B.C. Down-regulation of RNA helicase II/Gu results in the depletion of 18 and 28 S rRNAs in Xenopus oocyte. J. Biol. Chem. 2003;278:38847–38859. doi: 10.1074/jbc.M302258200. [DOI] [PubMed] [Google Scholar]
- 270.Calo E., Flynn R.A., Martin L., Spitale R.C., Chang H.Y., Wysocka J. RNA helicase DDX21 coordinates transcription and ribosomal RNA processing. Nature. 2015;518:249–253. doi: 10.1038/nature13923. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 271.Sloan K.E., Leisegang M.S., Doebele C., Ramírez A.S., Simm S., Safferthal C., Kretschmer J., Schorge T., Markoutsa S., Haag S., Karas M., Ebersberger I., Schleiff E., Watkins N.J., Bohnsack M.T. The association of late-acting snoRNPs with human pre-ribosomal complexes requires the RNA helicase DDX21. Nucleic Acids Res. 2015;43:553–564. doi: 10.1093/nar/gku1291. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 272.Zhang Z., Kim T., Bao M., Facchinetti V., Jung S.Y., Ghaffari A.A., Qin J., Cheng G., Liu Y.-J. DDX1, DDX21, and DHX36 Helicases Form a Complex with the Adaptor Molecule TRIF to Sense dsRNA in Dendritic Cells. Immunity. 2011;34:866–878. doi: 10.1016/j.immuni.2011.03.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 273.Chen G., Liu C.-H., Zhou L., Krug R.M. Cellular DDX21 RNA Helicase Inhibits Influenza A Virus Replication but Is Counteracted by the Viral NS1 Protein. Cell Host Microbe. 2014;15:484–493. doi: 10.1016/j.chom.2014.03.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 274.S. Naji, G. Ambrus, P. Cimermančič, J.R. Reyes, J.R. Johnson, R. Filbrandt, M.D. Huber, P. Vesely, N.J. Krogan, J.R. Yates, A.C. Saphire, L. Gerace, Host Cell Interactome of HIV-1 Rev Includes RNA Helicases Involved in Multiple Facets of Virus Production*, Mol. Cell. Proteomics. 11 (2012) M111.015313. doi: 10.1074/mcp.M111.015313. [DOI] [PMC free article] [PubMed]
- 275.H. Hao, T. Han, B. Xuan, Y. Sun, S. Tang, N. Yue, Z. Qian, Dissecting the Role of DDX21 in Regulating Human Cytomegalovirus Replication, J. Virol. 93 (2019) e01222-19. doi: 10.1128/JVI.01222-19. [DOI] [PMC free article] [PubMed]
- 276.Dong Y., Ye W., Yang J., Han P., Wang Y., Ye C., Weng D., Zhang F., Xu Z., Lei Y. DDX21 translocates from nucleus to cytoplasm and stimulates the innate immune response due to dengue virus infection. Biochem. Biophys. Res. Commun. 2016;473:648–653. doi: 10.1016/j.bbrc.2016.03.120. [DOI] [PubMed] [Google Scholar]
- 277.Mojzesz M., Klak K., Wojtal P., Adamek M., Podlasz P., Chmielewska-Krzesinska M., Matras M., Reichert M., Chadzinska M., Rakus K. Viral infection-induced changes in the expression profile of non-RLR DExD/H-box RNA helicases (DDX1, DDX3, DHX9, DDX21 and DHX36) in zebrafish and common carp. Fish Shellfish Immunol. 2020;104:62–73. doi: 10.1016/j.fsi.2020.06.010. [DOI] [PubMed] [Google Scholar]
- 278.Zhang H., Zhang Y., Chen C., Zhu X., Zhang C., Xia Y., Zhao Y., Andrisani O.M., Kong L. A double-negative feedback loop between DEAD-box protein DDX21 and Snail regulates epithelial-mesenchymal transition and metastasis in breast cancer. Cancer Lett. 2018;437:67–78. doi: 10.1016/j.canlet.2018.08.021. [DOI] [PubMed] [Google Scholar]
- 279.Wang K., Yu X., Tao B., Qu J. Downregulation of lncRNA HCP5 has inhibitory effects on gastric cancer cells by regulating DDX21 expression. Cytotechnology. 2021;73:1–11. doi: 10.1007/s10616-020-00429-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 280.Tanaka A., Wang J.Y., Shia J., Zhou Y., Ogawa M., Hendrickson R.C., Klimstra D.S., Roehrl M.H. DEAD-box RNA helicase protein DDX21 as a prognosis marker for early stage colorectal cancer with microsatellite instability. Sci. Rep. 2020;10:22085. doi: 10.1038/s41598-020-79049-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 281.Wang X., Wu Z., Qin W., Sun T., Lu S., Li Y., Wang Y., Hu X., Xu D., Wu Y., Chen Q., Yao W., Liu M., Wei M., Wu H. Long non-coding RNA ZFAS1 promotes colorectal cancer tumorigenesis and development through DDX21-POLR1B regulatory axis. Aging-Us. 2020;12:22656–22687. doi: 10.18632/aging.103875. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 282.Jung Y., Lee S., Choi H.-S., Kim S.-N., Lee E., Shin Y., Seo J., Kim B., Jung Y., Kim W.K., Chun H.-K., Lee W.Y., Kim J. Clinical Validation of Colorectal Cancer Biomarkers Identified from Bioinformatics Analysis of Public Expression Data. Clin. Cancer Res. 2011;17:700–709. doi: 10.1158/1078-0432.CCR-10-1300. [DOI] [PubMed] [Google Scholar]
- 283.Zhang Y., Baysac K.C., Yee L.-F., Saporita A.J., Weber J.D. Elevated DDX21 regulates c-Jun activity and rRNA processing in human breast cancers. Breast Cancer Res. 2014;16:449. doi: 10.1186/s13058-014-0449-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 284.Kim S., Kang N., Park S.H., Wells J., Hwang T., Ryu E., Kim B., Hwang S., Kim S., Kang S., Lee S., Stirling P., Myung K., Lee K. ATAD5 restricts R-loop formation through PCNA unloading and RNA helicase maintenance at the replication fork. Nucleic Acids Res. 2020;48:7218–7238. doi: 10.1093/nar/gkaa501. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 285.Argaud D., Boulanger M.-C., Chignon A., Mkannez G., Mathieu P. Enhancer-mediated enrichment of interacting JMJD3-DDX21 to ENPP2 locus prevents R-loop formation and promotes transcription. Nucleic Acids Res. 2019;47:8424–8438. doi: 10.1093/nar/gkz560. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 286.Santoriello C., Sporrij A., Yang S., Flynn R.A., Henriques T., Dorjsuren B., Greig E.C., McCall W., Stanhope M.E., Fazio M., Superdock M., Lichtig A., Adatto I., Abraham B.J., Kalocsay M., Jurynec M., Zhou Y., Adelman K., Calo E., Zon L. RNA helicase DDX21 mediates nucleotide stress responses in neural crest and melanoma cells. Nat. Cell Biol. 2020;22 doi: 10.1038/s41556-020-0493-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 287.McRae E.K.S., Booy E.P., Moya-Torres A., Ezzati P., Stetefeld J., McKenna S.A. Human DDX21 binds and unwinds RNA guanine quadruplexes. Nucleic Acids Res. 2017;45:6656–6668. doi: 10.1093/nar/gkx380. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 288.Asmari M., Ratih R., Alhazmi H.A., El Deeb S. Thermophoresis for characterizing biomolecular interaction. Methods San Diego Calif. 2018;146:107–119. doi: 10.1016/j.ymeth.2018.02.003. [DOI] [PubMed] [Google Scholar]
- 289.Mueller A.M., Breitsprecher D., Duhr S., Baaske P., Schubert T., Längst G. MicroScale Thermophoresis: A Rapid and Precise Method to Quantify Protein-Nucleic Acid Interactions in Solution. Methods Mol. Biol. Clifton NJ. 2017;1654:151–164. doi: 10.1007/978-1-4939-7231-9_10. [DOI] [PubMed] [Google Scholar]
- 290.Archer W.R., Schulz M.D. Isothermal titration calorimetry: practical approaches and current applications in soft matter. Soft Matter. 2020;16:8760–8774. doi: 10.1039/d0sm01345e. [DOI] [PubMed] [Google Scholar]
- 291.Martadinata H., Phan A.T. Structure of propeller-type parallel-stranded RNA G-quadruplexes, formed by human telomeric RNA sequences in K+ solution. J. Am. Chem. Soc. 2009;131:2570–2578. doi: 10.1021/ja806592z. [DOI] [PubMed] [Google Scholar]
- 292.McRae E.K.S., Davidson D.E., Dupas S.J., McKenna S.A. Insights into the RNA quadruplex binding specificity of DDX21. Biochim. Biophys. Acta BBA - Gen. Subj. 1862;2018:1973–1979. doi: 10.1016/j.bbagen.2018.06.009. [DOI] [PubMed] [Google Scholar]
- 293.K. Struhl, Ribonucleases, Curr. Protoc. Mol. Biol. Chapter 3 (2001) Unit3.13. 10.1002/0471142727.mb0313s08. [DOI] [PubMed]
- 294.Marcaida M.J., Kauzlaric A., Duperrex A., Sülzle J., Moncrieffe M.C., Adebajo D., Manley S., Trono D., Dal Peraro M. The Human RNA Helicase DDX21 Presents a Dimerization Interface Necessary for Helicase Activity. IScience. 2020;23 doi: 10.1016/j.isci.2020.101811. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 295.Tseng H.-Y., Chen L.H., Ye Y., Tay K.H., Jiang C.C., Guo S.T., Jin L., Hersey P., Zhang X.D. The melanoma-associated antigen MAGE-D2 suppresses TRAIL receptor 2 and protects against TRAIL-induced apoptosis in human melanoma cells. Carcinogenesis. 2012;33:1871–1881. doi: 10.1093/carcin/bgs236. [DOI] [PubMed] [Google Scholar]
- 296.Strekalova E., Malin D., Good D.M., Cryns V.L. Methionine Deprivation Induces a Targetable Vulnerability in Triple-Negative Breast Cancer Cells by Enhancing TRAIL Receptor-2 Expression. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2015;21:2780–2791. doi: 10.1158/1078-0432.CCR-14-2792. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 297.McRae E.K.S., Dupas S.J., Booy E.P., Piragasam R.S., Fahlman R.P., McKenna S.A. An RNA guanine quadruplex regulated pathway to TRAIL-sensitization by DDX21. RNA N. Y. N. 2020;26:44–57. doi: 10.1261/rna.072199.119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 298.Linder P., Jankowsky E. From unwinding to clamping — the DEAD box RNA helicase family. Nat. Rev. Mol. Cell Biol. 2011;12:505–516. doi: 10.1038/nrm3154. [DOI] [PubMed] [Google Scholar]
- 299.Shatsky I.N., Dmitriev S.E., Andreev D.E., Terenin I.M. Transcriptome-wide studies uncover the diversity of modes of mRNA recruitment to eukaryotic ribosomes. Crit. Rev. Biochem. Mol. Biol. 2014;49:164–177. doi: 10.3109/10409238.2014.887051. [DOI] [PubMed] [Google Scholar]
- 300.Parsyan A., Svitkin Y., Shahbazian D., Gkogkas C., Lasko P., Merrick W.C., Sonenberg N. mRNA helicases: the tacticians of translational control. Nat. Rev. Mol. Cell Biol. 2011;12:235–245. doi: 10.1038/nrm3083. [DOI] [PubMed] [Google Scholar]
- 301.Ray B.K., Lawson T.G., Kramer J.C., Cladaras M.H., Grifo J.A., Abramson R.D., Merrick W.C., Thach R.E. ATP-dependent unwinding of messenger RNA structure by eukaryotic initiation factors. J. Biol. Chem. 1985;260:7651–7658. [PubMed] [Google Scholar]
- 302.Raza F., Waldron J.A., Quesne J.L. Translational dysregulation in cancer: eIF4A isoforms and sequence determinants of eIF4A dependence. Biochem. Soc. Trans. 2015;43:1227–1233. doi: 10.1042/BST20150163. [DOI] [PubMed] [Google Scholar]
- 303.Marintchev A., Edmonds K.A., Marintcheva B., Hendrickson E., Oberer M., Suzuki C., Herdy B., Sonenberg N., Wagner G. Topology and regulation of the human eIF4A/4G/4H helicase complex in translation initiation. Cell. 2009;136:447–460. doi: 10.1016/j.cell.2009.01.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 304.Rozovsky N., Butterworth A.C., Moore M.J. Interactions between eIF4AI and its accessory factors eIF4B and eIF4H. RNA N. Y. N. 2008;14:2136–2148. doi: 10.1261/rna.1049608. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 305.Svitkin Y.V., Pause A., Haghighat A., Pyronnet S., Witherell G., Belsham G.J., Sonenberg N. The requirement for eukaryotic initiation factor 4A (elF4A) in translation is in direct proportion to the degree of mRNA 5’ secondary structure. RNA. 2001;7:382–394. doi: 10.1017/s135583820100108x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 306.Lu W.-T., Wilczynska A., Smith E., Bushell M. The diverse roles of the eIF4A family: you are the company you keep. Biochem. Soc. Trans. 2014;42:166–172. doi: 10.1042/BST20130161. [DOI] [PubMed] [Google Scholar]
- 307.Galicia-Vázquez G., Cencic R., Robert F., Agenor A.Q., Pelletier J. A cellular response linking eIF4AI activity to eIF4AII transcription. RNA N. Y. N. 2012;18:1373–1384. doi: 10.1261/rna.033209.112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 308.Lyu S., Lu J., Chen W., Huang W., Huang H., Xi S., Yan S. High expression of eIF4A2 is associated with a poor prognosis in esophageal squamous cell carcinoma. Oncol. Lett. 2020;20:1. doi: 10.3892/ol.2020.12038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 309.Chan K., Robert F., Oertlin C., Kapeller-Libermann D., Avizonis D., Gutierrez J., Handly-Santana A., Doubrovin M., Park J., Schoepfer C., Da Silva B., Yao M., Gorton F., Shi J., Thomas C.J., Brown L.E., Porco J.A., Pollak M., Larsson O., Pelletier J., Chio I.I.C. eIF4A supports an oncogenic translation program in pancreatic ductal adenocarcinoma. Nat. Commun. 2019;10:5151. doi: 10.1038/s41467-019-13086-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 310.Müller D., Shin S., de Rugy T.G., Samain R., Baer R., Strehaiano M., Masvidal-Sanz L., Guillermet-Guibert J., Jean C., Tsukumo Y., Sonenberg N., Marion F., Guilbaud N., Hoffmann J.-S., Larsson O., Bousquet C., Pyronnet S., Martineau Y. eIF4A inhibition circumvents uncontrolled DNA replication mediated by 4E-BP1 loss in pancreatic cancer. JCI Insight. 2019;4 doi: 10.1172/jci.insight.121951. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 311.Sridharan S., Robeson M., Bastihalli-Tukaramrao D., Howard C.M., Subramaniyan B., Tilley A.M.C., Tiwari A.K., Raman D. Targeting of the Eukaryotic Translation Initiation Factor 4A Against Breast Cancer Stemness. Front. Oncol. 2019;9 doi: 10.3389/fonc.2019.01311. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 312.Modelska A., Turro E., Russell R., Beaton J., Sbarrato T., Spriggs K., Miller J., Gräf S., Provenzano E., Blows F., Pharoah P., Caldas C., Le Quesne J. The malignant phenotype in breast cancer is driven by eIF4A1-mediated changes in the translational landscape. Cell Death Dis. 2015;6 doi: 10.1038/cddis.2014.542. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 313.Gerson-Gurwitz A., Goel V.K., Young N.P., Eam B., Stumpf C.R., Barrera M., Sung E., Staunton J., Chen J., Fish S., Chiang G.G., Thompson P.A. Dissection of cancer therapy combinations in RTK driven tumors using zotatifin (eFT226), a potent and selective eIF4A inhibitor. Eur. J. Cancer. 2020;138:S56. [Google Scholar]
- 314.Mohan P., Kim Y.M. Inhibiting eIF4A in liposarcoma to identify key regulators using ribosome profiling. Eur. J. Cancer. 2020;138:S29. [Google Scholar]
- 315.Singh K., Lin J., Lecomte N., Mohan P., Gokce A., Sanghvi V., Olivera G.-H., Jiang M., Burcul A., Stark S., Viale A., Romesser P.B., Destanchia E., Bagni R., Ratsch G., Leach S.D., Ouyang Z., Wendel H.-G. KRAS and RAS signaling network is co-regulated and can be therapeutically blocked by targeting eIF4A dependent translation program. Mol. Cancer Res. 2020;18:73–74. [Google Scholar]
- 316.Basak T., Dey A.K., Banerjee R., Paul S., Maiti T.K., Ain R. Sequestration of eIF4A by Angiomotin: A novel mechanism to restrict global protein synthesis in trophoblast cells. Stem Cells. 2021;39:210–226. doi: 10.1002/stem.3305. [DOI] [PubMed] [Google Scholar]
- 317.Montero H., Pérez-Gil G., Sampieri C.L. Eukaryotic initiation factor 4A (eIF4A) during viral infections. Virus Genes. 2019;55:267–273. doi: 10.1007/s11262-019-01641-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 318.Ferraiuolo M.A., Lee C.-S., Ler L.W., Hsu J.L., Costa-Mattioli M., Luo M.-J., Reed R., Sonenberg N. A nuclear translation-like factor eIF4AIII is recruited to the mRNA during splicing and functions in nonsense-mediated decay. Proc. Natl. Acad. Sci. U. S. A. 2004;101:4118–4123. doi: 10.1073/pnas.0400933101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 319.Ballut L., Marchadier B., Baguet A., Tomasetto C., Séraphin B., Le Hir H. The exon junction core complex is locked onto RNA by inhibition of eIF4AIII ATPase activity. Nat. Struct. Mol. Biol. 2005;12:861–869. doi: 10.1038/nsmb990. [DOI] [PubMed] [Google Scholar]
- 320.Chan C.C., Dostie J., Diem M.D., Feng W., Mann M., Rappsilber J., Dreyfuss G. eIF4A3 is a novel component of the exon junction complex. RNA. 2004;10:200–209. doi: 10.1261/rna.5230104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 321.Wolfe A.L., Singh K., Zhong Y., Drewe P., Rajasekhar V.K., Sanghvi V.R., Mavrakis K.J., Jiang M., Roderick J.E., Van der Meulen J., Schatz J.H., Rodrigo C.M., Zhao C., Rondou P., de Stanchina E., Teruya-Feldstein J., Kelliher M.A., Speleman F., Porco J.A., Pelletier J., Rätsch G., Wendel H.-G. RNA G-quadruplexes cause eIF4A-dependent oncogene translation in cancer. Nature. 2014;513:65–70. doi: 10.1038/nature13485. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 322.Rubio C.A., Weisburd B., Holderfield M., Arias C., Fang E., DeRisi J.L., Fanidi A. Transcriptome-wide characterization of the eIF4A signature highlights plasticity in translation regulation. Genome Biol. 2014;15:476. doi: 10.1186/s13059-014-0476-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 323.Sadlish H., Galicia-Vazquez G., Paris C.G., Aust T., Bhullar B., Chang L., Helliwell S.B., Hoepfner D., Knapp B., Riedl R., Roggo S., Schuierer S., Studer C., Porco J.A., Pelletier J., Movva N.R. Evidence for a functionally relevant rocaglamide binding site on the eIF4A-RNA complex. ACS Chem. Biol. 2013;8:1519–1527. doi: 10.1021/cb400158t. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 324.Waldron J.A., Raza F., Le Quesne J. eIF4A alleviates the translational repression mediated by classical secondary structures more than by G-quadruplexes. Nucleic Acids Res. 2018;46:3075–3087. doi: 10.1093/nar/gky108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 325.Waldron J.A., Tack D.C., Ritchey L.E., Gillen S.L., Wilczynska A., Turro E., Bevilacqua P.C., Assmann S.M., Bushell M., Le Quesne J. mRNA structural elements immediately upstream of the start codon dictate dependence upon eIF4A helicase activity. Genome Biol. 2019;20:300. doi: 10.1186/s13059-019-1901-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 326.Bordeleau M.-E., Mori A., Oberer M., Lindqvist L., Chard L.S., Higa T., Belsham G.J., Wagner G., Tanaka J., Pelletier J. Functional characterization of IRESes by an inhibitor of the RNA helicase eIF4A. Nat. Chem. Biol. 2006;2:213–220. doi: 10.1038/nchembio776. [DOI] [PubMed] [Google Scholar]
- 327.Weldon C., Behm-Ansmant I., Hurley L.H., Burley G.A., Branlant C., Eperon I.C., Dominguez C. Identification of G-quadruplexes in long functional RNAs using 7-deazaguanine RNA. Nat. Chem. Biol. 2017;13:18–20. doi: 10.1038/nchembio.2228. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 328.Ritchey L.E., Su Z., Tang Y., Tack D.C., Assmann S.M., Bevilacqua P.C. Structure-seq2: sensitive and accurate genome-wide profiling of RNA structure in vivo. Nucleic Acids Res. 2017;45 doi: 10.1093/nar/gkx533. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 329.J.-M. Garant, M.J. Luce, M.S. Scott, J.-P. Perreault, G4RNA: an RNA G-quadruplex database, Database J. Biol. Databases Curation. 2015 (2015) bav059. doi:10.1093/database/bav059. [DOI] [PMC free article] [PubMed]
- 330.Díaz-López I., Toribio R., Berlanga J.J., Ventoso I. An mRNA-binding channel in the ES6S region of the translation 48S-PIC promotes RNA unwinding and scanning. ELife. 2019;8 doi: 10.7554/eLife.48246. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 331.Dhapola P., Chowdhury S. QuadBase2: web server for multiplexed guanine quadruplex mining and visualization. Nucleic Acids Res. 2016;44:W277–W283. doi: 10.1093/nar/gkw425. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 332.Thompson P.A., Eam B., Young N.P., Fish S., Chen J., Barrera M., Howard H., Sung E., Parra A., Staunton J., Chiang G.G., Gerson-Gurwitz A., Wegerski C.J., Nevarez A., Clarine J., Sperry S., Xiang A., Nilewski C., Packard G.K., Michels T., Tran C., Sprengeler P.A., Ernst J.T., Reich S.H., Webster K.R. Targeting Oncogene mRNA Translation in B-Cell Malignancies with eFT226, a Potent and Selective Inhibitor of eIF4A. Mol. Cancer Ther. 2021;20:26–36. doi: 10.1158/1535-7163.MCT-19-0973. [DOI] [PubMed] [Google Scholar]
- 333.Hashimoto S., Furukawa S., Hashimoto A., Tsutaho A., Fukao A., Sakamura Y., Parajuli G., Onodera Y., Otsuka Y., Handa H., Oikawa T., Hata S., Nishikawa Y., Mizukami Y., Kodama Y., Murakami M., Fujiwara T., Hirano S., Sabe H. ARF6 and AMAP1 are major targets of KRAS and TP53 mutations to promote invasion, PD-L1 dynamics, and immune evasion of pancreatic cancer. Proc. Natl. Acad. Sci. 2019;116:17450–17459. doi: 10.1073/pnas.1901765116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 334.Pazo A., Pérez-González A., Oliveros J.C., Huarte M., Chavez J.P., Nieto A. hCLE/RTRAF-HSPC117-DDX1-FAM98B: A New Cap-Binding Complex That Activates mRNA Translation. Front. Physiol. 2019;10:92. doi: 10.3389/fphys.2019.00092. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 335.Zhong W., Li Z., Zhou M., Xu T., Wang Y. DDX1 regulates alternative splicing and insulin secretion in pancreatic β cells. Biochem. Biophys. Res. Commun. 2018;500:751–757. doi: 10.1016/j.bbrc.2018.04.147. [DOI] [PubMed] [Google Scholar]
- 336.Li Z., Zhou M., Cai Z., Liu H., Zhong W., Hao Q., Cheng D., Hu X., Hou J., Xu P., Xue Y., Zhou Y., Xu T. RNA-binding protein DDX1 is responsible for fatty acid-mediated repression of insulin translation. Nucleic Acids Res. 2018;46:12052–12066. doi: 10.1093/nar/gky867. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 337.Li L., Poon H.-Y., Hildebrandt M.R., Monckton E.A., Germain D.R., Fahlman R.P., Godbout R. Role for RIF1-interacting partner DDX1 in BLM recruitment to DNA double-strand breaks. DNA Repair. 2017;55:47–63. doi: 10.1016/j.dnarep.2017.05.001. [DOI] [PubMed] [Google Scholar]
- 338.Han C., Liu Y., Wan G., Choi H.J., Zhao L., Ivan C., He X., Sood A.K., Zhang X., Lu X. The RNA-binding protein DDX1 promotes primary microRNA maturation and inhibits ovarian tumor progression. Cell Rep. 2014;8:1447–1460. doi: 10.1016/j.celrep.2014.07.058. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 339.Hildebrandt M.R., Wang Y., Li L., Yasmin L., Glubrecht D.D., Godbout R. Cytoplasmic aggregation of DDX1 in developing embryos: Early embryonic lethality associated with Ddx1 knockout. Dev. Biol. 2019;455:420–433. doi: 10.1016/j.ydbio.2019.07.014. [DOI] [PubMed] [Google Scholar]
- 340.Germain D.R., Li L., Hildebrandt M.R., Simmonds A.J., Hughes S.C., Godbout R. Loss of the Drosophila melanogaster DEAD box protein Ddx1 leads to reduced size and aberrant gametogenesis. Dev. Biol. 2015;407:232–245. doi: 10.1016/j.ydbio.2015.09.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 341.Taschuk F., Cherry S. DEAD-Box Helicases: Sensors, Regulators, and Effectors for Antiviral Defense. Viruses. 2020;12 doi: 10.3390/v12020181. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 342.Tanaka K., Ikeda N., Miyashita K., Nuriya H., Hara T. DEAD box protein DDX1 promotes colorectal tumorigenesis through transcriptional activation of the LGR5 gene. Cancer Sci. 2018;109:2479–2489. doi: 10.1111/cas.13661. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 343.Akter K.A., Mansour M.A., Hyodo T., Senga T. FAM98A associates with DDX1-C14orf166-FAM98B in a novel complex involved in colorectal cancer progression. Int. J. Biochem. Cell Biol. 2017;84:1–13. doi: 10.1016/j.biocel.2016.12.013. [DOI] [PubMed] [Google Scholar]
- 344.de Almeida C.R., Dhir S., Dhir A., Moghaddam A.E., Sattentau Q., Meinhart A., Proudfoot N.J. RNA Helicase DDX1 Converts RNA G-Quadruplex Structures into R-Loops to Promote IgH Class Switch Recombination. Mol. Cell. 2018;70:650-+. doi: 10.1016/j.molcel.2018.04.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 345.Pennell C.A., Arnold L.W., Lutz P.M., LoCascio N.J., Willoughby P.B., Haughton G. Cross-reactive idiotypes and common antigen binding specificities expressed by a series of murine B-cell lymphomas: etiological implications. Proc. Natl. Acad. Sci. 1985;82:3799–3803. doi: 10.1073/pnas.82.11.3799. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 346.Gregersen L.H., Schueler M., Munschauer M., Mastrobuoni G., Chen W., Kempa S., Dieterich C., Landthaler M. MOV10 Is a 5’ to 3’ RNA helicase contributing to UPF1 mRNA target degradation by translocation along 3’ UTRs. Mol. Cell. 2014;54:573–585. doi: 10.1016/j.molcel.2014.03.017. [DOI] [PubMed] [Google Scholar]
- 347.Goodier J.L., Cheung L.E., Kazazian H.H. MOV10 RNA Helicase Is a Potent Inhibitor of Retrotransposition in Cells. PLoS Genet. 2012;8 doi: 10.1371/journal.pgen.1002941. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 348.Li X., Zhang J., Jia R., Cheng V., Xu X., Qiao W., Guo F., Liang C., Cen S. The MOV10 Helicase Inhibits LINE-1 Mobility. J. Biol. Chem. 2013;288:21148–21160. doi: 10.1074/jbc.M113.465856. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 349.Zheng K., Wang P.J. Blockade of Pachytene piRNA Biogenesis Reveals a Novel Requirement for Maintaining Post-Meiotic Germline Genome Integrity. PLOS Genet. 2012;8 doi: 10.1371/journal.pgen.1003038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 350.Fu Q., Pandey R.R., Leu N.A., Pillai R.S., Wang P.J. Mutations in the MOV10L1 ATP Hydrolysis Motif Cause piRNA Biogenesis Failure and Male Sterility in Mice1. Biol. Reprod. 2016;95 doi: 10.1095/biolreprod.116.142430. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 351.Vourekas A., Zheng K., Fu Q., Maragkakis M., Alexiou P., Ma J., Pillai R.S., Mourelatos Z., Wang P.J. The RNA helicase MOV10L1 binds piRNA precursors to initiate piRNA processing. Genes Dev. 2015;29:617–629. doi: 10.1101/gad.254631.114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 352.Kenny P.J., Zhou H., Kim M., Skariah G., Khetani R.S., Drnevich J., Arcila M.L., Kosik K.S., Ceman S. MOV10 and FMRP Regulate AGO2 Association with MicroRNA Recognition Elements. Cell Rep. 2014;9:1729–1741. doi: 10.1016/j.celrep.2014.10.054. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 353.Zhang X., Yu L., Ye S., Xie J., Huang X., Zheng K., Sun B. MOV10L1 Binds RNA G-Quadruplex in a Structure-Specific Manner and Resolves It More Efficiently Than MOV10. IScience. 2019;17:36–48. doi: 10.1016/j.isci.2019.06.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 354.Kenny P.J., Kim M., Skariah G., Nielsen J., Lannom M.C., Ceman S. The FMRP–MOV10 complex: a translational regulatory switch modulated by G-Quadruplexes. Nucleic Acids Res. 2020;48:862–878. doi: 10.1093/nar/gkz1092. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 355.Enemark E.J., Joshua-Tor L. On Helicases and other motor proteins. Curr. Opin. Struct. Biol. 2008;18:243–257. doi: 10.1016/j.sbi.2008.01.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 356.Gerstberger S., Hafner M., Tuschl T. A census of human RNA-binding proteins. Nat. Rev. Genet. 2014;15:829–845. doi: 10.1038/nrg3813. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 357.Rajavashisth T.B., Taylor A.K., Andalibi A., Svenson K.L., Lusis A.J. Identification of a zinc finger protein that binds to the sterol regulatory element. Science. 1989;245:640–643. doi: 10.1126/science.2562787. [DOI] [PubMed] [Google Scholar]
- 358.Armas P., Nasif S., Calcaterra N.B. Cellular nucleic acid binding protein binds G-rich single-stranded nucleic acids and may function as a nucleic acid chaperone. J. Cell. Biochem. 2008;103:1013–1036. doi: 10.1002/jcb.21474. [DOI] [PubMed] [Google Scholar]
- 359.Iadevaia V., Caldarola S., Tino E., Amaldi F., Loreni F. All translation elongation factors and the e, f, and h subunits of translation initiation factor 3 are encoded by 5′-terminal oligopyrimidine (TOP) mRNAs. RNA. 2008;14:1730–1736. doi: 10.1261/rna.1037108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 360.Huichalaf C., Schoser B., Schneider-Gold C., Jin B., Sarkar P., Timchenko L. Reduction of the Rate of Protein Translation in Patients with Myotonic Dystrophy 2. J. Neurosci. 2009;29:9042–9049. doi: 10.1523/JNEUROSCI.1983-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]