Nonsense-mediated messenger RNA decay (NMD) represents one of the main surveillance pathways used by eukaryotic cells to control the quality and abundance of mRNAs and to degrade viral RNA. NMD recognises mRNAs with a premature termination codon (PTC) and targets them to decay. Markers for a mRNA with a PTC, and thus NMD, are a long a 3′-untranslated region and the presence of an exon-junction complex (EJC) downstream of the stop codon. Here, we review our structural understanding of mammalian NMD factors and their functional interplay leading to a branched network of different interconnected but specialised mRNA decay pathways. We discuss recent insights into the potential impact of EJC composition on NMD pathway choice. We highlight the coexistence and function of different isoforms of up-frameshift protein 1 (UPF1) with an emphasis of their role at the endoplasmic reticulum and during stress, and the role of the paralogs UPF3B and UPF3A, underscoring that gene regulation by mammalian NMD is tightly controlled and context-dependent being conditional on developmental stage, tissue and cell types.

Nonsense-mediated messenger RNA decay (NMD) is a conserved pathway in eukaryotes that quality controls mRNAs and also acts to regulate transcript abundance. mRNAs containing a premature termination codon (PTC) are recognised by NMD, which targets them for degradation. Beyond mRNA surveillance, NMD plays an important role in controlling gene expression in mammalian cells, which affects cell cycle regulation, cell viability, response to DNA damage and defence against viral infection [1–6]. Dysregulation of the NMD pathway has been implicated in severe pathologies, including neuro-developmental disorders, cellular stress, and cancer [6–11]. Understanding the molecular mechanisms of NMD is therefore of paramount importance for the development of new treatment strategies aimed at modulating the function and activity of the proteins involved in NMD.

The best understood and characterised NMD substrates are (i) mRNAs containing a long 3′-untranslated region (3′-UTR) which results in a large distance between the terminating ribosome and translation termination-stimulating factors, such as the poly(A)-binding protein (PABP) bound to the 3′-poly(A)-tail [12–17]. (ii) The presence of one or more exon-junction complexes (EJCs) bound downstream of the PTC is a strong stimulator of NMD [18–21]. The EJC is a multi-subunit protein complex which is assembled 20–24 nucleotides upstream of exon–exon junctions during splicing [22]. mRNAs containing PTCs located 50–55 nucleotides upstream of an exon–exon junction are efficiently degraded by NMD. In contrast, normal stop codons are typically located in the last exon, and in this case translation of the mRNAs leads to disassembly of the associated EJCs such that normal mRNAs are EJC-free [23,24]. Remaining EJCs downstream of the PTC result in aberrant translation termination and allow the recruitment of NMD factors to the terminating ribosome which leads to activation of the NMD machinery [25–27].

Canonical NMD model

In humans, the canonical NMD machinery consists of the conserved up-frameshift (UPF) proteins (UPF1, UPF2, UPF3B), the suppressors with morphological effects on genitalia proteins (SMG1, SMG5, SMG6, SMG7, SMG8 and SMG9) and the EJC, with its subunits eIF4A3, MAGOH, RBM8A (also known as Y14) and CASC3 (also known as BTZ, MLN51), which will be described in detail below.

In the canonical EJC-dependent model for NMD, slowed translation termination at a PTC allows the formation of a ribosome-associated surveillance complex comprising the SMG1–8–9 kinase complex, UPF1, and the eukaryotic release factors eRF1 and eRF3 (SURF) [28]. Interaction of ribosome-bound UPF1 as part of the SURF complex with EJC-bound UPF2 and UPF3B leads to the formation of a decay-inducing complex (DECID) [28,29]. The DECID complex promotes activation of the SMG1–8–9 kinase complex to phosphorylate UPF1. UPF2 activates UPF1's ATPase and helicase functions which are essential for NMD [30,31]. Once activated and phosphorylated, UPF1 triggers the remodelling of the NMD complexes supporting post-termination ribosome recycling via ATP hydrolysis [32] and recruitment of mRNA decay factors to its phospho-sites, including the endonuclease SMG6 and SMG5–7 [33–35]. The SMG5–7 heterodimer interacts with factors responsible for mRNA deadenylation and mRNA 5′-cap removal (DCP2), and the resulting unprotected mRNA ends are then degraded by 5′–3′ exonuclease XRN1 and the exosome complex [33,36–38].

This review will describe recent insights into the molecular mechanisms of canonical NMD factors, with emphasis on structural data, the role of UPF1 isoforms, the function of UPF3B and UPF3A paralogs and the potential impact of EJC composition on NMD pathway choice in mammalian cells.

Exon junction complexes

EJC core proteins and EJC-associated proteins are intimately connected to mRNA metabolism including its correct splicing, export, mRNA localisation, translation, as well as mRNA turnover including NMD [39,40]. The DEAD-box RNA helicase eIF4A3 is the main RNA-binding component around which the EJC is assembled. eIF4A3 belongs to helicase superfamily 2, which is involved in different RNA-related processes and contains two globular domains (RecA1 and RecA2, Figure 1) with at least 12 characteristic highly conserved sequence motifs, whose functions include adenosine triphosphate (ATP)-binding and hydrolysis (Q, I, II, VI), RNA-binding (Ia, Ib, Ic, IV, IVa, V), and communication between ATP- and RNA-binding sites (III and Va) [41]. Motif II contains the conserved aspartate-glutamate-alanine-aspartate (DEAD) sequence, characteristic of RNA helicases [41]. In the presence of ATP, the two RecA domains adopt a closed conformation that binds the sugar phosphate backbone of six consecutive mRNA nucleotides (Figure 1A,D) [42,43]. MAGOH and RBM8A form a stable heterodimer [44], they lock eIF4A3 onto the mRNA ∼20–24 nucleotides upstream of exon–exon junctions through a tight, hydrophobic, and conserved interaction interface (Figure 1D) [22,42,43,45]. Formation of this EJC core complex inhibits eIF4A3's ATPase and helicase activity [44,46].

MAGOH folds into a single domain formed by six antiparallel β-strands, flanked on one side by two long and one short α-helices (Figure 1B) [42,43]. RBM8A comprises two conserved domains: the N-terminal domain and the RNA-binding-like domain (RBD) essential for MAGOH interaction (Figure 1). The RBD consists of a four-stranded antiparallel β-sheet on one side and two α-helices on the other face (Figure 1B) [42,43]. Atypically, the RBD of RBM8A binds MAGOH and not RNA, with the antiparallel β-sheet of RBM8A packing onto the α-helical surface of MAGOH (Figure 1B) [42].

For a long time, CASC3 (cancer susceptibility candidate gene 3) was considered the fourth EJC core component because it stably interacts with eIF4A3 (Figure 1A,D) and because it is required for EJC assembly in vitro. More recently, the role of CASC3 has been revisited because the absence of CASC3 does not affect EJC assembly during its early stages in the nucleus [47]. Based on this, CASC3 was suggested to function as a peripheral EJC protein, required for the degradation of EJC-dependent NMD substrates. Notably, while the EJC composition and EJC-dependent splicing remain virtually unchanged in the absence of CASC3, hundreds of alternatively spliced mRNAs targeted by NMD are up-regulated in CASC3 CRISPR–Cas9 knockout cells [47]. In agreement, when bound to reporter mRNAs by tethering it to the 3′-UTR or via complexing with the EJC, CASC3 stimulates SMG6-dependent endonucleolytic cleavage at the termination codon and thus promotes NMD [47]. Notably, the SMG6 endonuclease contains two conserved EJC-binding motifs (EBMs) and can directly associate with the EJC [48]. Taken together, CASC3 was shown to affect the efficiency by which NMD-sensitive mRNA isoforms are degraded [47].

UPF3B associates with the EJC in the nucleus and is exported with the spliced EJC-bound mRNA, while UPF2 is recruited by UPF3B to the EJC at the nuclear envelope after export [49]. Recent work suggests a role of CASC3 in promoting the association of UPF3B with the EJC and in enhancing UPF3B-dependent NMD [50,51]. This agrees with reports of a link between CASC3-containing EJCs and UPF3B [39,47]. Consistently, in CASC3 knockdown cells, UPF3B's association with the EJC was reduced, and vice versa CASC3 overexpression led to enhanced EJC–UPF3B association in HeLa cells [51]. The structure of EJC–UPF3B indicates that the interactions of CASC3 and UPF3B with the EJC core are limited to small regions of the two proteins (Figure 1E) [49,52]. However, it is possible that other, less-structured regions of CASC3 and UPF3B also contribute to their EJC association. Using co-immunoprecipitation (co-IP) and mass spectrometry analyses, Wallmeroth and colleagues reported that CASC3 also interacts with UPF2 in wild-type as well as in UPF3B knockout cells [50]. In conclusion, it appears that CASC3 is a major hub for NMD factor-binding to the EJC.

In addition to CASC3 and UPF3B, the core EJC complex is a dynamic binding platform for other peripheral factors, including the apoptosis- and splicing-associated protein complex (ASAP) composed of SAP18, RNPS1 and Acinus [53–55]. Of these, RNPS1 (RNA-binding protein with serine-rich domain 1) [56,57] has been shown to be able to activate NMD. In the nucleus, RNPS1 works together with the U1 and U2 snRNPs in splicing site recognition. RNPS1's serine-rich region at the N-terminus has been shown to be essential for NMD activation [58], and is followed by an RRM (RNA Recognition Motif) domain and a region containing serine- and arginine-rich repeats at the C-terminus (Figure 1C). Consistently, when tethered to the 3′-UTR of an mRNA [53] or when associated to the EJC during splicing [59,60], RNPS1 promotes mRNA decay.

Up-frameshift factor 1

Up-frameshift factor 1 is considered as the key NMD factor. It can bind non-specifically to mRNA and is displaced by the ribosome during translation from the coding sequence, but not from the 3′-UTR [61–64]. UPF1 is a ∼125 kDa protein which is highly conserved in eukaryotes. UPF1 harbours three major domains: a cysteine-histidine-rich domain (CH) and a helicase domain whose structures are well-characterised (Figure 2A,B) [31,65–68], while the serine-glutamine-rich domain (SQ domain) is found within the unstructured C-terminus. UPF1 is hyperphosphorylated by the SMG1–8–9 kinase complex in its disordered N-terminal region (T28) and in its SQ domain (S1078, S1096, S1116 among others) [69]. Recruitment of SMG6 by phospho-UPF1 occurs thanks to interactions with the helicase and C-terminus of UPF1, while SMG5–7 bind to the C-terminal phosphorylated SQ domain [70–72].

Structural characterisation of UPF1's CH domain showed that it forms a unique concatenation of two RING-box like modules (RING1 and RING2) (Figure 2A) [31,65]. The CH domain of UPF1 coordinates three structural zinc ions bound by two zinc-finger motifs located in RING1, and one zinc-finger motif located in RING2 (Figure 2A). Biochemical studies indicate that Upf1 has E3 ubiquitin ligase activity in yeast where it can self-ubiquitinate in the presence of Upf3 and the E2 ubiquitin conjugating enzyme Ubc3 [73]. Furthermore, human UPF1 has been shown to be responsible for the down-regulation of the transcription factor MYOD, a key regulator of myogenesis, not by affecting its mRNA stability, but by promoting its ubiquitination and subsequent degradation by the proteasome [74]. However, it remains to be shown if UPF1 is the E3 ubiquitin ligase directly responsible for MYOD ubiquitination [74]. To date, no other potential targets of the UPF1 E3 ubiquitin ligase have been identified. It is tempting to speculate that UPF1's E3 ubiquitin ligase activity could play a role in the degradation of C-terminally truncated and potentially misfolded polypeptides encoded by nonsense mRNAs. Notably, in other translational quality control pathways, stalled ribosomes trigger poly-ubiquitination of the nascent polypeptide by the E3 ubiquitin ligase LTN1/ Listerin for subsequent degradation by the proteasome [75,76]. However, the fate of the truncated polypeptides produced from PTC-containing mRNAs is currently unknown. Recently, a reporter system with normal and PTC-containing mRNAs encoding the same polypeptide showed specific destabilisation of the protein encoded by the nonsense mRNA which could be overcome by proteasome inhibition [77]. In support of this finding, whole genome knockdown and knockout screens identified the RING E3 ubiquitin ligase CNOT4 (CCR4–NOT transcription complex subunit 4), as a potential enzyme for ubiquitination of the nonsense protein, linking NMD to proteasome-mediated degradation of the encoded nascent polypeptide [78].

The helicase domain of UPF1 harbours the seven motifs distinctive of the DNA/RNA helicase superfamily 1 [30]. To date, several structures of the human UPF1 helicase domain have been solved in different states: in its apo-form [31], in its phosphate-, ADP- (adenosine diphosphate) or AMPPNP-(non-hydrolysable ATP analogue) bound forms [66], as well as in the presence of mRNA in a transition state of ATP hydrolysis with ADP:AlF4- [31]. The large array in conformations of the helicase domain highlights the dynamics of the RNA-binding and helicase activities of UPF1. The helicase domain comprises two tandem RecA-like modules (RecA1 and RecA2), with additional β-barrel and α-helical subdomains (Rec1B and Rec1C respectively) which stick out of RecA1 (Figure 2B). Rec1B is connected to RecA1 by a stalk formed by two α-helices. RecA1 and RecA2 face each other, forming a deep ATP-binding cleft [66], and the interface between RecA1, RecA2, Rec1B and Rec1C shapes a large RNA-binding channel (Figure 2B). In the presence of ATP (or ATP analogues), RNA-binding is impaired because of a constriction of the RNA-binding channel, indicating an allosteric effect of this nucleotide, whereas in the presence of ADP the channel becomes wider [30]. These findings corroborate previous observations where ATP-binding was hindering the binding of RNA [79] and ATP hydrolysis was required for UPF1's activity in NMD [80].

In a structure of yeast Upf1, a longer RNA-binding channel formed by the association of the CH and helicase domains was observed [31]. This is achieved by the presence of interactions between the CH domain RING1 module and the helicase stalk, and between the CH domain RING2 module and the helicase RecA2 module (Figure 2C). Thus, the CH domain is responsible for UPF1's improved RNA-clamping and lower RNA-unwinding activity [30]. This is further corroborated by the structure of UPF1 in complex with the C-terminus of UPF2, which illustrates how UPF2 activates UPF1's helicase activity (Figure 2D) [31,65,67]. Upon complex formation with the CH domain of UPF1, UPF2's UPF1-binding domain (U1BD) forms an α-helix that displaces the CH domain from the RecA2 module and a β-hairpin that binds the solvent-exposed side of UPF1's CH domain (Figure 2D) [31,65]. This displacement results in RNA-clamping being reduced and RNA-unwinding by UPF1's helicase activity being stimulated [31,67]. Taken together, these structural studies highlight the importance of the CH domain for the regulation of RNA-binding and helicase activity of UPF1.

In the cell, UPF1 isoform 2 (UniProtKB: Q92900-2) is the most abundant isoform and accordingly has been studied extensively. More recently however, interest in the less abundant isoform 1 of UPF1 (UniProtKB: Q92900-1) has increased [68,81]. The two UPF1 isoforms exist in mammals only and result from the usage of an alternative 5′ splice site in exon 7 of the Upf1 gene [68]. This alternative splicing adds 11 residues to a loop located in the Rec1B module of the helicase, resulting in UPF1 isoform 1 having a longer loop (UPF1LL) while UPF1 isoform 2 has a shorter loop (UPF1SL) (Figure 2). This loop plays an important role as it contributes to the RNA-binding channel and thus could regulate UPF1's RNA-binding and helicase activity differently in the two isoforms [68,82]. The shorter loop of UPF1SL is well-ordered in the absence of RNA and points towards the helicase core [31,66]. In the presence of RNA, this short loop is disordered and would likely clash with the bound RNA without a conformational change [31]. The longer regulatory loop of UPF1LL is partially disordered in the absence of RNA and points towards the solvent instead of towards the helicase core [68]. In this conformation, the longer regulatory loop would not interfere with RNA-binding. This is corroborated by studies showing that UPF1LL has higher RNA-binding affinity than UPF1SL in the presence of ATP [68]. While the length of the regulatory loop does not influence ATP-binding directly, it has been shown that UPF1's RNA-dependent ATPase activity is higher when this loop is either replaced with a glycine-serine-rich linker or extended (as in UPF1LL) [68]. Consistently, UPF1LL has higher RNA unwinding and translocation activities than UPF1SL [68]. Current findings indicate distinct roles for UPF1LL and UPF1SL in NMD [81]: protective RNA-binding proteins, such as polypyrimidine tract-binding protein-1 (PTBP1) or heterogeneous nuclear ribonucleoprotein L (hnRNP L), were shown to bind to mRNAs with long 3′-UTR and to inhibit translocation of UPF1SL helicase along the mRNA, thus promoting UPF1SL dissociation [83–85]. Because of its higher RNA-binding and unwinding activity, UPF1LL helicase is suggested to overcome this NMD-inhibition by protective RNA-binding proteins and thus promote (EJC-independent) NMD of mRNAs with very long 3′-UTRs [81].

Up-frameshift factor 2

Up-frameshift factor 2 is a conserved 1272 amino acid protein (∼148 kDa) containing three middle portion of eukaryotic initiation factor 4G domains (three MIF4G domains), flanked by N-terminal and C-terminal regions that are predicted to be inherently unstructured (Figure 2) [67]. The C-terminus harbours an aspartate- and glutamate-rich region, referred to as ‘acidic domain’, followed by the U1BD (Figure 2D). UPF2 is considered a scaffolding protein for the UPF-EJC complex, bringing together UPF1 via the U1BD (see above) and the EJC-bound UPF3B through interactions with its U1BD and third MIF4G domain (MIF4G-3), respectively [52,65,67,86]. Complex formation between UPF2 and UPF1 modulates UPF1's RNA-binding and promotes its ATPase and helicase activities (see above). Additionally, the interaction between UPF2 and the SMG1–8–9 complex is suggested to activate the SMG1 kinase to phosphorylate UPF1 [28,87].

Structures of UPF2's MIF4G-1 and MIF4G-2 domains have been solved separately (Figure 2E,F) [87,88], and the structure of its MIF4G-3 domain was solved in complex with UPF3B (Figure 2G) [86,89]. The domains adopt the canonical MIF4G fold comprised of ten core α-helices arranged into an N-terminal four α-helix bundle that interacts with two layers of three α-helices (Figure 2E–G). Moreover, UPF2's MIF4G-1 domain is stabilised by two additional N-terminal capping α-helices and an additional internal α-helix. Two of the core helices are also highly elongated forming a long, protruding coiled coil (Figure 2E) [87]. UPF2's MIF4G-2 domain displays an N-terminal extension forming two short capping α-helices associated with a third α-helix via a flexible linker (Figure 2F) [87]. The MIF4G-3 domain exhibits the canonical fold and has only one additional C-terminal α-helix (Figure 2G) [86]. In a low-resolution co-crystal structure of UPF2's MIF4G-2 and MIF4G-3, these domains are oriented perpendicularly [87]. Furthermore, low-resolution electron microscopy data indicate that UPF2's three MIF4G domains adopt a ring- or horseshoe-like structure [52,90].

UPF2's MIF4G-3 domain interacts with the RNA recognition like-motif domain (RRM-L) of UPF3B (Figure 2G), allowing the association of UPF2 with the EJC [49,86]. Notably, UPF2 has also been reported to bind the SMG1–8–9 kinase complex and eRF3 [28,87,90], with the MIF4G-3 domain of UPF2 being involved in both interactions. Interestingly, MIF4G-3-binding to eRF3 and to UPF3B are mutually exclusive [90]. In contrast, UPF2 can interact with the SMG1–8–9 kinase complex and UPF3B simultaneously to form a large complex in vitro, thereby activating SMG1 kinase activity [87]. A high-resolution structure of SMG1–8–9 in complex with UPF1 and UPF2 is required to shed light on the molecular mechanisms of SMG1 kinase activation by UPF2.

Up-frameshift factors 3A and 3B

UPF3 exists as two distinct paralogs, UPF3A and UPF3B, which are the products of an ancestral gene duplication event and are believed to be functionally distinct. UPF3B is considered as the main paralog in NMD and is better characterised than UPF3A. UPF3B comprises an N-terminal conserved RRM-L domain which lacks critical aromatic residues required for high-affinity RNA-binding by RRMs [86]. Accordingly, UPF3B's RRM-L only weakly interacts with RNA [86,89]. Instead, UPF3B binds the MIF4G-3 domain of UPF2 via the same binding surface which is used by other RRM's to bind RNA (Figure 2G) [86]. The RRM-L is followed by UPF3B’s middle-domain and a C-terminal EJC-binding motif (EBM) (Figure 2). The EBM of UPF3B interacts with the EJC core via the MAGOH–RBM8A heterodimer and via the RecA2 domain of eIF4A3, which remain virtually unaltered upon UPF3B-binding (Figure 1E) [49].

The middle-domain of UPF3B is poorly characterised. This part is predicted to be largely unstructured having a high density of positively charged residues, which may form two long coiled coil-like helices [89]. Biochemical characterisation indicated that the middle-domain of UPF3B is involved in binding UPF1, eRF3, RNA and ribosomes [91]. Moreover, a recent co-crystal structure of the MIF4G-3 domain of UPF2 in complex with the UPF3B's RRM-L followed by the N-terminal part of its middle-domain, revealed an intimate interaction between the two NMD factors, highlighting that the middle-domain of UPF3B contributes to UPF2-binding. In the structure, UPF2's MIF4G-3 domain is wedged between the RRM-L and the middle-domain's N-terminal portion, which adopts a NONA/paraspeckle-like linker followed by an α-helix [89]. This domain architecture is homologous to the family of paraspeckle (Drosophila behaviour/ human splicing) proteins [92], which serve as functional aggregators of nucleic acids in the nucleus.

The presence of the N-terminal portion of the middle-domain results in a significantly tighter interaction (∼200-fold) between UPF3B and UPF2 compared with the RRM-L alone [86,89]. Importantly, the structure of the UPF2-3B complex comprised UPF3B residue Y160 which was reported to be mutated to aspartate (Y160D) in patients with neurodevelopmental disorders [93,94]. UPF3B's middle domain residue Y160 is buried in a hydrophobic cleft formed by UPF2's MIF4G-3 domain. The Y160D mutation results in a ∼40-fold reduction in UPF2-binding affinity, as this residue is likely is displaced from UPF2's hydrophobic cleft due to its charge [89]. The finding that Y160 is phosphorylated in human cells highlights post-translational modification of UPF3B as a means to weaken complex formation with UPF2 and thus down-regulate NMD efficiency [89].

In patients with the UPF3B Y160D mutation, an up-regulation of the UPF3A paralog has been reported [94]. Similarly, knockout/ knockdown of UPF3B leads to increased levels of UPF3A [95,96]. UPF3A shares the domain architecture of UPF3B but its EBM has lower binding affinity to the EJC (Figure 2) [97]. In an UPF3B-deficient context, UPF3A has been shown to be an active NMD factor that can compensate for UPF3B [50,51,97] while in other contexts, UPF3A can function as a NMD inhibitor, protecting mRNAs from NMD [98] (see below).

Recently, Yi and colleagues observed that UPF3A can replace UPF3B in UPF3B knockout cells and then acts as a NMD activator on a similar set of mRNAs [51]. However, the use of a β-globin reporter mRNA with a PTC at codon 39 demonstrated that overexpression of UPF3A could not rescue NMD in the UPF3B knockout cells [51]. Interestingly, when the middle-domain of UPF3A was replaced with UPF3B's middle-domain, NMD was rescued for the β-globin reporter [51]. However, the molecular basis governing this functional difference in the middle-domains of the two paralogs remains to be explored. Consistently, Wallmeroth and colleagues found that UPF3A could compensate for the loss of UPF3B, and keep NMD activity virtually unaltered, while depletion of both UPF3 paralogs resulted in NMD inhibition and up-regulation of PTC-containing mRNAs [50].

UPF3B is proposed to function as a bridge between the EJC and the SURF complex in the EJC-dependent NMD model [30,50,51,97]. In contrast, the EBM of UPF3A has a weak affinity to the EJC and does not appear to be stably associated with the EJC [50,51,97], suggesting that UPF3A can activate NMD independently of the EJC.

Similarly, rescue experiments with UPF3B variants, performed in UPF3-double knockout cells, suggest UPF3B functions of which do not require EJC-SURF bridging [50]: the EBM or the UPF2-binding site in the RRM-L can be depleted or mutated, respectively, without affecting NMD efficiency. NMD was only impaired in the case of double mutants of RRM-L and EBM. Moreover, NMD was also inhibited if the middle domain of UPF3B was deleted in addition to deletion of the EBM or to mutation of the RRM-L [50]. Taken together, these findings highlight a role of UPF3B beyond bridging the EJC and the ribosome-bound SURF complex.

A possible EJC-independent function of UPF3B has been reported previously using a reconstituted mammalian translation system [91]. In this in vitro system, addition of UPF3B was shown to interfere with stop codon recognition and peptide release from the ribosome, thus reducing the efficiency of translation termination. In agreement with a role during translation termination at a PTC, UPF3B has been shown to bind eRF3a, UPF1 and ribosomes in pulldown and co-sedimentation experiments [91]. It was shown that the RRM-L and the middle-domain are required, and that the EBM is dispensable for UPF3B's inhibitory effect on translation termination [91].

In the presence of the EJC, Hauer and colleagues suggested that UPF3B participates in the positioning of the EJC and the NMD machinery, based on iCLIP experiments, where UPF3B was found to bind both the EJC core and mRNA [99].

An intriguing picture emerges where UPF3B and UPF3A levels fine-tune NMD efficiency and thus the transcriptome of cells. The relative and absolute levels of the UPF3 paralogs differ during development and in different cell types and tissues [98,100]. The levels of both paralogs appear to be tightly regulated and are of paramount importance during neurodevelopment, as evidenced by UPF3B mutations leading to decreased NMD efficiency and neurodevelopmental disorders in affected families [93,96]. For instance, in the case of the UPF3B Y160D mutation (see above), a slightly reduced NMD efficiency causes an up-regulation of nonsense mRNAs encoding the ATF4 transcription factor and the ARHGAP24 signalling factor. The consequence of this are changes in neuronal differentiation and a reduction in neurite branching, despite the up-regulation of UPF3A to compensate for the loss of UPF3B function [93,94]. Future work is required to shed light on the roles of UPF3B in translation termination and NMD, and to identify the functional differences between the paralogs UPF3A and UPF3B.

The SMG1–8–9 complex (SMG1C)

SMG1 kinase (∼410 kDa) belongs to the family of phosphatidylinositol-3-kinase-related kinases (PIKKs) [101]. SMG1 phosphorylates UPF1 at serine/threonine-glutamine (SQ) motifs present in the unstructured N- and C-terminal regions that flank the CH-helicase core (Figure 2) [102]. In addition, several UPF1 phosphorylation sites have been identified sharing a leucine-serine-glutamine (LSQ) consensus sequence which is identical with that found in ataxia telangiectasia mutated (ATM) kinase substrates [103].

In metazoans, SMG1 kinase is intimately associated with SMG8 and SMG9, forming the SMG1–8–9 complex (SMG1C) [29]. In SMG1, like in all PIKKs, the kinase domain consists of an N-terminal α-helical solenoid ‘arch’ and a compact C-terminal ‘head’ region (Figure 3A). The N-terminus is formed of (Huntington, Elongation factor 3, A subunit of PP2A, and TOR1) HEAT repeats which interact with SMG8 and SMG9 (Figure 3) [101]. The globular C-terminal ‘head’ region contains a (FRAP, ATM, and TRRAP) FAT domain, the kinase domain, an insertion domain of 1200 residues, and a short FAT C-terminal (FATC) domain (Figure 3) [101,104]. The insertion domain is unique to SMG1 and has been shown to negatively regulate kinase activity in combination with SMG8 and SMG9 (see below) [105].

SMG8 has an N-terminal G-like domain similar to that found in dynamin-like GTPases (Figure 3). It consists of a central eight-stranded β-sheet surrounded by seven α-helices, followed by a bundle of three α-helices to form a stem-like domain [106]. SMG9 contains a similar G-fold domain (Figure 3) which forms a stable pseudo-symmetric heterodimer with SMG8 [106]. SMG9 interacts with SMG1 through an extended coiled α-helical region and a peripheral β-sheet that forms hydrophobic contacts with the HEAT repeats of SMG1 (Figure 3B). Thus, SMG9 bridges SMG8 and SMG1 to form the SMG1C complex (Figure 3C) [29,106–108].

While SMG8 and SMG9 both contain G-like domains, only the SMG9 G-like domain contains a nucleotide-binding site required for NTP hydrolysis [106]. Consistently, a nucleotide bound to SMG9 was observed in crystal and cryo-EM structures [106,107,109]. Ion-pair high-performance liquid chromatography mass spectrometry (HPLC–MS) identified the co-purified nucleotide in SMG9 as ATP in SMG1C [107]. Comparison of the NTP-bound and NDP-bound structures shows a significant difference in the positions of SMG8 and SMG9 in the available structures, with a region of SMG8, most likely derived from the C-terminus, shifting by ∼10 Å and covering the substrate entry channel of SMG1's kinase domain when SMG9 is NTP-bound [106,109]. In agreement with a kinase-inhibitory role for this region, deletion of SMG8's C-terminus resulted in the hyperactivation of SMG1 kinase when compared with the wild-type complex [106]. Notably, the inhibition of SMG1 kinase activity by SMG8's C-terminus seems to be synergistic with the negative regulation by the C-terminal insertion domain of SMG1 [105,106]. Taken together, these findings suggest that UPF1's access to the catalytic site in SMG1C is tightly regulated, and UPF1 phosphorylation occurs after a reorganisation of both the SMG1 C-terminal insertion domain and the SMG8-9 heterodimer [105,106]. This structural rearrangement of SMG1C is suggested to be driven by ATP hydrolysis of SMG9.

UPF1 phosphorylation by SMG1C is the key step in NMD. Recently, Langer and colleagues revealed the molecular basis for phosphorylation site selection by SMG1 kinase by solving the cryo-EM structure of SMG1C with a peptide mimicking a part of UPF1's SQ domain (residues 1074–1084) (Figure 3D) [108]. In the kinase active site of SMG1, Q1079 at position +1 in the UPF1 SQ-domain peptide is positioned in a hydrophobic cage composed of the SMG1 activation segment and the FATC domain (Figure 3D) [108]. In agreement with the importance of Q1079, mutation of this glutamine inhibits phosphorylation. SMG1C's phosphorylation efficiency and specificity are further improved by a hydrophobic residue at position −1 (predominantly leucine) in the UPF1 SQ-domain peptide (Figure 3D) [108]. Taken together, these observations highlight why the LSQ motif found in UPF1 SQ-domain is optimal for binding the SMG1 kinase active site.

A cryo-EM structure of SMG1C in complex with a pyrimidine derivative inhibitor shows this compound bound between the N- and C-terminal lobes of the kinase domain where it interacts with residues P2213 and D2339 in the active site (Figure 3E) [110]. These proline and aspartate residues are unique to SMG1, which may explain the selectivity of the inhibitor for SMG1 compared with other PIKK family members [110]. The SMG1 inhibitor stabilises an autoinhibitory conformation of the SMG1 insertion domain (the 1200-residue region connecting the kinase and FATC domains of SMG1). Based on this finding, the authors suggest that the SMG1 insertion domain contains a PIKK-regulatory domain (PDR) that engages with and occludes access to the kinase active site, in analogy to PDRs identified in other PIKKs. Furthermore, the inhibitory conformation of SMG1's insertion domain is stabilised by interaction with SMG8's C-terminal domain, confirming that these domains work together to down-regulate SMG1 kinase activity [110].

SMG5, SMG6, SMG7

Hyperphosphorylated UPF1 recruits decay-inducing factors including SMG5, SMG6 and SMG7 [69,71,72,111,112]. These three proteins are structurally related and have similar tetratricopeptide repeat (TPR) domains, consisting of a 14-3-3-like fold formed by nine antiparallel α-helices, followed by an α-helical domain (Figure 4). [113]. The 14-3-3 fold binds phosphorylated serine/threonine residues (Figure 4A,D,E). However, the three proteins have different C-termini (Figure 4): SMG5 and SMG6 harbour PilT N-terminus (PIN) domains of which only the SMG6 domain has endonuclease activity (Figure 4B,C) [34,35,114,115] whereas SMG7 holds a proline-rich C-terminal (PC) domain [114].

The structure of the SMG5–7 TPR domains shows a heterodimer with an evolutionarily conserved interface between the two 14-3-3-like folds (Figure 4D) [113,116]. Complex formation with phospho-UPF1 leads to the recruitment of the CCR4–NOT deadenylation complex [117]. Deadenylation of the target mRNA is followed by mRNA decapping and 5′–3′ and/or 3′–5′ exonucleolytic degradation (Figure 5) [114,117].

In contrast with SMG5–7, SMG6 is a monomer, consisting of a 14-3-3-like fold domain, an EJC-binding domain, and a PIN domain with endonuclease activity (Figure 4A,B) [34,48,71,115]. SMG6 has been shown to cleave mRNA close to the PTC, followed by 5′–3′ degradation via the exonuclease XRN1 and 3′–5′ degradation by the exosome complex [35,37,38,118,119]. Co-IP assays showed that SMG6 associates via its two EBMs with MAGOH–RBM8A, competing with UPF3B for EJC-binding [48], and CASC3 was shown to stimulate SMG6 endonuclease activity [47]. In conclusion, the interaction of SMG6 with the EJC is suggested to facilitate its recruitment to nonsense mRNA and enzymatic activation [48].

SMG5 and SMG7 complementation experiments in knockout cells indicate that the SMG5–7 heterodimer is essential for SMG6's endonuclease activity [120]. SMG6 was unable to cleave nonsense mRNAs in the absence of SMG5–7, but a transient interaction between SMG6 and UPF1 could be detected in TurboID experiments nevertheless [120]. This finding agrees with pulldown assays showing that SMG6 can interact also with unphosphorylated UPF1 [71]. While previous work suggested that SMG6 and SMG5–7 promote the degradation of NMD targets in a redundant manner — either by endonucleolytic cleavage or by deadenylation and decapping (Figure 5) [113,121,122] — the work from Boehm and colleagues emphasises the interdependence of SMG6- and SMG5–7-mediated degradation paths [120]. The authors suggest a model, where SMG5–7 recruitment to phospho-UPF1 represents an authorising step required for SMG6-mediated degradation of the target transcript.

In mammalian cells, NMD functions extend beyond quality control of protein synthesis. Illustrative of this, NMD is a major contributor to mammalian gene regulation by targeting ∼10% of the transcripts [3,5,8,123]. Endogenous transcripts recognised by the NMD machinery are typically characterised by a long 3′-UTR, a 3′-UTR intron, or a short upstream open reading frame (ORF) in the 5′-UTR [3,5,124–126]. Moreover, NMD targets mRNAs encoding selenoproteins, small nucleolar RNAs and long non-coding RNAs [3,5,38,127]. This wide range of NMD substrates implies that more specific mechanisms have evolved to allow the NMD machinery to accurately select target RNAs, while having a highly regulated and fine-tuned activity. Consequently, mammalian NMD is not one pathway but specialised and subdivided into several branches. All NMD pathways share a common feature, namely the central role of UPF1 (reviewed in e.g. [5,128]) (Figure 5). The canonical NMD mechanism is EJC-dependent, involving the formation of the SURF and DECID complexes (see above). Furthermore, EJC-dependent NMD can branch off into UPF2-independent or UPF3B-independent pathways [58,121,129], and transcripts with long 3′-UTRs are degraded in an EJC-independent manner, involving UPF1, SMG1 as well as UPF2 and/or UPF3B [13,121].

In EJC-dependent NMD, further sub-branches exist due to varying EJC composition resulting from the presence of mutually exclusive factors (Figure 5) [39]. (i) In the presence of EJCs containing RNPS1, NMD has been shown to occur independently of UPF3B [58]. In this branch of NMD, the RNSP1-containing EJC can directly recruit UPF2 and elicit NMD (Figure 5). (ii) Inversely, UPF2-independent NMD can occur when CASC3 is bound in the EJC. CASC3 binds eIF4A3, UPF3B and UPF1, and UPF3B was shown to interact directly with UPF1 and eRF3a [39,59,91]. Therefore, CASC3 could enhance the interaction of UPF3B-bound EJC with UPF1 and thus promote NMD in the absence of UPF2 (Figure 5) [58]. Even though several studies support this CASC3-enhanced model [39,47,59], the mechanisms underlying the stimulation of UPF1's ATPase and helicase functions and downstream NMD activation in the absence of UPF2 remain unclear to date.

RNPS1- and CASC3-enhanced NMD pathways differ in the subcellular localisation and lifetime of the EJCs associated with these factors [39]. RNPS1 binds to mRNAs during co-transcriptional splicing, therefore RNSP1-containing EJCs are enriched in the nucleus and at early cytoplasmic EJC stages. In contrast, CASC3 engages with the EJC post-splicing following an EJC-compositional switch which occurs either before or during translation in the cytosol at later EJC stages [39,59]. In agreement, RNPS1-enhanced UPF3B-independent NMD is suggested to target mRNAs translated shortly after their export from the nucleus, directly in the nuclear periphery, while CASC3-enhanced UPF2-independent NMD could target mRNAs anytime during translation [39,58].

Splicing and NMD are linked in EJC-dependent NMD, and members of the serine- and arginine-rich (SR) protein family of splicing factors, which bind nascent mRNAs during co-transcriptional splicing, are suggested to enhance NMD [130–132]. Consistently, SR Splicing Factors 1 and 3 (SRSF1, SRSF3) strongly promote NMD when they are bound or tethered to a mRNA downstream of a PTC [131]. SRSF1 interacts in an RNA-independent manner with EJC-core components (eIF4A3, MAGOH and RBM8A), as well as other EJC-associated proteins including RNPS1 and UPF3B, but not CASC3 or UPF2 (Figure 5) [131]. However, SRSF1 interaction with the EJC is not required for its NMD-enhancing function [131]. Rather, SRSF1 interacts directly with UPF1 and appears to recruit UPF1 to the mRNA in the nucleus, and this process has been suggested to be responsible for SRSF1-enhanced NMD (Figure 5). Furthermore, recent reports suggest that in the cytoplasm, SRSF1 has an additional NMD-enhancing activity because co-IP assays show an interaction between SRSF1 and serine/threonine-protein phosphatase 2A (PP2A) and SMG7 [131,133]. Via these interactions, SRSF1 may promote the dephosphorylation of UPF1, which would result in the release of UPF1 from SMG6 and SMG5–7. Following on from this, the NMD factors are then recycled and ready for a new round of NMD [131,134,135]. The ability of SRSF1 to enhance NMD of intron-less transcripts when tethered downstream of a PTC supports the existence of an EJC-independent pathway in mammals, as it was suggested for long 3′-UTR transcripts [13,131].

Knock-down of several NMD factors showed that UPF1 and SMG1 are strictly required for EJC-independent NMD, while UPF3B or UPF2 can only be depleted separately [121], confirming their role of core NMD factors (Figure 5). The existence of different branches of NMD underscores a sophisticated fine-tuning of NMD for the regulation of gene expression in mammalian cells in a context-dependent manner. Examples for localised NMD responses have been reported for dendritic and axonal mRNAs in neurons and for mRNAs localised at the endoplasmic reticulum (ER) [136]. The ER is specialised for the translation and translocation of secreted and integral membrane proteins while also being the site of the unfolded protein response (UPR) [137,138]. Immunofluorescence microscopy indicates that UPF3B and SMG6 co-localise with the ER chaperone BiP and thus reside at the ER, and SMG6 is up-regulated upon induction of ER-stress [139]. Vice versa, NMD is involved in regulating the expression of UPR factors, including ATF-3, ATF-4, PERK and IRE1α [140,141]. Genome-wide RNA interference screens identified the neuroblastoma-amplified sequence protein (NBAS) and the helicase DHX34 as additional NMD factors. These proteins were found to regulate the stability of a subset of mRNAs encoding proteins predominantly involved in the response to cellular and environmental stress as well as NMD factors including UPF1 [136,141,142]. Previous reports show that NMD is inhibited by cellular stress, including amino acid starvation, up-regulation of the UPR and hypoxia [143]. More recently, NBAS was shown to bind mRNA and to interact with UPF1, SMG6 and SMG5–7 [136]. In addition, NBAS knockdown leads to a five-fold increase in membrane-associated mRNAs, in agreement with a role of NBAS at the ER [136]. Based on these findings, NBAS is suggested to recruit UPF1 to the ER membrane to facilitate an ER-associated NMD pathway. Taken together, NMD appears to respond to cellular stress and to be part of a feedback loop that activates the cellular stress response.

Different versions of the NMD core factors are used in context- and cell type-specific manners. The paralogs UPF3A and UPF3B are one example. Despite their homology, UPF3A is rapidly degraded in the cell and requires UPF2-binding for its stabilisation [95]. In most tissues, UPF3B is significantly more abundant than UPF3A, with one exception being the adult testis [98]. In an UPF3B-deficient context (see above), UPF3A is stabilised and activates NMD [50,51,97]. However, during male germ cell development, the X-linked Upf3B gene is transcriptionally silenced, due to meiotic sex cell chromosome inactivation, such that UPF3A is highly expressed [98]. Surprisingly, NMD was enhanced when Upf3A was removed from developing male germ cells, suggesting that UPF3A is a NMD repressor in these cells [98]. The presence of UPF3A and the absence of UPF3B was shown to stabilise mRNAs encoding proteins which are essential for spermatogenesis and male fertility [98]. At the same time, normal levels of NMD targeting other mRNAs with long 3′-UTRs were detected in UPF3A/3B-deficient cells, indicating down-regulation of one NMD branch in male germ cells [98,144].

A further example of branched NMD is UPF1 which exists in two splice variants in mammalian cells, UPF1LL and UPF1SL, with different functional roles and outcomes for NMD (see above). These variants are differentially expressed in human tissues: while UPF1LL represents ∼20% of expressed UPF1 in most tissues, it is up-regulated to ∼35% in muscles and down-regulated to ∼15% in the liver [81]. The specific knockdown of UPF1LL indicates that it participates in NMD under normal cellular conditions. Transcriptome analyses revealed that UPF1LL NMD targets were enriched in mRNAs encoding proteins preferentially translated at the ER, such as integral membrane proteins [81]. Interestingly, during ER stress, these UPF1LL NMD targets remained down-regulated. In contrast, mRNAs preferentially targeted by UPF1SL were up-regulated in agreement with stress-induced NMD inhibition [81]. This differential regulation of NMD is suggested to result from the down-regulation of translation which is induced by the integrated stress response and mediated by the phosphorylation of eIF2α. Similarly, a moderate slowdown of translation by treating cells with sub-inhibitory concentrations of the antibiotics Blasticidin S or cycloheximide was reported to repress NMD and resulted in higher expression levels of the protein encoded by the nonsense mRNA [145]. In this context, UPF1LL could detect rare aberrant translation termination events and induce NMD, possibly due to its increased residence time on mRNAs compared with UPF1SL [81].

Beyond translational control, NMD targets viral RNA transcripts and genomes, which often contain unique features that could elicit an NMD response. Accordingly, NMD plays a crucial role in controlling viral pathogen replication in both plants and animals [146–149]. Because of this, viruses have expanded their repertoire for infecting hosts by evolving new mechanisms for evading and inhibiting NMD. Research into these viral processes has provided vital new understandings into the complex protein–protein and RNA–protein interactions during NMD and highlighted two strategies employed by viruses to avoid NMD, known as cis and trans methods (reviewed e.g. in [150]). Cis methods are characterised by features of the viral RNA that either directly prevent NMD or recruit host factors that antagonise NMD. Trans methods are based on proteins encoded by the virus that are synthesised to directly interfere with NMD leading to a simultaneous up-regulation of host cell nonsense mRNAs [150].

An example for a cis-based mechanism is found in the Rous sarcoma virus (RSV). The retroviral mRNA of RSV contains an RNA stability element (RSE) which prevents UPF1 from initiating NMD. The RSE contains multiple polypyrimidine tracts which recruit PTBP1 (see above) [151,152]. PTBP1 prevents the translocation of UPF1 by acting as a physical barrier on the RNA promoting the dissociation of UPF1 from the RNA [83].

An example for a trans-acting mechanism is found in the human T-cell leukaemia virus type I (HTLV-1). HTLV-1 is a positive-sense RNA retrovirus that randomly integrates its genetic material with the host genome, causing aggressive adult T-cell leukaemia. The HTLV-1 genome contains >10 ORFs which are regulated by alternative splicing and programmed ribosomal frameshifting. The mRNAs of HTLV-1 should be targets for NMD due to various features including long 3′-UTRs of ∼4000 bases. However, HTLV-1 expresses two positive trans-regulatory proteins, Rex and Tax, which aid in NMD evasion [153–156]. Co-IP experiments against Tax from HTLV-1-infected cells, identified INT6 (one of the thirteen eIF3 subunits) and UPF1 as interaction partners [156]. NMD inhibition by Tax depends on its interaction with INT6, as Tax mutants that are unable to bind to INT6 were found to have no effect on NMD [157,158], suggesting an involvement of eIF3 in NMD of mRNAs with long 3′-UTRs. In addition, Tax interacts with the helicase domain of UPF1 and reduces its RNA-binding affinity by blocking the RNA entry site. Using single-molecule techniques, Tax was also shown to interact with RNA-bound UPF1, blocking its ATPase and helicase activities leading to the dissociation of UPF1 from the mRNA [156,158]. Taken together, Tax appears to inhibit NMD by binding UPF1 and INT6 and interfering with their interaction [156]. Co-IPs additionally indicate that HTLV-1 RNA is still able to interact with phosphorylated UPF1 which accumulates in P-bodies for degradation. P-bodies are cytoplasmic ribonucleoprotein granules that consist mostly of proteins associated with mRNA decay and translationally repressed proteins. When NMD was inhibited, an increased accumulation of UPF1 in P-bodies was observed [156,157]. In this inhibited NMD context, Tax was found to co-localise with UPF1 in P-bodies causing its hyperphosphorylation and leading to a decrease in free unphosphorylated UPF1 and further inhibition of NMD [157,159]. Thus, while Tax protects its encoding RNA from NMD, it also subjects the host's gene expression to dysregulation by stabilising endogenous NMD targets through global NMD inhibition [156,157].

The strategies deployed by viruses to evade NMD continue to be a rich field of discovery providing important new insights into basic NMD mechanisms as well as potential antiviral treatment strategies.

Research into NMD using a diverse set of model systems suggests that the NMD pathway has acquired an increasing importance over the course of evolution. Although Saccharomyces cerevisiae and Caenorhabditis elegans are viable without a functional NMD pathway [160,161], NMD is essential for viability in vertebrates [162–164]. This gain of importance with increasing organism complexity corroborates the expansion of NMD functions beyond translational control spanning into antiviral defence and, importantly, the global regulation of gene expression. In mammalian cells, the existence of multiple NMD branches expanded by factor variants and localised functions emerge, each acting on specific subsets of mRNA targets with fine-tuned efficiency, depending on the subcellar localisation, the tissue and cell types, as well as the developmental stage.

These diverse and vital roles played by NMD are highlighted by the medical importance of NMD, as ∼20% of human genetic diseases caused by single base pair mutations are related to NMD [165]. To develop new therapeutic strategies, an advanced understanding of the molecular mechanisms of NMD is needed, including high-resolution structures of complexes to provide insights into the interplay of NMD factors, and their interactions with the translation machinery and the EJC. In fact, crucial basic questions remain unanswered such as how a translation termination event is recognised as aberrant, how recruitment of NMD factors to target mRNAs is achieved, and what is their function in the different NMD branches. Future research in the NMD field will help in elucidating the molecular basis of the cross-talk between the NMD machinery and quality control mechanisms regulating cellular homeostasis, including ribosome quality control and the cellular stress response.

The authors declare that there are no competing interests associated with the manuscript.

We would like to thank Drs. Kyle Powers, George Orriss, Sathish Yadav Kadapalakere, Lingling Sun, and Jenn-Yeu Alvin Szeto for discussing the manuscript, critical reading and helpful comments. C.S. is Investigator of the Wellcome Trust (210701/Z/18/Z). M.V.V. and C.S. acknowledge support by a Marie Sklodowska-Curie Fellowship from the EC (101024558). For the purpose of Open Access, the authors have applied a CC BY public copyright license to any Author Accepted Manuscript version arising from this submission.

ADP

adenosine diphosphate

ATM

ataxia telangiectasia mutated

EBM

EJC-binding motif

EJC

exon-junction complex

ER

endoplasmic reticulum

HTLV-1

human T-cell leukaemia virus type I

LSQ

leucine-serine-glutamine

NBAS

neuroblastoma-amplified sequence protein

NMD

nonsense-mediated messenger RNA decay

PIKKs

phosphatidylinositol-3-kinase-related kinases

PTC

premature termination codon

RBD

RNA-binding-like domain

RRM

RNA Recognition Motif

RSE

RNA stability element

RSV

Rous sarcoma virus

SMG1C

SMG1–8–9 complex

TPR

tetratricopeptide repeat

UPF

up-frameshift protein

UPR

unfolded protein response

1
Nasif
,
S.
,
Contu
,
L.
and
Mühlemann
,
O.
(
2018
)
Beyond quality control: the role of nonsense-mediated mRNA decay (NMD) in regulating gene expression
.
Semin. Cell Dev. Biol.
75
,
78
87
2
Kim
,
Y.K.
and
Maquat
,
L.E.
(
2019
)
UPFront and center in RNA decay: UPF1 in nonsense-mediated mRNA decay and beyond
.
RNA
25
,
407
422
3
Lykke-Andersen
,
S.
and
Jensen
,
T.H.
(
2015
)
Nonsense-mediated mRNA decay: an intricate machinery that shapes transcriptomes
.
Nat. Rev. Mol. Cell Biol.
16
,
665
677
4
Kishor
,
A.
,
Fritz
,
S.E.
and
Hogg
,
J.R.
(
2019
)
Nonsense-mediated mRNA decay: the challenge of telling right from wrong in a complex transcriptome
.
Wiley Interdiscip. Rev. RNA
10
,
e1548
5
Kurosaki
,
T.
,
Popp
,
M.W.
and
Maquat
,
L.E.
(
2019
)
Quality and quantity control of gene expression by nonsense-mediated mRNA decay
.
Nat. Rev. Mol. Cell Biol.
20
,
406
420
6
Pawlicka
,
K.
,
Kalathiya
,
U.
and
Alfaro
,
J.
(
2020
)
Nonsense-Mediated mRNA decay: pathologies and the potential for novel therapeutics
.
Cancers (Basel)
12
,
765
7
Bhuvanagiri
,
M.
,
Schlitter
,
A.M.
,
Hentze
,
M.W.
and
Kulozik
,
A.E.
(
2010
)
NMD: RNA biology meets human genetic medicine
.
Biochem. J.
430
,
365
377
8
Ottens
,
F.
and
Gehring
,
N.H.
(
2016
)
Physiological and pathophysiological role of nonsense-mediated mRNA decay
.
Pflugers Arch.
468
,
1013
1028
9
Goetz
,
A.E.
and
Wilkinson
,
M.
(
2017
)
Stress and the nonsense-mediated RNA decay pathway
.
Cell. Mol. Life Sci.
74
,
3509
3531
10
Lindeboom
,
R.G.H.
,
Vermeulen
,
M.
,
Lehner
,
B.
and
Supek
,
F.
(
2019
)
The impact of nonsense-mediated mRNA decay on genetic disease, gene editing and cancer immunotherapy
.
Nat. Genet.
51
,
1645
1651
11
Fernandes
,
R.
,
Nogueira
,
G.
,
da Costa
,
P.J.
,
Pinto
,
F.
and
Romão
,
L.
(
2019
)
Nonsense-Mediated mRNA decay in development, stress and cancer
.
Adv. Exp. Med. Biol.
1157
,
41
83
12
Amrani
,
N.
,
Ganesan
,
R.
,
Kervestin
,
S.
,
Mangus
,
D.A.
,
Ghosh
,
S.
and
Jacobson
,
A.
(
2004
)
A faux 3'-UTR promotes aberrant termination and triggers nonsense-mediated mRNA decay
.
Nature
432
,
112
118
13
Bühler
,
M.
,
Steiner
,
S.
,
Mohn
,
F.
,
Paillusson
,
A.
and
Mühlemann
,
O.
(
2006
)
EJC-independent degradation of nonsense immunoglobulin-mu mRNA depends on 3′ UTR length
.
Nat. Struct. Mol. Biol.
13
,
462
464
14
Peixeiro
,
I.
,
Inácio
,
Â.
,
Barbosa
,
C.
,
Silva
,
A.L.
,
Liebhaber
,
S.A.
and
Romão
,
L.
(
2012
)
Interaction of PABPC1 with the translation initiation complex is critical to the NMD resistance of AUG-proximal nonsense mutations
.
Nucleic Acids Res.
40
,
1160
1173
15
Ivanov
,
A.
,
Mikhailova
,
T.
,
Eliseev
,
B.
,
Yeramala
,
L.
,
Sokolova
,
E.
,
Susorov
,
D.
et al (
2016
)
PABP enhances release factor recruitment and stop codon recognition during translation termination
.
Nucleic Acids Res.
44
,
7766
7776
16
Wu
,
C.
,
Roy
,
B.
,
He
,
F.
,
Yan
,
K.
and
Jacobson
,
A.
(
2020
)
Poly(A)-binding protein regulates the efficiency of translation termination
.
Cell Rep.
33
,
108399
17
Eberle
,
A.B.
,
Stalder
,
L.
,
Mathys
,
H.
,
Orozco
,
R.Z.
and
Mühlemann
,
O.
(
2008
)
Posttranscriptional gene regulation by spatial rearrangement of the 3′ untranslated region
.
PLoS Biol.
6
,
e92
18
Lindeboom
,
R.G.
,
Supek
,
F.
and
Lehner
,
B.
(
2016
)
The rules and impact of nonsense-mediated mRNA decay in human cancers
.
Nat. Genet.
48
,
1112
1118
19
Kim
,
V.N.
,
Kataoka
,
N.
and
Dreyfuss
,
G.
(
2001
)
Role of the nonsense-mediated decay factor hUpf3 in the splicing-dependent exon-exon junction complex
.
Science
293
,
1832
1836
20
Le Hir
,
H.
,
Gatfield
,
D.
,
Izaurralde
,
E.
and
Moore
,
M.J.
(
2001
)
The exon-exon junction complex provides a binding platform for factors involved in mRNA export and nonsense-mediated mRNA decay
.
EMBO J.
20
,
4987
4997
21
Lykke-Andersen
,
J.
,
Shu
,
M.D.
and
Steitz
,
J.A.
(
2000
)
Human Upf proteins target an mRNA for nonsense-mediated decay when bound downstream of a termination codon
.
Cell
103
,
1121
1131
22
Le Hir
,
H.
,
Izaurralde
,
E.
,
Maquat
,
L.E.
and
Moore
,
M.J.
(
2000
)
The spliceosome deposits multiple proteins 20-24 nucleotides upstream of mRNA exon-exon junctions
.
EMBO J.
19
,
6860
6869
23
Dostie
,
J.
and
Dreyfuss
,
G.
(
2002
)
Translation is required to remove Y14 from mRNAs in the cytoplasm
.
Curr. Biol.
12
,
1060
1067
24
Gehring
,
N.H.
,
Lamprinaki
,
S.
,
Kulozik
,
A.E.
and
Hentze
,
M.W.
(
2009
)
Disassembly of exon junction complexes by PYM
.
Cell
137
,
536
548
25
Raimondeau
,
E.
,
Bufton
,
J.C.
and
Schaffitzel
,
C.
(
2018
)
New insights into the interplay between the translation machinery and nonsense-mediated mRNA decay factors
.
Biochem. Soc. Trans.
46
,
503
512
26
Karousis
,
E.D.
and
Mühlemann
,
O.
(
2019
)
Nonsense-mediated mRNA decay begins where translation ends
.
Cold Spring Harb. Perspect. Biol.
11
,
a032862
27
Lavysh
,
D.
and
Neu-Yilik
,
G.
(
2020
)
UPF1-mediated RNA decay-Danse macabre in a cloud
.
Biomolecules
10
,
999
28
Kashima
,
I.
,
Yamashita
,
A.
,
Izumi
,
N.
,
Kataoka
,
N.
,
Morishita
,
R.
,
Hoshino
,
S.
et al (
2006
)
Binding of a novel SMG-1-Upf1-eRF1-eRF3 complex (SURF) to the exon junction complex triggers Upf1 phosphorylation and nonsense-mediated mRNA decay
.
Genes Dev.
20
,
355
367
29
Yamashita
,
A.
,
Izumi
,
N.
,
Kashima
,
I.
,
Ohnishi
,
T.
,
Saari
,
B.
,
Katsuhata
,
Y.
et al (
2009
)
SMG-8 and SMG-9, two novel subunits of the SMG-1 complex, regulate remodeling of the mRNA surveillance complex during nonsense-mediated mRNA decay
.
Genes Dev.
23
,
1091
1105
30
Chamieh
,
H.
,
Ballut
,
L.
,
Bonneau
,
F.
and
Le Hir
,
H.
(
2008
)
NMD factors UPF2 and UPF3 bridge UPF1 to the exon junction complex and stimulate its RNA helicase activity
.
Nat. Struct. Mol. Biol.
15
,
85
93
31
Chakrabarti
,
S.
,
Jayachandran
,
U.
,
Bonneau
,
F.
,
Fiorini
,
F.
,
Basquin
,
C.
,
Domcke
,
S.
et al (
2011
)
Molecular mechanisms for the RNA-dependent ATPase activity of Upf1 and its regulation by Upf2
.
Mol. Cell
41
,
693
703
32
Serdar
,
L.D.
,
Whiteside
,
D.L.
,
Nock
,
S.L.
,
McGrath
,
D.
and
Baker
,
K.E.
(
2020
)
Inhibition of post-termination ribosome recycling at premature termination codons in UPF1 ATPase mutants
.
ELife
9
,
e57834
33
Lejeune
,
F.
,
Li
,
X.
and
Maquat
,
L.E.
(
2003
)
Nonsense-mediated mRNA decay in mammalian cells involves decapping, deadenylating, and exonucleolytic activities
.
Mol. Cell
12
,
675
687
34
Huntzinger
,
E.
,
Kashima
,
I.
,
Fauser
,
M.
,
Saulière
,
J.
and
Izaurralde
,
E.
(
2008
)
SMG6 is the catalytic endonuclease that cleaves mRNAs containing nonsense codons in metazoan
.
RNA
14
,
2609
2617
35
Eberle
,
A.B.
,
Lykke-Andersen
,
S.
,
Mühlemann
,
O.
and
Jensen
,
T.H.
(
2009
)
SMG6 promotes endonucleolytic cleavage of nonsense mRNA in human cells
.
Nat. Struct. Mol. Biol.
16
,
49
55
36
Chen
,
C.Y.
and
Shyu
,
A.B.
(
2003
)
Rapid deadenylation triggered by a nonsense codon precedes decay of the RNA body in a mammalian cytoplasmic nonsense-mediated decay pathway
.
Mol. Cell Biol.
23
,
4805
4813
37
Boehm
,
V.
,
Haberman
,
N.
,
Ottens
,
F.
,
Ule
,
J.
and
Gehring
,
N.H.
(
2014
)
3′ UTR length and messenger ribonucleoprotein composition determine endocleavage efficiencies at termination codons
.
Cell Rep.
9
,
555
568
38
Lykke-Andersen
,
S.
,
Chen
,
Y.
,
Ardal
,
B.R.
,
Lilje
,
B.
,
Waage
,
J.
,
Sandelin
,
A.
et al (
2014
)
Human nonsense-mediated RNA decay initiates widely by endonucleolysis and targets snoRNA host genes
.
Genes Dev.
28
,
2498
2517
39
Mabin
,
J.W.
,
Woodward
,
L.A.
,
Patton
,
R.D.
,
Yi
,
Z.
,
Jia
,
M.
,
Wysocki
,
V.H.
et al (
2018
)
The exon junction complex undergoes a compositional switch that alters mRNP structure and nonsense-mediated mRNA decay activity
.
Cell Rep.
25
,
2431
2446.e2437
40
Viphakone
,
N.
,
Sudbery
,
I.
,
Griffith
,
L.
,
Heath
,
C.G.
,
Sims
,
D.
and
Wilson
,
S.A.
(
2019
)
Co-transcriptional loading of RNA export factors shapes the human transcriptome
.
Mol. Cell
75
,
310
323.e318
41
Linder
,
P.
and
Jankowsky
,
E.
(
2011
)
From unwinding to clamping - the DEAD box RNA helicase family
.
Nat. Rev. Mol. Cell Biol.
12
,
505
516
42
Andersen
,
C.B.
,
Ballut
,
L.
,
Johansen
,
J.S.
,
Chamieh
,
H.
,
Nielsen
,
K.H.
,
Oliveira
,
C.L.
et al (
2006
)
Structure of the exon junction core complex with a trapped DEAD-box ATPase bound to RNA
.
Science
313
,
1968
1972
43
Bono
,
F.
,
Ebert
,
J.
,
Lorentzen
,
E.
and
Conti
,
E.
(
2006
)
The crystal structure of the exon junction complex reveals how it maintains a stable grip on mRNA
.
Cell
126
,
713
725
44
Nielsen
,
K.H.
,
Chamieh
,
H.
,
Andersen
,
C.B.
,
Fredslund
,
F.
,
Hamborg
,
K.
,
Le Hir
,
H.
et al (
2009
)
Mechanism of ATP turnover inhibition in the EJC
.
RNA
15
,
67
75
45
Fribourg
,
S.
,
Gatfield
,
D.
,
Izaurralde
,
E.
and
Conti
,
E.
(
2003
)
A novel mode of RBD-protein recognition in the Y14-Mago complex
.
Nat. Struct. Biol.
10
,
433
439
46
Ballut
,
L.
,
Marchadier
,
B.
,
Baguet
,
A.
,
Tomasetto
,
C.
,
Séraphin
,
B.
and
Le Hir
,
H.
(
2005
)
The exon junction core complex is locked onto RNA by inhibition of eIF4AIII ATPase activity
.
Nat. Struct. Mol. Biol.
12
,
861
869
47
Gerbracht
,
J.V.
,
Boehm
,
V.
,
Britto-Borges
,
T.
,
Kallabis
,
S.
,
Wiederstein
,
J.L.
,
Ciriello
,
S.
et al (
2020
)
CASC3 promotes transcriptome-wide activation of nonsense-mediated decay by the exon junction complex
.
Nucleic Acids Res.
48
,
8626
8644
48
Kashima
,
I.
,
Jonas
,
S.
,
Jayachandran
,
U.
,
Buchwald
,
G.
,
Conti
,
E.
,
Lupas
,
A.N.
et al (
2010
)
SMG6 interacts with the exon junction complex via two conserved EJC-binding motifs (EBMs) required for nonsense-mediated mRNA decay
.
Genes Dev.
24
,
2440
2450
49
Buchwald
,
G.
,
Ebert
,
J.
,
Basquin
,
C.
,
Sauliere
,
J.
,
Jayachandran
,
U.
,
Bono
,
F.
et al (
2010
)
Insights into the recruitment of the NMD machinery from the crystal structure of a core EJC-UPF3b complex
.
Proc. Natl Acad. Sci. U.S.A.
107
,
10050
10055
50
Wallmeroth
,
D.
,
Lackmann
,
J.-W.
,
Kueckelmann
,
S.
,
Altmüller
,
J.
,
Dieterich
,
C.
,
Boehm
,
V.
et al (
2022
)
Human UPF3A and UPF3B enable fault-tolerant activation of nonsense-mediated mRNA decay
.
EMBO J.
, e109191
51
Yi
,
Z.
,
Arvola
,
R.M.
,
Myers
,
S.
,
Dilsavor
,
C.N.
,
Abu Alhasan
,
R.
,
Carter
,
B.N.
et al (
2022
)
Mammalian UPF3A and UPF3B can activate nonsense-mediated mRNA decay independently of their exon junction complex binding
.
EMBO J.
, e109202
52
Melero
,
R.
,
Buchwald
,
G.
,
Castaño
,
R.
,
Raabe
,
M.
,
Gil
,
D.
,
Lázaro
,
M.
et al (
2012
)
The cryo-EM structure of the UPF-EJC complex shows UPF1 poised toward the RNA 3′ end
.
Nat. Struct. Mol. Biol.
19
,
498
505
,
s491-492
53
Lykke-Andersen
,
J.
,
Shu
,
M.D.
and
Steitz
,
J.A.
(
2001
)
Communication of the position of exon-exon junctions to the mRNA surveillance machinery by the protein RNPS1
.
Science
293
,
1836
1839
54
Schwerk
,
C.
,
Prasad
,
J.
,
Degenhardt
,
K.
,
Erdjument-Bromage
,
H.
,
White
,
E.
,
Tempst
,
P.
et al (
2003
)
ASAP, a novel protein complex involved in RNA processing and apoptosis
.
Mol. Cell Biol.
23
,
2981
2990
55
Tange
,
T.
,
Shibuya
,
T.
,
Jurica
,
M.S.
and
Moore
,
M.J.
(
2005
)
Biochemical analysis of the EJC reveals two new factors and a stable tetrameric protein core
.
RNA
11
,
1869
1883
56
Tange
,
T.
,
Nott
,
A.
and
Moore
,
M.J.
(
2004
)
The ever-increasing complexities of the exon junction complex
.
Curr. Opin. Cell Biol.
16
,
279
284
57
Le Hir
,
H.
,
Saulière
,
J.
and
Wang
,
Z.
(
2016
)
The exon junction complex as a node of post-transcriptional networks
.
Nat. Rev. Mol. Cell Biol.
17
,
41
54
58
Gehring
,
N.H.
,
Kunz
,
J.B.
,
Neu-Yilik
,
G.
,
Breit
,
S.
,
Viegas
,
M.H.
,
Hentze
,
M.W.
et al (
2005
)
Exon-junction complex components specify distinct routes of nonsense-mediated mRNA decay with differential cofactor requirements
.
Mol. Cell
20
,
65
75
59
Gehring
,
N.H.
,
Lamprinaki
,
S.
,
Hentze
,
M.W.
and
Kulozik
,
A.E.
(
2009
)
The hierarchy of exon-junction complex assembly by the spliceosome explains key features of mammalian nonsense-mediated mRNA decay
.
PLoS Biol.
7
,
e1000120
60
Sakashita
,
E.
,
Tatsumi
,
S.
,
Werner
,
D.
,
Endo
,
H.
and
Mayeda
,
A.
(
2004
)
Human RNPS1 and its associated factors: a versatile alternative pre-mRNA splicing regulator in vivo
.
Mol. Cell Biol.
24
,
1174
1187
61
Hogg
,
J.R.
and
Goff
,
S.P.
(
2010
)
Upf1 senses 3'UTR length to potentiate mRNA decay
.
Cell
143
,
379
389
62
Zünd
,
D.
,
Gruber
,
A.R.
,
Zavolan
,
M.
and
Mühlemann
,
O.
(
2013
)
Translation-dependent displacement of UPF1 from coding sequences causes its enrichment in 3′ UTRs
.
Nat. Struct. Mol. Biol.
20
,
936
943
63
Kurosaki
,
T.
and
Maquat
,
L.E.
(
2013
)
Rules that govern UPF1 binding to mRNA 3′ UTRs
.
Proc. Natl Acad. Sci. U.S.A.
110
,
3357
3362
64
Hurt
,
J.A.
,
Robertson
,
A.D.
and
Burge
,
C.B.
(
2013
)
Global analyses of UPF1 binding and function reveal expanded scope of nonsense-mediated mRNA decay
.
Genome Res.
23
,
1636
1650
65
Kadlec
,
J.
,
Guilligay
,
D.
,
Ravelli
,
R.B.
and
Cusack
,
S.
(
2006
)
Crystal structure of the UPF2-interacting domain of nonsense-mediated mRNA decay factor UPF1
.
RNA
12
,
1817
1824
66
Cheng
,
Z.
,
Muhlrad
,
D.
,
Lim
,
M.K.
,
Parker
,
R.
and
Song
,
H.
(
2007
)
Structural and functional insights into the human Upf1 helicase core
.
EMBO J.
26
,
253
264
67
Clerici
,
M.
,
Mourão
,
A.
,
Gutsche
,
I.
,
Gehring
,
N.H.
,
Hentze
,
M.W.
,
Kulozik
,
A.
et al (
2009
)
Unusual bipartite mode of interaction between the nonsense-mediated decay factors, UPF1 and UPF2
.
EMBO J.
28
,
2293
2306
68
Gowravaram
,
M.
,
Bonneau
,
F.
,
Kanaan
,
J.
,
Maciej
,
V.D.
,
Fiorini
,
F.
,
Raj
,
S.
et al (
2018
)
A conserved structural element in the RNA helicase UPF1 regulates its catalytic activity in an isoform-specific manner
.
Nucleic Acids Res.
46
,
2648
2659
69
Durand
,
S.
,
Franks
,
T.M.
and
Lykke-Andersen
,
J.
(
2016
)
Hyperphosphorylation amplifies UPF1 activity to resolve stalls in nonsense-mediated mRNA decay
.
Nat. Commun.
7
,
12434
70
Okada-Katsuhata
,
Y.
,
Yamashita
,
A.
,
Kutsuzawa
,
K.
,
Izumi
,
N.
,
Hirahara
,
F.
and
Ohno
,
S.
(
2012
)
N- and C-terminal Upf1 phosphorylations create binding platforms for SMG-6 and SMG-5:SMG-7 during NMD
.
Nucleic Acids Res.
40
,
1251
1266
71
Chakrabarti
,
S.
,
Bonneau
,
F.
,
Schüssler
,
S.
,
Eppinger
,
E.
and
Conti
,
E.
(
2014
)
Phospho-dependent and phospho-independent interactions of the helicase UPF1 with the NMD factors SMG5-SMG7 and SMG6
.
Nucleic Acids Res.
42
,
9447
9460
72
Nicholson
,
P.
,
Josi
,
C.
,
Kurosawa
,
H.
,
Yamashita
,
A.
and
Mühlemann
,
O.
(
2014
)
A novel phosphorylation-independent interaction between SMG6 and UPF1 is essential for human NMD
.
Nucleic Acids Res.
42
,
9217
9235
73
Takahashi
,
S.
,
Araki
,
Y.
,
Ohya
,
Y.
,
Sakuno
,
T.
,
Hoshino
,
S.
,
Kontani
,
K.
et al (
2008
)
Upf1 potentially serves as a RING-related E3 ubiquitin ligase via its association with Upf3 in yeast
.
RNA
14
,
1950
1958
74
Feng
,
Q.
,
Jagannathan
,
S.
and
Bradley
,
R.K.
(
2017
)
The RNA surveillance factor UPF1 represses myogenesis via its E3 ubiquitin ligase activity
.
Mol. Cell
67
,
239
251.e236
75
Joazeiro
,
C.A.P.
(
2019
)
Mechanisms and functions of ribosome-associated protein quality control
.
Nat. Rev. Mol. Cell Biol.
20
,
368
383
76
Powers
,
K.T.
,
Szeto
,
J.A.
and
Schaffitzel
,
C.
(
2020
)
New insights into no-go, non-stop and nonsense-mediated mRNA decay complexes
.
Curr. Opin. Struct. Biol.
65
,
110
118
77
Chu
,
V.
,
Feng
,
Q.
,
Lim
,
Y.
and
Shao
,
S.
(
2021
)
Selective destabilization of polypeptides synthesized from NMD-targeted transcripts
.
Mol. Biol. Cell
32
,
ar38
78
Inglis
,
A.J.
,
Guna
,
A.
,
Merchán
,
Á.G
,
Pal
,
A.
,
Esantsi
,
T.K.
,
Keys
,
H.R.
et al (
2022
)
Coupled protein quality control during nonsense mediated mRNA decay
.
bioRxiv
79
Weng
,
Y.
,
Czaplinski
,
K.
and
Peltz
,
S.W.
(
1998
)
ATP is a cofactor of the Upf1 protein that modulates its translation termination and RNA binding activities
.
RNA
4
,
205
214
PMID:
[PubMed]
80
Franks
,
T.M.
,
Singh
,
G.
and
Lykke-Andersen
,
J.
(
2010
)
Upf1 ATPase-dependent mRNP disassembly is required for completion of nonsense- mediated mRNA decay
.
Cell
143
,
938
950
81
Fritz
,
S.E.
,
Ranganathan
,
S.
,
Wang
,
C.D.
and
Hogg
,
J.R.
(
2022
)
An alternative UPF1 isoform drives conditional remodeling of nonsense-mediated mRNA decay
.
EMBO J.
, e108898
82
Padariya
,
M.
,
Fahraeus
,
R.
,
Hupp
,
T.
and
Kalathiya
,
U.
(
2021
)
Molecular determinants and specificity of mRNA with alternatively-Spliced UPF1 isoforms, influenced by an insertion in the ‘regulatory loop’
.
Int. J. Mol. Sci.
22
,
12744
83
Ge
,
Z.
,
Quek
,
B.L.
,
Beemon
,
K.L.
and
Hogg
,
J.R.
(
2016
)
Polypyrimidine tract binding protein 1 protects mRNAs from recognition by the nonsense-mediated mRNA decay pathway
.
ELife
5
,
e11155
84
Kishor
,
A.
,
Ge
,
Z.
and
Hogg
,
J.R.
(
2019
)
hnRNP L-dependent protection of normal mRNAs from NMD subverts quality control in B cell lymphoma
.
EMBO J.
38
,
e99128
85
Kishor
,
A.
,
Fritz
,
S.E.
,
Haque
,
N.
,
Ge
,
Z.
,
Tunc
,
I.
,
Yang
,
W.
et al (
2020
)
Activation and inhibition of nonsense-mediated mRNA decay control the abundance of alternative polyadenylation products
.
Nucleic Acids Res.
48
,
7468
7482
86
Kadlec
,
J.
,
Izaurralde
,
E.
and
Cusack
,
S.
(
2004
)
The structural basis for the interaction between nonsense-mediated mRNA decay factors UPF2 and UPF3
.
Nat. Struct. Mol. Biol.
11
,
330
337
87
Clerici
,
M.
,
Deniaud
,
A.
,
Boehm
,
V.
,
Gehring
,
N.H.
,
Schaffitzel
,
C.
and
Cusack
,
S.
(
2014
)
Structural and functional analysis of the three MIF4G domains of nonsense-mediated decay factor UPF2
.
Nucleic Acids Res.
42
,
2673
2686
88
Fourati
,
Z.
,
Roy
,
B.
,
Millan
,
C.
,
Coureux
,
P.D.
,
Kervestin
,
S.
,
van Tilbeurgh
,
H.
et al (
2014
)
A highly conserved region essential for NMD in the Upf2 N-terminal domain
.
J. Mol. Biol.
426
,
3689
3702
89
Bufton
,
J.C.
,
Powers
,
K.T.
,
Szeto
,
J.-Y.A.
and
Schaffitzel
,
C.
(
2022
)
Molecular basis of neurodevelopmental disorder-causing mutation in nonsense-mediated mRNA decay factor UPF3B
.
bioRxiv
90
López-Perrote
,
A.
,
Castaño
,
R.
,
Melero
,
R.
,
Zamarro
,
T.
,
Kurosawa
,
H.
,
Ohnishi
,
T.
et al (
2016
)
Human nonsense-mediated mRNA decay factor UPF2 interacts directly with eRF3 and the SURF complex
.
Nucleic Acids Res.
44
,
1909
1923
91
Neu-Yilik
,
G.
,
Raimondeau
,
E.
,
Eliseev
,
B.
,
Yeramala
,
L.
,
Amthor
,
B.
,
Deniaud
,
A.
et al (
2017
)
Dual function of UPF3B in early and late translation termination
.
EMBO J.
36
,
2968
2986
92
Knott
,
G.J.
,
Bond
,
C.S.
and
Fox
,
A.H.
(
2016
)
The DBHS proteins SFPQ, NONO and PSPC1: a multipurpose molecular scaffold
.
Nucleic Acids Res.
44
,
3989
4004
93
Tarpey
,
P.S.
,
Raymond
,
F.L.
,
Nguyen
,
L.S.
,
Rodriguez
,
J.
,
Hackett
,
A.
,
Vandeleur
,
L.
et al (
2007
)
Mutations in UPF3B, a member of the nonsense-mediated mRNA decay complex, cause syndromic and nonsyndromic mental retardation
.
Nat. Genet.
39
,
1127
1133
94
Alrahbeni
,
T.
,
Sartor
,
F.
,
Anderson
,
J.
,
Miedzybrodzka
,
Z.
,
McCaig
,
C.
and
Müller
,
B.
(
2015
)
Full UPF3B function is critical for neuronal differentiation of neural stem cells
.
Mol. Brain
8
,
33
95
Chan
,
W.K.
,
Bhalla
,
A.D.
,
Le Hir
,
H.
,
Nguyen
,
L.S.
,
Huang
,
L.
,
Gécz
,
J.
et al (
2009
)
A UPF3-mediated regulatory switch that maintains RNA surveillance
.
Nat. Struct. Mol. Biol.
16
,
747
753
96
Nguyen
,
L.S.
,
Jolly
,
L.
,
Shoubridge
,
C.
,
Chan
,
W.K.
,
Huang
,
L.
,
Laumonnier
,
F.
et al (
2012
)
Transcriptome profiling of UPF3B/NMD-deficient lymphoblastoid cells from patients with various forms of intellectual disability
.
Mol. Psychiatry
17
,
1103
1115
97
Kunz
,
J.B.
,
Neu-Yilik
,
G.
,
Hentze
,
M.W.
,
Kulozik
,
A.E.
and
Gehring
,
N.H.
(
2006
)
Functions of hUpf3a and hUpf3b in nonsense-mediated mRNA decay and translation
.
RNA
12
,
1015
1022
98
Shum
,
E.Y.
,
Jones
,
S.H.
,
Shao
,
A.
,
Dumdie
,
J.
,
Krause
,
M.D.
,
Chan
,
W.K.
et al (
2016
)
The antagonistic gene paralogs Upf3a and Upf3b govern nonsense-mediated RNA decay
.
Cell
165
,
382
395
99
Hauer
,
C.
,
Sieber
,
J.
,
Schwarzl
,
T.
,
Hollerer
,
I.
,
Curk
,
T.
,
Alleaume
,
A.M.
et al (
2016
)
Exon junction complexes show a distributional bias toward alternatively spliced mRNAs and against mRNAs coding for ribosomal proteins
.
Cell Rep.
16
,
1588
1603
100
Zetoune
,
A.B.
,
Fontanière
,
S.
,
Magnin
,
D.
,
Anczuków
,
O.
,
Buisson
,
M.
,
Zhang
,
C.X.
et al (
2008
)
Comparison of nonsense-mediated mRNA decay efficiency in various murine tissues
.
BMC Genet.
9
,
83
101
Imseng
,
S.
,
Aylett
,
C.H.
and
Maier
,
T.
(
2018
)
Architecture and activation of phosphatidylinositol 3-kinase related kinases
.
Curr. Opin. Struct. Biol.
49
,
177
189
102
Yamashita
,
A.
,
Ohnishi
,
T.
,
Kashima
,
I.
,
Taya
,
Y.
and
Ohno
,
S.
(
2001
)
Human SMG-1, a novel phosphatidylinositol 3-kinase-related protein kinase, associates with components of the mRNA surveillance complex and is involved in the regulation of nonsense-mediated mRNA decay
.
Genes Dev.
15
,
2215
2228
103
Brumbaugh
,
K.M.
,
Otterness
,
D.M.
,
Geisen
,
C.
,
Oliveira
,
V.
,
Brognard
,
J.
,
Li
,
X.
et al (
2004
)
The mRNA surveillance protein hSMG-1 functions in genotoxic stress response pathways in mammalian cells
.
Mol. Cell
14
,
585
598
104
Arias-Palomo
,
E.
,
Yamashita
,
A.
,
Fernández
,
I.S.
,
Núñez-Ramírez
,
R.
,
Bamba
,
Y.
,
Izumi
,
N.
et al (
2011
)
The nonsense-mediated mRNA decay SMG-1 kinase is regulated by large-scale conformational changes controlled by SMG-8
.
Genes Dev.
25
,
153
164
105
Deniaud
,
A.
,
Karuppasamy
,
M.
,
Bock
,
T.
,
Masiulis
,
S.
,
Huard
,
K.
,
Garzoni
,
F.
et al (
2015
)
A network of SMG-8, SMG-9 and SMG-1 C-terminal insertion domain regulates UPF1 substrate recruitment and phosphorylation
.
Nucleic Acids Res.
43
,
7600
7611
106
Zhu
,
L.
,
Li
,
L.
,
Qi
,
Y.
,
Yu
,
Z.
and
Xu
,
Y.
(
2019
)
Cryo-EM structure of SMG1-SMG8-SMG9 complex
.
Cell Res.
29
,
1027
1034
107
Gat
,
Y.
,
Schuller
,
J.M.
,
Lingaraju
,
M.
,
Weyher
,
E.
,
Bonneau
,
F.
,
Strauss
,
M.
et al (
2019
)
Insp(6) binding to PIKK kinases revealed by the cryo-EM structure of an SMG1-SMG8-SMG9 complex
.
Nat. Struct. Mol. Biol.
26
,
1089
1093
108
Langer
,
L.M.
,
Gat
,
Y.
,
Bonneau
,
F.
and
Conti
,
E.
(
2020
)
Structure of substrate-bound SMG1-8-9 kinase complex reveals molecular basis for phosphorylation specificity
.
ELife
9
,
e57127
109
Li
,
L.
,
Lingaraju
,
M.
,
Basquin
,
C.
,
Basquin
,
J.
and
Conti
,
E.
(
2017
)
Structure of a SMG8-SMG9 complex identifies a G-domain heterodimer in the NMD effector proteins
.
RNA
23
,
1028
1034
110
Langer
,
L.M.
,
Bonneau
,
F.
,
Gat
,
Y.
and
Conti
,
E.
(
2021
)
Cryo-EM reconstructions of inhibitor-bound SMG1 kinase reveal an autoinhibitory state dependent on SMG8
.
ELife
10
,
e72353
111
Ohnishi
,
T.
,
Yamashita
,
A.
,
Kashima
,
I.
,
Schell
,
T.
,
Anders
,
K.R.
,
Grimson
,
A.
et al (
2003
)
Phosphorylation of hUPF1 induces formation of mRNA surveillance complexes containing hSMG-5 and hSMG-7
.
Mol. Cell
12
,
1187
1200
112
Kurosaki
,
T.
,
Li
,
W.
,
Hoque
,
M.
,
Popp
,
M.W.
,
Ermolenko
,
D.N.
,
Tian
,
B.
et al (
2014
)
A post-translational regulatory switch on UPF1 controls targeted mRNA degradation
.
Genes Dev.
28
,
1900
1916
113
Jonas
,
S.
,
Weichenrieder
,
O.
and
Izaurralde
,
E.
(
2013
)
An unusual arrangement of two 14-3-3-like domains in the SMG5-SMG7 heterodimer is required for efficient nonsense-mediated mRNA decay
.
Genes Dev.
27
,
211
225
114
Unterholzner
,
L.
and
Izaurralde
,
E.
(
2004
)
SMG7 acts as a molecular link between mRNA surveillance and mRNA decay
.
Mol. Cell
16
,
587
596
115
Glavan
,
F.
,
Behm-Ansmant
,
I.
,
Izaurralde
,
E.
and
Conti
,
E.
(
2006
)
Structures of the PIN domains of SMG6 and SMG5 reveal a nuclease within the mRNA surveillance complex
.
EMBO J.
25
,
5117
5125
116
Fukuhara
,
N.
,
Ebert
,
J.
,
Unterholzner
,
L.
,
Lindner
,
D.
,
Izaurralde
,
E.
and
Conti
,
E.
(
2005
)
SMG7 is a 14-3-3-like adaptor in the nonsense-mediated mRNA decay pathway
.
Mol. Cell
17
,
537
547
117
Loh
,
B.
,
Jonas
,
S.
and
Izaurralde
,
E.
(
2013
)
The SMG5-SMG7 heterodimer directly recruits the CCR4-NOT deadenylase complex to mRNAs containing nonsense codons via interaction with POP2
.
Genes Dev.
27
,
2125
2138
118
Gatfield
,
D.
and
Izaurralde
,
E.
(
2004
)
Nonsense-mediated messenger RNA decay is initiated by endonucleolytic cleavage in Drosophila
.
Nature
429
,
575
578
119
Schmidt
,
S.A.
,
Foley
,
P.L.
,
Jeong
,
D.H.
,
Rymarquis
,
L.A.
,
Doyle
,
F.
,
Tenenbaum
,
S.A.
et al (
2015
)
Identification of SMG6 cleavage sites and a preferred RNA cleavage motif by global analysis of endogenous NMD targets in human cells
.
Nucleic Acids Res.
43
,
309
323
120
Boehm
,
V.
,
Kueckelmann
,
S.
,
Gerbracht
,
J.V.
,
Kallabis
,
S.
,
Britto-Borges
,
T.
,
Altmüller
,
J.
et al (
2021
)
SMG5-SMG7 authorize nonsense-mediated mRNA decay by enabling SMG6 endonucleolytic activity
.
Nat. Commun.
12
,
3965
121
Metze
,
S.
,
Herzog
,
V.A.
,
Ruepp
,
M.D.
and
Mühlemann
,
O.
(
2013
)
Comparison of EJC-enhanced and EJC-independent NMD in human cells reveals two partially redundant degradation pathways
.
RNA
19
,
1432
1448
122
Colombo
,
M.
,
Karousis
,
E.D.
,
Bourquin
,
J.
,
Bruggmann
,
R.
and
Mühlemann
,
O.
(
2017
)
Transcriptome-wide identification of NMD-targeted human mRNAs reveals extensive redundancy between SMG6- and SMG7-mediated degradation pathways
.
RNA
23
,
189
201
123
Mendell
,
J.T.
,
Sharifi
,
N.A.
,
Meyers
,
J.L.
,
Martinez-Murillo
,
F.
and
Dietz
,
H.C.
(
2004
)
Nonsense surveillance regulates expression of diverse classes of mammalian transcripts and mutes genomic noise
.
Nat. Genet.
36
,
1073
1078
124
Kebaara
,
B.W.
and
Atkin
,
A.L.
(
2009
)
Long 3′-UTRs target wild-type mRNAs for nonsense-mediated mRNA decay in saccharomyces cerevisiae
.
Nucleic Acids Res.
37
,
2771
2778
125
McGlincy
,
N.J.
and
Smith
,
C.W.
(
2008
)
Alternative splicing resulting in nonsense-mediated mRNA decay: what is the meaning of nonsense?
Trends Biochem. Sci.
33
,
385
393
126
Bicknell
,
A.A.
,
Cenik
,
C.
,
Chua
,
H.N.
,
Roth
,
F.P.
and
Moore
,
M.J.
(
2012
)
Introns in UTRs: why we should stop ignoring them
.
Bioessays
34
,
1025
1034
127
Wilusz
,
J.E.
and
Wilusz
,
J.
(
2014
)
Nonsense-mediated RNA decay: at the ‘cutting edge’ of regulated snoRNA production
.
Genes Dev.
28
,
2447
2449
128
Yi
,
Z.
,
Sanjeev
,
M.
and
Singh
,
G.
(
2021
)
The branched nature of the nonsense-mediated mRNA decay pathway
.
Trends Genet.
37
,
143
159
129
Chan
,
W.K.
,
Huang
,
L.
,
Gudikote
,
J.P.
,
Chang
,
Y.F.
,
Imam
,
J.S.
,
MacLean
, II,
J.A.
et al (
2007
)
An alternative branch of the nonsense-mediated decay pathway
.
EMBO J.
26
,
1820
1830
130
Sato
,
H.
,
Hosoda
,
N.
and
Maquat
,
L.E.
(
2008
)
Efficiency of the pioneer round of translation affects the cellular site of nonsense-mediated mRNA decay
.
Mol. Cell
29
,
255
262
131
Aznarez
,
I.
,
Nomakuchi
,
T.T.
,
Tetenbaum-Novatt
,
J.
,
Rahman
,
M.A.
,
Fregoso
,
O.
,
Rees
,
H.
et al (
2018
)
Mechanism of nonsense-Mediated mRNA decay stimulation by splicing factor SRSF1
.
Cell Rep.
23
,
2186
2198
132
Zhang
,
Z.
and
Krainer
,
A.R.
(
2004
)
Involvement of SR proteins in mRNA surveillance
.
Mol. Cell
16
,
597
607
133
Michlewski
,
G.
,
Sanford
,
J.R.
and
Cáceres
,
J.F.
(
2008
)
The splicing factor SF2/ASF regulates translation initiation by enhancing phosphorylation of 4E-BP1
.
Mol. Cell
30
,
179
189
134
Chiu
,
S.Y.
,
Serin
,
G.
,
Ohara
,
O.
and
Maquat
,
L.E.
(
2003
)
Characterization of human Smg5/7a: a protein with similarities to caenorhabditis elegans SMG5 and SMG7 that functions in the dephosphorylation of Upf1
.
RNA
9
,
77
87
135
Anders
,
K.R.
,
Grimson
,
A.
and
Anderson
,
P.
(
2003
)
SMG-5, required for C.elegans nonsense-mediated mRNA decay, associates with SMG-2 and protein phosphatase 2A
.
EMBO J.
22
,
641
650
136
Longman
,
D.
,
Hug
,
N.
,
Keith
,
M.
,
Anastasaki
,
C.
,
Patton
,
E.E.
,
Grimes
,
G.
et al (
2013
)
DHX34 and NBAS form part of an autoregulatory NMD circuit that regulates endogenous RNA targets in human cells, zebrafish and Caenorhabditis elegans
.
Nucleic Acids Res.
41
,
8319
8331
137
Sakaki
,
K.
and
Kaufman
,
R.J.
(
2013
)
Interaction between quality control systems for ER protein folding and RNA biogenesis
.
Worm
2
,
e23005
138
Sieber
,
J.
,
Hauer
,
C.
,
Bhuvanagiri
,
M.
,
Leicht
,
S.
,
Krijgsveld
,
J.
,
Neu-Yilik
,
G.
et al (
2016
)
Proteomic analysis reveals branch-specific regulation of the unfolded protein response by nonsense-mediated mRNA decay
.
Mol. Cell. Proteomics
15
,
1584
1597
139
Sakaki
,
K.
,
Yoshina
,
S.
,
Shen
,
X.
,
Han
,
J.
,
DeSantis
,
M.R.
,
Xiong
,
M.
et al (
2012
)
RNA surveillance is required for endoplasmic reticulum homeostasis
.
Proc. Natl Acad. Sci. U.S.A.
109
,
8079
8084
140
Karam
,
R.
,
Lou
,
C.H.
,
Kroeger
,
H.
,
Huang
,
L.
,
Lin
,
J.H.
and
Wilkinson
,
M.F.
(
2015
)
The unfolded protein response is shaped by the NMD pathway
.
EMBO Rep.
16
,
599
609
141
Longman
,
D.
,
Jackson-Jones
,
K.A.
,
Maslon
,
M.M.
,
Murphy
,
L.C.
,
Young
,
R.S.
,
Stoddart
,
J.J.
et al (
2020
)
Identification of a localized nonsense-mediated decay pathway at the endoplasmic reticulum
.
Genes Dev.
34
,
1075
1088
142
Anastasaki
,
C.
,
Longman
,
D.
,
Capper
,
A.
,
Patton
,
E.E.
and
Cáceres
,
J.F.
(
2011
)
Dhx34 and nbas function in the NMD pathway and are required for embryonic development in zebrafish
.
Nucleic Acids Res.
39
,
3686
3694
143
Karam
,
R.
,
Wengrod
,
J.
,
Gardner
,
L.B.
and
Wilkinson
,
M.F.
(
2013
)
Regulation of nonsense-mediated mRNA decay: implications for physiology and disease
.
Biochim. Biophys. Acta
1829
,
624
633
144
Bao
,
J.
,
Tang
,
C.
,
Yuan
,
S.
,
Porse
,
B.T.
and
Yan
,
W.
(
2015
)
UPF2, a nonsense-mediated mRNA decay factor, is required for prepubertal sertoli cell development and male fertility by ensuring fidelity of the transcriptome
.
Development
142
,
352
362
145
Powers
,
K.T.
,
Stevenson-Jones
,
F.
,
Yadav
,
S.K.N.
,
Amthor
,
B.
,
Bufton
,
J.C.
,
Borucu
,
U.
et al (
2021
)
Blasticidin S inhibits mammalian translation and enhances production of protein encoded by nonsense mRNA
.
Nucleic Acids Res.
49
,
7665
7679
146
Garcia
,
D.
,
Garcia
,
S.
and
Voinnet
,
O.
(
2014
)
Nonsense-mediated decay serves as a general viral restriction mechanism in plants
.
Cell Host Microbe
16
,
391
402
147
Dickson
,
A.M.
and
Wilusz
,
J.
(
2011
)
Strategies for viral RNA stability: live long and prosper
.
Trends Genet.
27
,
286
293
148
Hogg
,
J.R.
(
2016
)
Viral evasion and manipulation of host RNA quality control pathways
.
J. Virol.
90
,
7010
7018
149
Leon
,
K.
and
Ott
,
M.
(
2021
)
An ‘Arms race’ between the nonsense-mediated mRNA decay pathway and viral infections
.
Semin. Cell Dev. Biol.
111
,
101
107
150
Popp
,
M.W.
,
Cho
,
H.
and
Maquat
,
L.E.
(
2020
)
Viral subversion of nonsense-mediated mRNA decay
.
RNA
26
,
1509
1518
151
Withers
,
J.B.
and
Beemon
,
K.L.
(
2010
)
Structural features in the Rous sarcoma virus RNA stability element are necessary for sensing the correct termination codon
.
Retrovirology
7
,
65
152
Balagopal
,
V.
and
Beemon
,
K.L.
(
2017
)
Rous sarcoma virus RNA stability element inhibits deadenylation of mRNAs with long 3′UTRs
.
Viruses
9
,
204
153
Desbois
,
C.
,
Rousset
,
R.
,
Bantignies
,
F.
and
Jalinot
,
P.
(
1996
)
Exclusion of Int-6 from PML nuclear bodies by binding to the HTLV-I Tax oncoprotein
.
Science
273
,
951
953
154
Hakata
,
Y.
,
Umemoto
,
T.
,
Matsushita
,
S.
and
Shida
,
H.
(
1998
)
Involvement of human CRM1 (exportin 1) in the export and multimerization of the Rex protein of human T-cell leukemia virus type 1
.
J. Virol.
72
,
6602
6607
155
Nakano
,
K.
and
Watanabe
,
T.
(
2016
)
HTLV-1 Rex tunes the cellular environment favorable for viral replication
.
Viruses
8
,
58
156
Fiorini
,
F.
,
Robin
,
J.P.
,
Kanaan
,
J.
,
Borowiak
,
M.
,
Croquette
,
V.
,
Le Hir
,
H.
et al (
2018
)
HTLV-1 Tax plugs and freezes UPF1 helicase leading to nonsense-mediated mRNA decay inhibition
.
Nat. Commun.
9
,
431
157
Mocquet
,
V.
,
Neusiedler
,
J.
,
Rende
,
F.
,
Cluet
,
D.
,
Robin
,
J.P.
,
Terme
,
J.M.
et al (
2012
)
The human T-lymphotropic virus type 1 tax protein inhibits nonsense-mediated mRNA decay by interacting with INT6/EIF3E and UPF1
.
J. Virol.
86
,
7530
7543
158
Prochasson
,
L.
,
Jalinot
,
P.
and
Mocquet
,
V.
(
2020
)
The complex relationship between HTLV-1 and nonsense-mediated mRNA decay (NMD)
.
Pathogens
9
,
287
159
Decker
,
C.J.
and
Parker
,
R.
(
2012
)
P-bodies and stress granules: possible roles in the control of translation and mRNA degradation
.
Cold Spring Harb. Perspect. Biol.
4
,
a012286
160
Pulak
,
R.
and
Anderson
,
P.
(
1993
)
mRNA surveillance by the caenorhabditis elegans smg genes
.
Genes Dev.
7
,
1885
1897
161
He
,
F.
,
Brown
,
A.H.
and
Jacobson
,
A.
(
1997
)
Upf1p, Nmd2p, and Upf3p are interacting components of the yeast nonsense-mediated mRNA decay pathway
.
Mol. Cell Biol.
17
,
1580
1594
162
Medghalchi
,
S.M.
,
Frischmeyer
,
P.A.
,
Mendell
,
J.T.
,
Kelly
,
A.G.
,
Lawler
,
A.M.
and
Dietz
,
H.C.
(
2001
)
Rent1, a trans-effector of nonsense-mediated mRNA decay, is essential for mammalian embryonic viability
.
Hum. Mol. Genet.
10
,
99
105
163
Weischenfeldt
,
J.
,
Damgaard
,
I.
,
Bryder
,
D.
,
Theilgaard-Mönch
,
K.
,
Thoren
,
L.A.
,
Nielsen
,
F.C.
et al (
2008
)
NMD is essential for hematopoietic stem and progenitor cells and for eliminating by-products of programmed DNA rearrangements
.
Genes Dev.
22
,
1381
1396
164
Wittkopp
,
N.
,
Huntzinger
,
E.
,
Weiler
,
C.
,
Saulière
,
J.
,
Schmidt
,
S.
,
Sonawane
,
M.
et al (
2009
)
Nonsense-mediated mRNA decay effectors are essential for zebrafish embryonic development and survival
.
Mol. Cell Biol.
29
,
3517
3528
165
Mort
,
M.
,
Ivanov
,
D.
,
Cooper
,
D.N.
and
Chuzhanova
,
N.A.
(
2008
)
A meta-analysis of nonsense mutations causing human genetic disease
.
Hum. Mutat.
29
,
1037
1047
166
Lau
,
C.K.
,
Diem
,
M.D.
,
Dreyfuss
,
G.
and
Van Duyne
,
G.D.
(
2003
)
Structure of the Y14-Magoh core of the exon junction complex
.
Curr. Biol.
13
,
933
941
167
Murachelli
,
A.G.
,
Ebert
,
J.
,
Basquin
,
C.
,
Le Hir
,
H.
and
Conti
,
E.
(
2012
)
The structure of the ASAP core complex reveals the existence of a Pinin-containing PSAP complex
.
Nat. Struct. Mol. Biol.
19
,
378
386

Author notes

*

These authors contributed equally to this work.

This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY).