Next Article in Journal
Michael Adduct of Sulfonamide Chalcone Targets Folate Metabolism in Brugia Malayi Parasite
Next Article in Special Issue
Identifying Potent Nonsense-Mediated mRNA Decay Inhibitors with a Novel Screening System
Previous Article in Journal
Overcoming Barriers to Wound Healing in a Neuropathic and Neuro-Ischaemic Diabetic Foot Cohort Using a Novel Bilayer Biodegradable Synthetic Matrix
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nonsense-Mediated mRNA Decay Factor Functions in Human Health and Disease

by
Lingling Sun
1,†,
Justine Mailliot
1,*,† and
Christiane Schaffitzel
1,2,*
1
School of Biochemistry, University of Bristol, University Walk, Bristol BS8 1TD, UK
2
Bristol Engineering Biology Centre BrisEngBio, 24 Tyndall Ave, Bristol BS8 1TQ, UK
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Biomedicines 2023, 11(3), 722; https://doi.org/10.3390/biomedicines11030722
Submission received: 1 February 2023 / Revised: 21 February 2023 / Accepted: 22 February 2023 / Published: 27 February 2023

Abstract

:
Nonsense-mediated mRNA decay (NMD) is a cellular surveillance mechanism that degrades mRNAs with a premature stop codon, avoiding the synthesis of C-terminally truncated proteins. In addition to faulty mRNAs, NMD recognises ~10% of endogenous transcripts in human cells and downregulates their expression. The up-frameshift proteins are core NMD factors and are conserved from yeast to human in structure and function. In mammals, NMD diversified into different pathways that target different mRNAs employing additional NMD factors. Here, we review our current understanding of molecular mechanisms and cellular roles of NMD pathways and the involvement of more specialised NMD factors. We describe the consequences of mutations in NMD factors leading to neurodevelopmental diseases, and the role of NMD in cancer. We highlight strategies of RNA viruses to evade recognition and decay by the NMD machinery.

1. Introduction

Nonsense-mediated messenger RNA decay (NMD) is a conserved eukaryotic pathway that quality controls protein synthesis by recognising and degrading mRNAs bearing premature termination codons (PTCs). One characteristic of canonical NMD substrates is a long 3′-untranslated region (3′-UTR). Long 3′-UTRs delay translation termination by keeping terminating ribosomes away from termination-stimulating factors, such as the poly(A)-binding protein (PABP), which binds to the poly(A) tail of mRNAs [1,2,3,4]. Another characteristic of canonical NMD substrates in mammalian cells is the presence of at least one exon-junction complex (EJC) downstream of the PTC [5,6,7]. During splicing, EJCs assemble 20–24 nucleotides (nt) upstream of exon-exon junctions [8]. As normal stop codons are typically situated in the last exon of a mRNA, EJCs are dissociated from transcripts during translation, resulting in EJC-free mRNAs [9]. When exon-exon junctions are situated more than 50–55 nt downstream of PTCs, the remaining EJCs give rise to abnormal translation termination and activate the NMD machinery by recruiting NMD factors to the terminating ribosomes [10,11,12].
NMD is also involved in the control of gene expression and targets ~10% of cellular transcripts [13,14]. Thereby, NMD takes part in a variety of cellular mechanisms, such as the regulation of the cell cycle, cell viability, DNA damage response and innate immune response to viral infections [13,14,15,16,17]. Moreover, dysregulation of the NMD pathway causes serious pathologies, such as cellular stress and cancer, and is associated with neurodevelopmental disorders (NDDs) [17,18,19]. Hence, elucidating the molecular mechanisms governing NMD is essential for the development of novel therapeutic strategies.
In this review, we describe the different mammalian NMD pathways identified to date, as well as the main NMD factors—the up-frameshift proteins (UPFs). Furthermore, we address the implication of NMD in human pathologies, including NDDs, cancer and viral infections.

2. Mammalian NMD Pathways

In mammals, NMD targets a wide range of RNAs, from mRNAs harbouring PTCs due to mutations, to mRNAs with long 3′-UTRs or short upstream open reading frames (uORFs) in their 5′-UTRs, as well as selenoprotein-encoding mRNAs, but also small nucleolar RNAs (snoRNAs) and long non-coding RNAs (lncRNAs) [20,21,22,23,24,25,26,27]. NMD depends on translation of the RNA transcript, with NMD functions spanning from translational quality control to gene regulation. Accordingly, mammalian NMD is tightly regulated. In fact, different NMD pathways were identified, each requiring distinct combinations of NMD factors [13,14]. All NMD pathways have in common that they involve UPF1, which, therefore, is considered as the key NMD factor (Figure 1) (reviewed in [14,28,29]).
In the canonical model of NMD, a PTC is defined by the presence of one or several EJCs associated with UPF2 and UPF3B in the 3′-UTR of the mRNA. In this model, translation termination at the PTC is aberrant and slow, allowing the assembly of the SURF surveillance complex on the terminating ribosome. SURF consists of the SMG1-8-9 kinase complex, UPF1 and the eukaryotic release factors eRF1 and eRF3 [30]. UPF1, as part of the ribosome-bound SURF complex, can interact with EJC-associated UPF2 and UPF3B, which gives rise to a so-called decay-inducing complex (DECID) [30,31]. Formation of the DECID complex leads to activation of UPF1 by SMG1-8-9-mediated hyper-phosphorylation [32]. Hyper-phosphorylated UPF1 triggers a series of events, ultimately leading to mRNA degradation which is suggested to take place in processing bodies (P-bodies) [33]. Notably, hyper-phosphorylated UPF1 promotes post-termination ribosome recycling. This step depends on UPF1’s ATP hydrolysis activity, which is activated by the UPF1-UPF2 interaction [34,35]. Hyper-phosphorylated UPF1 also binds eukaryotic translation initiation factor 3 (eIF3), which is required for ribosome recycling and translation re-initiation [36,37]. UPF1 binding to eIF3 is suggested to prevent new rounds of cap-dependent translation on nonsense mRNA [38]. Importantly, hyper-phosphorylated UPF1 is a scaffold for the recruitment of mRNA decay initiators SMG6 and SMG5-7 [39]. SMG6 endonuclease is activated by interaction with UPF1, and SMG5-7 and was shown to cleave mRNA close to the PTC [25,40,41]. SMG5-7 recruits additional factors for mRNA decapping (DCP2) and deadenylation (CCR4/NOT) [42,43]. Ultimate degradation of the unprotected mRNA is subsequently completed by exonuclease XRN1 and the exosome complex [43]. SMG5-7 also recruits the protein phosphatase 2A (PP2A) responsible for UPF1 dephosphorylation. Dephosphorylation of UPF1 results in the release of SMG6 and SMG5–7, and allows the recycling of UPF1 for novel rounds of NMD [33,44].
This canonical model is also referred to as EJC-dependent NMD (Figure 1). In fact, the presence of an EJC downstream of the stop codon is the strongest marker of NMD substrates [7]. More recently, variations of this NMD model have been described, varying in EJC composition. The core of the EJC is formed by eIF4A3, the protein mago nashi homolog (MAGOH), and the RNA-binding protein 8A (RBM8A, also known as Y14) [45,46,47]. Additional subunits can associate with the EJC core in a mutually exclusive manner. The cancer susceptibility candidate gene 3 (CASC3) protein (also known as BTZ or MLN51) is associated to the EJC in canonical NMD, and it has also been shown that CASC3-bound EJCs can mediate NMD, independently of UPF2 [47,48,49,50]. In this UPF2-independent ‘branch’ of EJC-dependent NMD (Figure 1), CASC3 is suggested to directly interact with eIF4A3, UPF3B as well as UPF1, with UPF3B binding to UPF1 and eRF3, thereby stabilising UPF1 in the vicinity of the EJC, despite of the lack of UPF2 [48,49,50]. In the canonical NMD pathway, UPF2 promotes UPF1’s ATPase and helicase activities and, thereby, supports NMD activation [51,52,53]. Recent studies have shed light on the mechanisms compensating for the absence of UPF2 in UPF2-independent NMD, revealing the involvement of serine/threonine-protein kinases AKT (Figure 1) [54,55]. AKT1 was first identified as a member of the NMD pathway in a screen of kinase inhibitors for putative NMD modulators [54]. In the same study, co-immunoprecipitation assays showed that AKT1 interacts with UPF1 and UPF3B, but not with UPF2, and that the interaction between AKT1 and UPF1 is dependent on UPF3B [54]. Subsequently, a genetic screen for NMD effectors demonstrated that AKT replaces UPF2 in CASC3-EJCs and is able to activate UPF1 [55].
UPF3B-independent NMD presents another branch of EJC-dependent NMD, relying on RNPS1 bound to EJCs (Figure 1) [48]. RNSP1-EJCs recruit UPF2 in the absence of UPF3B. It was shown that RNPS1-bound EJCs and CASC3-bound EJCs have distinct subcellular localisations and lifetimes [50]. RNPS1 associates to EJCs during co-transcriptional splicing. Consistently, RNPS1-EJCs are mostly found in the nucleus and in the cytoplasm shortly after mRNA export from the nucleus [50,56]. Therefore, RNPS1-enhanced NMD is suggested to target mRNAs translated in the nuclear periphery immediately after their export [50,56]. In contrast, CASC3 associates with EJCs in the cytoplasm, during a compositional switch from RNPS1 to CASC3 that is either pre- or co-translational. Thus, CASC3-enhanced NMD is suggested to occur during later stages of the mRNA lifetime [49,50].
During co-transcriptional splicing, factors belonging to the serine- and arginine-rich (SR) protein family bind to nascent mRNAs, and their presence is suggested to promote NMD [57,58,59]. In fact, SR Splicing Factors 1 (SRSF1) directly interacts with UPF1, recruiting it to mRNAs in the nucleus. Moreover, co-immunoprecipitation assays indicate that SRSF1 interacts with SMG7 and PP2A, which are both implicated in the dephosphorylation of UPF1 [59,60]. This suggests that SRSF1 could enhance NMD by promoting the recycling of this NMD factor [59,61,62].
The differentiation of NMD into several pathways or branches highlights the importance of NMD as a regulatory mechanism of gene expression in mammalian cells. Several cases of intricate modulation of NMD have been reported for specific cellular contexts. For example, mRNAs translated at the endoplasmic reticulum (ER), or in neuronal dendrites and axons, are targeted by localised NMD [63,64,65]. The ER membrane is the place for translation and translocation of secreted proteins and integral membrane proteins, and the ER is also where the unfolded protein response (UPR) takes place [66,67]. The expression of several UPR factors (e.g., ATF-3, ATF-4, PERK and IRE1α) was shown to be regulated by NMD [68,69]. Notably, the neuroblastoma-amplified sequence protein (NBAS) and the helicase DHX34 were identified as specialised NMD factors, targeting membrane-associated mRNAs and mRNAs encoding proteins of stress response pathways, as well as NMD factor-encoding mRNAs, including UPF1-mRNA [64,69]. The NBAS is suggested to assist in an ER-associated NMD pathway by recruiting UPF1 to the ER. The NBAS-mediated NMD is part of a feedback loop that responds to cellular stress, activates the cellular stress response and leads to downregulation of NMD [64]. The integrated stress response inhibits global NMD, mediated by UPF1’s isoform 2, which is most abundant in mammalian cells. This is achieved by translation inhibition via phosphorylation of the α-subunit of eukaryotic initiation factor 2 (eIF2α) [70].
In addition to the distinct factor requirements of the NMD branches, several NMD factors exist in different versions. One example is UPF3B and its paralog UPF3A (see below) [71,72,73,74,75,76]. UPF3B is the main form in most tissues, but UPF3A is most abundant in adult testis [74]. In male germ cells, meiotic sex chromosome inactivation causes transcriptional silencing of the UPF3B gene, which is located on the chromosome X. This downregulation of UPF3B results in an upregulation of UPF3A and the stabilisation of a specific subset of mRNAs encoding critical genes required for spermatogenesis and male fertility, which would otherwise be degraded by NMD [73,74]. A further example of sophisticated modulation of NMD is the existence of two UPF1 isoforms in mammals: UPF1-1 and UPF1-2 [77,78]. UPF1-2 is the main isoform in human tissues, and it is the main NMD factor in standard cellular conditions [78]. However, UPF1-1 was reported to target mRNAs translated at the ER, e.g., mRNAs encoding integral membrane proteins, even during ER stress when UPF1-2 mediated NMD is downregulated (see above) [78]. Despite down-regulated translation, UPF1-1 was shown to be able to activate NMD in stress conditions. This increased activity was linked to UPF1-1’s prolonged residence time on mRNAs compared to UPF1-2 (discussed below) [78].

3. Core NMD Factors

3.1. UPF1

UPF1 is highly conserved in eukaryotes. It comprises a central helicase domain, flanked by an upstream cysteine-histidine-rich (CH) domain and a downstream serine-glutamine-rich (SQ) domain (Figure 2) [29,51,52,77,79,80]. UPF1’s central role in NMD relies on its capacity to bind RNA, which is supported by its helicase domain. However, the CH domain and the N- and C-terminal regions of UPF1 were also shown to have regulatory functions of UPF1-mediated NMD activation [39,52,55,79].
UPF1 is a superfamily 1 (SF1) helicase, and its RNA-binding, translocase and unwinding activities are linked to ATP-binding and hydrolysis [81,82]. The helicase domain structure of UPF1 is well-characterised [52,80]. A wide range of conformations have been observed in structural studies, shedding light on the molecular dynamics linking UPF1’s RNA-associated catalytic functions to ATP-binding and hydrolysis (reviewed in [29]). ATP binding hinders the binding of RNA by UPF1 [81], and UPF1’s ATPase activity is essential to NMD [38,83,84]. The ATPase activity has been shown to be required for UPF1 to remodel ribonucleoprotein complexes (RNPs) by fuelling its RNA translocase and unwinding activities [34,85,86,87], promoting ribosome recycling at PTCs [34,86], displacing protective RNA-binding proteins [78,85] and disassembling NMD complexes for efficient RNA degradation [83]. A recent study using a range of UPF1 ATPase and helicase mutants has shown that ATP-hydrolysis stimulates dissociation from the RNA in a helicase-decoupled manner [84]. Intriguingly, UPF1 mutants with impaired helicase activity but intact ATPase function still support NMD [84]. In contrast, mutants deficient in the ATPase activity result in compromised NMD substrate selectivity [88,89]. Together, these results suggest that UPF1’s ATPase activity is essential for NMD, due to stimulation of RNA dissociation and concomitant release of non-NMD substrates, while UPF1’s helicase activity, which unwinds RNA and remodels RNPs, is not crucial for NMD [84].
Figure 2. Domain architecture (above) and AlphaFold2 structure prediction (below) of human UPF1 isoform 2. UPF2 interacts with UPF1’s CH domain (thick dashed line). UPF1 is phosphorylated at T28, S1078, S1096 and S1116 (blue) by SMG1, and T151 (orange) by AKT1 kinase. Missense mutations associated with neurodevelopmental disorder are found in the UPF1 helicase domain (red dashed lines in protein scheme; red spheres in the predicted protein structure) [90,91]. The positions of RNA and ATP were modelled by aligning the structure of UPF1’s helicase domain (PDB ID: 2XZO) to the AlphaFold2-predicted UPF1 model [52]. CH, cysteine-histidine-rich; SQ, serine-glutamine-rich.
Figure 2. Domain architecture (above) and AlphaFold2 structure prediction (below) of human UPF1 isoform 2. UPF2 interacts with UPF1’s CH domain (thick dashed line). UPF1 is phosphorylated at T28, S1078, S1096 and S1116 (blue) by SMG1, and T151 (orange) by AKT1 kinase. Missense mutations associated with neurodevelopmental disorder are found in the UPF1 helicase domain (red dashed lines in protein scheme; red spheres in the predicted protein structure) [90,91]. The positions of RNA and ATP were modelled by aligning the structure of UPF1’s helicase domain (PDB ID: 2XZO) to the AlphaFold2-predicted UPF1 model [52]. CH, cysteine-histidine-rich; SQ, serine-glutamine-rich.
Biomedicines 11 00722 g002
The importance of RNA dissociation and helicase functions of UPF1 for efficient NMD is supported by the existence of different regulatory features within UPF1. Firstly, biochemical studies show that the CH domain enhances RNA-clamping and reduces the RNA-unwinding activity of UPF1 [92]. This is supported by a structure of yeast Upf1, showing that the CH domain associates with the helicase domain in a closed conformation, forming a longer RNA-binding channel [52]. In the context of NMD, UPF2-binding to UPF1’s CH domain removes auto-inhibition of UPF1 helicase and of RNA dissociation. Structural data of UPF1, in complex with the C-terminus of UPF2, shows that the CH domain of UPF1 is displaced upon UPF2-binding into an open conformation [52,79], resulting in decreased RNA-clamping and increased RNA-unwinding by UPF1 [51,52]. Recent biochemical and biophysical studies have shown that binding of UPF2 to UPF1 directly promotes dissociation of UPF1 from RNA in a non-competitive manner [53]. A low-resolution cryo-EM structure of the EJC, in complex with RNA and the three UPF proteins, shows that UPF1 is found in the vicinity of the RNA 3′-end [93]. Based on this finding, it was suggested that, upon association with the EJC-UPF3B-UPF2 complex, UPF1 is activated and dissociates from the mRNA, but stays anchored to the mRNP via the EJC. More recently, a similar activating function was attributed to AKT [55]. In the absence of UPF2, AKT1 cooperates with EJCs containing CASC3 and UPF3B [54,55], and activates UPF1 helicase activity by phosphorylating threonine residue 151 in the CH domain [55]. Earlier structures of UPF1 CH show that T151 (T90 in yeast) is located at the interface between the CH and the helicase domains, where RNA is bound [52,79]. Accordingly, phosphorylation of T151 could lead to the repulsion of the CH domain from the RNA, inducing an open conformation [55]. Taken together, UPF2-binding and AKT1 phosphorylation of the CH domain are both suggested to induce a switch from a closed to an open conformation. This switch reduces the RNA-binding and increases helicase activities of UPF1, triggering NMD activation.
Besides the CH domain, RNA dissociation and helicase functions of UPF1 are suggested to be regulated by a loop located within its helicase domain. In the most abundant isoform of UPF1 (UPF1-2), this loop is 11-residues long. The alternatively spliced UPF1-1 isoform has an 11-residues insertion, resulting in a 22-residues long loop (Figure 2) [77]. Structural and biochemical data show that this loop is part of the RNA-binding channel and directly affects RNA-binding: while the shorter loop of UPF1-2 is oriented towards the helicase core and competes with RNA-binding, the longer loop of UPF1-1 is more flexible and oriented towards the solvent, thus not interfering with RNA-binding [52,77,80,94]. As a result, UPF1-1 showed increased RNA-binding in the presence of ATP, along with increased ATP hydrolysis in the presence of RNA, in comparison to the more abundant UPF1-2 isoform [77]. Consistently, RNA-unwinding and translocase activities were shown to be higher for UPF1-1 than for UPF1-2 [77]. In accordance with these differences in biochemical properties, recent biochemical and transcriptome-wide studies suggest distinct roles for the two UPF1 isoforms [78]. Proteins binding to mRNAs with long 3′-UTRs, such as the polypyrimidine tract-binding protein-1 (PTBP1) or the heterogeneous nuclear ribonucleoprotein L (hnRNP L), protect these mRNAs from NMD mediated by UPF1-2, by inhibiting UPF1-2′s translocase activity and promoting its dissociation from the mRNA [95,96,97]. However, due to the stronger translocase activity of UPF1-1, this isoform can displace protective RNA-binding proteins, such as PTBP1, and thereby promote NMD of mRNAs with very long 3′-UTRs [78], resulting in different target-specificities of the two UPF1 isoforms.
Phosphorylation of UPF1 is the decisive event triggering NMD, allowing it to proceed to the mRNA degradation step after its recognition by the NMD machinery. UPF1 is hyper-phosphorylated by the SMG1–8–9 kinase complex at serine/threonine-glutamine motifs located in its disordered N- and C-terminal regions (e.g., T28, S1078, S1096, S1116, Figure 2) [32,98]. Several of these phosphorylation sites share a consensus leucine-serine-glutamine (LSQ) sequence [99]. The structural study of the SMG1–8–9 kinase complex with a peptide mimicking UPF1 residues 1074–1084 (part of the SQ domain) reveals that these LSQ motifs bind within a hydrophobic cage in the vicinity of the SMG1 kinase active site [29,100]. These data highlight the importance of UPF1’s LSQ motifs for optimal phosphorylation specificity and efficiency of the SMG1–8–9 kinase complex. Hyper-phosphorylated UPF1 provides a platform to recruit SMG6 endonuclease and SMG5-SMG7 to the NMD substrate, initiating mRNA decay [32,39,44,88,101,102].
Despite the vital role of UPF1 in NMD, the UPF1 enzymatic functions specifically implicated in NMD are still enigmatic. UPF1’s CH domain is a good illustration of UPF1’s functions, which remain poorly understood. Structural studies of the CH domain of UPF1 showed a unique association of two RING-box-like modules [29,52,79], which are characteristic of some families of E3 ubiquitin ligase enzymes. Interestingly, biochemical studies showed that yeast Upf1 self-ubiquitinates when Upf3 and the E2 ubiquitin-conjugating enzyme Ubc3 are present [103]. In human cells, the transcription factor MYOD was shown to be down-regulated by UPF1 as a result of MYOD ubiquitination and proteasomal degradation rather than decay of MYOD-encoding mRNA [104], indicating that UPF1’s E3 ligase function is separate from the NMD function in this context. Further investigation into UPF1’s E3 ubiquitin ligase activity is required to identify the E2 ubiquitin conjugating partner for UPF1 and its substrate specificity in human cells. In fact, little is known about the fate of truncated and possibly misfolded proteins produced from NMD-targeted mRNAs. Therefore, it is tempting to speculate about a role for UPF1 in the ubiquitination and subsequent degradation of nonsense mRNA-encoded polypeptides. A study using normal and nonsense reporter mRNAs encoding the same protein showed a proteasome-dependent destabilisation of the newly synthesised polypeptide encoded by nonsense mRNA [105]. The role of the proteasome in NMD-related degradation of nonsense polypeptides is further supported by recent assays using a double-reporter mRNA allowing the decoupling of mRNA decay and protein quality-control in NMD [106]. In no-go and non-stop mRNA decay, the associated ribosome-associated quality-control (RQC) pathway involves E3 ubiquitin ligase LTN1/Listerin for nascent chain ubiquitination and subsequent proteasomal degradation [107,108]. In contrast, a recent study suggests that NMD-associated protein degradation does not require LTN1 or UPF1 as E3 ubiquitin ligases but involves the CCR4–NOT transcription complex subunit 4 (CNOT4). CNOT4 was identified in genome knockdown and knockout screens as being responsible for the ubiquitination signal leading to proteasomal degradation of the nonsense polypeptide [106]. However, the molecular mechanisms linking mRNA decay and nonsense polypeptide ubiquitination in NMD remain to be elucidated.

3.2. UPF2

UPF2 associates with the EJC in nucleoplasmic and cytoplasmic fractions of human cells, and is mostly found in the cytoplasmic perinuclear region [109]. UPF2 is composed of three tandem middle portions of eIF4G (MIF4G) domains (Figure 3) [110], which are suggested to form a ring-shaped structure [111]. The first two MIF4G domains (MIF4G-1 and MIF4G-2) are conserved from yeast to human but have an unknown function. The third MIF4G (MIF4G-3) domain plays a key role in mediating the assembly of the NMD machinery: MIF4G-3 interacts with SMG1 kinase and UPF3B and helps to activate SMG1 kinase activity [30,110]. In the complex with SMG1, UPF2′s MIF4G-3 and its C-terminal part were phosphorylated by SMG1 kinase, but the significance of this step remained elusive [110]. Furthermore, UPF2 is suggested to support dissociation of hyper-phosphorylated UPF1 from the SMG1-8-9 kinase complex [98].
UPF2′s C-terminal UPF1-binding domain (U1BD) (Figure 3) is natively disordered in free UPF2 protein, as determined in NMR studies [51]. In the presence of UPF1, UPF2′s C-terminal domain folds into an alpha-helical and a beta-hairpin structure which binds UPF1’s CH domain [51]. This U1BD-CH interaction induces an open UPF1 CH-helicase conformation with lower affinity for RNA and increased UPF1’s ATPase activity (see above) [51].
In a recent study, UPF1-UPF2 complexes were shown to have a low RNA affinity, despite the fact that both NMD factors have been shown to interact with RNA [53]. The authors performed fluorescence anisotropy experiments and reported affinities in the low nanomolar range for UPF2 MIF4G domains 1-2, as well as MIF4G domain 3 with the U1BD [53]. This corroborates previous work showing that UPF2′s MIF4G domain 3 binds RNA in electrophoretic mobility shift assays (EMSAs) via conserved basic and aromatic residues [112] and work showing ribosome and RNA binding of UPF2 in complex with UPF3B [113,114]. In complex with UPF1, UPF2 was shown to help UPF1-RNA dissociation, likely by binding to a region allosteric of UPF1’s RNA-binding site [53].
UPF2 was also shown to interact with eRF3 via its MIF4G-3 domain and C-terminal region [111]. However, UPF2 did not have an impact on the efficiency of translation termination in a reconstituted translation assay [113]. Therefore, this interaction is unlikely to contribute to the aberrant and slow translation termination at a PTC on an NMD substrate.
Figure 3. Domain architecture (above) and AlphaFold2 structure prediction (below) of human UPF2. UPF2 interaction partners are indicated by thick and dashed lines. V31L in UPF2 was found to be associated with neurodevelopmental disorder (red dashed line in protein scheme and red sphere in predicted UPF2 structure) [91,115]. MIF4G, middle portion of eIF4G; U1BD, UPF1 binding domain.
Figure 3. Domain architecture (above) and AlphaFold2 structure prediction (below) of human UPF2. UPF2 interaction partners are indicated by thick and dashed lines. V31L in UPF2 was found to be associated with neurodevelopmental disorder (red dashed line in protein scheme and red sphere in predicted UPF2 structure) [91,115]. MIF4G, middle portion of eIF4G; U1BD, UPF1 binding domain.
Biomedicines 11 00722 g003

3.3. UPF3A and UPF3B Paralogs

UPF3B is also called UPF3X, as it is encoded by the X chromosome. The transcription of the UPF3B gene is activated by the SATB2 transcription regulator by direct binding to the UPF3B gene promoter [116]. UPF3B is a nucleocytoplasmic shuttling protein which binds to the EJC in the nucleus and is then exported to the cytoplasm in complex with the EJC [5]. Human UPF3B comprises an RNA recognition motif-like (RRM-L) domain in its N-terminal, followed by a NONA/paraspeckle-like (NOPS-L) region, two predicted coiled-coil-like (CCL-1 and CCL-2) regions and a C-terminal EJC-binding motif (EBM) (Figure 4) [114].
UPF3B interacts with UPF2′s MIF4G-3 through its RRM-L domain and NOPS-L region [114]. The crystal structure of the complex revealed an intimate interaction where UPF2′s MIF4G-3 domain wedges between the RRM-L and NOPS-L domains of UPF3B. The NOPS-L-mediated interaction with UPF2′s MIF4G-3 is essential for high-affinity binding [112,114]. In fact, the UPF3B Y160D mutation causing neurodevelopmental disease is located in the UPF3B NOPS-L region. On the molecular level, aspartate residue 160 is displaced from a hydrophobic cleft formed by UPF2′s MIF4G-3, causing a ~40-fold decrease in UPF2-UPF3B interaction affinity [114]. This weakened affinity between UPF3B and UPF2 was shown to lead to an upregulation of UPF3A (see below) [117] and decreased NMD efficiency, as evidenced by increased mRNA levels of factors regulating neurodevelopment, such as ATF4 and ARHGAP24 [65]. UPF3B was also found to weakly interact with UPF1 in the absence of UPF2, which may be important in UPF2-independent NMD where UPF1-UPF3B complex formation may be stabilised by additional factors (Figure 1) [113].
More recently, UPF3B was shown to interact with single- and double-stranded RNA (ssRNA/dsRNA), as well as DNA, with a preference for dsRNA [114]. This agrees with a previous report showing that UPF3B interacts with 80S ribosomes which have dsRNA expansion segments [113]. UPF3B’s RRM-L, NOPS-L and the CCL-1 regions are required for high-affinity RNA interaction (Figure 4). Interestingly, an RNA-induced oligomerisation of UPF3B was observed which is reminiscent of paraspeckle protein behaviour [114]. The UPF3B-ribosome interaction agrees with the finding that UPF3B is enriched in the nucleolus where ribosomes are assembled in human cells [65]. A reconstituted mammalian translation system was used to show that UPF3B impacts the efficiency of translation termination [113]. UPF3B binds eRF1 and eRF3 and delays stop codon recognition and subsequent peptide release, slowing down translation termination [113]. After peptide release, UPF3B promotes the dissociation of post-termination 80S ribosomal complexes, releasing the bound mRNA [113].
The EBM of UPF3B (Figure 4) interacts with the EJC core composed of eIF4A3, RBM8A and MAGOH [92,118]. Mutation of residue R423 to alanine in the EBM inhibits UPF3B’s interaction with RBM8A and leads to decreased NMD efficiency, similar to a UPF3B EBM deletion mutation (UPF3B△421-434) [118,119].
UPF3A is a paralog of UPF3B evolutionary conserved in vertebrates [74]. In human cells, two UPF3A isoforms (UPF3AL and UPF3AS) were identified, caused by alternative splicing in exon 4 [71]. The longer isoform UPF3AL retains the exon 4, which encodes the β5 strand of the RRM-L domain and the NOPS-L region, and therefore can interact with UPF2. In contrast, UPF3AS lacks exon 4 and, thus, loses important residues in NOPS-L for UPF2 interaction [114]. UPF3A (UPF3AL) and UPF3B directly compete for binding UPF2 by binding to the same sites in MIF4G-3 [114]. Surprisingly, UPF2′s MIF4G-3 binds UPF3AL with ~10-fold higher affinity than UPF3B due to additional interactions between UPF3AL’s NOPS-L region and UPF2′s MIF4G-3 domain [114]. However, expression levels of UPF3B are significantly higher than UPF3A in most mammalian cells. Therefore, the downregulation of UPF3A in the presence of UPF3B is likely achieved because UPF3B can outcompete UPF3A for UPF2-binding, thereby negatively affecting UPF3A’s protein half-life [72]. The ‘free’ UPF3A (not stabilised by UPF2) was shown to be quickly degraded in cells [72]. Consistently, UPF2 overexpression leads to UPF3A stabilisation, even in the presence of UPF3B [72]. Vice versa, UPF2 knockdown prevents UPF3A stabilisation, even if the UPF3B expression is also inhibited [72].
Previously, UPF3A was described as a weak NMD factor or an NMD inhibitor [71,74]. UPF3A was suggested to interfere with the NMD machinery by sequestering UPF2, likely due to its weaker affinity for the EJC in comparison to UPF3B. Consistently, when replacing the human UPF3A’s EBM with UPF3B’s EBM, the mutated UPF3A turned from an NMD repressor into an NMD activator. Similarly, by deleting the EBM region in UPF3B, the mutated UPF3B became an NMD repressor [74]. More recently, UPF3A was shown to rescue NMD activity when UPF3B was knocked out, and NMD inhibition could only be achieved when UPF3A and UPF3B were co-depleted [75,76]. This indicates that UPF3A is an active NMD factor, similar to UPF3B. Taken together, it appears that UPF3A stabilises or destabilises nonsense mRNAs, depending on the specific substate and cellular context. For instance, it was observed that nonsense mRNAs stabilised by UPF3A tend to have a longer 3′-UTR than mRNAs destabilised by UPF3A [74].

4. Implication of NMD Factors in Human Disease

4.1. Neurodevelopmental Disorders (NDD)

As one of the post-translational regulation pathways in eukaryotic cells, NMD plays an important role in neurodevelopment by regulating gene expression [120]. Approximately 80% of all mRNAs linked to NDD are targeted by NMD [65,121]. During neural differentiation, NMD activity is downregulated. This is achieved partially by targeting mRNAs encoding UPF1 and CASC3 through microRNA-miR-128 [122]. However, despite downregulation of NMD factors, functional NMD is essential for neurodevelopment [65]. In agreement, NMD impairment is found to be associated with schizophrenia, intellectual disorder (ID) and autism spectrum disorder (ASD) [65].
Reduced NMD efficiency can be caused by mutations in UPF2 [123,124] and UPF3B [117,125,126,127,128,129,130,131]. Disease-associated nonsense and frameshift mutations in UPF2 or UPF3B introduce PTCs in their transcripts, such that their mRNAs are targeted by alternative NMD pathways (UFP2-independent NMD or UPF3B-independent NMD) (Figure 1). Resulting reduced levels of NMD factors UPF2 or UPF3B lead to a reduction of the efficacy of the canonical NMD pathway. Missense mutations in UPF3B (Figure 4 and Table 1) reduce the interactions with other NMD factors, including UPF2 [114] and RBM8A [119], leading to inefficient NMD (see above) [65,114,119]. A mutation in UPF2′s N-terminus (Figure 3 and Table 1) has been found in NDD patients (ClinVar accession VCV000996716.1) [90], but the molecular basis of pathogenesis remains to be elucidated. Mice lacking UPF2 in the forebrain were found to have deficits in spatial and contextual long-term memory (LTM) and long-term potentiation (LTP) [124]. This agrees with the learning and memory deficits found in NDD patients [124]. In addition, UPF2-deficient mice exhibited an increased neuroinflammatory response, and treatment with anti-inflammatory drugs was found to reduce brain inflammation and to improve their LTM and LTP [124]. UPF2 single allele deletion is associated with NDDs as well, leading to a more than two-fold transcriptome dysregulation with a very high (95%) similarity to disease-causing mutations in UPF3B [123].
In cell culture experiments, UPF3B missense mutations did not change UPF3B’s cellular location, but caused reduced NMD efficiency and impaired neuronal differentiation [65,120]. For example, ARHGAP24 mRNA, which encodes a GTPase-activating protein and influences neurite outgrowth, as well as neurite branching, is a canonical NMD target. ARHGAP24 mRNA was found to be significantly upregulated in some NDD patients, having mutations in UPF3B [65,117]. Similarly, transcription factor ATF4, known to be required for neuronal development and plasticity, is upregulated in UPF3B-mutated cells [65]. Importantly, in all NDD patients harbouring UPF3B mutations, the upregulation of UPF3A protein levels is correlated with the severity of patients’ disease phenotypes in a reverse way, i.e., less severe disease correlates with higher UPF3A protein levels [117,121].
Two mutations in the UPF1 helicase domain have been found in NDD patients (Figure 2 and Table 1). Homozygous deletion of UPF1 is embryonically lethal in mice, highlighting the importance of UPF1 [134]. Moreover, copy number variants in UPF2, UPF3A, eIF4A3, RBM8A, RNPS1 and SMG6 NMD factors have been identified in NDD patients, associating any imbalance of these NMD factors with NDD aetiologies [123]. Similarly, disease-causing SMG8 mutations are reported to interfere with interactions with SMG9 and/or SMG1, thereby influencing NMD efficiency [135,136].
In addition, hyper-activated NMD was recently reported to be associated with NDD [137]. Loss of the fragile X syndrome protein (FMRP) is a major aetiology of autism and ID [137]. Under normal conditions, FMRP is recruited to NMD substrates by direct interaction with UPF1, repressing NMD. FMRP deficiency increases UPF1 phosphorylation levels and leads to hyper-activated NMD in cells from fragile X syndrome (FXS) patients [137]. Hyper-activated NMD can be partially reversed by small molecules inhibiting NMD, as evidenced by the restoration of early and mature neuron marker expression and of enhanced neurite outgrowth [137].

4.2. Cancer

Cancer is assumed to start from abnormal clonal expansion of a single cell. The cancer genome is enriched with somatic point mutations and insertion/deletion (indel) mutations, many of which encode PTC-containing transcripts [138,139]. Analysis of cancer genomes and exomes revealed that PTC-containing tumour-suppressor transcripts are frequently degraded by NMD [7], whereas PTC-containing transcripts of oncogenes frequently encode dominant-negative proteins and evade NMD [7,139]. Accordingly, inhibition of NMD is proposed to benefit cancer treatment by promoting the expression of tumour-suppressor proteins. For example, mutations in the tumour-suppressor gene Tp53 are found in more than 50% of human cancers, and most mutations are nonsense mutations [140]. The function of p53 tumour-suppressor protein is to induce cell apoptosis. NMD inhibition through a combination treatment of NMDI14, which disrupts the UPF1-SMG7 interaction, and a stop codon read-through drug, G418, led to expression of full-length p53 in different cancer cell lines, ultimately leading to cell death (Table 2) [141]. Likewise, NMD inhibition in microsatellite instability colorectal cancer promoted nonsense mRNA expression, including HSP110DE9 which has a dominant negative activity against oncogenesis, and suppresses cell proliferation (Table 2) [139].
Since indel mutations in ORFs often lead to frameshift mutations, and NMD inhibition promotes expression of such proteins with truncated, mutated C-terminal, these new proteins have the potential to be processed into neoantigens [19,142]. Neoantigens derived from these aberrant proteins present tumour cell-specific antigens, enabling the immune system to recognise, attack and destroy the tumour cell [142]. Moreover, NMD inhibition can influence RNA alternative splicing by altering exon usage. This event can also lead to novel neoantigen expression [143]. Using a prostate-specific membrane antigen (PSMA) aptamer-SMG1 siRNA conjugate, colorectal carcinoma cell growth was significantly depressed in mice, and this therapeutic effect was more than additive when combined with 4-1BB aptamer treatment (Table 2) [142]. Similarly, transfecting B-cell lymphomas with a CD40-agonist aptamer SMG1-shRNA chimera inhibited NMD activity, due to SMG1 kinase knockdown, and improved survival of mice with B-cell lymphomas (Table 2) [144]. Moreover, UPF2 knockdown by aptamer-linked siRNA chimeras boosted CD8+ T-cell anti-tumour immunity and inhibited murine breast tumour growth (Table 2) [143]. Interestingly, in cancer cells having mutations in U2AF1 or SF3B1, two components of the U2 spliceosome, NMD is found to be attenuated [145]. Further weakening of the NMD activity by SMG1 kinase inhibition led to DNA damage and chromosomal instability in these cancer cells and specifically killed them, providing an alternative therapy for patients [145].
Table 2. Effects of NMD inhibition/activation on cancer cells.
Table 2. Effects of NMD inhibition/activation on cancer cells.
CancerPotential Therapeutic ApproachRef.
Cancer Type, Cell LineCharacteristicsStrategyMethodEffect
NMD inhibitionSmall-cell lung cancer, cell line N417
Breast cancer, cell line HDQP-I
PTC mutationsin
Tp53 gene
Promoting
expression
of tumour
suppressor
proteins
NMDI14 (disruption of UPF1-SMG7 interaction),
G418 (PTC read-through)
Restoration of full-length
p53 expression,
leading to cell death
[141]
Microsatellite instability
colorectal cancer
Overexpression of UPF1, UPF2, SMG1, SMG6 and SMG7Amlexanox
or UPF1 siRNA
Decrease in
cell proliferation rate
[139]
Murine colorectal carcinoma,
cell line CT26
n.d. aProducing
tumour-specific
neoantigens
that can be
recognised
by the human
immune system
PSMA b aptamer-SMG1 siRNA conjugate,
4-1BB aptamer
Suppression of tumour growth by PSMA-Smg1
± 4-1BB treatment
[142]
B-cell lymphoman.d. aCD40-agonist aptamer
SMG1-shRNA chimera
Tumour infiltration
by lymphocytes
Improved mice survival
[144]
Murine breast cancer,
cell line 4T1E
n.d. aEpCAM c aptamer-
UPF2 siRNA chimera
Improved CD8+ T-cell immunity
Inhibition of tumour growth
[143]
NMD activationProstate cancer, cell line PC3
Colon cancer, cell line HCT116
Melanoma, cell line A375
eIF2α
phosphorylation
Downregulating
tumorigenesis
transcripts
UPF1
overexpression
Inhibition of tumour growth[146]
Human gastric tumourInhibition of
UPF1 expression
Suppression of cell cycle
progression and proliferation
Promotion of cell apoptosis
Inhibition of EMT d and metastasis
[147]
a n.d., not defined. b PSMA, prostate-specific membrane antigen, introduced by exogenous transduction of PSMA expression vector into the cell. c EpCAM, epithelial cell adhesion molecule. d EMT, epithelial-mesenchymal transition.
However, in other cases, enhancement of NMD might be beneficial to cancer treatment. In breast cancer, prostate cancer and melanoma, for instance, eIF2α is phosphorylated because of nutrient deprivation, hypoxia and other cellular stresses [146]. NMD efficiency is therefore suppressed, leading to upregulation of many transcripts, some involved in tumorigenesis [146]. Reactivation of NMD by overexpressing UPF1 inhibited tumour growth, as shown in prostate cancer, colon cancer and melanoma cell lines, but not in all cell lines tested (Table 2) [146]. Moreover, UPF1 overexpression suppresses cell proliferation and progression, promotes cell apoptosis, and inhibits cell epithelial-mesenchymal transition (EMT) and metastasis in gastric tumours (Table 2) [147].
Recent bioinformatic studies suggest than some of the NMD factors can be used as prognostic markers for cancers. For instance, UPF3B and SMG5 are highly expressed in hepatocellular carcinoma (HCC) tissue and are associated with poor prognosis in these patients [148,149]. Similarly, UPF3B expression was highly upregulated in colorectal cancer patients at late clinical stages [150].

4.3. Viral Infections

Viruses have evolved to maximise the coding capacity of their genomes. Viral evolutionary strategies include alternative splicing of viral transcripts or polycistronic genomes, resulting in viral RNAs with stop codons upstream of EJCs or very long 3′-UTRs, both features that trigger NMD [14]. In this context, NMD acts as a cellular antiviral defence mechanism. In response, viruses have developed different mechanisms to avoid or inhibit NMD. The distinctive characteristic of positive-sense single-stranded RNA (+ssRNA) viruses is the presence of membrane-bound replication factories, from which several protein synthesis factors are excluded, and that were demonstrated to protect viral RNA from RNases [151]. These replication factories are suggested to also exclude NMD factors [151]. Riboviria viruses have developed different mechanisms to avoid NMD, relying on inherent RNA features, making its genome NMD-resistant (cis-strategies), and viral proteins that inhibit NMD (trans-strategies; reviewed in [152]). Interestingly, NMD restriction of viral replication and vice versa mechanisms, by which viruses circumvent NMD, have been described mainly for Riboviria viruses, including +ssRNA viruses, a few retroviruses and double-stranded RNA (dsRNA) viruses, but also for a double-stranded DNA (dsDNA) virus (Table 3, Table 4 and Table 5, reviewed in [153,154]).

4.3.1. NMD Restriction of Viral Replication

To date, most viruses that have been identified as being sensitive to NMD are +ssRNA viruses (Table 3). This could be explained by the fact that the replication of +ssRNA viruses takes place solely in the cytosol, with their genomic RNA (gRNA) being translated into non-structural viral proteins directly after entry into the host cell. As some +ssRNA virus gRNAs are polycistronic, they act as mRNAs with long 3′-UTRs and are thus targeted by the NMD machinery. NMD was first found to inhibit viral infection when increased infectivity of Sindbis virus (SINV)-like particles was observed in a transgenic Drosophila strain, where NMD was abrogated by co-expression of a dominant negative UPF1 mutant [155]. Another study using a genome-wide siRNA screen in HeLa cells showed that Semliki forest virus (SFV) viral genome replication is suppressed early in infection, with involvement of UPF1, SMG5 and SMG7, but not SMG6 [156]. Shortening of the SFV gRNA 3′-UTR did not impact the sensitivity of viral replication to UPF1 presence/depletion, suggesting that the 3′-UTR length might not be the only NMD-triggering feature of the RNA [156]. NMD inhibition in mouse fibroblasts transfected with mouse hepatitis virus (MHV) gRNA resulted in higher virus titres, demonstrating that MHV replication is also targeted by the NMD pathway [157]. Additionally, quantitative reverse-transcription–polymerase chain reaction (qRT-PCR) analyses showed that NMD targets MHV gRNA and subgenomic RNAs (sgRNAs) with long 3’-UTRs early in infection [157]. A similar infection restriction by NMD-mediated degradation of viral RNA was observed for Zika virus (ZIKV), using UPF1-depleted human neural progenitor cells (NPCs) [158].
In the case of human immunodeficiency virus-1 (HIV-1), several knockdown studies in human cell lines showed that UPF2 and SMG6 are detrimental to viral protein expression and RNA replication, as they promote viral RNA decay [159,160,161,162]. In contrast, UPF1 knockdown results in impaired viral RNA expression [159,160,163] suggesting that UPF1 is a positive regulator of HIV-1 replication, with roles promoting reverse-transcription and viral RNA export (see below). These functions are probably independent of UPF1’s NMD function [159,160,163].
Kaposi’s sarcoma-associated herpesvirus (KSHV) is a dsDNA virus whose replication is restricted by NMD [164]. The KSHV transcriptome includes numerous potential NMD targets with long 3′-UTRs or intron-containing 3′-UTRs. In agreement, UPF1 or UPF3B depletion was shown to enhance viral gene expression and virion production of KSHV [164]. Formaldehyde crosslinking RNA immunoprecipitation (fRIP)-coupled to high-throughput sequencing identified ORF48 and ORF50 transcripts as being enriched in phospho-UPF1 and targeted by NMD [164]. ORF50 mRNA encodes a transcription factor required for KSHV reactivation. Interestingly, NMD also specifically targets the spliced RNA encoding the active form of the X-box binding protein 1 (XBP1) [164]. XBP1 is a transcription factor activated during the unfolded protein response (UPR), which transactivates the promoter of ORF50 [165,166]. Taken together, NMD is suggested to restrict the reactivation of KSHV by regulating ORF50 mRNA levels, both at transcription and post-transcription levels.
Table 3. Sensitivity of animal viruses to NMD factors.
Table 3. Sensitivity of animal viruses to NMD factors.
Virus ClassificationEffect of NMD Factors
Name aRealmGenome bFamilyGenusAnti-ViralNonePro-ViralRef.
MHVRiboviria+ssRNACoronaviridaeBetacoronavirusUPF1n.d. cn.d. c[157]
UPF2
SMG5
SMG6
SFVRiboviria+ssRNATogaviridaeAlphavirusUPF1SMG6n.d. c[156]
SMG5
SMG7
SINVRiboviria+ssRNATogaviridaeAlphavirusUPF1n.d. cn.d. c[155]
ZIKVRiboviria+ssRNAFlaviviridaeFlavivirusUPF1n.d. cn.d. c[158]
UUKVRiboviria-ssRNAPhenuiviridaePhlebovirusn.d. cUPF1n.d. c[167]
HRSVRiboviria-ssRNAPneumoviridaeOrthopneumovirusn.d. cUPF1n.d. c[156]
HIV-1RiboviriassRNA-RTRetroviridaeLentivirusUPF2n.d. cUPF1 d[159,160,161,162,163]
SMG6
KSHVDuplodnaviriadsDNAHerpesviridaeRhadinovirusUPF1n.d. cn.d. c[164]
UPF3B
a Abbreviated virus names: MHV, mouse hepatitis virus; SFV, Semliki forest virus; SINV, Sindbis virus; ZIKV, Zika virus; UUKV, Uukuniemi virus; HRSV, human respiratory syncytial virus; HIV-1, human immunodeficiency virus-1; KSHV, Kaposi’s sarcoma-associated herpesvirus. b Abbreviated genomic groups: +ssRNA, positive-sense single-stranded RNA; -ssRNA, negative-sense single-stranded RNA; ssRNA-RT, single-stranded RNA reverse transcribing; dsDNA, double-stranded DNA. c n.d., no data. d Pro-viral activity of UPF1 can be due to one or several of its functions, including roles in NMD, reverse transcription and mRNA transport.

4.3.2. Viral NMD Evasion cis-Strategies: NMD Resistance Conferred by Inherent RNA Features

Viruses have evolved mechanisms to overcome NMD of their RNAs (Table 4 and Table 5). Several polycistronic viruses depend on programmed ribosome readthrough or frameshifting to produce their replicase proteins. The Moloney murine leukaemia virus (MoMLV), a member of the Retroviridae family, has a readthrough-enhancing pseudoknot immediately downstream of the gag termination codon, allowing the production of the Gag-Pol polyprotein. Consistently, reporter assays showed that placing the MoMLV readthrough-enhancing pseudoknot immediately downstream of the reporter ORF prevents decay of the otherwise NMD-sensitive reporter mRNA [168,169].
Similarly, a hairpin structure from the dsRNA Colorado tick fever virus (CTFV) was shown to stabilise a dual-reporter transcript, due to NMD avoidance [169]. Inherent viral readthrough- and frameshift-enhancing elements lead to UPF1 displacement by ribosomes and thus stabilisation of NMD-sensitive viral polycistronic RNAs [169].
Table 4. NMD evasion cis-strategies of animal viruses.
Table 4. NMD evasion cis-strategies of animal viruses.
NMD-Resistant
RNA Feature
Virus Classification
RealmGenome aFamilyGenusName bRef.
Programmed ribosome readthrough or frameshiftingRiboviriassRNA-RTRetroviridaeGammaretrovirusMoMLV[168]
RiboviriadsRNAReoviridaeColtivirusCTFV[169]
Regulated RNA splicing and exportRiboviriassRNA-RTRetroviridaeLentivirusHIV-1[170,171]
3′-UTR RSE bound by PTBP1RiboviriassRNA-RTRetroviridaeAlpharetrovirusRSV[95,172]
a Abbreviated genome groups: ssRNA-RT, single-stranded RNA reverse transcribing; dsRNA, double-stranded RNA. b Abbreviated virus names: MoMLV, Moloney murine leukemia virus; CTFV, Colorado tick fever virus; HIV-1, human immunodeficiency virus-1; RSV, Rous sarcoma virus.
The ssRNA-RT virus has a complex genome that is integrated to the host cell genome before being transcribed and spliced by the cellular machinery. All retroviruses, including HIV-1, have primary retroviral transcripts characterised by a major 5′ splicing site in their 5′-UTR, upstream of the gag gene, and a 3′ splicing site downstream of the pol gene, turning gag-pol into an intron [173]. The HIV-1 genome can produce >30 different mRNAs, with 3 types of splicing patterns: unspliced, once-spliced, and completely spliced [174,175]. The first genes to be expressed during HIV-1 replication (tat, rev and nef) arise from completely spliced mRNAs that are not vulnerable to NMD because they contain no EJC downstream of the termination codon [171]. Subsequently, the transactivating Rev protein binds to incompletely spliced HIV-1 RNAs and mediates their export to the cytoplasm, where they are translated with no susceptibility to NMD [170,176]. During this process, Rev is suggested to interact with UPF1’s CH domain [160]. Immunoprecipitation assays indicate a competition between Rev and UPF2 for UPF1 binding suggesting that UPF2 is excluded from HIV RNPs while UPF1 is present [159,160]. Additionally, overexpression of UPF2 appears to prevent the export of gRNA from the nucleus [159,160].
Some viruses have evolved internal RNA sequences rendering them NMD-resistant: The Rous sarcoma virus (RSV) comprises the RNA stability element (RSE) [95,172,177,178]. The RSV RSE is a 400-nt long region, with a 150-nt minimal functional element, containing a ~30-nt AU-rich stretch at its 5′ end associated to several stem-loop structures [172,177,178]. The RSE is located immediately downstream of the gag termination codon in RSV unspliced RNA, which can serve both as a gRNA and RNA template for the synthesis of Gag and Gag-Pol polyproteins. Translation of only the gag gene results in a 7-kb long 3′-UTR that would normally be recognised by the NMD machinery [172,177,178]. However, RSV RSE was shown to recruit PTBP1, and thereby exclude UPF1 from RSV unspliced RNAs, preventing NMD (see above) [95].

4.3.3. Viral NMD Evasion trans-Strategies: NMD Inhibition by Viral Proteins

In addition to viral RNA features that prevent NMD, viruses can avoid NMD by interacting directly with NMD factors. This results in global NMD inhibition and concomitant upregulation of cellular NMD targets. In the case of coronaviruses (CoV), several studies support an NMD-antagonist role of the nucleocapsid (N) protein (Table 5) [157,179,180]. In agreement, UPF1 was found to interact with the N protein in interactome-mapping studies of the avian infectious bronchitis virus (IBV) and severe acute respiratory syndrome CoV 2 (SARS-CoV-2) [179,180]. However, the molecular mechanism of UPF1 inhibition by the N-protein remains to be elucidated.
With regard to Alphaviruses, UPF1 depletion increased the half-life of viral RNAs in HeLa cells infected with a SFV variant [156]. Similarly, UPF1 suppression resulted in higher virus release from HeLa cells infected with a SFV mutant [156]. It was suggested that SFV replication machinery could inhibit NMD by displacing UPF1 from the positive-sense gRNA [156]. A recent virus-host protein interactome investigation of SFV in HeLa cells identified UPF1 as an antiviral factor interacting with SFV capsid protein [181]. Consistently, expression of the capsid protein induced increased levels of NMD-target RNAs, indicating that it interferes with cellular NMD [181]. However, the molecular mechanism of UPF1 inhibition by SFV’s capsid protein is still enigmatic.
In Flaviviridae, RNAs normally targeted by NMD were shown to be stabilised during infection for hepatitis C virus (HCV), West Nile virus (WNV), dengue virus (DENV) and ZIKV [158,182,183]. Immunoprecipitation experiments in HCV-infected cells showed that this capsid protein interacts with PYM1 (Partner of Y14 and MAGOH homologue 1) [182]. PYM1 is known to interact with both RBM8A (also known as Y14) and MAGOH, and to recycle RNA-bound EJCs [9,184,185]. PYM1 suppression in hepatoma cells leads to less HCV replication, suggesting that PYM1 is a pro-viral factor of HCV [182]. Immunoprecipitation experiments in HCV-infected hepatoma cells indicated a reduced interaction of PYM1 with RBM8A and MAGOH, suggesting that the capsid protein of HCV might interfere with this interaction. Similarly, PYM1 was found to interact with the capsid proteins from DENV, WNV and ZIKV, suggesting that this interaction is conserved among Flaviviridae family members [158,182,183]. Intriguingly and in contrast to HCV, PYM1, along with MAGOH and UPF1, was identified as an anti-viral factor of WNV, DENV and ZIKV (who are from a different Flaviviridae genus than HCV) because knockdown of PYM1 increased viral infection and/or RNA levels in HEKs cells [183]. Fractionation experiments suggest that, by sequestrating PYM1, the capsid protein of Flaviviruses alters the location of RBM8A and MAGOH. Cross-linking immunoprecipitation (CLIP) in WNV-infected cells shows a specific interaction between RBM8A and viral RNA, which is reduced when PYM1 or MAGOH are suppressed [183]. This suggests that RBM8A, associated to other EJC factors, marks WNV RNA as a NMD substrate, and that Flaviviruses capsid protein interacts with PYM1 to impede EJC-RNA-binding and subsequent NMD.
Table 5. NMD evasion trans-strategies of animal viruses.
Table 5. NMD evasion trans-strategies of animal viruses.
NMD InhibitionVirus Classification
Viral ProteinMechanismRealmGenome aFamilyGenusName bRef.
N
(nucleocapsid)
Interacts with UPF1Riboviria+ssRNACoronaviridaeGammacoronavirusIBV[179]
Riboviria+ssRNACoronaviridaeBetacoronavirusMHV[157]
Riboviria+ssRNACoronaviridaeBetacoronavirusSARS-CoV-2[180]
C
(capsid)
Interacts with UPF1
Interacts with PYM, preventing its interaction with MAGOH and RBM8A
Riboviria+ssRNATogaviridaeAlphavirusSFV[181]
Riboviria+ssRNAFlaviviridaeHepacivirusHCV[182]
Riboviria+ssRNAFlaviviridaeFlavivirusWNV[183]
Riboviria+ssRNAFlaviviridaeFlavivirusDENV[183]
Riboviria+ssRNAFlaviviridaeFlavivirusZIKV[158]
Interacts with UPF1, promoting its degradation in the nucleusRiboviria+ssRNAFlaviviridaeFlavivirusZIKV[186]
Rex
(regulatory)
Unknown mechanismRiboviriassRNA-RTRetroviridaeDeltaretrovirusHTLV-1[187]
Tax
(regulatory)
Interacts with INT6 and UPF1, impeding UPF1 association with RNA and promoting phospho-UPF1 sequestration in P-bodiesRiboviriassRNA-RTRetroviridaeDeltaretrovirusHTLV-1[188,189]
a Abbreviated genome groups: +ssRNA, positive-sense single-stranded RNA; ssRNA-RT, single-stranded RNA reverse transcribing. b Abbreviated virus names: IBV, infectious bronchitis virus; MHV, mouse hepatitis virus; SARS-CoV-2, severe acute respiratory syndrome coronavirus-2; SFV, Semliki forest virus; HCV, hepatitis C virus; WNV, West Nile virus; DENV, dengue virus; ZIKV, Zika virus; HTLV-1, human T-cell leukaemia virus type 1.
In addition to its interaction with PYM1, the capsid protein of ZIKV was shown to interact with UPF1 and UPF3B in ZIKV-infected HEK cells [158]. UPF1 depletion in ZIKV-infected NPCs resulted in higher infectious titres and viral RNA levels [158]. Intriguingly, lower protein levels of UPF1, but not of UPF3B, were observed in ZIKV-infected cells [158]. Consistently, in HEK cells expressing ZIKV capsid protein, UPF1 levels were considerably decreased in the nucleus, whereas they were unaltered in the cytoplasm [158]. Nuclear UPF1 levels in cells transfected with the ZIKV capsid protein were rescued by inhibition of the proteasome, indicating that ZIKV capsid protein promotes UPF1 degradation in the nucleus [158]. Expression of ZIKV capsid protein in hepatoma cells showed an accumulation of polyadenylated RNAs in the nucleus, which could also be reproduced by depleting UPF1 from the cells [186]. Taken together, these experiments suggest that UPF1 is involved in mRNA transport to the cytoplasm, and that ZIKV capsid protein disturbs this process by promoting degradation of nuclear UPF1 [186]. Notably, ZIKV infection of NPCs particularly impacts the expression of FREM2 (FRAS1-related extracellular matrix protein 2), which is a critical neurodevelopmental gene [186]. Knockdown of FREM2 in NPCs resulted in reduced expression of pluripotency marker Sox2 and increased expression of neuronal lineage marker βIII-Tubulin, mimicking human ZIKV infection, and potentially contributing to Zika syndrome [186,190].
Regarding retroviruses, human T-cell leukaemia virus type 1 (HTLV-1) was shown to globally inhibit NMD [187,188,189]. Two different viral regulatory proteins were identified: Rex and Tax [187,188,189]. Rex is an RNA-binding protein, responsible for the export of HTLV-1 unspliced and singly spliced mRNAs to the cytoplasm [191]. Rex can suppress NMD independently from its mRNA export function [187], but the mechanism of NMD-inhibition by Rex remains enigmatic. In contrast, the viral transcription transactivator Tax was shown to directly interact with INT6 (integrator complex subunit 6, also known as eIF3 subunit e) and with UPF1 [188,189]. Co-immunoprecipitation experiments in HEK cells indicated that Tax binding prevents the interaction between UPF1 and INT6, an interaction which is important for NMD [188,192]. Tax was shown to bind to UPF1’s helicase domain and impedes its association with RNA, its ATP hydrolysis, unwinding and translocation activities [189,192]. In addition, immunofluorescence experiments indicate that, in HeLa cells expressing Tax, hyper-phosphorylated UPF1 is sequestered in P-bodies [188]. It was therefore suggested that Tax may inhibit NMD by preventing UPF1’s dephosphorylation and recycling [33,44,188]. Even though Tax interaction with INT6 was shown to be required for Tax-mediated accumulation of UPF1 in P-bodies, the role of INT6 in this process remains unknown [188].

5. Concluding Remarks

In recent years, a picture of NMD pathways emerged involving a set of core NMD factors and an increasing number of additional proteins that finetune NMD for specific mRNA transcripts and cellular contexts. These include UPF1 isoforms, UPF3 paralogs, different EJC compositions and specialised NMD factors, such as NBAS and DHX34 during the cellular stress response. This additional complexity, compared to yeast NMD, shows that gene regulation by NMD is a major function of NMD factors in mammalian cells. This function was acquired during evolution in addition to mRNA quality control, the main NMD function in yeast. Mutations in NMD factors (in particular UPF3B) and NMD dysregulation can be associated with neurodevelopmental disease and cancer [193,194]. While NMD functions as an anti-viral defence mechanism against RNA viruses, these viruses have developed diverse strategies to evade NMD, often by directly targeting NMD factors [152,154]. An improved mechanistic understating of NMD will therefore inform the development of novel therapeutic strategies. For instance, tissue-directed gene knockdown by RNA aptamers to inhibit NMD locally has been successfully applied in tumour immunotherapy [143,144]. However, NMD stimulation may be required for other cancer therapies [195].
Interestingly, NMD core factor knockout is embryonically lethal in mice (except for UPF3B), indicating that NMD factors have a role during cell differentiation in early mammalian development [196]. However, research involving knockout or knockdown of NMD factors is complicated by NMD-independent roles of these factors in the nucleus, which affect genome stability, mRNA splicing and export [143,197]. Specifically, UPF1, UPF2, SMG1, SMG5-7 and SMG6 are reported to have functions in DNA damage response and maintenance of telomere integrity, processes which are vital for cells [196]. In particular, UPF1 has numerous roles independent of NMD, e.g., UPF1 regulates cell cycle S-phase progression [197] and replication-dependent decay of histone mRNAs [198], promotes R-loop formation in dsDNA break repair [199], has E3 ubiquitin ligase activity [104], and is involved in other RNA decay pathways, including Staufen-mediated mRNA decay [16]. Taken together, the additional roles and secondary effects of knockdown all complicate the study of molecular mechanisms of NMD factors in cells, requiring innovative approaches to investigate the specific NMD-related functions.

Author Contributions

Conceptualisation, writing, and editing: L.S., J.M. and C.S.; Figure Preparation: L.S. and J.M. All authors have read and agreed to the published version of the manuscript.

Funding

L.S. is funded by the Chinese Scholarship Council. C.S. is Investigator of the Wellcome Trust (210701/Z/18/Z).

Acknowledgments

We would like to thank all members of the Schaffitzel team for discussing the manuscript, critical reading and helpful comments.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Amrani, N.; Ganesan, R.; Kervestin, S.; Mangus, D.A.; Ghosh, S.; Jacobson, A. A faux 3′-UTR promotes aberrant termination and triggers nonsense-mediated mRNA decay. Nature 2004, 432, 112–118. [Google Scholar] [CrossRef]
  2. Eberle, A.B.; Stalder, L.; Mathys, H.; Orozco, R.Z.; Mühlemann, O. Posttranscriptional gene regulation by spatial rearrangement of the 3′ untranslated region. PLoS Biol. 2008, 6, e92. [Google Scholar] [CrossRef]
  3. Peixeiro, I.; Inácio, Â.; Barbosa, C.; Silva, A.L.; Liebhaber, S.A.; Romão, L. Interaction of PABPC1 with the translation initiation complex is critical to the NMD resistance of AUG-proximal nonsense mutations. Nucleic Acids Res. 2012, 40, 1160–1173. [Google Scholar] [CrossRef] [PubMed]
  4. Ivanov, A.; Mikhailova, T.; Eliseev, B.; Yeramala, L.; Sokolova, E.; Susorov, D.; Shuvalov, A.; Schaffitzel, C.; Alkalaeva, E. PABP enhances release factor recruitment and stop codon recognition during translation termination. Nucleic Acids Res. 2016, 44, 7766–7776. [Google Scholar] [CrossRef] [PubMed]
  5. Lykke-Andersen, J.; Shu, M.D.; Steitz, J.A. Human Upf proteins target an mRNA for nonsense-mediated decay when bound downstream of a termination codon. Cell 2000, 103, 1121–1131. [Google Scholar] [CrossRef]
  6. Le Hir, H.; Gatfield, D.; Izaurralde, E.; Moore, M.J. The exon-exon junction complex provides a binding platform for factors involved in mRNA export and nonsense-mediated mRNA decay. EMBO J. 2001, 20, 4987–4997. [Google Scholar] [CrossRef]
  7. Lindeboom, R.G.; Supek, F.; Lehner, B. The rules and impact of nonsense-mediated mRNA decay in human cancers. Nat. Genet. 2016, 48, 1112–1118. [Google Scholar] [CrossRef] [PubMed]
  8. Le Hir, H.; Izaurralde, E.; Maquat, L.E.; Moore, M.J. The spliceosome deposits multiple proteins 20-24 nucleotides upstream of mRNA exon-exon junctions. EMBO J. 2000, 19, 6860–6869. [Google Scholar] [CrossRef]
  9. Gehring, N.H.; Lamprinaki, S.; Kulozik, A.E.; Hentze, M.W. Disassembly of exon junction complexes by PYM. Cell 2009, 137, 536–548. [Google Scholar] [CrossRef]
  10. Raimondeau, E.; Bufton, J.C.; Schaffitzel, C. New insights into the interplay between the translation machinery and nonsense-mediated mRNA decay factors. Biochem. Soc. Trans. 2018, 46, 503–512. [Google Scholar] [CrossRef]
  11. Karousis, E.D.; Mühlemann, O. Nonsense-Mediated mRNA Decay Begins Where Translation Ends. Cold Spring Harb. Perspect. Biol. 2019, 11, a032862. [Google Scholar] [CrossRef] [PubMed]
  12. Lavysh, D.; Neu-Yilik, G. UPF1-Mediated RNA Decay-Danse Macabre in a Cloud. Biomolecules 2020, 10, 999. [Google Scholar] [CrossRef] [PubMed]
  13. Lykke-Andersen, S.; Jensen, T.H. Nonsense-mediated mRNA decay: An intricate machinery that shapes transcriptomes. Nat. Rev. Mol. Cell Biol. 2015, 16, 665–677. [Google Scholar] [CrossRef] [PubMed]
  14. Kurosaki, T.; Popp, M.W.; Maquat, L.E. Quality and quantity control of gene expression by nonsense-mediated mRNA decay. Nat. Rev. Mol. Cell Biol. 2019, 20, 406–420. [Google Scholar] [CrossRef]
  15. Nasif, S.; Contu, L.; Mühlemann, O. Beyond quality control: The role of nonsense-mediated mRNA decay (NMD) in regulating gene expression. Semin. Cell Dev. Biol. 2018, 75, 78–87. [Google Scholar] [CrossRef]
  16. Kim, Y.K.; Maquat, L.E. UPFront and center in RNA decay: UPF1 in nonsense-mediated mRNA decay and beyond. RNA 2019, 25, 407–422. [Google Scholar] [CrossRef]
  17. Pawlicka, K.; Kalathiya, U.; Alfaro, J. Nonsense-Mediated mRNA Decay: Pathologies and the Potential for Novel Therapeutics. Cancers 2020, 12, 765. [Google Scholar] [CrossRef]
  18. Fernandes, R.; Nogueira, G.; Da Costa, P.J.; Pinto, F.; Romão, L. Nonsense-Mediated mRNA Decay in Development, Stress and Cancer. Adv. Exp. Med. Biol. 2019, 1157, 41–83. [Google Scholar] [CrossRef]
  19. Lindeboom, R.G.H.; Vermeulen, M.; Lehner, B.; Supek, F. The impact of nonsense-mediated mRNA decay on genetic disease, gene editing and cancer immunotherapy. Nat. Genet. 2019, 51, 1645–1651. [Google Scholar] [CrossRef]
  20. Muhlrad, D.; Parker, R. Aberrant mRNAs with extended 3′ UTRs are substrates for rapid degradation by mRNA surveillance. RNA 1999, 5, 1299–1307. [Google Scholar] [CrossRef]
  21. Ruiz-Echevarría, M.J.; Peltz, S.W. The RNA binding protein Pub1 modulates the stability of transcripts containing upstream open reading frames. Cell 2000, 101, 741–751. [Google Scholar] [CrossRef] [PubMed]
  22. Yepiskoposyan, H.; Aeschimann, F.; Nilsson, D.; Okoniewski, M.; Mühlemann, O. Autoregulation of the nonsense-mediated mRNA decay pathway in human cells. RNA 2011, 17, 2108–2118. [Google Scholar] [CrossRef] [PubMed]
  23. Usuki, F.; Yamashita, A.; Fujimura, M. Post-transcriptional defects of antioxidant selenoenzymes cause oxidative stress under methylmercury exposure. J. Biol. Chem. 2011, 286, 6641–6649. [Google Scholar] [CrossRef] [PubMed]
  24. Bicknell, A.A.; Cenik, C.; Chua, H.N.; Roth, F.P.; Moore, M.J. Introns in UTRs: Why we should stop ignoring them. Bioessays 2012, 34, 1025–1034. [Google Scholar] [CrossRef]
  25. Lykke-Andersen, S.; Chen, Y.; Ardal, B.R.; Lilje, B.; Waage, J.; Sandelin, A.; Jensen, T.H. Human nonsense-mediated RNA decay initiates widely by endonucleolysis and targets snoRNA host genes. Genes Dev. 2014, 28, 2498–2517. [Google Scholar] [CrossRef] [PubMed]
  26. Wilusz, J.E.; Wilusz, J. Nonsense-mediated RNA decay: At the ‘cutting edge’ of regulated snoRNA production. Genes Dev. 2014, 28, 2447–2449. [Google Scholar] [CrossRef]
  27. Andjus, S.; Morillon, A.; Wery, M. From Yeast to Mammals, the Nonsense-Mediated mRNA Decay as a Master Regulator of Long Non-Coding RNAs Functional Trajectory. Noncoding RNA 2021, 7, 44. [Google Scholar] [CrossRef]
  28. Yi, Z.; Sanjeev, M.; Singh, G. The Branched Nature of the Nonsense-Mediated mRNA Decay Pathway. Trends Genet. 2021, 37, 143–159. [Google Scholar] [CrossRef]
  29. Mailliot, J.; Vivoli-Vega, M.; Schaffitzel, C. No-nonsense: Insights into the functional interplay of nonsense-mediated mRNA decay factors. Biochem. J. 2022, 479, 973–993. [Google Scholar] [CrossRef]
  30. Kashima, I.; Yamashita, A.; Izumi, N.; Kataoka, N.; Morishita, R.; Hoshino, S.; Ohno, M.; Dreyfuss, G.; Ohno, S. 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. 2006, 20, 355–367. [Google Scholar] [CrossRef] [PubMed]
  31. Yamashita, A.; Izumi, N.; Kashima, I.; Ohnishi, T.; Saari, B.; Katsuhata, Y.; Muramatsu, R.; Morita, T.; Iwamatsu, A.; Hachiya, T.; et al. 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. 2009, 23, 1091–1105. [Google Scholar] [CrossRef]
  32. Durand, S.; Franks, T.M.; Lykke-Andersen, J. Hyperphosphorylation amplifies UPF1 activity to resolve stalls in nonsense-mediated mRNA decay. Nat. Commun. 2016, 7, 12434. [Google Scholar] [CrossRef] [PubMed]
  33. Durand, S.; Cougot, N.; Mahuteau-Betzer, F.; Nguyen, C.H.; Grierson, D.S.; Bertrand, E.; Tazi, J.; Lejeune, F. Inhibition of nonsense-mediated mRNA decay (NMD) by a new chemical molecule reveals the dynamic of NMD factors in P-bodies. J. Cell Biol. 2007, 178, 1145–1160. [Google Scholar] [CrossRef] [PubMed]
  34. Serdar, L.D.; Whiteside, D.L.; Nock, S.L.; McGrath, D.; Baker, K.E. Inhibition of post-termination ribosome recycling at premature termination codons in UPF1 ATPase mutants. eLife 2020, 9, e57834. [Google Scholar] [CrossRef] [PubMed]
  35. Ganesan, R.; Mangkalaphiban, K.; Baker, R.E.; He, F.; Jacobson, A. Ribosome-bound Upf1 forms distinct 80S complexes and conducts mRNA surveillance. RNA 2022, 28, 1621–1642. [Google Scholar] [CrossRef]
  36. Pisarev, A.V.; Hellen, C.U.; Pestova, T.V. Recycling of eukaryotic posttermination ribosomal complexes. Cell 2007, 131, 286–299. [Google Scholar] [CrossRef]
  37. Skabkin, M.A.; Skabkina, O.V.; Hellen, C.U.; Pestova, T.V. Reinitiation and other unconventional posttermination events during eukaryotic translation. Mol. Cell 2013, 51, 249–264. [Google Scholar] [CrossRef] [PubMed]
  38. Isken, O.; Kim, Y.K.; Hosoda, N.; Mayeur, G.L.; Hershey, J.W.; Maquat, L.E. Upf1 phosphorylation triggers translational repression during nonsense-mediated mRNA decay. Cell 2008, 133, 314–327. [Google Scholar] [CrossRef]
  39. Okada-Katsuhata, Y.; Yamashita, A.; Kutsuzawa, K.; Izumi, N.; Hirahara, F.; Ohno, S. N- and C-terminal Upf1 phosphorylations create binding platforms for SMG-6 and SMG-5:SMG-7 during NMD. Nucleic Acids Res. 2012, 40, 1251–1266. [Google Scholar] [CrossRef]
  40. Eberle, A.B.; Lykke-Andersen, S.; Mühlemann, O.; Jensen, T.H. SMG6 promotes endonucleolytic cleavage of nonsense mRNA in human cells. Nat. Struct. Mol. Biol. 2009, 16, 49–55. [Google Scholar] [CrossRef]
  41. Boehm, V.; Haberman, N.; Ottens, F.; Ule, J.; Gehring, N.H. 3′ UTR length and messenger ribonucleoprotein composition determine endocleavage efficiencies at termination codons. Cell Rep. 2014, 9, 555–568. [Google Scholar] [CrossRef] [PubMed]
  42. Chen, C.Y.; Shyu, A.B. Rapid deadenylation triggered by a nonsense codon precedes decay of the RNA body in a mammalian cytoplasmic nonsense-mediated decay pathway. Mol. Cell. Biol. 2003, 23, 4805–4813. [Google Scholar] [CrossRef]
  43. Lejeune, F.; Li, X.; Maquat, L.E. Nonsense-mediated mRNA decay in mammalian cells involves decapping, deadenylating, and exonucleolytic activities. Mol. Cell 2003, 12, 675–687. [Google Scholar] [CrossRef]
  44. Ohnishi, T.; Yamashita, A.; Kashima, I.; Schell, T.; Anders, K.R.; Grimson, A.; Hachiya, T.; Hentze, M.W.; Anderson, P.; Ohno, S. Phosphorylation of hUPF1 induces formation of mRNA surveillance complexes containing hSMG-5 and hSMG-7. Mol. Cell 2003, 12, 1187–1200. [Google Scholar] [CrossRef] [PubMed]
  45. Bono, F.; Ebert, J.; Lorentzen, E.; Conti, E. The crystal structure of the exon junction complex reveals how it maintains a stable grip on mRNA. Cell 2006, 126, 713–725. [Google Scholar] [CrossRef] [PubMed]
  46. Andersen, C.B.; Ballut, L.; Johansen, J.S.; Chamieh, H.; Nielsen, K.H.; Oliveira, C.L.; Pedersen, J.S.; Séraphin, B.; Le Hir, H.; Andersen, G.R. Structure of the exon junction core complex with a trapped DEAD-box ATPase bound to RNA. Science 2006, 313, 1968–1972. [Google Scholar] [CrossRef]
  47. Gerbracht, J.V.; Boehm, V.; Britto-Borges, T.; Kallabis, S.; Wiederstein, J.L.; Ciriello, S.; Aschemeier, D.U.; Krüger, M.; Frese, C.K.; Altmüller, J.; et al. CASC3 promotes transcriptome-wide activation of nonsense-mediated decay by the exon junction complex. Nucleic Acids Res. 2020, 48, 8626–8644. [Google Scholar] [CrossRef]
  48. Gehring, N.H.; Kunz, J.B.; Neu-Yilik, G.; Breit, S.; Viegas, M.H.; Hentze, M.W.; Kulozik, A.E. Exon-junction complex components specify distinct routes of nonsense-mediated mRNA decay with differential cofactor requirements. Mol. Cell 2005, 20, 65–75. [Google Scholar] [CrossRef]
  49. Gehring, N.H.; Lamprinaki, S.; Hentze, M.W.; Kulozik, A.E. The hierarchy of exon-junction complex assembly by the spliceosome explains key features of mammalian nonsense-mediated mRNA decay. PLoS Biol. 2009, 7, e1000120. [Google Scholar] [CrossRef]
  50. Mabin, J.W.; Woodward, L.A.; Patton, R.D.; Yi, Z.; Jia, M.; Wysocki, V.H.; Bundschuh, R.; Singh, G. The Exon Junction Complex Undergoes a Compositional Switch that Alters mRNP Structure and Nonsense-Mediated mRNA Decay Activity. Cell Rep. 2018, 25, 2431–2446. [Google Scholar] [CrossRef]
  51. Clerici, M.; Mourão, A.; Gutsche, I.; Gehring, N.H.; Hentze, M.W.; Kulozik, A.; Kadlec, J.; Sattler, M.; Cusack, S. Unusual bipartite mode of interaction between the nonsense-mediated decay factors, UPF1 and UPF2. EMBO J. 2009, 28, 2293–2306. [Google Scholar] [CrossRef]
  52. Chakrabarti, S.; Jayachandran, U.; Bonneau, F.; Fiorini, F.; Basquin, C.; Domcke, S.; Le Hir, H.; Conti, E. Molecular mechanisms for the RNA-dependent ATPase activity of Upf1 and its regulation by Upf2. Mol. Cell 2011, 41, 693–703. [Google Scholar] [CrossRef] [PubMed]
  53. Xue, G.; Maciej, V.D.; Machado de Amorim, A.; Pak, M.; Jayachandran, U.; Chakrabarti, S. Modulation of RNA-binding properties of the RNA helicase UPF1 by its activator UPF2. RNA 2023, 29, 178–187. [Google Scholar] [CrossRef] [PubMed]
  54. Palma, M.; Leroy, C.; Salomé-Desnoulez, S.; Werkmeister, E.; Kong, R.; Mongy, M.; Le Hir, H.; Lejeune, F. A role for AKT1 in nonsense-mediated mRNA decay. Nucleic Acids Res. 2021, 49, 11022–11037. [Google Scholar] [CrossRef] [PubMed]
  55. Cho, H.; Abshire, E.T.; Popp, M.W.; Pröschel, C.; Schwartz, J.L.; Yeo, G.W.; Maquat, L.E. AKT constitutes a signal-promoted alternative exon-junction complex that regulates nonsense-mediated mRNA decay. Mol. Cell 2022, 82, 2779–2796.e2710. [Google Scholar] [CrossRef] [PubMed]
  56. Lykke-Andersen, J.; Shu, M.D.; Steitz, J.A. Communication of the position of exon-exon junctions to the mRNA surveillance machinery by the protein RNPS1. Science 2001, 293, 1836–1839. [Google Scholar] [CrossRef]
  57. Zhang, Z.; Krainer, A.R. Involvement of SR proteins in mRNA surveillance. Mol. Cell 2004, 16, 597–607. [Google Scholar] [CrossRef]
  58. Sato, H.; Hosoda, N.; Maquat, L.E. Efficiency of the pioneer round of translation affects the cellular site of nonsense-mediated mRNA decay. Mol. Cell 2008, 29, 255–262. [Google Scholar] [CrossRef]
  59. Aznarez, I.; Nomakuchi, T.T.; Tetenbaum-Novatt, J.; Rahman, M.A.; Fregoso, O.; Rees, H.; Krainer, A.R. Mechanism of Nonsense-Mediated mRNA Decay Stimulation by Splicing Factor SRSF1. Cell Rep. 2018, 23, 2186–2198. [Google Scholar] [CrossRef]
  60. Michlewski, G.; Sanford, J.R.; Cáceres, J.F. The splicing factor SF2/ASF regulates translation initiation by enhancing phosphorylation of 4E-BP1. Mol. Cell 2008, 30, 179–189. [Google Scholar] [CrossRef]
  61. Chiu, S.Y.; Serin, G.; Ohara, O.; Maquat, L.E. Characterization of human Smg5/7a: A protein with similarities to Caenorhabditis elegans SMG5 and SMG7 that functions in the dephosphorylation of Upf1. RNA 2003, 9, 77–87. [Google Scholar] [CrossRef]
  62. Anders, K.R.; Grimson, A.; Anderson, P. SMG-5, required for C.elegans nonsense-mediated mRNA decay, associates with SMG-2 and protein phosphatase 2A. EMBO J. 2003, 22, 641–650. [Google Scholar] [CrossRef] [PubMed]
  63. Colak, D.; Ji, S.J.; Porse, B.T.; Jaffrey, S.R. Regulation of axon guidance by compartmentalized nonsense-mediated mRNA decay. Cell 2013, 153, 1252–1265. [Google Scholar] [CrossRef] [PubMed]
  64. Longman, D.; Hug, N.; Keith, M.; Anastasaki, C.; Patton, E.E.; Grimes, G.; Cáceres, J.F. 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. 2013, 41, 8319–8331. [Google Scholar] [CrossRef]
  65. Alrahbeni, T.; Sartor, F.; Anderson, J.; Miedzybrodzka, Z.; McCaig, C.; Müller, B. Full UPF3B function is critical for neuronal differentiation of neural stem cells. Mol. Brain 2015, 8, 33. [Google Scholar] [CrossRef] [PubMed]
  66. Sakaki, K.; Kaufman, R.J. Interaction between quality control systems for ER protein folding and RNA biogenesis. Worm 2013, 2, e23005. [Google Scholar] [CrossRef]
  67. Sieber, J.; Hauer, C.; Bhuvanagiri, M.; Leicht, S.; Krijgsveld, J.; Neu-Yilik, G.; Hentze, M.W.; Kulozik, A.E. Proteomic Analysis Reveals Branch-specific Regulation of the Unfolded Protein Response by Nonsense-mediated mRNA Decay. Mol. Cell. Proteomics 2016, 15, 1584–1597. [Google Scholar] [CrossRef]
  68. Karam, R.; Lou, C.H.; Kroeger, H.; Huang, L.; Lin, J.H.; Wilkinson, M.F. The unfolded protein response is shaped by the NMD pathway. EMBO Rep. 2015, 16, 599–609. [Google Scholar] [CrossRef] [PubMed]
  69. Longman, D.; Jackson-Jones, K.A.; Maslon, M.M.; Murphy, L.C.; Young, R.S.; Stoddart, J.J.; Hug, N.; Taylor, M.S.; Papadopoulos, D.K.; Cáceres, J.F. Identification of a localized nonsense-mediated decay pathway at the endoplasmic reticulum. Genes Dev. 2020, 34, 1075–1088. [Google Scholar] [CrossRef]
  70. Karam, R.; Wengrod, J.; Gardner, L.B.; Wilkinson, M.F. Regulation of nonsense-mediated mRNA decay: Implications for physiology and disease. Biochim. Biophys. Acta 2013, 1829, 624–633. [Google Scholar] [CrossRef]
  71. Kunz, J.B.; Neu-Yilik, G.; Hentze, M.W.; Kulozik, A.E.; Gehring, N.H. Functions of hUpf3a and hUpf3b in nonsense-mediated mRNA decay and translation. RNA 2006, 12, 1015–1022. [Google Scholar] [CrossRef] [PubMed]
  72. Chan, W.K.; Bhalla, A.D.; Le Hir, H.; Nguyen, L.S.; Huang, L.; Gécz, J.; Wilkinson, M.F. A UPF3-mediated regulatory switch that maintains RNA surveillance. Nat. Struct. Mol. Biol. 2009, 16, 747–753. [Google Scholar] [CrossRef]
  73. Bao, J.; Tang, C.; Yuan, S.; Porse, B.T.; Yan, W. UPF2, a nonsense-mediated mRNA decay factor, is required for prepubertal Sertoli cell development and male fertility by ensuring fidelity of the transcriptome. Development 2015, 142, 352–362. [Google Scholar] [CrossRef]
  74. Shum, E.Y.; Jones, S.H.; Shao, A.; Dumdie, J.; Krause, M.D.; Chan, W.K.; Lou, C.H.; Espinoza, J.L.; Song, H.W.; Phan, M.H.; et al. The Antagonistic Gene Paralogs Upf3a and Upf3b Govern Nonsense-Mediated RNA Decay. Cell 2016, 165, 382–395. [Google Scholar] [CrossRef] [PubMed]
  75. Wallmeroth, D.; Lackmann, J.W.; Kueckelmann, S.; Altmüller, J.; Dieterich, C.; Boehm, V.; Gehring, N.H. Human UPF3A and UPF3B enable fault-tolerant activation of nonsense-mediated mRNA decay. EMBO J. 2022, 41, e109191. [Google Scholar] [CrossRef] [PubMed]
  76. Yi, Z.; Arvola, R.M.; Myers, S.; Dilsavor, C.N.; Abu Alhasan, R.; Carter, B.N.; Patton, R.D.; Bundschuh, R.; Singh, G. Mammalian UPF3A and UPF3B can activate nonsense-mediated mRNA decay independently of their exon junction complex binding. EMBO J. 2022, 41, e109202. [Google Scholar] [CrossRef]
  77. Gowravaram, M.; Bonneau, F.; Kanaan, J.; Maciej, V.D.; Fiorini, F.; Raj, S.; Croquette, V.; Le Hir, H.; Chakrabarti, S. A conserved structural element in the RNA helicase UPF1 regulates its catalytic activity in an isoform-specific manner. Nucleic Acids Res. 2018, 46, 2648–2659. [Google Scholar] [CrossRef]
  78. Fritz, S.E.; Ranganathan, S.; Wang, C.D.; Hogg, J.R. An alternative UPF1 isoform drives conditional remodeling of nonsense-mediated mRNA decay. EMBO J. 2022, 41, e108898. [Google Scholar] [CrossRef]
  79. Kadlec, J.; Guilligay, D.; Ravelli, R.B.; Cusack, S. Crystal structure of the UPF2-interacting domain of nonsense-mediated mRNA decay factor UPF1. RNA 2006, 12, 1817–1824. [Google Scholar] [CrossRef]
  80. Cheng, Z.; Muhlrad, D.; Lim, M.K.; Parker, R.; Song, H. Structural and functional insights into the human Upf1 helicase core. EMBO J. 2007, 26, 253–264. [Google Scholar] [CrossRef]
  81. Weng, Y.; Czaplinski, K.; Peltz, S.W. ATP is a cofactor of the Upf1 protein that modulates its translation termination and RNA binding activities. RNA 1998, 4, 205–214. [Google Scholar]
  82. Bhattacharya, A.; Czaplinski, K.; Trifillis, P.; He, F.; Jacobson, A.; Peltz, S.W. Characterization of the biochemical properties of the human Upf1 gene product that is involved in nonsense-mediated mRNA decay. RNA 2000, 6, 1226–1235. [Google Scholar] [CrossRef]
  83. Franks, T.M.; Singh, G.; Lykke-Andersen, J. Upf1 ATPase-dependent mRNP disassembly is required for completion of nonsense-mediated mRNA decay. Cell 2010, 143, 938–950. [Google Scholar] [CrossRef]
  84. Chapman, J.H.; Craig, J.M.; Wang, C.D.; Gundlach, J.H.; Neuman, K.C.; Hogg, J.R. UPF1 mutants with intact ATPase but deficient helicase activities promote efficient nonsense-mediated mRNA decay. Nucleic Acids Res. 2022, 50, 11876–11894. [Google Scholar] [CrossRef] [PubMed]
  85. Fiorini, F.; Bagchi, D.; Le Hir, H.; Croquette, V. Human Upf1 is a highly processive RNA helicase and translocase with RNP remodelling activities. Nat. Commun. 2015, 6, 7581. [Google Scholar] [CrossRef] [PubMed]
  86. Serdar, L.D.; Whiteside, D.L.; Baker, K.E. ATP hydrolysis by UPF1 is required for efficient translation termination at premature stop codons. Nat. Commun. 2016, 7, 14021. [Google Scholar] [CrossRef] [PubMed]
  87. Fritz, S.E.; Ranganathan, S.; Wang, C.D.; Hogg, J.R. The RNA-binding protein PTBP1 promotes ATPase-dependent dissociation of the RNA helicase UPF1 to protect transcripts from nonsense-mediated mRNA decay. J. Biol. Chem. 2020, 295, 11613–11625. [Google Scholar] [CrossRef]
  88. Kurosaki, T.; Li, W.; Hoque, M.; Popp, M.W.; Ermolenko, D.N.; Tian, B.; Maquat, L.E. A post-translational regulatory switch on UPF1 controls targeted mRNA degradation. Genes Dev. 2014, 28, 1900–1916. [Google Scholar] [CrossRef]
  89. Lee, S.R.; Pratt, G.A.; Martinez, F.J.; Yeo, G.W.; Lykke-Andersen, J. Target Discrimination in Nonsense-Mediated mRNA Decay Requires Upf1 ATPase Activity. Mol. Cell 2015, 59, 413–425. [Google Scholar] [CrossRef]
  90. National Center for Biotechnology Information. ClinVar. Available online: https://www.ncbi.nlm.nih.gov/clinvar (accessed on 13 January 2023).
  91. Jumper, J.; Evans, R.; Pritzel, A.; Green, T.; Figurnov, M.; Ronneberger, O.; Tunyasuvunakool, K.; Bates, R.; Žídek, A.; Potapenko, A.; et al. Highly accurate protein structure prediction with AlphaFold. Nature 2021, 596, 583–589. [Google Scholar] [CrossRef]
  92. Chamieh, H.; Ballut, L.; Bonneau, F.; Le Hir, H. NMD factors UPF2 and UPF3 bridge UPF1 to the exon junction complex and stimulate its RNA helicase activity. Nat. Struct. Mol. Biol. 2008, 15, 85–93. [Google Scholar] [CrossRef] [PubMed]
  93. Melero, R.; Buchwald, G.; Castaño, R.; Raabe, M.; Gil, D.; Lázaro, M.; Urlaub, H.; Conti, E.; Llorca, O. The cryo-EM structure of the UPF-EJC complex shows UPF1 poised toward the RNA 3′ end. Nat. Struct. Mol. Biol. 2012, 19, 498–505. [Google Scholar] [CrossRef] [PubMed]
  94. Padariya, M.; Fahraeus, R.; Hupp, T.; Kalathiya, U. Molecular Determinants and Specificity of mRNA with Alternatively-Spliced UPF1 Isoforms, Influenced by an Insertion in the ‘Regulatory Loop’. Int. J. Mol. Sci. 2021, 22, 2744. [Google Scholar] [CrossRef] [PubMed]
  95. Ge, Z.; Quek, B.L.; Beemon, K.L.; Hogg, J.R. Polypyrimidine tract binding protein 1 protects mRNAs from recognition by the nonsense-mediated mRNA decay pathway. eLife 2016, 5, e11155. [Google Scholar] [CrossRef] [PubMed]
  96. Kishor, A.; Ge, Z.; Hogg, J.R. hnRNP L-dependent protection of normal mRNAs from NMD subverts quality control in B cell lymphoma. EMBO J. 2019, 38, e99128. [Google Scholar] [CrossRef]
  97. Kishor, A.; Fritz, S.E.; Haque, N.; Ge, Z.; Tunc, I.; Yang, W.; Zhu, J.; Hogg, J.R. Activation and inhibition of nonsense-mediated mRNA decay control the abundance of alternative polyadenylation products. Nucleic Acids Res. 2020, 48, 7468–7482. [Google Scholar] [CrossRef] [PubMed]
  98. Deniaud, A.; Karuppasamy, M.; Bock, T.; Masiulis, S.; Huard, K.; Garzoni, F.; Kerschgens, K.; Hentze, M.W.; Kulozik, A.E.; Beck, M.; et al. A network of SMG-8, SMG-9 and SMG-1 C-terminal insertion domain regulates UPF1 substrate recruitment and phosphorylation. Nucleic Acids Res. 2015, 43, 7600–7611. [Google Scholar] [CrossRef]
  99. Brumbaugh, K.M.; Otterness, D.M.; Geisen, C.; Oliveira, V.; Brognard, J.; Li, X.; Lejeune, F.; Tibbetts, R.S.; Maquat, L.E.; Abraham, R.T. The mRNA surveillance protein hSMG-1 functions in genotoxic stress response pathways in mammalian cells. Mol. Cell 2004, 14, 585–598. [Google Scholar] [CrossRef]
  100. Langer, L.M.; Gat, Y.; Bonneau, F.; Conti, E. Structure of substrate-bound SMG1-8-9 kinase complex reveals molecular basis for phosphorylation specificity. eLife 2020, 9, e57127. [Google Scholar] [CrossRef]
  101. Chakrabarti, S.; Bonneau, F.; Schüssler, S.; Eppinger, E.; Conti, E. Phospho-dependent and phospho-independent interactions of the helicase UPF1 with the NMD factors SMG5-SMG7 and SMG6. Nucleic Acids Res. 2014, 42, 9447–9460. [Google Scholar] [CrossRef]
  102. Nicholson, P.; Josi, C.; Kurosawa, H.; Yamashita, A.; Mühlemann, O. A novel phosphorylation-independent interaction between SMG6 and UPF1 is essential for human NMD. Nucleic Acids Res. 2014, 42, 9217–9235. [Google Scholar] [CrossRef]
  103. Takahashi, S.; Araki, Y.; Ohya, Y.; Sakuno, T.; Hoshino, S.; Kontani, K.; Nishina, H.; Katada, T. Upf1 potentially serves as a RING-related E3 ubiquitin ligase via its association with Upf3 in yeast. RNA 2008, 14, 1950–1958. [Google Scholar] [CrossRef] [PubMed]
  104. Feng, Q.; Jagannathan, S.; Bradley, R.K. The RNA Surveillance Factor UPF1 Represses Myogenesis via Its E3 Ubiquitin Ligase Activity. Mol. Cell 2017, 67, 239–251.e236. [Google Scholar] [CrossRef]
  105. Chu, V.; Feng, Q.; Lim, Y.; Shao, S. Selective destabilization of polypeptides synthesized from NMD-targeted transcripts. Mol. Biol. Cell 2021, 32, ar38. [Google Scholar] [CrossRef]
  106. Inglis, A.J.; Guna, A.; Merchán, Á.G.; Pal, A.; Esantsi, T.K.; Keys, H.R.; Frenkel, E.M.; Oania, R.; Weissman, J.S.; Voorhees, R.M. Coupled protein quality control during nonsense mediated mRNA decay. bioRxiv 2022. [Google Scholar] [CrossRef]
  107. Joazeiro, C.A.P. Mechanisms and functions of ribosome-associated protein quality control. Nat. Rev. Mol. Cell Biol. 2019, 20, 368–383. [Google Scholar] [CrossRef]
  108. Powers, K.T.; Szeto, J.A.; Schaffitzel, C. New insights into no-go, non-stop and nonsense-mediated mRNA decay complexes. Curr. Opin. Struct. Biol. 2020, 65, 110–118. [Google Scholar] [CrossRef]
  109. Tatsuno, T.; Nakamura, Y.; Ma, S.; Tomosugi, N.; Ishigaki, Y. Nonsense-mediated mRNA decay factor Upf2 exists in both the nucleoplasm and the cytoplasm. Mol. Med. Rep. 2016, 14, 655–660. [Google Scholar] [CrossRef] [PubMed]
  110. Clerici, M.; Deniaud, A.; Boehm, V.; Gehring, N.H.; Schaffitzel, C.; Cusack, S. Structural and functional analysis of the three MIF4G domains of nonsense-mediated decay factor UPF2. Nucleic Acids Res. 2014, 42, 2673–2686. [Google Scholar] [CrossRef]
  111. López-Perrote, A.; Castaño, R.; Melero, R.; Zamarro, T.; Kurosawa, H.; Ohnishi, T.; Uchiyama, A.; Aoyagi, K.; Buchwald, G.; Kataoka, N.; et al. Human nonsense-mediated mRNA decay factor UPF2 interacts directly with eRF3 and the SURF complex. Nucleic Acids Res. 2016, 44, 1909–1923. [Google Scholar] [CrossRef] [PubMed]
  112. Kadlec, J.; Izaurralde, E.; Cusack, S. The structural basis for the interaction between nonsense-mediated mRNA decay factors UPF2 and UPF3. Nat. Struct. Mol. Biol. 2004, 11, 330–337. [Google Scholar] [CrossRef]
  113. Neu-Yilik, G.; Raimondeau, E.; Eliseev, B.; Yeramala, L.; Amthor, B.; Deniaud, A.; Huard, K.; Kerschgens, K.; Hentze, M.W.; Schaffitzel, C.; et al. Dual function of UPF3B in early and late translation termination. EMBO J. 2017, 36, 2968–2986. [Google Scholar] [CrossRef]
  114. Bufton, J.C.; Powers, K.T.; Szeto, J.A.; Toelzer, C.; Berger, I.; Schaffitzel, C. Structures of nonsense-mediated mRNA decay factors UPF3B and UPF3A in complex with UPF2 reveal molecular basis for competitive binding and for neurodevelopmental disorder-causing mutation. Nucleic Acids Res. 2022, 50, 5934–5947. [Google Scholar] [CrossRef]
  115. Varadi, M.; Anyango, S.; Deshpande, M.; Nair, S.; Natassia, C.; Yordanova, G.; Yuan, D.; Stroe, O.; Wood, G.; Laydon, A.; et al. AlphaFold Protein Structure Database: Massively expanding the structural coverage of protein-sequence space with high-accuracy models. Nucleic Acids Res. 2022, 50, D439–D444. [Google Scholar] [CrossRef] [PubMed]
  116. Leoyklang, P.; Suphapeetiporn, K.; Srichomthong, C.; Tongkobpetch, S.; Fietze, S.; Dorward, H.; Cullinane, A.R.; Gahl, W.A.; Huizing, M.; Shotelersuk, V. Disorders with similar clinical phenotypes reveal underlying genetic interaction: SATB2 acts as an activator of the UPF3B gene. Hum. Genet. 2013, 132, 1383–1393. [Google Scholar] [CrossRef] [PubMed]
  117. Nguyen, L.S.; Jolly, L.; Shoubridge, C.; Chan, W.K.; Huang, L.; Laumonnier, F.; Raynaud, M.; Hackett, A.; Field, M.; Rodriguez, J.; et al. Transcriptome profiling of UPF3B/NMD-deficient lymphoblastoid cells from patients with various forms of intellectual disability. Mol. Psychiatry 2012, 17, 1103–1115. [Google Scholar] [CrossRef]
  118. Buchwald, G.; Ebert, J.; Basquin, C.; Sauliere, J.; Jayachandran, U.; Bono, F.; Le Hir, H.; Conti, E. Insights into the recruitment of the NMD machinery from the crystal structure of a core EJC-UPF3b complex. Proc. Natl. Acad. Sci. USA 2010, 107, 10050–10055. [Google Scholar] [CrossRef]
  119. Gehring, N.H.; Neu-Yilik, G.; Schell, T.; Hentze, M.W.; Kulozik, A.E. Y14 and hUpf3b form an NMD-activating complex. Mol. Cell 2003, 11, 939–949. [Google Scholar] [CrossRef] [PubMed]
  120. Huang, L.; Shum, E.Y.; Jones, S.H.; Lou, C.H.; Chousal, J.; Kim, H.; Roberts, A.J.; Jolly, L.A.; Espinoza, J.L.; Skarbrevik, D.M.; et al. A Upf3b-mutant mouse model with behavioral and neurogenesis defects. Mol. Psychiatry 2018, 23, 1773–1786. [Google Scholar] [CrossRef] [PubMed]
  121. Domingo, D.; Nawaz, U.; Corbett, M.; Espinoza, J.L.; Tatton-Brown, K.; Coman, D.; Wilkinson, M.F.; Gecz, J.; Jolly, L.A. A synonymous UPF3B variant causing a speech disorder implicates NMD as a regulator of neurodevelopmental disorder gene networks. Hum. Mol. Genet. 2020, 29, 2568–2578. [Google Scholar] [CrossRef]
  122. Bruno, I.G.; Karam, R.; Huang, L.; Bhardwaj, A.; Lou, C.H.; Shum, E.Y.; Song, H.W.; Corbett, M.A.; Gifford, W.D.; Gecz, J.; et al. Identification of a microRNA that activates gene expression by repressing nonsense-mediated RNA decay. Mol. Cell 2011, 42, 500–510. [Google Scholar] [CrossRef]
  123. Nguyen, L.S.; Kim, H.G.; Rosenfeld, J.A.; Shen, Y.; Gusella, J.F.; Lacassie, Y.; Layman, L.C.; Shaffer, L.G.; Gécz, J. Contribution of copy number variants involving nonsense-mediated mRNA decay pathway genes to neuro-developmental disorders. Hum. Mol. Genet. 2013, 22, 1816–1825. [Google Scholar] [CrossRef]
  124. Johnson, J.L.; Stoica, L.; Liu, Y.; Zhu, P.J.; Bhattacharya, A.; Buffington, S.A.; Huq, R.; Eissa, N.T.; Larsson, O.; Porse, B.T.; et al. Inhibition of Upf2-Dependent Nonsense-Mediated Decay Leads to Behavioral and Neurophysiological Abnormalities by Activating the Immune Response. Neuron 2019, 104, 665–679.e668. [Google Scholar] [CrossRef] [PubMed]
  125. Tarpey, P.S.; Raymond, F.L.; Nguyen, L.S.; Rodriguez, J.; Hackett, A.; Vandeleur, L.; Smith, R.; Shoubridge, C.; Edkins, S.; Stevens, C.; et al. Mutations in UPF3B, a member of the nonsense-mediated mRNA decay complex, cause syndromic and nonsyndromic mental retardation. Nat. Genet. 2007, 39, 1127–1133. [Google Scholar] [CrossRef] [PubMed]
  126. Laumonnier, F.; Shoubridge, C.; Antar, C.; Nguyen, L.S.; Van Esch, H.; Kleefstra, T.; Briault, S.; Fryns, J.P.; Hamel, B.; Chelly, J.; et al. Mutations of the UPF3B gene, which encodes a protein widely expressed in neurons, are associated with nonspecific mental retardation with or without autism. Mol. Psychiatry 2010, 15, 767–776. [Google Scholar] [CrossRef] [PubMed]
  127. Addington, A.M.; Gauthier, J.; Piton, A.; Hamdan, F.F.; Raymond, A.; Gogtay, N.; Miller, R.; Tossell, J.; Bakalar, J.; Inoff-Germain, G.; et al. A novel frameshift mutation in UPF3B identified in brothers affected with childhood onset schizophrenia and autism spectrum disorders. Mol. Psychiatry 2011, 16, 238–239. [Google Scholar] [CrossRef]
  128. Szyszka, P.; Sharp, S.I.; Dedman, A.; Gurling, H.M.; McQuillin, A. A nonconservative amino acid change in the UPF3B gene in a patient with schizophrenia. Psychiatr. Genet. 2012, 22, 150–151. [Google Scholar] [CrossRef]
  129. Lynch, S.A.; Nguyen, L.S.; Ng, L.Y.; Waldron, M.; McDonald, D.; Gecz, J. Broadening the phenotype associated with mutations in UPF3B: Two further cases with renal dysplasia and variable developmental delay. Eur. J. Med. Genet. 2012, 55, 476–479. [Google Scholar] [CrossRef]
  130. Xu, X.; Zhang, L.; Tong, P.; Xun, G.; Su, W.; Xiong, Z.; Zhu, T.; Zheng, Y.; Luo, S.; Pan, Y.; et al. Exome sequencing identifies UPF3B as the causative gene for a Chinese non-syndrome mental retardation pedigree. Clin. Genet. 2013, 83, 560–564. [Google Scholar] [CrossRef]
  131. Tejada, M.I.; Villate, O.; Ibarluzea, N.; De la Hoz, A.B.; Martínez-Bouzas, C.; Beristain, E.; Martínez, F.; Friez, M.J.; Sobrino, B.; Barros, F. Molecular and Clinical Characterization of a Novel Nonsense Variant in Exon 1 of the UPF3B Gene Found in a Large Spanish Basque Family (MRX82). Front. Genet. 2019, 10, 1074. [Google Scholar] [CrossRef]
  132. Guo, H.; Duyzend, M.H.; Coe, B.P.; Baker, C.; Hoekzema, K.; Gerdts, J.; Turner, T.N.; Zody, M.C.; Beighley, J.S.; Murali, S.C.; et al. Genome sequencing identifies multiple deleterious variants in autism patients with more severe phenotypes. Genet. Med. 2019, 21, 1611–1620. [Google Scholar] [CrossRef] [PubMed]
  133. Grozeva, D.; Carss, K.; Spasic-Boskovic, O.; Tejada, M.I.; Gecz, J.; Shaw, M.; Corbett, M.; Haan, E.; Thompson, E.; Friend, K.; et al. Targeted Next-Generation Sequencing Analysis of 1,000 Individuals with Intellectual Disability. Hum. Mutat. 2015, 36, 1197–1204. [Google Scholar] [CrossRef] [PubMed]
  134. Medghalchi, S.M.; Frischmeyer, P.A.; Mendell, J.T.; Kelly, A.G.; Lawler, A.M.; Dietz, H.C. Rent1, a trans-effector of nonsense-mediated mRNA decay, is essential for mammalian embryonic viability. Hum. Mol. Genet. 2001, 10, 99–105. [Google Scholar] [CrossRef]
  135. Alzahrani, F.; Kuwahara, H.; Long, Y.; Al-Owain, M.; Tohary, M.; AlSayed, M.; Mahnashi, M.; Fathi, L.; Alnemer, M.; Al-Hamed, M.H.; et al. Recessive, Deleterious Variants in SMG8 Expand the Role of Nonsense-Mediated Decay in Developmental Disorders in Humans. Am. J. Hum. Genet. 2020, 107, 1178–1185. [Google Scholar] [CrossRef]
  136. Abdel-Salam, G.M.H.; Duan, R.; Abdel-Hamid, M.S.; Sayed, I.S.M.; Jhangiani, S.N.; Khan, Z.; Du, H.; Gibbs, R.A.; Posey, J.E.; Marafi, D.; et al. Expanding the phenotypic and allelic spectrum of SMG8: Clinical observations reveal overlap with SMG9-associated disease trait. Am. J. Med. Genet. A 2022, 188, 648–657. [Google Scholar] [CrossRef] [PubMed]
  137. Kurosaki, T.; Imamachi, N.; Pröschel, C.; Mitsutomi, S.; Nagao, R.; Akimitsu, N.; Maquat, L.E. Loss of the fragile X syndrome protein FMRP results in misregulation of nonsense-mediated mRNA decay. Nat. Cell Biol. 2021, 23, 40–48. [Google Scholar] [CrossRef]
  138. Martincorena, I.; Campbell, P.J. Somatic mutation in cancer and normal cells. Science 2015, 349, 1483–1489. [Google Scholar] [CrossRef]
  139. Bokhari, A.; Jonchere, V.; Lagrange, A.; Bertrand, R.; Svrcek, M.; Marisa, L.; Buhard, O.; Greene, M.; Demidova, A.; Jia, J.; et al. Targeting nonsense-mediated mRNA decay in colorectal cancers with microsatellite instability. Oncogenesis 2018, 7, 70. [Google Scholar] [CrossRef]
  140. Zhu, G.; Pan, C.; Bei, J.X.; Li, B.; Liang, C.; Xu, Y.; Fu, X. Mutant p53 in Cancer Progression and Targeted Therapies. Front. Oncol. 2020, 10, 595187. [Google Scholar] [CrossRef]
  141. Martin, L.; Grigoryan, A.; Wang, D.; Wang, J.; Breda, L.; Rivella, S.; Cardozo, T.; Gardner, L.B. Identification and characterization of small molecules that inhibit nonsense-mediated RNA decay and suppress nonsense p53 mutations. Cancer Res. 2014, 74, 3104–3113. [Google Scholar] [CrossRef]
  142. Pastor, F.; Kolonias, D.; Giangrande, P.H.; Gilboa, E. Induction of tumour immunity by targeted inhibition of nonsense-mediated mRNA decay. Nature 2010, 465, 227–230. [Google Scholar] [CrossRef] [PubMed]
  143. Zhang, Y.; Xie, X.; Yeganeh, P.N.; Lee, D.J.; Valle-Garcia, D.; Meza-Sosa, K.F.; Junqueira, C.; Su, J.; Luo, H.R.; Hide, W.; et al. Immunotherapy for breast cancer using EpCAM aptamer tumor-targeted gene knockdown. Proc. Natl. Acad. Sci. USA 2021, 118, e2022830118. [Google Scholar] [CrossRef] [PubMed]
  144. Soldevilla, M.M.; Villanueva, H.; Bendandi, M.; Inoges, S.; López-Díaz de Cerio, A.; Pastor, F. 2-fluoro-RNA oligonucleotide CD40 targeted aptamers for the control of B lymphoma and bone-marrow aplasia. Biomaterials 2015, 67, 274–285. [Google Scholar] [CrossRef] [PubMed]
  145. Cheruiyot, A.; Li, S.; Nonavinkere Srivatsan, S.; Ahmed, T.; Chen, Y.; Lemacon, D.S.; Li, Y.; Yang, Z.; Wadugu, B.A.; Warner, W.A.; et al. Nonsense-Mediated RNA Decay Is a Unique Vulnerability of Cancer Cells Harboring SF3B1 or U2AF1 Mutations. Cancer Res. 2021, 81, 4499–4513. [Google Scholar] [CrossRef] [PubMed]
  146. Wang, D.; Zavadil, J.; Martin, L.; Parisi, F.; Friedman, E.; Levy, D.; Harding, H.; Ron, D.; Gardner, L.B. Inhibition of nonsense-mediated RNA decay by the tumor microenvironment promotes tumorigenesis. Mol. Cell. Biol. 2011, 31, 3670–3680. [Google Scholar] [CrossRef]
  147. Li, L.; Geng, Y.; Feng, R.; Zhu, Q.; Miao, B.; Cao, J.; Fei, S. The Human RNA Surveillance Factor UPF1 Modulates Gastric Cancer Progression by Targeting Long Non-Coding RNA MALAT1. Cell. Physiol. Biochem. 2017, 42, 2194–2206. [Google Scholar] [CrossRef]
  148. Huang, Y.; Chen, S.; Qin, W.; Wang, Y.; Li, L.; Li, Q.; Yuan, X. A Novel RNA Binding Protein-Related Prognostic Signature for Hepatocellular Carcinoma. Front. Oncol. 2020, 10, 580513. [Google Scholar] [CrossRef]
  149. Man, Z.; Chen, Y.; Gao, L.; Xei, G.; Li, Q.; Lu, Q.; Yan, J. A Prognostic Model Based on RNA Binding Protein Predicts Clinical Outcomes in Hepatocellular Carcinoma Patients. Front. Oncol. 2020, 10, 613102. [Google Scholar] [CrossRef]
  150. Zhang, Z.; Wang, L.; Wang, Q.; Zhang, M.; Wang, B.; Jiang, K.; Ye, Y.; Wang, S.; Shen, Z. Molecular Characterization and Clinical Relevance of RNA Binding Proteins in Colorectal Cancer. Front. Genet. 2020, 11, 580149. [Google Scholar] [CrossRef]
  151. Paul, D.; Bartenschlager, R. Architecture and biogenesis of plus-strand RNA virus replication factories. World J. Virol. 2013, 2, 32–48. [Google Scholar] [CrossRef]
  152. Popp, M.W.; Cho, H.; Maquat, L.E. Viral subversion of nonsense-mediated mRNA decay. RNA 2020, 26, 1509–1518. [Google Scholar] [CrossRef] [PubMed]
  153. Contu, L.; Steiner, S.; Thiel, V.; Mühlemann, O. The Role of Stress Granules and the Nonsense-mediated mRNA Decay Pathway in Antiviral Defence. Chimia 2019, 73, 374–379. [Google Scholar] [CrossRef]
  154. Leon, K.; Ott, M. An ‘Arms Race’ between the Nonsense-mediated mRNA Decay Pathway and Viral Infections. Semin. Cell Dev. Biol. 2021, 111, 101–107. [Google Scholar] [CrossRef] [PubMed]
  155. Wernet, M.F.; Klovstad, M.; Clandinin, T.R. Generation of infectious virus particles from inducible transgenic genomes. Curr. Biol. 2014, 24, R107–R108. [Google Scholar] [CrossRef] [PubMed]
  156. Balistreri, G.; Horvath, P.; Schweingruber, C.; Zünd, D.; McInerney, G.; Merits, A.; Mühlemann, O.; Azzalin, C.; Helenius, A. The host nonsense-mediated mRNA decay pathway restricts Mammalian RNA virus replication. Cell Host Microbe 2014, 16, 403–411. [Google Scholar] [CrossRef] [PubMed]
  157. Wada, M.; Lokugamage, K.G.; Nakagawa, K.; Narayanan, K.; Makino, S. Interplay between coronavirus, a cytoplasmic RNA virus, and nonsense-mediated mRNA decay pathway. Proc. Natl. Acad. Sci. USA 2018, 115, E10157–E10166. [Google Scholar] [CrossRef]
  158. Fontaine, K.A.; Leon, K.E.; Khalid, M.M.; Tomar, S.; Jimenez-Morales, D.; Dunlap, M.; Kaye, J.A.; Shah, P.S.; Finkbeiner, S.; Krogan, N.J.; et al. The Cellular NMD Pathway Restricts Zika Virus Infection and Is Targeted by the Viral Capsid Protein. mBio 2018, 9, e02126-18. [Google Scholar] [CrossRef] [PubMed]
  159. Ajamian, L.; Abrahamyan, L.; Milev, M.; Ivanov, P.V.; Kulozik, A.E.; Gehring, N.H.; Mouland, A.J. Unexpected roles for UPF1 in HIV-1 RNA metabolism and translation. RNA 2008, 14, 914–927. [Google Scholar] [CrossRef] [PubMed]
  160. Ajamian, L.; Abel, K.; Rao, S.; Vyboh, K.; García-de-Gracia, F.; Soto-Rifo, R.; Kulozik, A.E.; Gehring, N.H.; Mouland, A.J. HIV-1 Recruits UPF1 but Excludes UPF2 to Promote Nucleocytoplasmic Export of the Genomic RNA. Biomolecules 2015, 5, 2808–2839. [Google Scholar] [CrossRef]
  161. Rao, S.; Amorim, R.; Niu, M.; Temzi, A.; Mouland, A.J. The RNA surveillance proteins UPF1, UPF2 and SMG6 affect HIV-1 reactivation at a post-transcriptional level. Retrovirology 2018, 15, 42. [Google Scholar] [CrossRef]
  162. Rao, S.; Amorim, R.; Niu, M.; Breton, Y.; Tremblay, M.J.; Mouland, A.J. Host mRNA decay proteins influence HIV-1 replication and viral gene expression in primary monocyte-derived macrophages. Retrovirology 2019, 16, 3. [Google Scholar] [CrossRef] [PubMed]
  163. Serquiña, A.K.; Das, S.R.; Popova, E.; Ojelabi, O.A.; Roy, C.K.; Göttlinger, H.G. UPF1 is crucial for the infectivity of human immunodeficiency virus type 1 progeny virions. J. Virol. 2013, 87, 8853–8861. [Google Scholar] [CrossRef] [PubMed]
  164. Zhao, Y.; Ye, X.; Shehata, M.; Dunker, W.; Xie, Z.; Karijolich, J. The RNA quality control pathway nonsense-mediated mRNA decay targets cellular and viral RNAs to restrict KSHV. Nat. Commun. 2020, 11, 3345. [Google Scholar] [CrossRef] [PubMed]
  165. Wilson, S.J.; Tsao, E.H.; Webb, B.L.; Ye, H.; Dalton-Griffin, L.; Tsantoulas, C.; Gale, C.V.; Du, M.Q.; Whitehouse, A.; Kellam, P. X box binding protein XBP-1s transactivates the Kaposi’s sarcoma-associated herpesvirus (KSHV) ORF50 promoter, linking plasma cell differentiation to KSHV reactivation from latency. J. Virol. 2007, 81, 13578–13586. [Google Scholar] [CrossRef] [PubMed]
  166. Yu, F.; Feng, J.; Harada, J.N.; Chanda, S.K.; Kenney, S.C.; Sun, R. B cell terminal differentiation factor XBP-1 induces reactivation of Kaposi’s sarcoma-associated herpesvirus. FEBS Lett. 2007, 581, 3485–3488. [Google Scholar] [CrossRef] [PubMed]
  167. Meier, R.; Franceschini, A.; Horvath, P.; Tetard, M.; Mancini, R.; von Mering, C.; Helenius, A.; Lozach, P.Y. Genome-wide small interfering RNA screens reveal VAMP3 as a novel host factor required for Uukuniemi virus late penetration. J. Virol. 2014, 88, 8565–8578. [Google Scholar] [CrossRef]
  168. Tang, X.; Zhu, Y.; Baker, S.L.; Bowler, M.W.; Chen, B.J.; Chen, C.; Hogg, J.R.; Goff, S.P.; Song, H. Structural basis of suppression of host translation termination by Moloney Murine Leukemia Virus. Nat. Commun. 2016, 7, 12070. [Google Scholar] [CrossRef]
  169. Baker, S.L.; Hogg, J.R. A system for coordinated analysis of translational readthrough and nonsense-mediated mRNA decay. PLoS ONE 2017, 12, e0173980. [Google Scholar] [CrossRef]
  170. Fischer, U.; Huber, J.; Boelens, W.C.; Mattaj, I.W.; Lührmann, R. The HIV-1 Rev activation domain is a nuclear export signal that accesses an export pathway used by specific cellular RNAs. Cell 1995, 82, 475–483. [Google Scholar] [CrossRef]
  171. Bohne, J.; Wodrich, H.; Kräusslich, H.G. Splicing of human immunodeficiency virus RNA is position-dependent suggesting sequential removal of introns from the 5′ end. Nucleic Acids Res. 2005, 33, 825–837. [Google Scholar] [CrossRef]
  172. Weil, J.E.; Beemon, K.L. A 3′ UTR sequence stabilizes termination codons in the unspliced RNA of Rous sarcoma virus. RNA 2006, 12, 102–110. [Google Scholar] [CrossRef]
  173. Beemon, K.L. Retroviral RNA Processing. Viruses 2022, 14, 1113. [Google Scholar] [CrossRef] [PubMed]
  174. Schwartz, S.; Felber, B.K.; Benko, D.M.; Fenyö, E.M.; Pavlakis, G.N. Cloning and functional analysis of multiply spliced mRNA species of human immunodeficiency virus type 1. J. Virol. 1990, 64, 2519–2529. [Google Scholar] [CrossRef] [PubMed]
  175. Purcell, D.F.; Martin, M.A. Alternative splicing of human immunodeficiency virus type 1 mRNA modulates viral protein expression, replication, and infectivity. J. Virol. 1993, 67, 6365–6378. [Google Scholar] [CrossRef]
  176. Daly, T.J.; Cook, K.S.; Gray, G.S.; Maione, T.E.; Rusche, J.R. Specific binding of HIV-1 recombinant Rev protein to the Rev-responsive element in vitro. Nature 1989, 342, 816–819. [Google Scholar] [CrossRef] [PubMed]
  177. Weil, J.E.; Hadjithomas, M.; Beemon, K.L. Structural characterization of the Rous sarcoma virus RNA stability element. J. Virol. 2009, 83, 2119–2129. [Google Scholar] [CrossRef] [PubMed]
  178. Withers, J.B.; Beemon, K.L. Structural features in the Rous sarcoma virus RNA stability element are necessary for sensing the correct termination codon. Retrovirology 2010, 7, 65. [Google Scholar] [CrossRef] [PubMed]
  179. Emmott, E.; Munday, D.; Bickerton, E.; Britton, P.; Rodgers, M.A.; Whitehouse, A.; Zhou, E.M.; Hiscox, J.A. The cellular interactome of the coronavirus infectious bronchitis virus nucleocapsid protein and functional implications for virus biology. J. Virol. 2013, 87, 9486–9500. [Google Scholar] [CrossRef]
  180. Gordon, D.E.; Jang, G.M.; Bouhaddou, M.; Xu, J.; Obernier, K.; White, K.M.; O’Meara, M.J.; Rezelj, V.V.; Guo, J.Z.; Swaney, D.L.; et al. A SARS-CoV-2 protein interaction map reveals targets for drug repurposing. Nature 2020, 583, 459–468. [Google Scholar] [CrossRef]
  181. Contu, L.; Balistreri, G.; Domanski, M.; Uldry, A.C.; Mühlemann, O. Characterisation of the Semliki Forest Virus-host cell interactome reveals the viral capsid protein as an inhibitor of nonsense-mediated mRNA decay. PLoS Pathog. 2021, 17, e1009603. [Google Scholar] [CrossRef]
  182. Ramage, H.R.; Kumar, G.R.; Verschueren, E.; Johnson, J.R.; Von Dollen, J.; Johnson, T.; Newton, B.; Shah, P.; Horner, J.; Krogan, N.J.; et al. A combined proteomics/genomics approach links hepatitis C virus infection with nonsense-mediated mRNA decay. Mol. Cell 2015, 57, 329–340. [Google Scholar] [CrossRef]
  183. Li, M.; Johnson, J.R.; Truong, B.; Kim, G.; Weinbren, N.; Dittmar, M.; Shah, P.S.; Von Dollen, J.; Newton, B.W.; Jang, G.M.; et al. Identification of antiviral roles for the exon-junction complex and nonsense-mediated decay in flaviviral infection. Nat. Microbiol. 2019, 4, 985–995. [Google Scholar] [CrossRef]
  184. Bono, F.; Ebert, J.; Unterholzner, L.; Güttler, T.; Izaurralde, E.; Conti, E. Molecular insights into the interaction of PYM with the Mago-Y14 core of the exon junction complex. EMBO Rep. 2004, 5, 304–310. [Google Scholar] [CrossRef]
  185. Diem, M.D.; Chan, C.C.; Younis, I.; Dreyfuss, G. PYM binds the cytoplasmic exon-junction complex and ribosomes to enhance translation of spliced mRNAs. Nat. Struct. Mol. Biol. 2007, 14, 1173–1179. [Google Scholar] [CrossRef]
  186. Leon, K.E.; Khalid, M.M.; Flynn, R.A.; Fontaine, K.A.; Nguyen, T.T.; Kumar, G.R.; Simoneau, C.R.; Tomar, S.; Jimenez-Morales, D.; Dunlap, M.; et al. Nuclear accumulation of host transcripts during Zika Virus Infection. PLoS Pathog. 2023, 19, e1011070. [Google Scholar] [CrossRef]
  187. Nakano, K.; Ando, T.; Yamagishi, M.; Yokoyama, K.; Ishida, T.; Ohsugi, T.; Tanaka, Y.; Brighty, D.W.; Watanabe, T. Viral interference with host mRNA surveillance, the nonsense-mediated mRNA decay (NMD) pathway, through a new function of HTLV-1 Rex: Implications for retroviral replication. Microbes Infect. 2013, 15, 491–505. [Google Scholar] [CrossRef] [PubMed]
  188. Mocquet, V.; Neusiedler, J.; Rende, F.; Cluet, D.; Robin, J.P.; Terme, J.M.; Duc Dodon, M.; Wittmann, J.; Morris, C.; Le Hir, H.; et al. The human T-lymphotropic virus type 1 tax protein inhibits nonsense-mediated mRNA decay by interacting with INT6/EIF3E and UPF1. J. Virol. 2012, 86, 7530–7543. [Google Scholar] [CrossRef] [PubMed]
  189. Fiorini, F.; Robin, J.P.; Kanaan, J.; Borowiak, M.; Croquette, V.; Le Hir, H.; Jalinot, P.; Mocquet, V. HTLV-1 Tax plugs and freezes UPF1 helicase leading to nonsense-mediated mRNA decay inhibition. Nat. Commun. 2018, 9, 431. [Google Scholar] [CrossRef]
  190. Ferraris, P.; Cochet, M.; Hamel, R.; Gladwyn-Ng, I.; Alfano, C.; Diop, F.; Garcia, D.; Talignani, L.; Montero-Menei, C.N.; Nougairède, A.; et al. Zika virus differentially infects human neural progenitor cells according to their state of differentiation and dysregulates neurogenesis through the Notch pathway. Emerg. Microbes Infect. 2019, 8, 1003–1016. [Google Scholar] [CrossRef] [PubMed]
  191. Nakano, K.; Watanabe, T. Tuning Rex rules HTLV-1 pathogenesis. Front. Immunol. 2022, 13, 959962. [Google Scholar] [CrossRef]
  192. Morris, C.; Wittmann, J.; Jäck, H.M.; Jalinot, P. Human INT6/eIF3e is required for nonsense-mediated mRNA decay. EMBO Rep. 2007, 8, 596–602. [Google Scholar] [CrossRef] [PubMed]
  193. Supek, F.; Lehner, B.; Lindeboom, R.G.H. To NMD or Not To NMD: Nonsense-Mediated mRNA Decay in Cancer and Other Genetic Diseases. Trends Genet. 2021, 37, 657–668. [Google Scholar] [CrossRef] [PubMed]
  194. Lee, P.J.; Yang, S.; Sun, Y.; Guo, J.U. Regulation of nonsense-mediated mRNA decay in neural development and disease. J. Mol. Cell Biol. 2021, 13, 269–281. [Google Scholar] [CrossRef] [PubMed]
  195. Tan, K.; Stupack, D.G.; Wilkinson, M.F. Nonsense-mediated RNA decay: An emerging modulator of malignancy. Nat. Rev. Cancer 2022, 22, 437–451. [Google Scholar] [CrossRef] [PubMed]
  196. Han, X.; Wei, Y.; Wang, H.; Wang, F.; Ju, Z.; Li, T. Nonsense-mediated mRNA decay: A ‘nonsense’ pathway makes sense in stem cell biology. Nucleic Acids Res. 2018, 46, 1038–1051. [Google Scholar] [CrossRef] [PubMed]
  197. Azzalin, C.M.; Lingner, J. The human RNA surveillance factor UPF1 is required for S phase progression and genome stability. Curr. Biol. 2006, 16, 433–439. [Google Scholar] [CrossRef]
  198. Choe, J.; Ahn, S.H.; Kim, Y.K. The mRNP remodeling mediated by UPF1 promotes rapid degradation of replication-dependent histone mRNA. Nucleic Acids Res. 2014, 42, 9334–9349. [Google Scholar] [CrossRef]
  199. Ngo, G.H.P.; Grimstead, J.W.; Baird, D.M. UPF1 promotes the formation of R loops to stimulate DNA double-strand break repair. Nat. Commun. 2021, 12, 3849. [Google Scholar] [CrossRef]
Figure 1. Scheme showing different factor requirements for the main and alternative pathways leading to the activation of the key NMD factor UPF1. Upon inefficient translation termination at a premature termination codon (PTC), NMD is triggered by two-step activation of UPF1: transition from its closed to its open conformation, and hyper-phosphorylation of its N- and C-termini by SMG1-8-9 kinase. In the canonical EJC-dependent NMD pathway (black arrows), CASC3, UPF3B and UPF2 are required for UPF1 activation. UPF2 promotes the change in UPF1’s conformation by directly interacting with its CH domain. In the CASC3- and UPF3B-independent NMD pathway (dashed aquamarine arrows), RNPS1 directly recruits UPF2 to activate UPF1. In the UPF2-independent NMD pathway (dashed orange arrows), CASC3, UPF3B and AKT1 are required for the activation of UPF1. AKT1 promotes the change in UPF1’s conformation by phosphorylating its CH domain. SRSF1 enhances all NMD pathways by interacting with the EJC and NMD factors.
Figure 1. Scheme showing different factor requirements for the main and alternative pathways leading to the activation of the key NMD factor UPF1. Upon inefficient translation termination at a premature termination codon (PTC), NMD is triggered by two-step activation of UPF1: transition from its closed to its open conformation, and hyper-phosphorylation of its N- and C-termini by SMG1-8-9 kinase. In the canonical EJC-dependent NMD pathway (black arrows), CASC3, UPF3B and UPF2 are required for UPF1 activation. UPF2 promotes the change in UPF1’s conformation by directly interacting with its CH domain. In the CASC3- and UPF3B-independent NMD pathway (dashed aquamarine arrows), RNPS1 directly recruits UPF2 to activate UPF1. In the UPF2-independent NMD pathway (dashed orange arrows), CASC3, UPF3B and AKT1 are required for the activation of UPF1. AKT1 promotes the change in UPF1’s conformation by phosphorylating its CH domain. SRSF1 enhances all NMD pathways by interacting with the EJC and NMD factors.
Biomedicines 11 00722 g001
Figure 4. Domain architecture (above) and AlphaFold2 structure prediction (below) of human UPF3B isoform 2. UPF3B interaction regions with other proteins are indicated by thick and dashed lines. Missense mutations related to neurodevelopmental disorders are highlighted by red dashed lines in UPF3B protein scheme and red spheres in predicted UPF3B structure [90,91]. RRM-L, RNA recognition motif-like domain; NOPS-L, NONA/paraspeckle-like region; CCL, coiled-coil-like region; EBM, EJC-binding motif.
Figure 4. Domain architecture (above) and AlphaFold2 structure prediction (below) of human UPF3B isoform 2. UPF3B interaction regions with other proteins are indicated by thick and dashed lines. Missense mutations related to neurodevelopmental disorders are highlighted by red dashed lines in UPF3B protein scheme and red spheres in predicted UPF3B structure [90,91]. RRM-L, RNA recognition motif-like domain; NOPS-L, NONA/paraspeckle-like region; CCL, coiled-coil-like region; EBM, EJC-binding motif.
Biomedicines 11 00722 g004
Table 1. Missense mutations in UPF proteins found in patients with neurodevelopmental disorders.
Table 1. Missense mutations in UPF proteins found in patients with neurodevelopmental disorders.
ProteinDomainGene
Variation
Isoform 1Isoform 2Related DiseaseClinVar
Accession [90]
Ref.
UPF1Helicasec.2381C > TT805MT794MAutism, developmental delay, joint
hypermobility, hypotonia
VCV000996716.1-
c.2489A > GQ841RQ830RIntellectual disabilityVCV000930204.1-
UPF2N-terminusc.91G > TV31L Autism spectrum disorderVCV000996716.1[132]
UPF3BN-terminusc.52A > CT18PT18PSyndromic X-linked intellectual disability 14VCV000287042.5-
c.53C > TT18IT18ISyndromic X-linked intellectual disability 14VCV000698418.4-
RRM-Lc.277A > GM93VM93VSyndromic X-linked intellectual disability 14VCV001012886.1-
NOPS-Lc.470A > GD157GD157GSyndromic X-linked intellectual disability 14VCV001485374.3-
c.478T > GY160DY160DSyndromic X-linked intellectual disability 14VCV000011401.3[125]
Interdomainc.520A > GM174VM174VSyndromic X-linked intellectual disability 14VCV001718760.1-
CCL-1c.667A > GI223VI223VSyndromic X-linked intellectual disability 14VCV000985848.4[133]
c.672A > CE224DE224DIntellectual disabilityVCV000981400.1-
c.758T > CI253TI253TCataract, Microcephaly, Severe global
developmental delay
VCV000284394.21-
c.763A > GR255GR255GSyndromic X-linked intellectual disability 14VCV001029871.1-
c.764G > AR255KR255KSyndromic X-linked intellectual disability 14VCV000536851.9-
Interdomainc.962T > CL321SL308SSyndromic X-linked intellectual disability 14VCV000212548.9-
CCL-2c.1049G > AR350QR337QSyndromic X-linked intellectual disability 14VCV001469888.3-
c.1072C > TR358CR345CSyndromic X-linked intellectual disability 14VCV000700934.5-
c.1101G > CK367NK354NSyndromic X-linked intellectual disability 14VCV000804079.4-
c.1102C > TR368WR355WSyndromic X-linked intellectual disability 14VCV001041099.5-
c.1103G > AR368QR355QBorderline intellectual disability with autism-[126]
c.1117C > TR373CR360CSyndromic X-linked intellectual disability 14VCV001560636.4-
c.1118G > AR373HR360HSyndromic X-linked intellectual disability 14VCV000547850.2-
c.1121G > AR374HR361HSyndromic X-linked intellectual disability 14VCV000212546.17-
c.1136G > AR379HR366HSyndromic X-linked intellectual disability 14VCV001597732.4[126]
c.1201C > TR401WR388WSyndromic X-linked intellectual disability 14VCV001601042.3-
c.1202G > AR401QR388QSyndromic X-linked intellectual disability 14VCV000581972.4-
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sun, L.; Mailliot, J.; Schaffitzel, C. Nonsense-Mediated mRNA Decay Factor Functions in Human Health and Disease. Biomedicines 2023, 11, 722. https://doi.org/10.3390/biomedicines11030722

AMA Style

Sun L, Mailliot J, Schaffitzel C. Nonsense-Mediated mRNA Decay Factor Functions in Human Health and Disease. Biomedicines. 2023; 11(3):722. https://doi.org/10.3390/biomedicines11030722

Chicago/Turabian Style

Sun, Lingling, Justine Mailliot, and Christiane Schaffitzel. 2023. "Nonsense-Mediated mRNA Decay Factor Functions in Human Health and Disease" Biomedicines 11, no. 3: 722. https://doi.org/10.3390/biomedicines11030722

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

Article Metrics

Back to TopTop