Next Article in Journal
Nanoliposomal Delivery of MicroRNA-203 Suppresses Migration of Triple-Negative Breast Cancer through Distinct Target Suppression
Next Article in Special Issue
Interdependent Transcription of a Natural Sense/Antisense Transcripts Pair (SLC34A1/PFN3)
Previous Article in Journal
A Phylogenetic Approach to Structural Variation in Organization of Nuclear Group I Introns and Their Ribozymes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

From Yeast to Mammals, the Nonsense-Mediated mRNA Decay as a Master Regulator of Long Non-Coding RNAs Functional Trajectory

1
ncRNA, Epigenetic and Genome Fluidity, Institut Curie, PSL University, Sorbonne Université, CNRS UMR3244, 26 Rue d’Ulm, CEDEX 05, F-75248 Paris, France
2
ncRNA, Epigenetic and Genome Fluidity, Institut Curie, Sorbonne Université, CNRS UMR3244, 26 Rue d’Ulm, CEDEX 05, F-75248 Paris, France
*
Authors to whom correspondence should be addressed.
Submission received: 15 June 2021 / Revised: 22 July 2021 / Accepted: 25 July 2021 / Published: 27 July 2021

Abstract

:
The Nonsense-Mediated mRNA Decay (NMD) has been classically viewed as a translation-dependent RNA surveillance pathway degrading aberrant mRNAs containing premature stop codons. However, it is now clear that mRNA quality control represents only one face of the multiple functions of NMD. Indeed, NMD also regulates the physiological expression of normal mRNAs, and more surprisingly, of long non-coding (lnc)RNAs. Here, we review the different mechanisms of NMD activation in yeast and mammals, and we discuss the molecular bases of the NMD sensitivity of lncRNAs, considering the functional roles of NMD and of translation in the metabolism of these transcripts. In this regard, we describe several examples of functional micropeptides produced from lncRNAs. We propose that translation and NMD provide potent means to regulate the expression of lncRNAs, which might be critical for the cell to respond to environmental changes.

1. Introduction

The accurate transmission of the genetic information is crucial for the cell, and several surveillance mechanisms have evolved to monitor the distinct steps of gene expression. RNA surveillance pathways are responsible for detecting and eliminating RNA intermediates that lack integrity or functionality [1,2,3]. Such transcripts can arise due to deleterious or genomic frameshift mutations or inappropriate processing, and the subsequent failure to produce functional proteins may result in disease.
If an mRNA is devoid of a stop codon (for instance, in the case of truncation or premature 3′-end cleavage and polyadenylation), it will cause the ribosome to progress to its 3′ extremity and stall. Such aberrant mRNAs are rapidly degraded through a process termed non-stop decay [4,5,6]. In contrast, the presence of stable structures or damaged nucleotides within an open reading frame (ORF) can impede ribosome progression, resulting into ribosome stalling upstream of the stop codon. In this case, the transcript is targeted to the degradation by the no-go decay pathway [7,8].
The Nonsense-Mediated mRNA Decay (NMD) is another quality control pathway targeting transcripts that terminate translation prematurely [9,10], such as mRNAs harboring a premature termination codon (PTC) within the ORF [11], as well as PTC-less mRNAs displaying long 3′ untranslated regions (UTRs) [12,13] or short upstream ORFs [13,14,15,16]. The NMD-targeted mRNAs are rapidly degraded [17,18], thus preventing the production of truncated, possibly deleterious proteins [19,20,21].
Here, we review the different mechanisms of NMD activation in yeast and mammals. We discuss the recent evidence showing that NMD also targets and regulates the expression of long non-coding (lnc)RNAs, including antisense (as)lncRNAs in yeast, indicating that translation is part of the metabolism of transcripts initially thought to be devoid of coding potential. Supporting this idea, we describe several examples of functional micropeptides produced from small (sm)ORFs of lncRNAs. We propose that NMD and translation take part in the metabolism of lncRNAs, regulating their expression and providing the opportunity to produce micropeptides which might have a role in the cellular response to environmental changes.

2. Discovery, Conservation and Functions of NMD

NMD is a translation-dependent RNA decay pathway [22,23,24], which has been evolutionarily conserved [10,25]. It was originally discovered in the budding yeast Saccharomyces cerevisiae by Losson and Lacroute, when they observed that the presence of nonsense mutations reduces the level of a mutant mRNA without affecting its synthesis rate [26]. It was discovered afterwards in humans in the context of β0-thalassemia, where it was observed that β-globin mRNAs levels dramatically decrease when carrying nonsense mutations [27,28].
Upstream frameshift proteins (Upfs) 1, 2 and 3 constitute the conserved core components of NMD [29] and were initially identified in S. cerevisiae [30,31,32].
Upf1 is a monomeric, highly regulated superfamily 1 helicase. Its ATPase and helicase activities are essential for NMD [33,34]. Upf1 has the ability to translocate slowly but with high processivity on nucleic acids and to unwind long double-stranded (ds)RNA structures [35]. Upf2 is the second core NMD factor and functions as a bridge between Upf1 and Upf3 [36,37,38]. Its interaction with Upf1 is a prerequisite for the phosphorylation of Upf1 [39]. However, NMD can be activated independently of Upf2 [40,41]. Upf3 is the least conserved of the three core NMD factors [42]. Vertebrates have two Upf3 paralogs, Upf3A and Upf3B [37]; in human cells, Upf3B seems to be the main contributor to NMD [43]. Like Upf2, Upf3 stimulates the ATPase and helicase activity of Upf1 in vitro [36]. In metazoans, NMD requires four additional factors: Smg1, Smg5, Smg6, and Smg7 [10,20,44,45]. Interestingly, there is a correlation between the organism complexity and the dependency on NMD; while Upf1 is essential in Arabidopsis, Drosophila and vertebrates [46,47,48,49], NMD-deficient mutants in yeast and C. elegans are viable [20,30,32,50].
At present, it has become clear that the mRNA quality control represents only one face of the multiple functions of NMD [51,52,53,54,55]. In yeast, almost half of protein-coding genes can generate NMD-sensitive mRNA isoforms, including truncated mRNAs for which transcription initiation occurs downstream of the canonical translation initiation site [56]. NMD also targets intron-containing pre-mRNAs that have escaped splicing and were exported to the cytoplasm [14]. In addition, NMD regulates 3–10% of physiological, non-mutated mRNAs in yeast, Drosophila and humans, including mRNAs with small upstream ORFs [13,14,15,16,57,58,59,60,61], long 3′ UTRs [12,13], as well as mRNAs displaying low translational efficiency and average codon optimality [14]; considered together, NMD provides a significant contribution to the post-transcriptional regulation of gene expression [55].
Numerous physiological processes rely on the capacity of the cell to adjust NMD activity at global and/or transcript specific levels. NMD factors are essential for embryonic development in vertebrates, as disrupted expression of core NMD factors confers lethality at an early embryonic stage [47,62]. NMD is also crucial for the maintenance of hematopoietic stem and progenitor cells [62], the maturation of T cells [62], as well as for liver development, function and regeneration in mice [63]. Furthermore, NMD is important for the response to multiple stresses [64,65,66], being itself regulated in response to stresses such as hypoxia [67] and amino acid deprivation [68]. In fact, many stress-related mRNAs are targeted by NMD under normal physiological conditions but are stabilized upon stress, due to the inhibition of NMD activity [69]. However, as Upf1 is also involved in diverse RNA decay pathways distinct from NMD, including staufen-mediated mRNA decay, replication-dependent histone mRNA decay, glucocorticoid receptor-mediated mRNA decay, regnase 1-mediated mRNA decay, and tudor-staphylococcal/micrococcal-like nuclease-mediated microRNA decay [70], it remains possible that some of the phenotypes associated with mutants of Upf1 do not reflect the loss of NMD per se. Finally, seven NMD factors (Upf1, Upf2, Upf3B, Smg1, Smg5, Smg6, and Smg7) have been found to be NMD targets in mouse and human cells, revealing the existence of a regulatory feedback network between NMD factors, which is critical for the maintenance of physiological NMD activity and RNA homeostasis [71].

3. Molecular Bases of NMD Activation

In many organisms, NMD has been coupled to pre-mRNA splicing [24,72,73,74,75,76,77]. The Exon Junction complex (EJC) is deposited by the spliceosome at the level of the junction between two exons [78], and it is normally removed from the coding regions by the translating ribosomes [79]. The EJC is formed around four core components: the DEAD-box RNA helicase eIF4A3, MLN51, and the Magoh/Y14 heterodimer [80]. The presence of an EJC downstream of a stop codon is recognized as an abnormal situation and enhances the association and activity of Upf1 [81]. In the EJC-enhanced NMD model (Figure 1a), premature translation termination involves the SURF (Smg1–Upf1–eRF1–eRF3) complex, which consists of the Smg1 kinase, Upf1 and the eukaryotic release factors eRF1 and eRF3, and associates with the ribosome stalled at the PTC [82]. Upf2 and Upf3 are then recruited to SURF via the proximal EJC, leading to the formation of the DECID (DECay InDucing) complex [82]. The interaction with Upf2 induces a conformational change in Upf1, allowing its phosphorylation by Smg1 and its activation [82]. The activated Upf1 recruits the Smg6 endonuclease [83] and the Smg5–7 heterodimer [84], which in turn activates RNA deadenylation and decapping. In addition, phosphorylated Upf1 also prevents new translation initiation events by interacting with the translation initiation factor eIF3, inhibiting the formation of a competent translation initiation complex [85]. Finally, protein phosphatase 2 (PP2A) dephosphorylates Upf1, allowing it to return to its unphosphorylated state for another NMD cycle [84].
In addition to the EJC-enhanced NMD, examples of EJC-independent NMD have been described in human cells [86,87], as well as in fission yeast [88], C. elegans [89], Drosophila [90] and plants [72], all of which have orthologs of EJC factors. In contrast, in S. cerevisiae, not only is the proportion of intron-containing genes low (4%) [91], but EJC factors are absent, with the exception of eIF4A3 (Fal1), which acts in pre-rRNA processing in yeast [92].
The EJC-independent NMD targets RNAs with extended 3′ UTR but lacking EJC downstream of the translation termination codon [77,87,93,94,95,96]. Indeed, RNAs where long EJC-free sequences are inserted downstream of a stop codon show reduced levels due to accelerated degradation by NMD [93,95]. This EJC-independent NMD might be a vestige of an ancestral NMD mechanism associated with an abnormally long 3′ UTR, referred to as “faux 3′ UTR”, which is still present in S. cerevisiae [97]. In this model, a compromised interaction between the polyadenylate-binding protein Pab1 and the prematurely terminating ribosome results in less efficient termination and enhanced interaction between Upf1 and eRF1/eRF3, triggering NMD (Figure 1b). In this context, a recent proteomics-based analysis in yeast characterized the composition of two distinct NMD complexes associated with Upf1 named Upf1-23 (Upf1, Upf2, Upf3) and Upf1-decapping [98]. The latter contained the decapping enzyme Dcp2 and its co-factor Dcp1, the decapping activator Ebs3, and two poorly characterized proteins, Nmd4 and Ebs1. The Upf1-23 complex is recruited and assembled on the RNA substrate, and then a complete re-organization leads to the replacement of the Upf2/3 heterodimer by Nmd4, Ebs1, Dcp2 and its co-factors (Figure 1b). Nmd4 and Ebs1 are accessory factors for NMD and could be functional homologues of human Smg6 and Smg5/7, respectively [98,99]. The discovery of these new factors in yeast suggests that NMD mechanisms could be more conserved than previously thought. However, how the switch from the “Upf1-23” to an “Upf1-decapping” complex occurs remains unclear.
The polyadenylate-binding protein 1 (PABPC1 in mammals, Pab1 in yeast) is known to stimulate translation termination efficiency by recruiting the release factors to the ribosome [100]. A long distance between the PTC and Pab1/PABPC1 triggers NMD in all studied species [87,93,95,101,102], while tethering it close to the PTC suppresses the NMD sensitivity of the PTC-containing transcripts in yeast [97] and Drosophila cells [101]. Mechanistically, it has been proposed that the long 3′ UTR would act by impeding the efficient interaction between Pab1/PABPC1 and eRF1/eRF3, favoring the recruitment of Upf1 by the latter and the formation of a SURF complex at the level of the PTC.
Currently, several questions remain open regarding Upf1 recruitment to the target transcripts. Until recently, the classical view was that Upf1 is recruited at the level of the nonsense codon by the stalled ribosome through an interaction with eRF1/eRF3. However, it has been shown that substrate discrimination by NMD can occur independently of Pab1/PABPC1 or its interaction with eRF3 [103,104], indicating that other features contribute to RNA recognition by NMD. In addition, if Upf1 preferentially binds NMD-targeted transcripts [61,105,106,107], with a marked enrichment in the 3′ UTR [81,108,109,110,111], it is redistributed into the coding sequence upon translation inhibition [109,110,111]. This suggests that Upf1 can bind the RNA independently of translation as well as to NMD non-targets and is pushed away from the coding region by the elongating ribosomes (Figure 2). This means that NMD substrate selection occurs after Upf1 association with the RNA. In this regard, NMD substrate discrimination was shown to rely on a faster dissociation of Upf1 from non-target mRNAs, and this depends on its ATPase activity [106,112]. ATP hydrolysis by Upf1 is also required for ribosome release and recycling and efficient RNA degradation [113,114].

4. Long Non-Coding RNAs: An Unexpected Class of NMD Substrates

Unexpectedly, recent transcriptome-wide analyses of RNA binding sites of Upf1 in human and yeast cells revealed that, in addition to mRNAs, Upf1 can also bind lncRNAs [111,115,116].
LncRNAs are a prominent class of transcripts that play important roles in multiple cellular processes, including chromatin modification and regulation of gene expression [117,118,119,120]. They were a priori presumed to be devoid of coding potential [121]. However, this initial assumption has been challenged over recent years by a number of analyses showing that transcripts produced from non-coding regions of the genome, including intergenic regions and sequences antisense to protein-coding genes, associate with the translation machinery in different models, including S. cerevisiae [122,123,124,125,126], fission yeast [127,128], plant [129], Drosophila [129,130], zebrafish [129,131,132], mouse [129,133] and human cells [129,131,134,135,136]. Thus, not only could the ribosome constitute a default destination for cytoplasmic lncRNAs [136], but the smORFs they carry are likely to be translated into micropeptides [129]. Furthermore, the observation that translation elongation inhibitors results in the stabilization of polysomal lncRNAs in human (K562) cells indicates that translation also determines the degradation of cytoplasmic lncRNAs [136].
In budding and fission yeasts, cytoplasmic lncRNAs are extensively degraded by the 5′-3′ exoribonuclease Xrn1/Exo2 [124,137,138,139]. Inactivation of Xrn1 leads to the stabilization of Xrn1-sensititve Unstable Transcripts (XUTs), the majority of which are antisense to protein-coding genes [124,138,139,140]. Strikingly, in S. cerevisiae, 70% of these XUTs are targeted to Xrn1 through NMD [56,124], indicating that most XUTs are translated and that translation constitutes a prerequisite for their degradation. In fact, NMD-sensitive XUTs display ribosome footprints restricted to their 5′ regions, followed by long downstream ribosome-free regions [124]. Conversely, antisense (as)XUTs were found to form dsRNA structures with their paired-sense mRNAs, thus modulating their sensitivity to NMD [124]. This suggests that unless blocked by dsRNA structures, ribosomes could rapidly bind smORFs in the 5′ region of cytoplasmic lncRNAs (Figure 3). The detection of a long 3′ UTR would trigger NMD, leading to the decapping of the transcript and its degradation by Xrn1. Alternatively, but not exclusively, dsRNA could also interfere with the recruitment of NMD factors to asXUTs. Given the current view of Upf1 binding to the RNA (Figure 2) and the observation that Upf1 physically interacts with yeast lncRNAs [115], we propose that Upf1 binds XUTs in a promiscuous manner, independently of translation, regardless of whether or not the transcript will be targeted to Xrn1 through NMD. As proposed for mRNAs, Upf1 would be displaced from the smORF of XUTs by the translating ribosomes and would accumulate on the 3′ UTR. Since NMD-sensitive XUTs are globally longer than NMD-insensitive ones [124], we speculate that longer XUTs carry longer 3′ UTRs, which will be more likely to impede the interaction between eRF1/eRF3 and Pab1. Instead, this situation would favor the Upf1–eRF1/eRF3 interaction, enclosing the XUT as an NMD target through a mechanism similar to the “faux 3′ UTR” model. Supporting this idea, XUTs are poly-adenylated [124,139], and their poly(A) tail is likely to be bound by Pab1.
We speculate that cytoplasmic smORFs-bearing lncRNAs, reminiscent of the yeast NMD-sensitive XUTs, could be targeted by NMD in other eukaryotic cells. Consistent with this idea, NMD inactivation in mouse embryonic stem cells results in the stabilization of a subset of lncRNAs [123]. Moreover, the levels of lncRNAs, including Natural Antisense Transcripts, are also modulated by NMD in Arabidopsis [141]. Further support comes from the observation that the growth arrest-specific 5 (GAS5) lncRNA is targeted by NMD and accumulates in Upf1-depleted human cells [142].

5. Functional Importance of NMD and Translation in lncRNA Metabolism

NMD could be seen as an additional pathway contributing to the clearance of unproductive and potentially harmful spurious non-coding transcripts. However, we believe that this view is too reductive and that there might be more behind the involvement of NMD in the metabolism of lncRNAs (Figure 4). For example, as NMD is a cytoplasmic process, it could ensure that regulatory lncRNAs exhibit their functions exclusively in the nucleus by limiting their accumulation outside the nucleus. In addition, NMD could limit the accumulation of nonfunctional and potentially deleterious peptides from cytoplasmic lncRNAs during the de novo gene birth [125,126,143]. Additionally, the peptides produced from NMD-sensitive lncRNAs could be functional and important, despite their low levels. Even if this remains completely speculative for NMD-sensitive lncRNAs, we note that antigens of the MHC class I pathway are produced from PTC-containing mRNA [144], raising the question of how this process might be generalized for “cryptic” lncRNAs.
NMD could also specifically modulate the levels of regulatory lncRNAs. The apoptotic lncRNA GAS5 has been proposed to act in an NMD-based circuit, which is critical in response to serum starvation [142]. In normal conditions, NMD restricts the constitutive GAS5 expression to low levels. However, in stress conditions associated with NMD inhibition (such as serum starvation), GAS5 expression is up-regulated and binds the glucocorticoid receptor, perturbing its function as a transcription activator in the anti-apoptotic program [142].
More globally, by targeting regulatory cytoplasmic aslncRNAs, NMD could contribute to regulate gene expression. For instance, stabilization of subsets of Xrn1-sensitive aslncRNAs, most of which are NMD-sensitive [124,128], correlates with the transcriptional attenuation of the paired-sense genes, in budding and fission yeasts [137,139]. Interestingly, two independent studies in zebrafish embryos reported that NMD factors cycle to the nucleus to trigger transcriptional adaptation of genes with a sequence complementarity to the PTC-containing RNA in a mechanism called genetic compensation [145,146].
Finally, coupling translation to aslncRNA degradation via NMD could be important for cell recovery upon translation inhibitory stress. In such a condition, NMD-sensitive aslncRNAs are expected to be stabilized and form duplexes with their paired-sense mRNAs. By analogy with the protective effect on the aslncRNA [124], this interaction could also prevent the degradation of the mRNA partner, since local dsRNA formation correlates with higher mRNA stability [147]. After stress, the protected sense/as transcripts would be rapidly released upon the action of RNA helicases, thereby providing a pool of mRNAs in the cytoplasm that can be translated, while NMD-sensitive aslncRNAs would be rapidly degraded.
Together, the observations reported above support the idea that NMD is able to target lncRNAs, and that this might be important for the maintenance of RNA homeostasis, the regulation of gene expression, and for a robust response to several stress conditions. It also challenges the initial assumption that such transcripts are devoid of coding potential.

6. Insight into the Coding Potential of “Non-Coding” Transcripts

The accumulating evidence that cytoplasmic lncRNAs interact with the translation machinery raises the question of their coding potential. Numerous methods have been developed to assess this possibility [148,149,150]. Additionally, a growing body of experimental data indicate that “non-coding” RNAs can indeed be translated [151], including not only lncRNAs but also circular RNAs [152,153,154] and primary microRNAs transcripts [155,156]; moreover, these translation events can produce functional peptides [134,151,157,158,159].
“Non-coding” RNAs contain one or more smORFs that can be translated into micropeptides (i.e., peptides not exceeding 100 amino acids in length.) Previously, such smORFs were ignored as the traditional gene annotation process filtered out ORFs shorter than 100 codons, considering them as noise or false positives. However, as ribosome profiling techniques and proteomics are growing in popularity and increasing in sensitivity, accuracy and efficiency, it is becoming clear that at least a fraction of short ribosome-bound sequences of (l)ncRNAs represent genuine smORFs.
Importantly, a recent work from Weissman’s lab provided a catalog of smORFs and functional peptides derived from human lncRNAs, which included the identification of >800 novel lncRNA-associated smORFs and the observation that, for 91 of them, CRISPR-mediated knockout of the smORF resulted in a growth phenotype [134], indicating that the corresponding peptides are important for cell growth. Other studies previously showed that lncRNA-derived micropeptides are involved in the regulation of RNA decapping [160], in embryonic development [161,162], in muscle development [163,164,165], regeneration [166,167] or contraction [168,169,170], and in tumor development [154,171,172] (see Table 1).
Mechanistically, if global information about the mode of action of lncRNA-derived peptides is still lacking, pioneer studies revealed that they can act by binding other proteins and regulate their activity [162,168,169,176], or as signaling pathway molecules [177]. We anticipate that future works will reveal additional modes of action.
In the light of the observations that micropeptides produced from lncRNAs can be biologically important, it is tempting to speculate that aberrant expression of endogenous lncRNA-derived peptides could be associated with diseases, including cancer [154,171]. In addition to providing a new perspective on pathogenicity, lncRNA-derived peptides could also constitute promising targets for targeted therapy [178], including tumor immunotherapy [179]. In this respect, it is interesting to note that a recent characterization of different murine cell lines and cancer patient samples showed that non-coding regions constitute the major source of tumor specific neo-antigens [179], which could be pivotal for the development of future immunological treatments and cancer vaccines [180].

7. Conclusions

Today, NMD is extending far beyond its original definition assigning it only to the clearance of aberrant “nonsense” transcripts. The current research has revealed that it provides potent means to regulate the expression of many mRNAs and lncRNAs, as well as contributes to the establishment of suitable cellular responses to environmental changes, including adaptation, differentiation or apoptosis. The accumulating biochemical and transcriptomic evidence showing that NMD targets lncRNAs implores us to reconsider the idea that lncRNAs are devoid of coding potential, and challenges us to address how translation of smORFs could not only affect their stability, but also could be used to produce functional micropeptides. Revealing the possibility for a “dark peptidome” to arise from the “dark non-coding side of the genome” (i.e., “the dark side of the dark matter”) is one of the challenges in the RNA field for the coming years and will open exciting perspectives regarding the roles of lncRNA-derived peptides.

Author Contributions

Conceptualization, S.A., A.M. and M.W.; writing—original draft preparation, S.A. and M.W.; writing—review and editing, S.A., A.M. and M.W.; funding acquisition, A.M. All authors have read and agreed to the published version of the manuscript.

Funding

Research in Morillon’s lab has been supported by the “DNA-life” (ANR-15-CE12-0007) grant from the ANR (Agence Nationale de la Recherche) and the “DARK” consolidator grant from the European Research Council (ERC).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We are grateful to all members of the lab for discussions and to our colleague Michael Schertzer for his careful reading of this manuscript and English revisions. Sara Andjus is a recipient of a PhD fellowship from PSL university.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wolin, S.L.; Maquat, L.E. Cellular RNA surveillance in health and disease. Science 2019, 366, 822–827. [Google Scholar] [CrossRef] [PubMed]
  2. Parker, R. RNA degradation in Saccharomyces cerevisae. Genetics 2012, 191, 671–702. [Google Scholar] [CrossRef] [Green Version]
  3. Isken, O.; Maquat, L.E. Quality control of eukaryotic mRNA: Safeguarding cells from abnormal mRNA function. Genes Dev. 2007, 21, 1833–1856. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Frischmeyer, P.A.; van Hoof, A.; O’Donnell, K.; Guerrerio, A.L.; Parker, R.; Dietz, H.C. An mRNA surveillance mechanism that eliminates transcripts lacking termination codons. Science 2002, 295, 2258–2261. [Google Scholar] [CrossRef] [Green Version]
  5. Van Hoof, A.; Frischmeyer, P.A.; Dietz, H.C.; Parker, R. Exosome-mediated recognition and degradation of mRNAs lacking a termination codon. Science 2002, 295, 2262–2264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Klauer, A.A.; van Hoof, A. Degradation of mRNAs that lack a stop codon: A decade of nonstop progress. Wiley Interdiscip. Rev. RNA 2012, 3, 649–660. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Doma, M.K.; Parker, R. Endonucleolytic cleavage of eukaryotic mRNAs with stalls in translation elongation. Nature 2006, 440, 561–564. [Google Scholar] [CrossRef] [Green Version]
  8. Harigaya, Y.; Parker, R. No-go decay: A quality control mechanism for RNA in translation. Wiley Interdiscip. Rev. RNA 2010, 1, 132–141. [Google Scholar] [CrossRef] [PubMed]
  9. Chang, Y.F.; Imam, J.S.; Wilkinson, M.F. The nonsense-mediated decay RNA surveillance pathway. Annu. Rev. Biochem. 2007, 76, 51–74. [Google Scholar] [CrossRef] [Green Version]
  10. Behm-Ansmant, I.; Kashima, I.; Rehwinkel, J.; Sauliere, J.; Wittkopp, N.; Izaurralde, E. mRNA quality control: An ancient machinery recognizes and degrades mRNAs with nonsense codons. FEBS Lett. 2007, 581, 2845–2853. [Google Scholar] [CrossRef] [Green Version]
  11. Muhlrad, D.; Parker, R. Premature translational termination triggers mRNA decapping. Nature 1994, 370, 578–581. [Google Scholar] [CrossRef]
  12. 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] [PubMed] [Green Version]
  13. Yepiskoposyan, H.; Aeschimann, F.; Nilsson, D.; Okoniewski, M.; Muhlemann, O. Autoregulation of the nonsense-mediated mRNA decay pathway in human cells. RNA 2011, 17, 2108–2118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Celik, A.; Baker, R.; He, F.; Jacobson, A. High-resolution profiling of NMD targets in yeast reveals translational fidelity as a basis for substrate selection. RNA 2017, 23, 735–748. [Google Scholar] [CrossRef] [Green Version]
  15. Oliveira, C.C.; McCarthy, J.E. The relationship between eukaryotic translation and mRNA stability. A short upstream open reading frame strongly inhibits translational initiation and greatly accelerates mRNA degradation in the yeast Saccharomyces cerevisiae. J. Biol. Chem. 1995, 270, 8936–8943. [Google Scholar] [CrossRef] [Green Version]
  16. Ruiz-Echevarria, 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] [Green Version]
  17. Nicholson, P.; Muhlemann, O. Cutting the nonsense: The degradation of PTC-containing mRNAs. BioChem. Soc. Trans. 2010, 38, 1615–1620. [Google Scholar] [CrossRef] [Green Version]
  18. Rebbapragada, I.; Lykke-Andersen, J. Execution of nonsense-mediated mRNA decay: What defines a substrate? Curr. Opin. Cell Biol. 2009, 21, 394–402. [Google Scholar] [CrossRef]
  19. Bhuvanagiri, M.; Schlitter, A.M.; Hentze, M.W.; Kulozik, A.E. NMD: RNA biology meets human genetic medicine. BioChem. J. 2010, 430, 365–377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Pulak, R.; Anderson, P. mRNA surveillance by the Caenorhabditis elegans smg genes. Genes Dev. 1993, 7, 1885–1897. [Google Scholar] [CrossRef] [Green Version]
  21. Hall, G.W.; Thein, S. Nonsense codon mutations in the terminal exon of the beta-globin gene are not associated with a reduction in beta-mRNA accumulation: A mechanism for the phenotype of dominant beta-thalassemia. Blood 1994, 83, 2031–2037. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Zhang, S.; Welch, E.M.; Hogan, K.; Brown, A.H.; Peltz, S.W.; Jacobson, A. Polysome-associated mRNAs are substrates for the nonsense-mediated mRNA decay pathway in Saccharomyces cerevisiae. RNA 1997, 3, 234–244. [Google Scholar]
  23. Carter, M.S.; Doskow, J.; Morris, P.; Li, S.; Nhim, R.P.; Sandstedt, S.; Wilkinson, M.F. A regulatory mechanism that detects premature nonsense codons in T-cell receptor transcripts in vivo is reversed by protein synthesis inhibitors in vitro. J. Biol. Chem. 1995, 270, 28995–29003. [Google Scholar] [CrossRef] [Green Version]
  24. Thermann, R.; Neu-Yilik, G.; Deters, A.; Frede, U.; Wehr, K.; Hagemeier, C.; Hentze, M.W.; Kulozik, A.E. Binary specification of nonsense codons by splicing and cytoplasmic translation. EMBO J. 1998, 17, 3484–3494. [Google Scholar] [CrossRef] [PubMed]
  25. Causier, B.; Li, Z.; De Smet, R.; Lloyd, J.P.B.; Van de Peer, Y.; Davies, B. Conservation of Nonsense-Mediated mRNA Decay Complex Components Throughout Eukaryotic Evolution. Sci. Rep. 2017, 7, 16692. [Google Scholar] [CrossRef]
  26. Losson, R.; Lacroute, F. Interference of nonsense mutations with eukaryotic messenger RNA stability. Proc. Natl. Acad. Sci. USA 1979, 76, 5134–5137. [Google Scholar] [CrossRef] [Green Version]
  27. Maquat, L.E.; Kinniburgh, A.J.; Rachmilewitz, E.A.; Ross, J. Unstable beta-globin mRNA in mRNA-deficient beta o thalassemia. Cell 1981, 27, 543–553. [Google Scholar] [CrossRef]
  28. Baserga, S.J.; Benz, E.J., Jr. Nonsense mutations in the human beta-globin gene affect mRNA metabolism. Proc. Natl. Acad. Sci. USA 1988, 85, 2056–2060. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Culbertson, M.R.; Underbrink, K.M.; Fink, G.R. Frameshift suppression Saccharomyces cerevisiae. II. Genetic properties of group II suppressors. Genetics 1980, 95, 833–853. [Google Scholar] [CrossRef]
  30. Leeds, P.; Peltz, S.W.; Jacobson, A.; Culbertson, M.R. The product of the yeast UPF1 gene is required for rapid turnover of mRNAs containing a premature translational termination codon. Genes Dev. 1991, 5, 2303–2314. [Google Scholar] [CrossRef]
  31. Leeds, P.; Wood, J.M.; Lee, B.S.; Culbertson, M.R. Gene products that promote mRNA turnover in Saccharomyces cerevisiae. Mol. Cell Biol. 1992, 12, 2165–2177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Cui, Y.; Hagan, K.W.; Zhang, S.; Peltz, S.W. Identification and characterization of genes that are required for the accelerated degradation of mRNAs containing a premature translational termination codon. Genes Dev. 1995, 9, 423–436. [Google Scholar] [CrossRef] [Green Version]
  33. 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] [Green Version]
  34. Weng, Y.; Czaplinski, K.; Peltz, S.W. Genetic and biochemical characterization of mutations in the ATPase and helicase regions of the Upf1 protein. Mol. Cell Biol. 1996, 16, 5477–5490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. 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] [Green Version]
  36. 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]
  37. 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] [Green Version]
  38. Melero, R.; Buchwald, G.; Castano, R.; Raabe, M.; Gil, D.; Lazaro, 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, S491–S492. [Google Scholar] [CrossRef] [Green Version]
  39. 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] [Green Version]
  40. 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] [PubMed] [Green Version]
  41. 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] [PubMed]
  42. Culbertson, M.R.; Leeds, P.F. Looking at mRNA decay pathways through the window of molecular evolution. Curr. Opin. Genet. Dev. 2003, 13, 207–214. [Google Scholar] [CrossRef]
  43. 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] [Green Version]
  44. Yamashita, A.; Ohnishi, T.; Kashima, I.; Taya, Y.; Ohno, S. Human SMG-1, a novel phosphatidylinositol 3-kinase-related protein kinase, associates with components of the mRNA surveillance complex and is involved in the regulation of nonsense-mediated mRNA decay. Genes Dev. 2001, 15, 2215–2228. [Google Scholar] [CrossRef] [Green Version]
  45. Cali, B.M.; Kuchma, S.L.; Latham, J.; Anderson, P. smg-7 is required for mRNA surveillance in Caenorhabditis elegans. Genetics 1999, 151, 605–616. [Google Scholar] [CrossRef]
  46. Wittkopp, N.; Huntzinger, E.; Weiler, C.; Sauliere, J.; Schmidt, S.; Sonawane, M.; Izaurralde, E. Nonsense-mediated mRNA decay effectors are essential for zebrafish embryonic development and survival. Mol. Cell Biol. 2009, 29, 3517–3528. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. 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]
  48. Metzstein, M.M.; Krasnow, M.A. Functions of the nonsense-mediated mRNA decay pathway in Drosophila development. PLoS Genet. 2006, 2, e180. [Google Scholar] [CrossRef] [Green Version]
  49. Yoine, M.; Nishii, T.; Nakamura, K. Arabidopsis UPF1 RNA helicase for nonsense-mediated mRNA decay is involved in seed size control and is essential for growth. Plant. Cell Physiol. 2006, 47, 572–580. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Hodgkin, J.; Papp, A.; Pulak, R.; Ambros, V.; Anderson, P. A new kind of informational suppression in the nematode Caenorhabditis elegans. Genetics 1989, 123, 301–313. [Google Scholar] [CrossRef]
  51. Nasif, S.; Contu, L.; Muhlemann, 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] [PubMed]
  52. 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] [Green Version]
  53. 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] [PubMed]
  54. Peccarelli, M.; Kebaara, B.W. Regulation of natural mRNAs by the nonsense-mediated mRNA decay pathway. Eukaryot. Cell 2014, 13, 1126–1135. [Google Scholar] [CrossRef]
  55. Smith, J.E.; Baker, K.E. Nonsense-mediated RNA decay—A switch and dial for regulating gene expression. Bioessays 2015, 37, 612–623. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Malabat, C.; Feuerbach, F.; Ma, L.; Saveanu, C.; Jacquier, A. Quality control of transcription start site selection by nonsense-mediated-mRNA decay. eLife 2015, 4, e06722. [Google Scholar] [CrossRef]
  57. Gaba, A.; Jacobson, A.; Sachs, M.S. Ribosome occupancy of the yeast CPA1 upstream open reading frame termination codon modulates nonsense-mediated mRNA decay. Mol. Cell 2005, 20, 449–460. [Google Scholar] [CrossRef]
  58. Arribere, J.A.; Gilbert, W.V. Roles for transcript leaders in translation and mRNA decay revealed by transcript leader sequencing. Genome Res. 2013, 23, 977–987. [Google Scholar] [CrossRef] [Green Version]
  59. Guan, Q.; Zheng, W.; Tang, S.; Liu, X.; Zinkel, R.A.; Tsui, K.W.; Yandell, B.S.; Culbertson, M.R. Impact of nonsense-mediated mRNA decay on the global expression profile of budding yeast. PLoS Genet. 2006, 2, e203. [Google Scholar] [CrossRef] [PubMed]
  60. He, F.; Li, X.; Spatrick, P.; Casillo, R.; Dong, S.; Jacobson, A. Genome-wide analysis of mRNAs regulated by the nonsense-mediated and 5’ to 3’ mRNA decay pathways in yeast. Mol. Cell 2003, 12, 1439–1452. [Google Scholar] [CrossRef]
  61. Johansson, M.J.; He, F.; Spatrick, P.; Li, C.; Jacobson, A. Association of yeast Upf1p with direct substrates of the NMD pathway. Proc. Natl. Acad. Sci. USA 2007, 104, 20872–20877. [Google Scholar] [CrossRef] [Green Version]
  62. Weischenfeldt, J.; Damgaard, I.; Bryder, D.; Theilgaard-Monch, K.; Thoren, L.A.; Nielsen, F.C.; Jacobsen, S.E.; Nerlov, C.; Porse, B.T. NMD is essential for hematopoietic stem and progenitor cells and for eliminating by-products of programmed DNA rearrangements. Genes Dev. 2008, 22, 1381–1396. [Google Scholar] [CrossRef] [Green Version]
  63. Thoren, L.A.; Norgaard, G.A.; Weischenfeldt, J.; Waage, J.; Jakobsen, J.S.; Damgaard, I.; Bergstrom, F.C.; Blom, A.M.; Borup, R.; Bisgaard, H.C.; et al. UPF2 is a critical regulator of liver development, function and regeneration. PLoS ONE 2010, 5, e11650. [Google Scholar] [CrossRef]
  64. Goetz, A.E.; Wilkinson, M. Stress and the nonsense-mediated RNA decay pathway. Cell Mol. Life Sci. 2017, 74, 3509–3531. [Google Scholar] [CrossRef]
  65. 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]
  66. Rodriguez-Gabriel, M.A.; Watt, S.; Bahler, J.; Russell, P. Upf1, an RNA helicase required for nonsense-mediated mRNA decay, modulates the transcriptional response to oxidative stress in fission yeast. Mol. Cell Biol. 2006, 26, 6347–6356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Gardner, L.B. Hypoxic inhibition of nonsense-mediated RNA decay regulates gene expression and the integrated stress response. Mol. Cell Biol. 2008, 28, 3729–3741. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Mendell, J.T.; Sharifi, N.A.; Meyers, J.L.; Martinez-Murillo, F.; Dietz, H.C. Nonsense surveillance regulates expression of diverse classes of mammalian transcripts and mutes genomic noise. Nat. Genet. 2004, 36, 1073–1078. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Usuki, F.; Yamashita, A.; Fujimura, M. Environmental stresses suppress nonsense-mediated mRNA decay (NMD) and affect cells by stabilizing NMD-targeted gene expression. Sci. Rep. 2019, 9, 1279. [Google Scholar] [CrossRef] [Green Version]
  70. 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] [PubMed] [Green Version]
  71. Huang, L.; Lou, C.H.; Chan, W.; Shum, E.Y.; Shao, A.; Stone, E.; Karam, R.; Song, H.W.; Wilkinson, M.F. RNA homeostasis governed by cell type-specific and branched feedback loops acting on NMD. Mol. Cell 2011, 43, 950–961. [Google Scholar] [CrossRef] [Green Version]
  72. Kerenyi, Z.; Merai, Z.; Hiripi, L.; Benkovics, A.; Gyula, P.; Lacomme, C.; Barta, E.; Nagy, F.; Silhavy, D. Inter-kingdom conservation of mechanism of nonsense-mediated mRNA decay. EMBO J. 2008, 27, 1585–1595. [Google Scholar] [CrossRef] [Green Version]
  73. Woodward, L.A.; Mabin, J.W.; Gangras, P.; Singh, G. The exon junction complex: A lifelong guardian of mRNA fate. Wiley Interdiscip. Rev. RNA 2017, 8. [Google Scholar] [CrossRef]
  74. 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] [PubMed] [Green Version]
  75. Sun, X.; Moriarty, P.M.; Maquat, L.E. Nonsense-mediated decay of glutathione peroxidase 1 mRNA in the cytoplasm depends on intron position. EMBO J. 2000, 19, 4734–4744. [Google Scholar] [CrossRef] [Green Version]
  76. Sun, X.; Maquat, L.E. mRNA surveillance in mammalian cells: The relationship between introns and translation termination. RNA 2000, 6, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Zhang, J.; Sun, X.; Qian, Y.; Maquat, L.E. Intron function in the nonsense-mediated decay of beta-globin mRNA: Indications that pre-mRNA splicing in the nucleus can influence mRNA translation in the cytoplasm. RNA 1998, 4, 801–815. [Google Scholar] [CrossRef]
  78. 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] [Green Version]
  79. Dostie, J.; Dreyfuss, G. Translation is required to remove Y14 from mRNAs in the cytoplasm. Curr. Biol. 2002, 12, 1060–1067. [Google Scholar] [CrossRef] [Green Version]
  80. Tange, T.O.; Nott, A.; Moore, M.J. The ever-increasing complexities of the exon junction complex. Curr. Opin. Cell Biol. 2004, 16, 279–284. [Google Scholar] [CrossRef] [PubMed]
  81. Kurosaki, T.; Maquat, L.E. Rules that govern UPF1 binding to mRNA 3′ UTRs. Proc. Natl. Acad. Sci. USA 2013, 110, 3357–3362. [Google Scholar] [CrossRef] [Green Version]
  82. 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] [Green Version]
  83. Eberle, A.B.; Lykke-Andersen, S.; Muhlemann, 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]
  84. 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]
  85. 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] [PubMed] [Green Version]
  86. Metze, S.; Herzog, V.A.; Ruepp, M.D.; Muhlemann, O. Comparison of EJC-enhanced and EJC-independent NMD in human cells reveals two partially redundant degradation pathways. RNA 2013, 19, 1432–1448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Buhler, M.; Steiner, S.; Mohn, F.; Paillusson, A.; Muhlemann, O. EJC-independent degradation of nonsense immunoglobulin-mu mRNA depends on 3’ UTR length. Nat. Struct. Mol. Biol. 2006, 13, 462–464. [Google Scholar] [CrossRef] [PubMed]
  88. Wen, J.; Brogna, S. Splicing-dependent NMD does not require the EJC in Schizosaccharomyces pombe. EMBO J. 2010, 29, 1537–1551. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Longman, D.; Plasterk, R.H.; Johnstone, I.L.; Caceres, J.F. Mechanistic insights and identification of two novel factors in the C. elegans NMD pathway. Genes Dev. 2007, 21, 1075–1085. [Google Scholar] [CrossRef] [Green Version]
  90. Gatfield, D.; Unterholzner, L.; Ciccarelli, F.D.; Bork, P.; Izaurralde, E. Nonsense-mediated mRNA decay in Drosophila: At the intersection of the yeast and mammalian pathways. EMBO J. 2003, 22, 3960–3970. [Google Scholar] [CrossRef] [PubMed]
  91. Goffeau, A.; Barrell, B.G.; Bussey, H.; Davis, R.W.; Dujon, B.; Feldmann, H.; Galibert, F.; Hoheisel, J.D.; Jacq, C.; Johnston, M.; et al. Life with 6000 genes. Science 1996, 274, 546–567. [Google Scholar] [CrossRef] [Green Version]
  92. Alexandrov, A.; Colognori, D.; Steitz, J.A. Human eIF4AIII interacts with an eIF4G-like partner, NOM1, revealing an evolutionarily conserved function outside the exon junction complex. Genes Dev. 2011, 25, 1078–1090. [Google Scholar] [CrossRef] [Green Version]
  93. Eberle, A.B.; Stalder, L.; Mathys, H.; Orozco, R.Z.; Muhlemann, O. Posttranscriptional gene regulation by spatial rearrangement of the 3’ untranslated region. PLoS Biol. 2008, 6, e92. [Google Scholar] [CrossRef] [Green Version]
  94. Matsuda, D.; Hosoda, N.; Kim, Y.K.; Maquat, L.E. Failsafe nonsense-mediated mRNA decay does not detectably target eIF4E-bound mRNA. Nat. Struct. Mol. Biol. 2007, 14, 974–979. [Google Scholar] [CrossRef] [PubMed]
  95. Singh, G.; Rebbapragada, I.; Lykke-Andersen, J. A competition between stimulators and antagonists of Upf complex recruitment governs human nonsense-mediated mRNA decay. PLoS Biol. 2008, 6, e111. [Google Scholar] [CrossRef]
  96. Wang, J.; Gudikote, J.P.; Olivas, O.R.; Wilkinson, M.F. Boundary-independent polar nonsense-mediated decay. EMBO Rep. 2002, 3, 274–279. [Google Scholar] [CrossRef] [Green Version]
  97. 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] [PubMed]
  98. Dehecq, M.; Decourty, L.; Namane, A.; Proux, C.; Kanaan, J.; Le Hir, H.; Jacquier, A.; Saveanu, C. Nonsense-mediated mRNA decay involves two distinct Upf1-bound complexes. EMBO J. 2018, 37. [Google Scholar] [CrossRef] [PubMed]
  99. Luke, B.; Azzalin, C.M.; Hug, N.; Deplazes, A.; Peter, M.; Lingner, J. Saccharomyces cerevisiae Ebs1p is a putative ortholog of human Smg7 and promotes nonsense-mediated mRNA decay. Nucleic Acids Res. 2007, 35, 7688–7697. [Google Scholar] [CrossRef] [Green Version]
  100. 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]
  101. Behm-Ansmant, I.; Gatfield, D.; Rehwinkel, J.; Hilgers, V.; Izaurralde, E. A conserved role for cytoplasmic poly(A)-binding protein 1 (PABPC1) in nonsense-mediated mRNA decay. EMBO J. 2007, 26, 1591–1601. [Google Scholar] [CrossRef] [Green Version]
  102. Silva, A.L.; Ribeiro, P.; Inacio, A.; Liebhaber, S.A.; Romao, L. Proximity of the poly(A)-binding protein to a premature termination codon inhibits mammalian nonsense-mediated mRNA decay. RNA 2008, 14, 563–576. [Google Scholar] [CrossRef] [Green Version]
  103. Meaux, S.; van Hoof, A.; Baker, K.E. Nonsense-mediated mRNA decay in yeast does not require PAB1 or a poly(A) tail. Mol. Cell 2008, 29, 134–140. [Google Scholar] [CrossRef] [Green Version]
  104. Roque, S.; Cerciat, M.; Gaugue, I.; Mora, L.; Floch, A.G.; de Zamaroczy, M.; Heurgue-Hamard, V.; Kervestin, S. Interaction between the poly(A)-binding protein Pab1 and the eukaryotic release factor eRF3 regulates translation termination but not mRNA decay in Saccharomyces cerevisiae. RNA 2015, 21, 124–134. [Google Scholar] [CrossRef] [Green Version]
  105. Johns, L.; Grimson, A.; Kuchma, S.L.; Newman, C.L.; Anderson, P. Caenorhabditis elegans SMG-2 selectively marks mRNAs containing premature translation termination codons. Mol. Cell Biol. 2007, 27, 5630–5638. [Google Scholar] [CrossRef] [Green Version]
  106. 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] [PubMed] [Green Version]
  107. Hwang, J.; Sato, H.; Tang, Y.; Matsuda, D.; Maquat, L.E. UPF1 association with the cap-binding protein, CBP80, promotes nonsense-mediated mRNA decay at two distinct steps. Mol. Cell 2010, 39, 396–409. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Hogg, J.R.; Goff, S.P. Upf1 senses 3’UTR length to potentiate mRNA decay. Cell 2010, 143, 379–389. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Gregersen, L.H.; Schueler, M.; Munschauer, M.; Mastrobuoni, G.; Chen, W.; Kempa, S.; Dieterich, C.; Landthaler, M. MOV10 Is a 5’ to 3’ RNA helicase contributing to UPF1 mRNA target degradation by translocation along 3′ UTRs. Mol. Cell 2014, 54, 573–585. [Google Scholar] [CrossRef] [Green Version]
  110. Hurt, J.A.; Robertson, A.D.; Burge, C.B. Global analyses of UPF1 binding and function reveal expanded scope of nonsense-mediated mRNA decay. Genome Res. 2013, 23, 1636–1650. [Google Scholar] [CrossRef] [Green Version]
  111. Zund, D.; Gruber, A.R.; Zavolan, M.; Muhlemann, O. Translation-dependent displacement of UPF1 from coding sequences causes its enrichment in 3′ UTRs. Nat. Struct. Mol. Biol. 2013, 20, 936–943. [Google Scholar] [CrossRef]
  112. 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] [Green Version]
  113. 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]
  114. 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. [Google Scholar] [CrossRef]
  115. Sohrabi-Jahromi, S.; Hofmann, K.B.; Boltendahl, A.; Roth, C.; Gressel, S.; Baejen, C.; Soeding, J.; Cramer, P. Transcriptome maps of general eukaryotic RNA degradation factors. eLife 2019, 8. [Google Scholar] [CrossRef]
  116. Colombo, M.; Karousis, E.D.; Bourquin, J.; Bruggmann, R.; Muhlemann, O. Transcriptome-wide identification of NMD-targeted human mRNAs reveals extensive redundancy between SMG6- and SMG7-mediated degradation pathways. RNA 2017, 23, 189–201. [Google Scholar] [CrossRef] [Green Version]
  117. Jarroux, J.; Morillon, A.; Pinskaya, M. History, Discovery, and Classification of lncRNAs. Adv. Exp. Med. Biol. 2017, 1008, 1–46. [Google Scholar] [CrossRef]
  118. Yao, R.W.; Wang, Y.; Chen, L.L. Cellular functions of long noncoding RNAs. Nat. Cell Biol. 2019, 21, 542–551. [Google Scholar] [CrossRef] [PubMed]
  119. Statello, L.; Guo, C.J.; Chen, L.L.; Huarte, M. Gene regulation by long non-coding RNAs and its biological functions. Nat. Rev. Mol. Cell Biol. 2021, 22, 96–118. [Google Scholar] [CrossRef] [PubMed]
  120. Wery, M.; Kwapisz, M.; Morillon, A. Noncoding RNAs in gene regulation. Wiley Interdiscip. Rev. Syst. Biol. Med. 2011, 3, 728–738. [Google Scholar] [CrossRef] [PubMed]
  121. Verheggen, K.; Volders, P.J.; Mestdagh, P.; Menschaert, G.; Van Damme, P.; Gevaert, K.; Martens, L.; Vandesompele, J. Noncoding after All: Biases in Proteomics Data Do Not Explain Observed Absence of lncRNA Translation Products. J. Proteome Res. 2017, 16, 2508–2515. [Google Scholar] [CrossRef]
  122. Brar, G.A.; Yassour, M.; Friedman, N.; Regev, A.; Ingolia, N.T.; Weissman, J.S. High-resolution view of the yeast meiotic program revealed by ribosome profiling. Science 2012, 335, 552–557. [Google Scholar] [CrossRef] [Green Version]
  123. Smith, J.E.; Alvarez-Dominguez, J.R.; Kline, N.; Huynh, N.J.; Geisler, S.; Hu, W.; Coller, J.; Baker, K.E. Translation of small open reading frames within unannotated RNA transcripts in Saccharomyces cerevisiae. Cell Rep. 2014, 7, 1858–1866. [Google Scholar] [CrossRef] [Green Version]
  124. Wery, M.; Descrimes, M.; Vogt, N.; Dallongeville, A.S.; Gautheret, D.; Morillon, A. Nonsense-Mediated Decay Restricts LncRNA Levels in Yeast Unless Blocked by Double-Stranded RNA Structure. Mol. Cell 2016, 61, 379–392. [Google Scholar] [CrossRef] [Green Version]
  125. Carvunis, A.R.; Rolland, T.; Wapinski, I.; Calderwood, M.A.; Yildirim, M.A.; Simonis, N.; Charloteaux, B.; Hidalgo, C.A.; Barbette, J.; Santhanam, B.; et al. Proto-genes and de novo gene birth. Nature 2012, 487, 370–374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Wilson, B.A.; Masel, J. Putatively noncoding transcripts show extensive association with ribosomes. Genome Biol. Evol. 2011, 3, 1245–1252. [Google Scholar] [CrossRef]
  127. Duncan, C.D.; Mata, J. The translational landscape of fission-yeast meiosis and sporulation. Nat. Struct. Mol. Biol. 2014, 21, 641–647. [Google Scholar] [CrossRef] [Green Version]
  128. Atkinson, S.R.; Marguerat, S.; Bitton, D.A.; Rodriguez-Lopez, M.; Rallis, C.; Lemay, J.F.; Cotobal, C.; Malecki, M.; Smialowski, P.; Mata, J.; et al. Long noncoding RNA repertoire and targeting by nuclear exosome, cytoplasmic exonuclease, and RNAi in fission yeast. RNA 2018, 24, 1195–1213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Ruiz-Orera, J.; Messeguer, X.; Subirana, J.A.; Alba, M.M. Long non-coding RNAs as a source of new peptides. eLife 2014, 3, e03523. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Aspden, J.L.; Eyre-Walker, Y.C.; Phillips, R.J.; Amin, U.; Mumtaz, M.A.; Brocard, M.; Couso, J.P. Extensive translation of small Open Reading Frames revealed by Poly-Ribo-Seq. eLife 2014, 3, e03528. [Google Scholar] [CrossRef] [PubMed]
  131. Bazzini, A.A.; Johnstone, T.G.; Christiano, R.; Mackowiak, S.D.; Obermayer, B.; Fleming, E.S.; Vejnar, C.E.; Lee, M.T.; Rajewsky, N.; Walther, T.C.; et al. Identification of small ORFs in vertebrates using ribosome footprinting and evolutionary conservation. EMBO J. 2014, 33, 981–993. [Google Scholar] [CrossRef] [Green Version]
  132. Chew, G.L.; Pauli, A.; Rinn, J.L.; Regev, A.; Schier, A.F.; Valen, E. Ribosome profiling reveals resemblance between long non-coding RNAs and 5’ leaders of coding RNAs. Development 2013, 140, 2828–2834. [Google Scholar] [CrossRef] [Green Version]
  133. Ingolia, N.T.; Lareau, L.F.; Weissman, J.S. Ribosome profiling of mouse embryonic stem cells reveals the complexity and dynamics of mammalian proteomes. Cell 2011, 147, 789–802. [Google Scholar] [CrossRef] [Green Version]
  134. Chen, J.; Brunner, A.D.; Cogan, J.Z.; Nunez, J.K.; Fields, A.P.; Adamson, B.; Itzhak, D.N.; Li, J.Y.; Mann, M.; Leonetti, M.D.; et al. Pervasive functional translation of noncanonical human open reading frames. Science 2020, 367, 1140–1146. [Google Scholar] [CrossRef]
  135. Van Heesch, S.; van Iterson, M.; Jacobi, J.; Boymans, S.; Essers, P.B.; de Bruijn, E.; Hao, W.; MacInnes, A.W.; Cuppen, E.; Simonis, M. Extensive localization of long noncoding RNAs to the cytosol and mono- and polyribosomal complexes. Genome Biol. 2014, 15, R6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Carlevaro-Fita, J.; Rahim, A.; Guigo, R.; Vardy, L.A.; Johnson, R. Cytoplasmic long noncoding RNAs are frequently bound to and degraded at ribosomes in human cells. RNA 2016, 22, 867–882. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Wery, M.; Gautier, C.; Descrimes, M.; Yoda, M.; Migeot, V.; Hermand, D.; Morillon, A. Bases of antisense lncRNA-associated regulation of gene expression in fission yeast. PLoS Genet. 2018, 14, e1007465. [Google Scholar] [CrossRef] [PubMed]
  138. Szachnowski, U.; Andjus, S.; Foretek, D.; Morillon, A.; Wery, M. Endogenous RNAi pathway evolutionarily shapes the destiny of the antisense lncRNAs transcriptome. Life Sci. Alliance 2019, 2, e201900407. [Google Scholar] [CrossRef] [Green Version]
  139. Van Dijk, E.L.; Chen, C.L.; d’Aubenton-Carafa, Y.; Gourvennec, S.; Kwapisz, M.; Roche, V.; Bertrand, C.; Silvain, M.; Legoix-Né, P.; Loeillet, S.; et al. XUTs are a class of Xrn1-sensitive antisense regulatory non coding RNA in yeast. Nature 2011, 475, 114–117. [Google Scholar] [CrossRef]
  140. Wery, M.; Gautier, C.; Descrimes, M.; Yoda, M.; Vennin-Rendos, H.; Migeot, V.; Gautheret, D.; Hermand, D.; Morillon, A. Native elongating transcript sequencing reveals global anti-correlation between sense and antisense nascent transcription in fission yeast. RNA 2018, 24, 196–208. [Google Scholar] [CrossRef] [Green Version]
  141. Kurihara, Y.; Matsui, A.; Hanada, K.; Kawashima, M.; Ishida, J.; Morosawa, T.; Tanaka, M.; Kaminuma, E.; Mochizuki, Y.; Matsushima, A.; et al. Genome-wide suppression of aberrant mRNA-like noncoding RNAs by NMD in Arabidopsis. Proc. Natl. Acad. Sci. USA 2009, 106, 2453–2458. [Google Scholar] [CrossRef] [Green Version]
  142. Tani, H.; Torimura, M.; Akimitsu, N. The RNA degradation pathway regulates the function of GAS5 a non-coding RNA in mammalian cells. PLoS ONE 2013, 8, e55684. [Google Scholar] [CrossRef]
  143. Blevins, W.R.; Ruiz-Orera, J.; Messeguer, X.; Blasco-Moreno, B.; Villanueva-Canas, J.L.; Espinar, L.; Diez, J.; Carey, L.B.; Alba, M.M. Uncovering de novo gene birth in yeast using deep transcriptomics. Nat. Commun. 2021, 12, 604. [Google Scholar] [CrossRef] [PubMed]
  144. Apcher, S.; Daskalogianni, C.; Lejeune, F.; Manoury, B.; Imhoos, G.; Heslop, L.; Fahraeus, R. Major source of antigenic peptides for the MHC class I pathway is produced during the pioneer round of mRNA translation. Proc. Natl. Acad. Sci. USA 2011, 108, 11572–11577. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Ma, Z.; Zhu, P.; Shi, H.; Guo, L.; Zhang, Q.; Chen, Y.; Chen, S.; Zhang, Z.; Peng, J.; Chen, J. PTC-bearing mRNA elicits a genetic compensation response via Upf3a and COMPASS components. Nature 2019, 568, 259–263. [Google Scholar] [CrossRef]
  146. El-Brolosy, M.A.; Kontarakis, Z.; Rossi, A.; Kuenne, C.; Gunther, S.; Fukuda, N.; Kikhi, K.; Boezio, G.L.M.; Takacs, C.M.; Lai, S.L.; et al. Genetic compensation triggered by mutant mRNA degradation. Nature 2019, 568, 193–197. [Google Scholar] [CrossRef]
  147. Geisberg, J.V.; Moqtaderi, Z.; Fan, X.; Ozsolak, F.; Struhl, K. Global analysis of mRNA isoform half-lives reveals stabilizing and destabilizing elements in yeast. Cell 2014, 156, 812–824. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Mishra, S.K.; Wang, H. Computational Analysis Predicts Hundreds of Coding lncRNAs in Zebrafish. Biology 2021, 10, 371. [Google Scholar] [CrossRef]
  149. Makarewich, C.A.; Olson, E.N. Mining for Micropeptides. Trends Cell Biol. 2017, 27, 685–696. [Google Scholar] [CrossRef]
  150. Choi, S.W.; Kim, H.W.; Nam, J.W. The small peptide world in long noncoding RNAs. Brief. Bioinform. 2019, 20, 1853–1864. [Google Scholar] [CrossRef] [Green Version]
  151. Wei, L.H.; Guo, J.U. Coding functions of “noncoding” RNAs. Science 2020, 367, 1074–1075. [Google Scholar] [CrossRef]
  152. Legnini, I.; Di Timoteo, G.; Rossi, F.; Morlando, M.; Briganti, F.; Sthandier, O.; Fatica, A.; Santini, T.; Andronache, A.; Wade, M.; et al. Circ-ZNF609 Is a Circular RNA that Can Be Translated and Functions in Myogenesis. Mol. Cell 2017, 66, 22–37.e9. [Google Scholar] [CrossRef] [Green Version]
  153. Yang, Y.; Gao, X.; Zhang, M.; Yan, S.; Sun, C.; Xiao, F.; Huang, N.; Yang, X.; Zhao, K.; Zhou, H.; et al. Novel Role of FBXW7 Circular RNA in Repressing Glioma Tumorigenesis. J. Natl. Cancer Inst. 2018, 110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Zhang, M.; Zhao, K.; Xu, X.; Yang, Y.; Yan, S.; Wei, P.; Liu, H.; Xu, J.; Xiao, F.; Zhou, H.; et al. A peptide encoded by circular form of LINC-PINT suppresses oncogenic transcriptional elongation in glioblastoma. Nat. Commun. 2018, 9, 4475. [Google Scholar] [CrossRef] [Green Version]
  155. Montigny, A.; Tavormina, P.; Duboe, C.; San Clemente, H.; Aguilar, M.; Valenti, P.; Lauressergues, D.; Combier, J.P.; Plaza, S. Drosophila primary microRNA-8 encodes a microRNA-encoded peptide acting in parallel of miR-8. Genome Biol. 2021, 22, 118. [Google Scholar] [CrossRef] [PubMed]
  156. Lauressergues, D.; Couzigou, J.M.; Clemente, H.S.; Martinez, Y.; Dunand, C.; Becard, G.; Combier, J.P. Primary transcripts of microRNAs encode regulatory peptides. Nature 2015, 520, 90–93. [Google Scholar] [CrossRef] [PubMed]
  157. Slavoff, S.A.; Mitchell, A.J.; Schwaid, A.G.; Cabili, M.N.; Ma, J.; Levin, J.Z.; Karger, A.D.; Budnik, B.A.; Rinn, J.L.; Saghatelian, A. Peptidomic discovery of short open reading frame-encoded peptides in human cells. Nat. Chem. Biol. 2013, 9, 59–64. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Matsumoto, A.; Nakayama, K.I. Hidden Peptides Encoded by Putative Noncoding RNAs. Cell Struct. Funct. 2018, 43, 75–83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Van Heesch, S.; Witte, F.; Schneider-Lunitz, V.; Schulz, J.F.; Adami, E.; Faber, A.B.; Kirchner, M.; Maatz, H.; Blachut, S.; Sandmann, C.L.; et al. The Translational Landscape of the Human Heart. Cell 2019, 178, 242–260.e29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. D’Lima, N.G.; Ma, J.; Winkler, L.; Chu, Q.; Loh, K.H.; Corpuz, E.O.; Budnik, B.A.; Lykke-Andersen, J.; Saghatelian, A.; Slavoff, S.A. A human microprotein that interacts with the mRNA decapping complex. Nat. Chem. Biol. 2017, 13, 174–180. [Google Scholar] [CrossRef] [PubMed]
  161. Kondo, T.; Plaza, S.; Zanet, J.; Benrabah, E.; Valenti, P.; Hashimoto, Y.; Kobayashi, S.; Payre, F.; Kageyama, Y. Small peptides switch the transcriptional activity of Shavenbaby during Drosophila embryogenesis. Science 2010, 329, 336–339. [Google Scholar] [CrossRef] [Green Version]
  162. Zanet, J.; Benrabah, E.; Li, T.; Pelissier-Monier, A.; Chanut-Delalande, H.; Ronsin, B.; Bellen, H.J.; Payre, F.; Plaza, S. Pri sORF peptides induce selective proteasome-mediated protein processing. Science 2015, 349, 1356–1358. [Google Scholar] [CrossRef] [PubMed]
  163. Bi, P.; Ramirez-Martinez, A.; Li, H.; Cannavino, J.; McAnally, J.R.; Shelton, J.M.; Sanchez-Ortiz, E.; Bassel-Duby, R.; Olson, E.N. Control of muscle formation by the fusogenic micropeptide myomixer. Science 2017, 356, 323–327. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Zhang, Q.; Vashisht, A.A.; O’Rourke, J.; Corbel, S.Y.; Moran, R.; Romero, A.; Miraglia, L.; Zhang, J.; Durrant, E.; Schmedt, C.; et al. The microprotein Minion controls cell fusion and muscle formation. Nat. Commun. 2017, 8, 15664. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Lin, Y.F.; Xiao, M.H.; Chen, H.X.; Meng, Y.; Zhao, N.; Yang, L.; Tang, H.; Wang, J.L.; Liu, X.; Zhu, Y.; et al. A novel mitochondrial micropeptide MPM enhances mitochondrial respiratory activity and promotes myogenic differentiation. Cell Death Dis. 2019, 10, 528. [Google Scholar] [CrossRef]
  166. Matsumoto, A.; Pasut, A.; Matsumoto, M.; Yamashita, R.; Fung, J.; Monteleone, E.; Saghatelian, A.; Nakayama, K.I.; Clohessy, J.G.; Pandolfi, P.P. mTORC1 and muscle regeneration are regulated by the LINC00961-encoded SPAR polypeptide. Nature 2017, 541, 228–232. [Google Scholar] [CrossRef]
  167. Bi, P.; McAnally, J.R.; Shelton, J.M.; Sanchez-Ortiz, E.; Bassel-Duby, R.; Olson, E.N. Fusogenic micropeptide Myomixer is essential for satellite cell fusion and muscle regeneration. Proc. Natl. Acad. Sci. USA 2018, 115, 3864–3869. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Nelson, B.R.; Makarewich, C.A.; Anderson, D.M.; Winders, B.R.; Troupes, C.D.; Wu, F.; Reese, A.L.; McAnally, J.R.; Chen, X.; Kavalali, E.T.; et al. A peptide encoded by a transcript annotated as long noncoding RNA enhances SERCA activity in muscle. Science 2016, 351, 271–275. [Google Scholar] [CrossRef] [Green Version]
  169. Makarewich, C.A.; Munir, A.Z.; Schiattarella, G.G.; Bezprozvannaya, S.; Raguimova, O.N.; Cho, E.E.; Vidal, A.H.; Robia, S.L.; Bassel-Duby, R.; Olson, E.N. The DWORF micropeptide enhances contractility and prevents heart failure in a mouse model of dilated cardiomyopathy. eLife 2018, 7. [Google Scholar] [CrossRef]
  170. Magny, E.G.; Pueyo, J.I.; Pearl, F.M.; Cespedes, M.A.; Niven, J.E.; Bishop, S.A.; Couso, J.P. Conserved regulation of cardiac calcium uptake by peptides encoded in small open reading frames. Science 2013, 341, 1116–1120. [Google Scholar] [CrossRef]
  171. Polycarpou-Schwarz, M.; Gross, M.; Mestdagh, P.; Schott, J.; Grund, S.E.; Hildenbrand, C.; Rom, J.; Aulmann, S.; Sinn, H.P.; Vandesompele, J.; et al. The cancer-associated microprotein CASIMO1 controls cell proliferation and interacts with squalene epoxidase modulating lipid droplet formation. Oncogene 2018, 37, 4750–4768. [Google Scholar] [CrossRef]
  172. Zhu, S.; Wang, J.Z.; Chen, D.; He, Y.T.; Meng, N.; Chen, M.; Lu, R.X.; Chen, X.H.; Zhang, X.L.; Yan, G.R. An oncopeptide regulates m(6)A recognition by the m(6)A reader IGF2BP1 and tumorigenesis. Nat. Commun. 2020, 11, 1685. [Google Scholar] [CrossRef] [Green Version]
  173. Huang, J.Z.; Chen, M.; Chen, D.E.; Gao, X.C.; Zhu, S.; Huang, H.; Hu, M.; Zhu, H.; Yan, G.R. A Peptide Encoded by a Putative lncRNA HOXB-AS3 Suppresses Colon Cancer Growth. Mol. Cell 2017, 68, 171–184.e6. [Google Scholar] [CrossRef] [Green Version]
  174. Lewandowski, J.P.; Dumbovic, G.; Watson, A.R.; Hwang, T.; Jacobs-Palmer, E.; Chang, N.; Much, C.; Turner, K.M.; Kirby, C.; Rubinstein, N.D.; et al. The Tug1 lncRNA locus is essential for male fertility. Genome Biol. 2020, 21, 237. [Google Scholar] [CrossRef] [PubMed]
  175. Stein, C.S.; Jadiya, P.; Zhang, X.; McLendon, J.M.; Abouassaly, G.M.; Witmer, N.H.; Anderson, E.J.; Elrod, J.W.; Boudreau, R.L. Mitoregulin: A lncRNA-Encoded Microprotein that Supports Mitochondrial Supercomplexes and Respiratory Efficiency. Cell Rep. 2018, 23, 3710–3720.e3718. [Google Scholar] [CrossRef]
  176. Anderson, D.M.; Anderson, K.M.; Chang, C.L.; Makarewich, C.A.; Nelson, B.R.; McAnally, J.R.; Kasaragod, P.; Shelton, J.M.; Liou, J.; Bassel-Duby, R.; et al. A micropeptide encoded by a putative long noncoding RNA regulates muscle performance. Cell 2015, 160, 595–606. [Google Scholar] [CrossRef] [Green Version]
  177. Pauli, A.; Norris, M.L.; Valen, E.; Chew, G.L.; Gagnon, J.A.; Zimmerman, S.; Mitchell, A.; Ma, J.; Dubrulle, J.; Reyon, D.; et al. Toddler: An embryonic signal that promotes cell movement via Apelin receptors. Science 2014, 343, 1248636. [Google Scholar] [CrossRef] [PubMed]
  178. Zhu, S.; Wang, J.; He, Y.; Meng, N.; Yan, G.R. Peptides/Proteins Encoded by Non-coding RNA: A Novel Resource Bank for Drug Targets and Biomarkers. Front. Pharm. 2018, 9, 1295. [Google Scholar] [CrossRef] [Green Version]
  179. Laumont, C.M.; Vincent, K.; Hesnard, L.; Audemard, E.; Bonneil, E.; Laverdure, J.P.; Gendron, P.; Courcelles, M.; Hardy, M.P.; Cote, C.; et al. Noncoding regions are the main source of targetable tumor-specific antigens. Sci. Transl. Med. 2018, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Lancaster, E.M.; Jablons, D.; Kratz, J.R. Applications of Next-Generation Sequencing in Neoantigen Prediction and Cancer Vaccine Development. Genet. Test. Mol. Biomark. 2020, 24, 59–66. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Models of NMD activation mechanisms in mammals and yeast. (a) Mammalian EJC-enhanced NMD. When an EJC remains bound to the RNA downstream of a termination codon, translation termination is inefficient as the EJC interferes with the interaction between PABPC1 and the eukaryotic release factors (eRF1/eRF3). Instead, a SURF complex (Smg1–Upf1–eRF1–eRF3) forms at the level of the PTC. Upf2 and Upf3 are then recruited by the downstream EJC and associate with SURF to form the decay-inducing (DECID) complex. Smg1 phosphorylates Upf1 (P), activating it. Phosphorylated Upf1 promotes RNA decay via Smg6-dependent endonucleolytic cleavage and the Smg5–Smg7-dependent triggering of mRNA deadenylation and decapping. ATP hydrolysis by Upf1 allows the dissociation of the termination complex and the release of the transcript, which can be degraded. Upf1 dephosphorylation by protein phosphatase 2A (PP2A) allows it to return to a dephosphorylated state. The coding region and the UTRs of the mRNA are represented as a large blue arrow and thin blue lines, respectively. See the key for the other symbols. (b) “Faux” 3′ UTR model in yeast. A long 3′ UTR results in inefficient translation termination and Upf1 interaction with eRF1/eRF3, promoting the formation of the Upf1-23 complex (Upf1, Upf2, Upf3) at the level of the terminating ribosome. The Upf2-Upf3 heterodimer is then replaced by Nmd4, Ebs1, the decapping enzyme Dcp2 and its co-factors Dcp1 and Edc3 in the Upf1-decapping complex, leading to RNA decapping. ATP hydrolysis by Upf1 promotes the disassembly of the mRNA/ribosome/Upf1-decapping complex, leading to the release of the transcript which can finally be degraded by Xrn1.
Figure 1. Models of NMD activation mechanisms in mammals and yeast. (a) Mammalian EJC-enhanced NMD. When an EJC remains bound to the RNA downstream of a termination codon, translation termination is inefficient as the EJC interferes with the interaction between PABPC1 and the eukaryotic release factors (eRF1/eRF3). Instead, a SURF complex (Smg1–Upf1–eRF1–eRF3) forms at the level of the PTC. Upf2 and Upf3 are then recruited by the downstream EJC and associate with SURF to form the decay-inducing (DECID) complex. Smg1 phosphorylates Upf1 (P), activating it. Phosphorylated Upf1 promotes RNA decay via Smg6-dependent endonucleolytic cleavage and the Smg5–Smg7-dependent triggering of mRNA deadenylation and decapping. ATP hydrolysis by Upf1 allows the dissociation of the termination complex and the release of the transcript, which can be degraded. Upf1 dephosphorylation by protein phosphatase 2A (PP2A) allows it to return to a dephosphorylated state. The coding region and the UTRs of the mRNA are represented as a large blue arrow and thin blue lines, respectively. See the key for the other symbols. (b) “Faux” 3′ UTR model in yeast. A long 3′ UTR results in inefficient translation termination and Upf1 interaction with eRF1/eRF3, promoting the formation of the Upf1-23 complex (Upf1, Upf2, Upf3) at the level of the terminating ribosome. The Upf2-Upf3 heterodimer is then replaced by Nmd4, Ebs1, the decapping enzyme Dcp2 and its co-factors Dcp1 and Edc3 in the Upf1-decapping complex, leading to RNA decapping. ATP hydrolysis by Upf1 promotes the disassembly of the mRNA/ribosome/Upf1-decapping complex, leading to the release of the transcript which can finally be degraded by Xrn1.
Ncrna 07 00044 g001
Figure 2. Model of translation-dependent displacement of Upf1 from the mRNA coding region. Upf1 binds promiscuously to any accessible RNA (including NMD non-targets), independently of translation. ATP hydrolysis promotes Upf1 dissociation from non-target RNA regions. Upf1 is also displaced from the coding region by the translating ribosome. This model implies that NMD substrate selection occurs after Upf1 associates with the RNA.
Figure 2. Model of translation-dependent displacement of Upf1 from the mRNA coding region. Upf1 binds promiscuously to any accessible RNA (including NMD non-targets), independently of translation. ATP hydrolysis promotes Upf1 dissociation from non-target RNA regions. Upf1 is also displaced from the coding region by the translating ribosome. This model implies that NMD substrate selection occurs after Upf1 associates with the RNA.
Ncrna 07 00044 g002
Figure 3. Model of yeast cytoplasmic aslncRNA degradation by NMD. Once in the cytoplasm, an aslncRNA (red) would rapidly be bound by ribosomes, unless in a dsRNA structure with its paired-sense mRNA (blue). This dsRNA might also interfere with Upf1 binding to the aslncRNAs and could be removed by the action of RNA helicases. The detection of a long 3′ UTR would trigger NMD by a mechanism similar to the “faux 3′ UTR”. A Upf1-23 complex would form at the level of the termination codon thanks to the interaction between Upf1 and eRF1/eRF3. The subsequent formation of the Upf1-decapping complex would lead to the decapping of the aslncRNA by Dcp2. Upon ATP-dependent disassembly of the complex, the decapped aslncRNA is degraded by Xrn1. The mRNA and the aslncRNA are represented in blue and red, respectively. Large arrows and thin lines represent the coding regions and the UTRs, respectively. The ribosome and NMD/decapping factors are represented as in Figure 1.
Figure 3. Model of yeast cytoplasmic aslncRNA degradation by NMD. Once in the cytoplasm, an aslncRNA (red) would rapidly be bound by ribosomes, unless in a dsRNA structure with its paired-sense mRNA (blue). This dsRNA might also interfere with Upf1 binding to the aslncRNAs and could be removed by the action of RNA helicases. The detection of a long 3′ UTR would trigger NMD by a mechanism similar to the “faux 3′ UTR”. A Upf1-23 complex would form at the level of the termination codon thanks to the interaction between Upf1 and eRF1/eRF3. The subsequent formation of the Upf1-decapping complex would lead to the decapping of the aslncRNA by Dcp2. Upon ATP-dependent disassembly of the complex, the decapped aslncRNA is degraded by Xrn1. The mRNA and the aslncRNA are represented in blue and red, respectively. Large arrows and thin lines represent the coding regions and the UTRs, respectively. The ribosome and NMD/decapping factors are represented as in Figure 1.
Ncrna 07 00044 g003
Figure 4. Possible roles of translation and NMD in the metabolism of (as)lncRNAs. Schematic representation of the functional importance of NMD and translation in lncRNAs metabolism (see main text for details). The chain of blue balls represents the micropeptide produced upon the translation of the smORF (red arrow) of the (as)lncRNA. The ribosome, the (as)lncRNA and the NMD/decay factors are represented as above.
Figure 4. Possible roles of translation and NMD in the metabolism of (as)lncRNAs. Schematic representation of the functional importance of NMD and translation in lncRNAs metabolism (see main text for details). The chain of blue balls represents the micropeptide produced upon the translation of the smORF (red arrow) of the (as)lncRNA. The ribosome, the (as)lncRNA and the NMD/decay factors are represented as above.
Ncrna 07 00044 g004
Table 1. Examples of functional lncRNA-derived micropeptides.
Table 1. Examples of functional lncRNA-derived micropeptides.
MicropeptideSpeciesTargetFunction(s)Ref.
NoBoDyHumanmRNA decapping factorsRegulation of mRNA turnover and P-body numbers[160]
CASIMO1HumanSqualene epoxidaseCarcinogenesis; cell lipid homeostasis[171]
PINT87aaHumanPolymerase associated factor complex (PAF1c)Oncogene transcriptional inhibition; tumor suppressive effect[154]
HOXB-AS3HumanhnRNP A1 splicing factorColon cancer growth suppression[173]
RBRPHumanm6A reader IGF2BP1Regulation of m6A recognition by IGF2BP1 on c-Myc mRNA; tumorigenesis[172]
Minion/
Myomixer
Human, mouseUnknownMyoblast fusion; muscle formation and development[163,164,167]
SPARHuman, mouseLysosomal v-ATPaseRegulation of mTORC1 signaling pathway; muscle regeneration[166]
TUG1-BOATHuman, mouseUnknownUnknown; alters mitochondrial membrane potential when overexpressed[174]
MtlnHuman, mouseCardiolipinIncrease of mitochondrial functions[175]
DWORFMouseSERCASERCA (sarcoplasmic reticulum Ca2+-ATPase) activation[168,169]
MLNMouseSERCASERCA inhibition[176]
ToddlerZebrafishUnknownPromoting cell migration during embryogenesis[174]
PriDrosophilaUbr3 E3 ubiquitin ligaseProteasome-dependent processing of the developmental Svb transcription factor[162]
SclDrosophilaCa-P60A SERCACalcium transport regulation[170]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

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. Non-Coding RNA 2021, 7, 44. https://0-doi-org.brum.beds.ac.uk/10.3390/ncrna7030044

AMA Style

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. Non-Coding RNA. 2021; 7(3):44. https://0-doi-org.brum.beds.ac.uk/10.3390/ncrna7030044

Chicago/Turabian Style

Andjus, Sara, Antonin Morillon, and Maxime Wery. 2021. "From Yeast to Mammals, the Nonsense-Mediated mRNA Decay as a Master Regulator of Long Non-Coding RNAs Functional Trajectory" Non-Coding RNA 7, no. 3: 44. https://0-doi-org.brum.beds.ac.uk/10.3390/ncrna7030044

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