Skip to content
Publicly Available Published by De Gruyter October 29, 2019

Ribosome recycling in mRNA translation, quality control, and homeostasis

  • Elina Nürenberg-Goloub and Robert Tampé EMAIL logo
From the journal Biological Chemistry

Abstract

Protein biosynthesis is a conserved process, essential for life. Ongoing research for four decades has revealed the structural basis and mechanistic details of most protein biosynthesis steps. Numerous pathways and their regulation have recently been added to the translation system describing protein quality control and messenger ribonucleic acid (mRNA) surveillance, ribosome-associated protein folding and post-translational modification as well as human disorders associated with mRNA and ribosome homeostasis. Thus, translation constitutes a key regulatory process placing the ribosome as a central hub at the crossover of numerous cellular pathways. Here, we describe the role of ribosome recycling by ATP-binding cassette sub-family E member 1 (ABCE1) as a crucial regulatory step controlling the biogenesis of functional proteins and the degradation of aberrant nascent chains in quality control processes.

Introduction

The origin of translation is inextricably related to the universal genetic code. Evolution has assembled large ribonucleoprotein (RNP) complexes to catalyze translation and developed a sophisticated network to regulate and survey it. Genetic information is stored in large deoxyribonucleic acid (DNA) molecules termed chromosomes. Highly regulated cellular mechanisms allow the transcription of a requested gene into messenger ribonucleic acid (mRNA). Decoding of the mRNA triplet code into a polypeptide chain is coordinated and catalyzed within the ribosome, a macromolecular factory comprising ribosomal RNA (rRNA) and proteins (rps) that form the translation machinery (70S ribosomes in Bacteria and Archaea; 80S ribosomes in Eukarya) out of a small (30S/40S) and a large (50S/60S) subunit. Transfer RNAs (tRNAs) act as adapter units, carrying an antisense triplet to decode the mRNA, and deliver the correct amino acid to elongate the polypeptide chain accordingly. The tightly coupled assembly, translocation, and disassembly of the ribosome as well as the delivery of tRNA are managed by translation factors and divided into four phases: initiation, elongation, termination, and ribosome recycling (Hellen, 2018; Shirokikh and Preiss, 2018; Stein and Frydman, 2019). However, the biogenesis of functional proteins requires precise reconcilement of translation with protein folding, chemical modification, and eventually membrane insertion or translocation. Thus, these processes are likewise coordinated by ribosome-associated factors (Kramer et al., 2019; Waudby et al., 2019). Ribosome heterogeneity is debated to alter key functions of single ribosomes (Ferretti and Karbstein, 2019). Additionally, quality control pathways, partly triggered at macromolecular assemblies of multiple ribosomes, dynamically regulate mRNA and protein homeostasis (Joazeiro, 2019). By fine-tuning the translation machinery, cells can switch between global operation modes such as survival, proliferation, differentiation, and apoptosis (Holcik and Sonenberg, 2005; Chang and Stanford, 2008). Imbalances in this sensitive system can therefore lead to various diseases, including neurodegeneration, cancer, or inherited ribosomopathies, for reasons we are just beginning to understand (Gao et al., 2017; Robichaud et al., 2018; Tahmasebi et al., 2018). Here, we discuss the structural and regulatory aspects of translation with an emphasis on ribosome recycling, a crucial step at the intersection of numerous translational pathways.

Features of the ribosome

Ribosomal subunits consist of universally conserved core rRNA and rps but also of peripheral elements specific for each phyla, organism, and even organelle (Bieri et al., 2018; Rodnina, 2018; Jobe et al., 2019). Assembled ribosomes are highly dynamic and undergo large conformational rearrangements both among the two subunits against each other as well as within each subunit (Prabhakar et al., 2017; Rodnina, 2018; Jobe et al., 2019). The interface composes three ribosomal tRNA binding sites for aminoacylated tRNA (A-site), peptidyl-tRNA (P-site), and deacylated tRNA (E-site, exit site) (Rheinberger et al., 1981; Agrawal et al., 1996; Stark et al., 1997). The conserved helix 44 (h44) of the 16S/18S rRNA maintains subunit association during translocation and contributes to translation fidelity (Jenner et al., 2010; Qin et al., 2012; Liu and Fredrick, 2016). Catalysis takes place on the large ribosomal subunit, in the peptidyl transferase center (PTC) constituted by the 23S/28S rRNA component. The emerging polypeptide moves through the exit tunnel, which is ~10 nm long and allows folding of the nascent chain into secondary structures in a shielded environment (Nilsson et al., 2015; Komar, 2018). On the solvent-exposed side of the large subunit, the N terminus of the partially folded polypeptide encounters various regulatory and processing factors like folding chaperones, signal recognition particles for trafficking, and enzymes for co-translational modification (Kramer et al., 2019). The sarcin-ricin loop (SRL) in the 23S/28S rRNA and the P-stalk control translational factors, which orchestrate each phase of protein synthesis (Voorhees and Ramakrishnan, 2013).

Ribosome synthesis requires over 200 additional proteins and is one of the most energy-consuming processes in the cell (Kressler et al., 2017; Pena et al., 2017). Disorders in ribosome biogenesis are connected to various human diseases referred to as ribosomopathies (Narla and Ebert, 2010; Mills and Green, 2017; Tahmasebi et al., 2018). On the other hand, deactivation of ribosomes immediately arrests the synthesis of all proteins and can protect cells from viral assault, starvation, and endoplasmic reticulum (ER) stress by unfolded proteins. In these critical situations, cells monitor definite ribosome loss and cleave rRNA by RNAse L in innate immunity, RNAse PH during nutrient shortage, and inositol requiring enzyme-1β (IRE1β) in the context of the unfolded protein response (UPR) (Wreschner et al., 1981; Iwawaki et al., 2001; Basturea et al., 2011). Thus, ribosome homeostasis is critical during cell stress, development, and proliferation (Mills and Green, 2017; de la Cruz et al., 2018).

Translation termination

Canonical termination involves recognition of the stop codon in the ribosomal A-site and peptide release by the cooperative action of class I and class II release factors (RFs). Termination in Bacteria differs significantly from the eukaryotic mechanism, while Archaea follow the eukaryotic route, albeit with a fractional set of factors. Eukaryotic/archaeal (e/a) class I e/aRF1 comprises three flexible domains and structurally mimics a tRNA molecule (Song et al., 2000). e/aRF1 is delivered to the ribosomal A-site in a stable ternary complex with the translational GTPases eRF3 (Zhouravleva et al., 1995; Alkalaeva et al., 2006) or aEF1α (Kobayashi et al., 2012). In Eukarya, guanosine triphosphate (GTP) hydrolysis and dissociation of eRF3 allows the full accommodation of eRF1 in the A-site, leading to peptide release (Figure 1). The decisive fidelity and efficiency of decoding factor accommodation during elongation (aa-tRNA), termination (eRF1), and mRNA surveillance (ePelota) depends on a complex interplay between (i) decoding and (de-)stabilization of mRNA in the A-site, (ii) conformational changes of the small ribosomal subunit, in bacteria also referred to as domain closure (Ogle et al., 2002), (iii) activation of the respective delivery GTPase, and (iv) structural rearrangements within the decoding factor (Alkalaeva et al., 2006; Shao et al., 2016). The conserved NIKS motif in the N-terminal (N) domain of eRF1 recognizes the stop codon in a precise geometric arrangement, which includes mRNA compaction and stacking interaction with the A1825 base (rabbit) in the 40S ribosomal h44. The additional conserved GTS and YxCxxxF motifs in eRF1 discriminate against sense codons (Brown et al., 2015). Upon successful decoding, GTP hydrolysis within eRF3 is activated via the SRL, and eRF3·guanosine diphosphate (GDP) dissociates. Thereby, the middle (M) domain of eRF1 is liberated and swings into the PTC, where the universally conserved GGQ motif assists the nucleophilic attack of a water molecule to release the nascent chain (Frolova et al., 1999; Brown et al., 2015). Notably, these structural rearrangements in class I RFs, rather than peptide release, mark the formation of a post-termination complex (post-TC) ready for ribosome recycling.

Figure 1: Molecular mechanism of translation termination and ribosome recycling.
(A) The pre-termination complex (PDB 5LZT) comprises 80S ribosomes, mRNA, deacylated E-site tRNA, peptidyl-tRNA in the P-site, eRF1 in the A-site, and eRF3-GTP at the GTPase center. The conserved NIKS and GTS motifs in the N-terminal domain (NTD) of eRF1 recognize the stop-codon in the decoding center. The middle-domain (MD) with the GGQ motif is locked by eRF3-GTP, thus the nascent chain in the PTC has not yet been released. (B) After GTP hydrolysis and eRF3 dissociation, ABCE1 occupies the GTPase center forming the 80S pre-splitting complex (PDB 5LZV). The MD of eRF1 is in the post-termination state, swung out into the PTC (black arrow), where a water molecule coordinated by the GGQ motif can release the nascent chain. The N-terminal FeS domain (FeSD) of ABCE1 is associated with the C-terminal domain (CTD) of eRF1. The control site II is in a semi-closed state, with NBD1 and NBD2 slightly approaching each other while site I remains open. (C) Occlusion of two ATP molecules by ABCE1 splits the ribosome apart by a 150° rotation of the FeSD (black arrow) toward helix 44 and ABCE1 remains bound at the 40S subunit, reconstituting the post-splitting complex.
Figure 1:

Molecular mechanism of translation termination and ribosome recycling.

(A) The pre-termination complex (PDB 5LZT) comprises 80S ribosomes, mRNA, deacylated E-site tRNA, peptidyl-tRNA in the P-site, eRF1 in the A-site, and eRF3-GTP at the GTPase center. The conserved NIKS and GTS motifs in the N-terminal domain (NTD) of eRF1 recognize the stop-codon in the decoding center. The middle-domain (MD) with the GGQ motif is locked by eRF3-GTP, thus the nascent chain in the PTC has not yet been released. (B) After GTP hydrolysis and eRF3 dissociation, ABCE1 occupies the GTPase center forming the 80S pre-splitting complex (PDB 5LZV). The MD of eRF1 is in the post-termination state, swung out into the PTC (black arrow), where a water molecule coordinated by the GGQ motif can release the nascent chain. The N-terminal FeS domain (FeSD) of ABCE1 is associated with the C-terminal domain (CTD) of eRF1. The control site II is in a semi-closed state, with NBD1 and NBD2 slightly approaching each other while site I remains open. (C) Occlusion of two ATP molecules by ABCE1 splits the ribosome apart by a 150° rotation of the FeSD (black arrow) toward helix 44 and ABCE1 remains bound at the 40S subunit, reconstituting the post-splitting complex.

Ribosome recycling

Within the canonical translation cycle, terminated ribosomes are recycled into free subunits. Ribosome recycling is tightly regulated and represents a key process in translational control. In Eukarya and Archaea, ribosome recycling is orchestrated by the essential and conserved ATP-binding cassette sub-family E member 1 (ABCE1) in collaboration with the class I RFs in the A-site of the post-TC (Pisarev et al., 2010; Barthelme et al., 2011; Shoemaker and Green, 2011). ABCE1 is a non-membrane-associated ABC-type ATPase (70 kDa) and represents the only member of ABC subfamily E. This ribosome recycling factor is evolutionarily conserved and essential in all organisms except Bacteria. Like in a typical ABC protein, the two nucleotide-binding domains (NBDs) of ABCE1 are arranged head-to-tail and harbor two composite nucleotide-binding sites at their interface (Karcher et al., 2005, 2008; Barthelme et al., 2011). ATP binding in both sites leads to tight dimerization (closure) of the NBDs, while subsequent ATP hydrolysis and release of inorganic phosphate allow their disengagement. Thus, the NBDs perform a tweezer-like motion, which is mechanically coupled to associated domains to either transport substrates across the membrane or remodel nucleoprotein complexes (Hopfner, 2016; Thomas and Tampé, 2018). Additionally, ABCE1 possesses a unique N-terminal iron-sulfur (FeS) cluster domain, which harbors two diamagnetic [4Fe-4S]2+ clusters, a helix-loop-helix insertion in NBD1, and a small and flexible hinge region, which connects the two NBDs and supports their orientation (Karcher et al., 2005; Barthelme et al., 2007, 2011). Mutations in conserved functional motifs of ABCE1 are lethal at early embryonic stage and have asymmetric effects on the overall ATPase activity of the protein (Andersen and Leevers, 2007; Barthelme et al., 2011; Nürenberg-Goloub et al., 2018). Similar to translational GTPases, ABCE1 engages with the ribosomal P-stalk of 70S/80S ribosomes (Imai et al., 2018) before it attaches near the GTPase activating center and contacts the C-terminal (C) domain of the respective class I RF thereby forming the pre-splitting complex (Figure 1) (Becker et al., 2012; Shao et al., 2016). Interestingly, ABCE1 promotes eRF1-mediated release of short peptides in the absence of eRF3, though with lower efficiency, suggesting a yet uncharacterized function in translation termination (Shoemaker and Green, 2011). The two structurally and functionally distinct sites allow ABCE1 to distinguish splitting-competent ribosomes (Nürenberg-Goloub et al., 2018). More precisely, a control site with an intrinsically low ATP-turnover rate holds ABCE1 at the pre-splitting complex while the other site probes for a conformation, which would induce ribosome splitting. If ABCE1 fails to achieve this conformation, it can be released from the pre-splitting complex and bind another ribosome (Nürenberg-Goloub et al., 2018; Gouridis et al., 2019). In case the complex comprises a splitting-competent ribosome, ATP-dependent conformational changes of the NBDs and the FeS cluster domain induce a translocation-like ribosome destabilization (Pisarev et al., 2010). Finally, ATP occlusion and closure of both sites displaces the FeS domain, which protrudes between the ribosomal subunits causing their dissociation (Heuer et al., 2017; Nürenberg-Goloub et al., 2018). After splitting, ABCE1 remains at the small ribosomal subunit, establishing the post-splitting complex (post-SC, Figure 1) (Barthelme et al., 2011; Kiosze-Becker et al., 2016; Heuer et al., 2017). ATP hydrolysis in both sites of ABCE1 is downregulated within the post-SC (Nürenberg-Goloub et al., 2018), where it prevents rejoining of the large ribosomal subunit (Heuer et al., 2017). Deacylated tRNA and mRNA are removed from separated 40S subunits by the translation initiation factors (IFs) eIF1, eIF1A, and eIF3 (Pisarev et al., 2010), ligatin, or multiple copies in T-cell lymphoma-1/density regulated protein (Skabkin et al., 2010) in an ABCE1-independent manner in vitro. The additional presence of eIF2 and initiator Met-tRNAMet promotes re-initiation of translation at up- and downstream start-codons on the same mRNA prior to its ejection and the addition of eIF4F ensures 3′-directionality of re-initiation (Skabkin et al., 2013).

Within the post-SC, ABCE1 can link ribosome recycling to translation initiation (Heuer et al., 2017; Mancera-Martinez et al., 2017; Gerovac and Tampé, 2019) and has been proposed to act in 5′-cap and poly-A independent mRNA circularization (Afonina and Shirokov, 2018). First indications on the role of ABCE1 in initiation was gained from human, fruit fly, and yeast before its definite assignment to ribosome recycling (Dong et al., 2004; Yarunin et al., 2005; Chen et al., 2006; Andersen and Leevers, 2007). Notably, the non-essential initiation factor eIF3j (Hcr1 in yeast) assists ABCE1 during recycling of terminated 80S ribosomes at stop codons in vivo (Young and Guydosh, 2019). In cooperation with other IFs, Hcr1 controls stringent start codon selection during translation initiation and assists eRF3 ejection from 80S ribosomes after translation termination (Elantak et al., 2010; Beznoskova et al., 2013). Both ABCE1 and Hcr1 are associated within the multifactor complex free of ribosomes, rising the assumption of a pre-formed unit functional in termination, ribosome recycling, and (re-)initiation (Asano et al., 2000; Dong et al., 2004). Therein, ABCE1 not only restocks the cellular pool of free ribosomal subunits, but schedules the selective recycling of post-TCs and gates small ribosomal subunits toward translation initiation (Heuer et al., 2017; Nürenberg-Goloub et al., 2018).

Apart from its multiple essential roles related to ribosome recycling, ABCE1 has vaguely defined functions as RNAse L inhibitor (Bisbal et al., 1995) and in the assembly of numerous virions including those of human immunodeficiency virus (Dooher et al., 2007; Anderson et al., 2019). Thus, the molecular mechanism of this versatile protein is of high interest in diverse fields of research, including therapeutics against human diseases like rabies or cancer (Huang et al., 2010; Lingappa et al., 2013).

mRNA surveillance and ribosome-based quality control

Elongation and termination can fail for numerous reasons, resulting in stalled ribosomes occupied by faulty mRNA and non-functional, potentially harmful polypeptides. The cellular quality control machinery aims to eliminate mRNA, polypeptides, and/or damaged ribosomes at the earliest time point, while they are still unambiguously connected to each other. Thus, quality control directly takes place at the ribosome, which allows the cell to simultaneously target aberrant polypeptides to the proteasome, degrade the respective mRNA, and activate stress response signaling (Brandman and Hegde, 2016; Joazeiro, 2019). Bacteria utilize three different systems to resolve the stalled ribosomal complex and eventually target mRNA and the nascent chain for degradation. Trans-translation of the trapped mRNA (tmRNA) molecule appends a specific sequence to the aberrant peptide that serves as a degradation signal. Notably, bacteria also employ tmRNA to monitor co-translational processes (Hayes and Keiler, 2010) and during starvation to quickly adapt to environmental conditions (Keiler, 2008). Alternative rescue factors (ArfA and ArfB) promote translation termination, peptide release, and ribosome recycling without tagging the nascent chain (Keiler et al., 1996; Keiler, 2015; Huter et al., 2017). Finally, the bacterial ribosome-based quality control (RQC) factor RqcH recognizes peptidyl-tRNA bound to 50S ribosomal subunits and induces nascent chain degradation by appending a poly-alanine tail, suggesting the presence of another yet elusive recycling pathway for stalled ribosomes (Lytvynenko et al., 2019). The three bacterial quality control systems are essential or redundant in some bacterial species but their high evolutional distribution points toward a substantial contribution to cellular fitness during environmental stress (Keiler, 2015; Lytvynenko et al., 2019).

The eukaryotic mRNA surveillance and RQC system comprises multistep response pathways for a variety of translational errors induced by (i) truncated or highly structured mRNAs (no-go decay, NGD), (ii) open reading frames (ORFs) either lacking (no-stop decay, NSD) or containing a (premature) in-frame stop codon (nonsense-mediated decay, NMD), (iii) an aa-tRNA shortage, (iv) polypeptides blocking the ribosomal exit tunnel, or (v) defective ribosomes (18S-NRD) (Brandman and Hegde, 2016; Buskirk and Green, 2017; Joazeiro, 2019). Notably, the NMD pathway is tightly coupled to translation termination and employs the canonical eRF1-eRF3 complex, which interacts with up-frameshift (UPF) translational regulators to induce mRNA and protein degradation (Karousis and Muhlemann, 2019; Kurosaki et al., 2019). In contrast, the hallmark of all other quality control pathways is the stalled 80S ribosome, with either no mRNA or a sense codon in the decoding site. Stalled 80S ribosomes recruit the stop codon-independent class I RF ePelota (Dom34 in yeast), delivered by GTPase Hbs1 (Chen et al., 2010; Becker et al., 2011). Hbs1 senses the lengths of 3′-mRNA overhangs at the 80S (Shoemaker and Green, 2011) and recruits the superkiller (SKI) complex involved in fast degradation of the aberrant mRNA, in concert with the exosome and other nucleases (Saito et al., 2013; Joazeiro, 2019). Subsequently, Hbs1 dissociates, leaving a post-TC with intact peptidyl-tRNA in the P-site and ePelota in the A-site. This complex is a substrate for ABCE1 (Pisarev et al., 2010; Shoemaker and Green, 2011; Becker et al., 2012; Nürenberg and Tampé, 2013). The dynamic regulation of ribosome abundance by ePelota and ABCE1 plays a crucial role during erythropoiesis (Mills et al., 2016; Khajuria et al., 2018). Erroneous peptide targeting and stress signaling take place downstream of recycling at 60S subunits, harboring the intact peptidyl-tRNA and involve numerous factors with yet incompletely defined roles and operation modes, including a homolog of the bacterial RqcH (Brandman and Hegde, 2016; Joazeiro, 2019). Especially, the NMD pathway is tightly coupled to post-transcriptional mRNA processing events, mRNA and protein homeostasis, and deterministic regulation of gene expression (He et al., 2003; Yap and Makeyev, 2013). Thus, it is not surprising that failure of NMD leads to various human diseases (Jaffrey and Wilkinson, 2018; Kurosaki et al., 2019).

Information about archaeal mRNA surveillance and RQC pathways is scarce, but appears to include eukaryotic features: the stop codon-independent RF aPelota is delivered to stalled ribosomes by aEF1α (Kobayashi et al., 2010) and recycling of stalled post-TCs likewise depends on ABCE1 (Becker et al., 2012). Archaea possess a yet uncharacterized RqcH homolog (Lytvynenko et al., 2019). It remains largely elusive how Archaea deal with translation and transcription errors, but a simplified eukaryotic system appears to be most likely. Furthermore, it remains to be clarified how archaeal co-translational quality control pathways are involved in global cellular regulation.

Supramolecular ribosome assemblies

Supramolecular assemblies reaching from di-ribosomes to various topologies of polyribosomes can alter translation activity. Formation of defined di-ribosomes by collisions depend on the availability of translation and ribosome recycling factors as well as the sequence of the mRNA. Such collided di-ribosomes are subject to NGD and/or RQC (Simms et al., 2017; Juszkiewicz et al., 2018; Ikeuchi et al., 2019). Polysomes dynamically reorganize during progressive ribosome loading onto an mRNA molecule, establishing contacts between conserved sites on the ribosomal surface (Afonina and Shirokov, 2018; Gohara and Yap, 2018). As inefficient late-stage polysomes were discussed to be collided ribosomes, their formation possibly depends on the availability of translation factors. Thus, ribosome recycling is likewise involved in translation regulation by ribosome assemblies.

Defined eukaryotic di-ribosomes (Figure 2) are formed in the context of RQC and mRNA surveillance pathways (Simms et al., 2017; Juszkiewicz et al., 2018; Ikeuchi et al., 2019). If the rate of translation initiation on an mRNA exceeds its elongation, termination, and/or ribosome recycling capacity, ribosomes likely collide within or at the end of the ORF, respectively. Collided ribosomes have a conserved dimerization interface, which is recognized by the E3 ubiquitin ligases ZNF598/Hel2 in rabbit/yeast. The minimal recognition motif is a di-ribosome. Upon ZNF598 binding, the stalled and collided ribosomes are sequentially ubiquitinated at ribosomal proteins uS10, eS10, and uS3. Elongation is arrested by a yet unknown mechanism. Ribosomes are recycled by ePelota and ABCE1, and aberrant nascent chains and mRNA are degraded via NGD and RQC in an NGDRQC+ response (Simms et al., 2017; Sundaramoorthy et al., 2017; Juszkiewicz et al., 2018; Ikeuchi et al., 2019). Interestingly, a distinct ubiquitination pathway of ribosomal protein eS7 by the E3 ligase Not4 induces an NGDRQC− response, which does not feature degradation of the nascent chain (Ikeuchi et al., 2019). Strikingly, Not4 also mediates ABCE1 ubiquitination in ribosome-associated control of mitophagy (Wu et al., 2018), illustrating the intricate interplay of ribosome recycling and quality control in decisive cellular events. Thus, higher-order ribosome architecture can induce multiple quality control pathways, explaining how cells may distinguish between intended ribosome pausing and situations necessitating intervention to prevent synthesis of aberrant products. As an interesting consequence, the cells’ definition of ‘aberrant’ mRNA and nascent chains becomes dependent on the cellular context, e.g. proliferation or differentiation status and environmental conditions (Juszkiewicz et al., 2018).

Figure 2: Supramolecular ribosome assemblies.
(A) Collided yeast and mammalian di-ribosomes establish a specific interface, mainly involving 40S subunits (PDB 6I7O). This interface is recognized by E3 ubiquitin ligases Hel2 (ZNF598 in mammals) and Not4, which specifically ubiquitinate ribosomal proteins uS3, uS10, and eS7, thereby initiating NGD and RQC. The ribosomal protein Asc1 (RACK1 in mammals) is not ubiquitinated but essential for interface recognition. (B) In helical polysomes (PDB 4V3P), 40S subunits (white) and mRNA are sequestered inside the helix, while the 60S subunits (blue) and nascent chains are exposed to the cytosol. The ribosomal P-stalks (red) are accessible to translation factors. Individual disome units in the helix resemble the stalled di-ribosome in panel (A), with a similar orientation and interface between individual ribosomes. Thus, specific ribosome ubiquitination might explain the reduced translation activity of helical polysomes.
Figure 2:

Supramolecular ribosome assemblies.

(A) Collided yeast and mammalian di-ribosomes establish a specific interface, mainly involving 40S subunits (PDB 6I7O). This interface is recognized by E3 ubiquitin ligases Hel2 (ZNF598 in mammals) and Not4, which specifically ubiquitinate ribosomal proteins uS3, uS10, and eS7, thereby initiating NGD and RQC. The ribosomal protein Asc1 (RACK1 in mammals) is not ubiquitinated but essential for interface recognition. (B) In helical polysomes (PDB 4V3P), 40S subunits (white) and mRNA are sequestered inside the helix, while the 60S subunits (blue) and nascent chains are exposed to the cytosol. The ribosomal P-stalks (red) are accessible to translation factors. Individual disome units in the helix resemble the stalled di-ribosome in panel (A), with a similar orientation and interface between individual ribosomes. Thus, specific ribosome ubiquitination might explain the reduced translation activity of helical polysomes.

In general, higher-order ribosome arrangement modulates protein biosynthesis rates and mRNA stability and enables spatial and temporal control over translation as shown for neurons (Graber et al., 2013; Sossin and Costa-Mattioli, 2019). Interaction of poly-A binding proteins at the 3′-end and the eIF4 cap-binding complex at the 5′-end of juvenile cytosolic mRNA induce a pseudo-circular architecture, which is a widely accepted model (Hinnebusch, 2017). During initial rounds of translation, the spatial proximity of 3′- and 5′-ends allows rapid initiation of recycled ribosomes on the same mRNA. Thus, translation is most efficient within this initial phase, and polysomes have a ring-type shape corresponding to circular mRNA (Kopeina et al., 2008; Behrmann et al., 2015). Notably, approximately 5% of human 80S ribosomes in the actively translating polysome fraction are pre-splitting complexes resembling eRF1 in the post-termination state and ABCE1 (Behrmann et al., 2015). Strikingly, ring-type polysomes were equally observed on synthetic mRNAs that lacked a poly-A tail or 5′-cap, suggesting a distinct or more general mechanism of mRNA circularization (Afonina et al., 2015). As translation initiation occurs at the 5′-end of the mRNA and ribosome recycling follows termination at the 3′-end of the ORF, the multifactor complex including ABCE1 may spatially and temporarily connect these fundamental processes at the post-SC (Heuer et al., 2017; Gerovac and Tampé, 2019).

With progressive ribosome load, eukaryotic polysomes acquire a three-dimensional (3D) helical structure (Figure 2), and the translation efficiency constantly drops (Kopeina et al., 2008; Brandt et al., 2010; Afonina et al., 2015). Interactions between the ribosomes involve 40S and 60S ribosomal subunits, including the ribosomal P-stalk, and rRNA expansion segments (ES) (Brandt et al., 2010; Myasnikov et al., 2014). The mRNA is shielded from the cytosol, and the peptide-exit sites are exposed. Interestingly, the previously enigmatic function of the rRNA ES was suggested to stabilize 3D polysomes. However, rRNA ES are also crucial for recruitment of the N-terminal acyltransferase NatA (Knorr et al., 2019) and possibly other nascent chain processing factors. It has been speculated that the severe decrease of translation efficiency in 3D helical polysomes originates from ZNF598-mediated ribosome arrest (Juszkiewicz et al., 2018) (Figure 2). Therefore, it remains a challenging objective to investigate the mechanism of translational arrest and resumption versus quality control within 3D helical polysomes in light of the diverse in vivo situations.

Translation regulation

Transcription and translation determine the protein composition and thereby define the function and fate of a cell. The levels of mRNA and protein only correlate with approximately half of eukaryotic genes (Vogel and Marcotte, 2012). The reason for this discrepancy lies within the dynamic interplay of regulatory mechanisms involving the mRNA, additional factors, and the ribosome itself (Figure 3). We would like to introduce the basics of eukaryotic translation regulation using examples of recent structural insights and discuss the impact of ribosome recycling and ribosome heterogeneity on the biosynthesis of functional proteins.

Figure 3: Translational control.
(A) Translation of the pseudo-circular eukaryotic mRNA is regulated by cis (orange) and trans (blue) factors and largely depends on the ribosome-recycling activity of ABCE1. Translation is initiated at the small ribosomal subunit, which can be delivered by ABCE1 within the post-splitting complex. A pre-initiation complex scans the mRNA in 5′ to 3′ direction to find a start codon, where initiation is accomplished and elongation begins. Ribosome recycling by ABCE1 promotes re-initiation on the main ORF start codon after translation of an uORF. Terminated, stalled, unrecycled, and hibernating 80S ribosomes are split by ABCE1 before individual subunits can re-enter the translation cycle. (B) The CrPV IRES (PDB 6D9J) mimics parts of the E-site tRNA and mRNA to promote eEF2-mediated ribosome translocation and to initiate translation of viral proteins, independent of the cellular translation initiation machinery. (C) IFRD2 prevents translation in hibernating ribosomes by blocking the P-site and the mRNA channel. Z-site tRNA has recently been visualized in hibernating ribosomes in the presence and absence of IFRD2 (PDB 6MTC). Its detailed role remains to be elucidated.
Figure 3:

Translational control.

(A) Translation of the pseudo-circular eukaryotic mRNA is regulated by cis (orange) and trans (blue) factors and largely depends on the ribosome-recycling activity of ABCE1. Translation is initiated at the small ribosomal subunit, which can be delivered by ABCE1 within the post-splitting complex. A pre-initiation complex scans the mRNA in 5′ to 3′ direction to find a start codon, where initiation is accomplished and elongation begins. Ribosome recycling by ABCE1 promotes re-initiation on the main ORF start codon after translation of an uORF. Terminated, stalled, unrecycled, and hibernating 80S ribosomes are split by ABCE1 before individual subunits can re-enter the translation cycle. (B) The CrPV IRES (PDB 6D9J) mimics parts of the E-site tRNA and mRNA to promote eEF2-mediated ribosome translocation and to initiate translation of viral proteins, independent of the cellular translation initiation machinery. (C) IFRD2 prevents translation in hibernating ribosomes by blocking the P-site and the mRNA channel. Z-site tRNA has recently been visualized in hibernating ribosomes in the presence and absence of IFRD2 (PDB 6MTC). Its detailed role remains to be elucidated.

Eukaryotic mRNA translation is in general controlled by cis and trans regulators (Kozak, 2005; Sonenberg and Hinnebusch, 2009; Hershey et al., 2012; Leppek et al., 2018). Trans regulators include all translation factors, mRNA-binding proteins but also functional cellular RNAs such as micro- and long non-coding RNAs. Examples of trans factors inhibiting translation are interferon-related developmental regulator 2 (IFRD2) and Z-site tRNA specifically bound in proximity of the ribosomal E-site (Brown et al., 2018) (Figure 3). Cis regulators are secondary structures or specific sequences on the mRNA itself, e.g. stem loops, internal ribosomal entry sites (IRES), and upstream ORFs (uORFs). Thereof, a cryo-electron microscopy (cryo-EM) study demonstrated that ribosome hijacking by cricket paralysis virus (CrPV) IRES involves structural mimicry of E-site tRNA (Figure 3) (Muhs et al., 2015). Strikingly, epigenetic imprinting of mRNA during transcription adds another level to translational cis-regulation (Slobodin et al., 2017). uORFs have versatile effects on eukaryotic gene expression relying on cis- and trans-regulatory elements (Wethmar, 2014; Weisser et al., 2017), and are readily exploited in cancer cells (Sriram et al., 2018). Ribosome recycling by ABCE1 precedes binding of (re-)initiation factors to the intersubunit side of 40S ribosomes after translation termination on uORFs (Skabkin et al., 2010, 2013; Lomakin et al., 2017). In the absence of ABCE1, 80S post-TCs can re-initiate on nearby codons cognate with the deacylated P-site tRNA (Skabkin et al., 2013; Young et al., 2015), a mechanism enhanced by viral mRNA containing the cis-regulatory ‘termination upstream ribosomal binding site’ (TURBS) motif (Zinoviev et al., 2015). Thus, ribosome recycling by ABCE1 is a decisive branch point between 80S and 40S re-initiation. Interestingly, various subunits of the eIF3 complex have distinct effects on translation re-initiation suggesting a regulatory role for the ABCE1-eIF3 tandem. Thus, eIF3j/Hcr1, which assists ABCE1 during 80S recycling, reduces the efficiency of translation re-initiation in vitro (Skabkin et al., 2013). In contrast, eIF3h significantly promotes re-initiation in human cells (Hronova et al., 2017) and plants (Roy et al., 2010), while it strongly depends on eIF3a in yeast (Munzarova et al., 2011). Consistently, ribosome recycling is regarded as a regulatory gateway in canonical and aberrant translation (Dever and Green, 2012; Buskirk and Green, 2017; Gerovac and Tampé, 2019) and is strongly connected to ribosome homeostasis (Young et al., 2015; Mills et al., 2016).

However, the pathologic phenotypes of certain mutations involving ribosomal proteins or ribosome biogenesis factors cannot be explained by neither cis- nor trans-regulatory elements of translation (Tahmasebi et al., 2018). The most prominent example is Diamond-Blackfan anemia (DBA), which affects different tissues depending on which ribosomal protein is defective (Draptchinskaia et al., 1999; Gazda et al., 2008; Boria et al., 2010; Gazda et al., 2012). These findings shift the ribosome into focus. The cellular pool of ribosomes is heterogenous and distinct populations vary in rRNA and ribosomal protein composition as well as their respective chemical modifications (Slavov et al., 2015; Ferretti et al., 2017; Shi et al., 2017). However, it remains controversial whether ribosome heterogeneity is actively regulated within the cell and reflects functional specialization leading to populations of ‘specialized ribosomes’ (Gilbert, 2011; Xue and Barna, 2012; Emmott et al., 2019; Ferretti and Karbstein, 2019).

Between 200 and 400 copies of rDNA exist in the human genome. Some of them carry mutations, which lie within the sequence of mature rRNA and exhibit tissue-specific expression (Parks et al., 2018). Additionally, over 200 post-transcriptional modifications concentrate in most important functional rRNA regions and were shown to participate in mRNA and tRNA binding as well as subunit association in Bacteria and Eukarya (Polikanov et al., 2015; Sharma and Lafontaine, 2015; Natchiar et al., 2017; Roundtree et al., 2017). They can alter the local structure of ribosomes, influence mRNA selectivity (Schosserer et al., 2015; Sloan et al., 2017; Sharma et al., 2018), and are thereby involved in the development of cancer (Bellodi et al., 2010; Truitt and Ruggero, 2017) as well as several hereditary diseases (Doll and Grzeschik, 2001; Armistead et al., 2009). Ribosomal subsets differing in protein composition can arise from mutations (Draptchinskaia et al., 1999), haploinsufficiency (Ebert et al., 2008), or tissue-specific expression of paralogous genes (Williams and Sussex, 1995; Marygold et al., 2007), prevalently in diseases, during development (Whittle and Krochko, 2009) or stress response (Zhang and Lu, 2009). Numerous ribosomal proteins affect specific translation of mRNA subpools (Kondrashov et al., 2011; Ferretti et al., 2017; Shi et al., 2017; Yamada et al., 2019) as reviewed elsewhere in more detail (Shi and Barna, 2015; Emmott et al., 2019). Additionally, over 2500 post-translational modifications of ribosomal proteins are known (Emmott et al., 2019), some of which contribute to translational selectivity, e.g. during mitosis (Imami et al., 2018) or in Parkinson’s disease (Martin et al., 2014). Considering the sophisticated biogenesis and degradation pathways, ribosome heterogeneity may also arise from stable ribosome-turnover intermediates (Ferretti and Karbstein, 2019). If so, it must be elucidated whether these intermediates are functional and to what extent their retention is regulated within the cell.

However, not only (i) mRNA selectivity of structurally diverse ribosomes may be crucial to induce a physiological effect but also their (ii) localization, (iii) accuracy, (iv) processivity, and (v) speed (Dinman, 2016). Additionally, the biosynthesis of functional proteins depends on co-translational recruitment of auxiliary factors (e.g. chaperones and protein modification enzymes), which is likewise affected by ribosome composition (Simsek et al., 2017). The most prominent example is RACK1/Asc1, which conducts various roles in translation and bridges to numerous cell signaling pathways (Gallo and Manfrini, 2015). Importantly, RACK1/Asc1 is crucial to trigger mRNA surveillance and RQC pathways on stalled ribosomes (Ikeuchi and Inada, 2016; Sitron et al., 2017; Sundaramoorthy et al., 2017; Juszkiewicz et al., 2018; Ikeuchi et al., 2019). During starvation, RACK1/Asc1 is depleted from ribosomes (Baum et al., 2004). Thus, the cell responds to environmental conditions and may control pathways downstream of translation via ribosome composition.

In contrast to the specialized ribosomes, the ribosome abundance hypothesis describes the selective translation of certain subsets of mRNA as an effect of ribosome amount and concentration (Lodish, 1974; Mills and Green, 2017) that can be dynamically regulated by the ribosome recycling factors ePelota and ABCE1 during cellular differentiation (Mills et al., 2016; Khajuria et al., 2018). Further, ribosomes occupy approximately 20% of the cytosolic volume and were shown to regulate cellular biochemistry by low-affinity interactions with major metabolic enzymes but also molecular crowding and phase separation in the cytosol (Delarue et al., 2018). The ribosome abundance hypothesis is equally justified, especially in regard to translation regulation by supramolecular ribosome assemblies in polysomes and quality control pathways triggered by collided ribosomes. In respect of the latter, the abundance and activity of ribosome recycling factors is putatively crucial for the fate of the nascent chain and mRNA (see previous section). Effictive mRNA decay by the NGD pathway as well as ribosome ubiquitination and RQC depend on the presence of at least two collided ribosomes (Simms et al., 2017; Ikeuchi et al., 2019). However, stalled ribosomes are efficiently split in the presence of ePelota (Dom34) in vivo (Sitron et al., 2017). Thus, we assume that reduced ribosome recycling activity (either global or local) would lead to ribosome collisions at transient pause sites and enhance NGD and RQC. In contrast, elevated levels of ribosome recycling activity would protect mRNA from effective decay by resolving stalled ribosomes before NGD is triggered. During oxidative stress, the supply of ABCE1 with the essential FeS clusters is inhibited, which globally affects translation (Alhebshi et al., 2012). Consistently, ePelota as well as multiple factors of the NGD and NSD pathways are essential for oxidative stress tolerance, possibly reflecting the increased requirement to resolve stalled ribosomes (Jamar et al., 2017). Another convincing example of translation regulation by ribosome recycling is found during erythrocyte maturation. ABCE1 depletion causes 80S build-up in the 3′ untranslated region (3′UTR) of the transcriptome (which is rescued by elevated levels of ePelota at initial differentiation stage) (Mills et al., 2016). Consistently, the di-ribosome recognition factor ZNF598 is absent from rabbit reticulocyte lysate, which allowed the formation of stable di-ribosomes for structural assessment (Juszkiewicz et al., 2018). Presuming that human and rabbit proteome changes in a similar fashion during erythropoiesis, NGD and/or RQC induced by ribosome collisions upon specific downregulation of ribosome recycling would interfere with cellular function and were thus abolished by loss of ZNF598.

Finally, the two major approaches explaining the regulatory role of the ribosome are not mutually exclusive, and a sophisticated interplay of both with the translational cis and trans regulators most likely shapes the cellular protein composition with all its consequences (Tahmasebi et al., 2018; Dalla Venezia et al., 2019).

Future perspectives

Beyond ribosomal complexes with a defined architecture, a concept for translation regulation in membrane-less organelles emerges as a mechanism for spatial and temporal control of protein biosynthesis (Protter and Parker, 2016; Aguilera-Gomez and Rabouille, 2017; Franzmann and Alberti, 2019). Biochemical and physicochemical processes underlying cytosolic heterogeneity and their biological consequences are not yet understood and remain one of the most exciting future perspectives. Thus, translation regulation, mRNA homeostasis and co-translational modification, translocation, and folding of proteins must be studied in the light of the diverse cytosolic formations such as P-bodies, stress granules, or Sec bodies (Zacharogianni et al., 2014; Acosta-Alvear et al., 2018; Van Treeck and Parker, 2018). As the composition, stoichiometry, and organization of membrane-less organelles are not well defined, structural studies of cellular ensembles by cryo-electron tomography in combination with high-resolution cryo-EM will obviously be most valuable. Additionally, membrane-less organelles are highly dynamic, which is why biophysical methods with spatial and temporal resolution are inevitable in this novel field of research. Strikingly, RNA-protein granules are linked to development, neurodegenerative disorder, and viral infection (Protter and Parker, 2016; Wolozin and Ivanov, 2019), emphasizing their physiological significance. In our current models, only mRNA and individual factors bind to isolated ribosomes and form defined complexes. In contrast to this, we tend to speculate that ribosomes create a molecular sociology around them and that factors therein compete with each other, thus establishing dynamic regulatory mechanisms including (anti-) synergistic effects. The molecular surrounding depends on structure, composition, and availability of the ribosomes, their localization, overall cellular condition, and many more yet elusive parameters. This model is in line with the hypotheses of the specialized ribosome (Xue and Barna, 2012) and ribosome abundance (Mills and Green, 2017) as well as the concept of molecular phase separation (Turoverov et al., 2019), and awaits validation or refutation by accurate in vivo experiments and visualization on single-molecule basis.

Award Identifier / Grant number: Nürenberg-Goloub

Award Identifier / Grant number: Nüsslein-Volhard/L’Oréal

Funding statement: We thank Inga Nold and Andrea Pott for editing and Simon Trowitzsch, Lukas Susac, Holger Heinemann, Stefan Brüchert, and Bianca Hetzert for helpful discussions. E.N.G. was supported by the Christiane Nüsslein-Volhard Foundation, L’Oréal, and the United Nations Educational, Scientific and Cultural Organization (UNESCO) (funder Id: http://dx.doi.org/10.13039/100005243), Grant Number: Nürenberg-Goloub (Nüsslein-Volhard/L’Oréal Grant). The German Research Foundation (DFG) SFB.902 ‘Molecular mechanisms of RNA-based regulation’ and the Cluster of Excellence EXC115 funded this work (to R.T.) (funder Id: http://dx.doi.org/10.13039/501100001659).

References

Acosta-Alvear, D., Karagoz, G.E., Frohlich, F., Li, H., Walther, T.C., and Walter, P. (2018). The unfolded protein response and endoplasmic reticulum protein targeting machineries converge on the stress sensor IRE1. eLife 7, e43036.10.7554/eLife.43036.025Search in Google Scholar

Afonina, Z.A. and Shirokov, V.A. (2018). Three-dimensional organization of polyribosomes – a modern approach. Biochemistry (Mosc) 83, S48–S55.10.1134/S0006297918140055Search in Google Scholar PubMed

Afonina, Z.A., Myasnikov, A.G., Shirokov, V.A., Klaholz, B.P., and Spirin, A.S. (2015). Conformation transitions of eukaryotic polyribosomes during multi-round translation. Nucleic Acids Res. 43, 618–628.10.1093/nar/gku1270Search in Google Scholar PubMed PubMed Central

Agrawal, R.K., Penczek, P., Grassucci, R.A., Li, Y., Leith, A., Nierhaus, K.H., and Frank, J. (1996). Direct visualization of A-, P-, and E-site transfer RNAs in the Escherichia coli ribosome. Science 271, 1000–1002.10.1142/9789813234864_0019Search in Google Scholar

Aguilera-Gomez, A. and Rabouille, C. (2017). Membrane-bound organelles versus membrane-less compartments and their control of anabolic pathways in Drosophila. Dev. Biol. 428, 310–317.10.1016/j.ydbio.2017.03.029Search in Google Scholar PubMed

Alhebshi, A., Sideri, T.C., Holland, S.L., and Avery, S.V. (2012). The essential iron-sulfur protein Rli1 is an important target accounting for inhibition of cell growth by reactive oxygen species. Mol. Biol. Cell. 23, 3582–3590.10.1091/mbc.e12-05-0413Search in Google Scholar

Alkalaeva, E.Z., Pisarev, A.V., Frolova, L.Y., Kisselev, L.L., and Pestova, T.V. (2006). In vitro reconstitution of eukaryotic translation reveals cooperativity between release factors eRF1 and eRF3. Cell 125, 1125–1136.10.1016/j.cell.2006.04.035Search in Google Scholar PubMed

Andersen, D.S. and Leevers, S.J. (2007). The essential Drosophila ATP-binding cassette domain protein, pixie, binds the 40 S ribosome in an ATP-dependent manner and is required for translation initiation. J. Biol. Chem. 282, 14752–14760.10.1074/jbc.M701361200Search in Google Scholar PubMed

Anderson, D.E., Pfeffermann, K., Kim, S.Y., Sawatsky, B., Pearson, J., Kovtun, M., Corcoran, D.L., Krebs, Y., Sigmundsson, K., Jamison, S.F., et al. (2019). Comparative loss-of-function screens reveal ABCE1 as an essential cellular host factor for efficient translation of Paramyxoviridae and Pneumoviridae. MBio 10, e00826–e00819.10.1128/mBio.00826-19Search in Google Scholar PubMed PubMed Central

Armistead, J., Khatkar, S., Meyer, B., Mark, B.L., Patel, N., Coghlan, G., Lamont, R.E., Liu, S., Wiechert, J., Cattini, P.A., et al. (2009). Mutation of a gene essential for ribosome biogenesis, EMG1, causes Bowen-Conradi syndrome. Am. J. Hum. Genet. 84, 728–739.10.1016/j.ajhg.2009.04.017Search in Google Scholar PubMed PubMed Central

Asano, K., Clayton, J., Shalev, A., and Hinnebusch, A.G. (2000). A multifactor complex of eukaryotic initiation factors, eIF1, eIF2, eIF3, eIF5, and initiator tRNA(Met) is an important translation initiation intermediate in vivo. Genes Dev. 14, 2534–2546.10.1101/gad.831800Search in Google Scholar PubMed PubMed Central

Barthelme, D., Scheele, U., Dinkelaker, S., Janoschka, A., Macmillan, F., Albers, S.V., Driessen, A.J., Stagni, M.S., Bill, E., Meyer-Klaucke, W., et al. (2007). Structural organization of essential iron-sulfur clusters in the evolutionarily highly conserved ATP-binding cassette protein ABCE1. J. Biol. Chem. 282, 14598–14607.10.1074/jbc.M700825200Search in Google Scholar PubMed

Barthelme, D., Dinkelaker, S., Albers, S.V., Londei, P., Ermler, U., and Tampé, R. (2011). Ribosome recycling depends on a mechanistic link between the FeS cluster domain and a conformational switch of the twin-ATPase ABCE1. Proc. Natl. Acad. Sci. USA 108, 3228–3233.10.1073/pnas.1015953108Search in Google Scholar PubMed PubMed Central

Basturea, G.N., Zundel, M.A., and Deutscher, M.P. (2011). Degradation of ribosomal RNA during starvation: comparison to quality control during steady-state growth and a role for RNase PH. RNA 17, 338–345.10.1261/rna.2448911Search in Google Scholar PubMed PubMed Central

Baum, S., Bittins, M., Frey, S., and Seedorf, M. (2004). Asc1p, a WD40-domain containing adaptor protein, is required for the interaction of the RNA-binding protein Scp160p with polysomes. Biochem. J. 380, 823–830.10.1042/bj20031962Search in Google Scholar PubMed PubMed Central

Becker, T., Armache, J.P., Jarasch, A., Anger, A.M., Villa, E., Sieber, H., Motaal, B.A., Mielke, T., Berninghausen, O., and Beckmann, R. (2011). Structure of the no-go mRNA decay complex Dom34-Hbs1 bound to a stalled 80S ribosome. Nat. Struct. Mol. Biol. 18, 715–720.10.1038/nsmb.2057Search in Google Scholar PubMed

Becker, T., Franckenberg, S., Wickles, S., Shoemaker, C.J., Anger, A.M., Armache, J.P., Sieber, H., Ungewickell, C., Berninghausen, O., Daberkow, I., et al. (2012). Structural basis of highly conserved ribosome recycling in eukaryotes and archaea. Nature 482, 501–506.10.1038/nature10829Search in Google Scholar PubMed PubMed Central

Behrmann, E., Loerke, J., Budkevich, T.V., Yamamoto, K., Schmidt, A., Penczek, P.A., Vos, M.R., Burger, J., Mielke, T., Scheerer, P., et al. (2015). Structural snapshots of actively translating human ribosomes. Cell 161, 845–857.10.1016/j.cell.2015.03.052Search in Google Scholar PubMed PubMed Central

Bellodi, C., Kopmar, N., and Ruggero, D. (2010). Deregulation of oncogene-induced senescence and p53 translational control in X-linked dyskeratosis congenita. EMBO J. 29, 1865–1876.10.1038/emboj.2010.83Search in Google Scholar PubMed PubMed Central

Beznoskova, P., Cuchalova, L., Wagner, S., Shoemaker, C.J., Gunisova, S., von der Haar, T., and Valasek, L.S. (2013). Translation initiation factors eIF3 and HCR1 control translation termination and stop codon read-through in yeast cells. PLoS Genet. 9, e1003962.10.1371/journal.pgen.1003962Search in Google Scholar PubMed PubMed Central

Bieri, P., Greber, B.J., and Ban, N. (2018). High-resolution structures of mitochondrial ribosomes and their functional implications. Curr. Opin. Struct. Biol. 49, 44–53.10.1016/j.sbi.2017.12.009Search in Google Scholar PubMed

Bisbal, C., Martinand, C., Silhol, M., Lebleu, B., and Salehzada, T. (1995). Cloning and characterization of a RNAse L inhibitor. A new component of the interferon-regulated 2–5A pathway. J. Biol. Chem. 270, 13308–13317.10.1074/jbc.270.22.13308Search in Google Scholar PubMed

Boria, I., Garelli, E., Gazda, H.T., Aspesi, A., Quarello, P., Pavesi, E., Ferrante, D., Meerpohl, J.J., Kartal, M., Da Costa, L., et al. (2010). The ribosomal basis of Diamond-Blackfan Anemia: mutation and database update. Hum. Mutat. 31, 1269–1279.10.1002/humu.21383Search in Google Scholar PubMed PubMed Central

Brandman, O. and Hegde, R.S. (2016). Ribosome-associated protein quality control. Nat. Struct. Mol. Biol. 23, 7–15.10.1038/nsmb.3147Search in Google Scholar PubMed PubMed Central

Brandt, F., Carlson, L.A., Hartl, F.U., Baumeister, W., and Grunewald, K. (2010). The three-dimensional organization of polyribosomes in intact human cells. Mol. Cell. 39, 560–569.10.1016/j.molcel.2010.08.003Search in Google Scholar PubMed

Brown, A., Shao, S., Murray, J., Hegde, R.S., and Ramakrishnan, V. (2015). Structural basis for stop codon recognition in eukaryotes. Nature 524, 493–496.10.1038/nature14896Search in Google Scholar PubMed PubMed Central

Brown, A., Baird, M.R., Yip, M.C., Murray, J., and Shao, S. (2018). Structures of translationally inactive mammalian ribosomes. eLife 7, e40486.10.7554/eLife.40486Search in Google Scholar PubMed PubMed Central

Buskirk, A.R. and Green, R. (2017). Ribosome pausing, arrest and rescue in bacteria and eukaryotes. Philos. Trans. R Soc. Lond. B Biol. Sci. 372, 20160183.10.1098/rstb.2016.0183Search in Google Scholar PubMed PubMed Central

Chang, W.Y. and Stanford, W.L. (2008). Translational control: a new dimension in embryonic stem cell network analysis. Cell Stem Cell 2, 410–412.10.1016/j.stem.2008.04.009Search in Google Scholar PubMed

Chen, Z.Q., Dong, J., Ishimura, A., Daar, I., Hinnebusch, A.G., and Dean, M. (2006). The essential vertebrate ABCE1 protein interacts with eukaryotic initiation factors. J. Biol. Chem. 281, 7452–7457.10.1074/jbc.M510603200Search in Google Scholar PubMed

Chen, L., Muhlrad, D., Hauryliuk, V., Cheng, Z., Lim, M.K., Shyp, V., Parker, R., and Song, H. (2010). Structure of the Dom34-Hbs1 complex and implications for no-go decay. Nat. Struct. Mol. Biol. 17, 1233–1240.10.1038/nsmb.1922Search in Google Scholar PubMed

Dalla Venezia, N., Vincent, A., Marcel, V., Catez, F., and Diaz, J.J. (2019). Emerging role of eukaryote ribosomes in translational control. Int. J. Mol. Sci. 20, E1226.10.3390/ijms20051226Search in Google Scholar PubMed PubMed Central

de la Cruz, J., Gomez-Herreros, F., Rodriguez-Galan, O., Begley, V., de la Cruz Munoz-Centeno, M., and Chavez, S. (2018). Feedback regulation of ribosome assembly. Curr. Genet. 64, 393–404.10.1007/s00294-017-0764-xSearch in Google Scholar PubMed

Delarue, M., Brittingham, G.P., Pfeffer, S., Surovtsev, I.V., Pinglay, S., Kennedy, K.J., Schaffer, M., Gutierrez, J.I., Sang, D., Poterewicz, G., et al. (2018). mTORC1 controls phase separation and the biophysical properties of the cytoplasm by tuning crowding. Cell 174, 338–349 e320.10.1016/j.cell.2018.05.042Search in Google Scholar PubMed

Dever, T.E. and Green, R. (2012). The elongation, termination, and recycling phases of translation in eukaryotes. Cold Spring Harb. Perspect. Biol. 4, a013706.10.1101/cshperspect.a013706Search in Google Scholar PubMed PubMed Central

Dinman, J.D. (2016). Pathways to specialized ribosomes: the Brussels lecture. J. Mol. Biol. 428, 2186–2194.10.1016/j.jmb.2015.12.021Search in Google Scholar PubMed PubMed Central

Doll, A. and Grzeschik, K.H. (2001). Characterization of two novel genes, WBSCR20 and WBSCR22, deleted in Williams-Beuren syndrome. Cytogenet. Cell Genet. 95, 20–27.10.1159/000057012Search in Google Scholar PubMed

Dong, J., Lai, R., Nielsen, K., Fekete, C.A., Qiu, H., and Hinnebusch, A.G. (2004). The essential ATP-binding cassette protein RLI1 functions in translation by promoting preinitiation complex assembly. J. Biol. Chem. 279, 42157–42168.10.1074/jbc.M404502200Search in Google Scholar PubMed

Dooher, J.E., Schneider, B.L., Reed, J.C., and Lingappa, J.R. (2007). Host ABCE1 is at plasma membrane HIV assembly sites and its dissociation from Gag is linked to subsequent events of virus production. Traffic 8, 195–211.10.1111/j.1600-0854.2006.00524.xSearch in Google Scholar PubMed PubMed Central

Draptchinskaia, N., Gustavsson, P., Andersson, B., Pettersson, M., Willig, T.N., Dianzani, I., Ball, S., Tchernia, G., Klar, J., Matsson, H., et al. (1999). The gene encoding ribosomal protein S19 is mutated in Diamond-Blackfan anaemia. Nat. Genet. 21, 169–175.10.1038/5951Search in Google Scholar PubMed

Ebert, B.L., Pretz, J., Bosco, J., Chang, C.Y., Tamayo, P., Galili, N., Raza, A., Root, D.E., Attar, E., Ellis, S.R., et al. (2008). Identification of RPS14 as a 5q- syndrome gene by RNA interference screen. Nature 451, 335–339.10.1038/nature06494Search in Google Scholar PubMed PubMed Central

Elantak, L., Wagner, S., Herrmannova, A., Karaskova, M., Rutkai, E., Lukavsky, P.J., and Valasek, L. (2010). The indispensable N-terminal half of eIF3j/HCR1 cooperates with its structurally conserved binding partner eIF3b/PRT1-RRM and with eIF1A in stringent AUG selection. J. Mol. Biol. 396, 1097–1116.10.1016/j.jmb.2009.12.047Search in Google Scholar PubMed PubMed Central

Emmott, E., Jovanovic, M., and Slavov, N. (2019). Ribosome stoichiometry: from form to function. Trends Biochem. Sci. 44, 95–109.10.1016/j.tibs.2018.10.009Search in Google Scholar PubMed PubMed Central

Ferretti, M.B. and Karbstein, K. (2019). Does functional specialization of ribosomes really exist? RNA 25, 521–538.10.1261/rna.069823.118Search in Google Scholar PubMed PubMed Central

Ferretti, M.B., Ghalei, H., Ward, E.A., Potts, E.L., and Karbstein, K. (2017). Rps26 directs mRNA-specific translation by recognition of Kozak sequence elements. Nat. Struct. Mol. Biol. 24, 700–707.10.1038/nsmb.3442Search in Google Scholar PubMed PubMed Central

Franzmann, T.M. and Alberti, S. (2019). Protein phase separation as a stress survival strategy. Cold Spring Harb. Perspect. Biol. 11, a034058.10.1101/cshperspect.a034058Search in Google Scholar PubMed PubMed Central

Frolova, L.Y., Tsivkovskii, R.Y., Sivolobova, G.F., Oparina, N.Y., Serpinsky, O.I., Blinov, V.M., Tatkov, S.I., and Kisselev, L.L. (1999). Mutations in the highly conserved GGQ motif of class 1 polypeptide release factors abolish ability of human eRF1 to trigger peptidyl-tRNA hydrolysis. RNA 5, 1014–1020.10.1017/S135583829999043XSearch in Google Scholar

Gallo, S. and Manfrini, N. (2015). Working hard at the nexus between cell signaling and the ribosomal machinery: an insight into the roles of RACK1 in translational regulation. Translation 3, e1120382.10.1080/21690731.2015.1120382Search in Google Scholar PubMed PubMed Central

Gao, F.B., Richter, J.D., and Cleveland, D.W. (2017). Rethinking unconventional translation in neurodegeneration. Cell 171, 994–1000.10.1016/j.cell.2017.10.042Search in Google Scholar PubMed PubMed Central

Gazda, H.T., Sheen, M.R., Vlachos, A., Choesmel, V., O’Donohue, M.F., Schneider, H., Darras, N., Hasman, C., Sieff, C.A., Newburger, P.E., et al. (2008). Ribosomal protein L5 and L11 mutations are associated with cleft palate and abnormal thumbs in Diamond-Blackfan anemia patients. Am. J. Hum. Genet. 83, 769–780.10.1016/j.ajhg.2008.11.004Search in Google Scholar PubMed PubMed Central

Gazda, H.T., Preti, M., Sheen, M.R., O’Donohue, M.F., Vlachos, A., Davies, S.M., Kattamis, A., Doherty, L., Landowski, M., Buros, C., et al. (2012). Frameshift mutation in p53 regulator RPL26 is associated with multiple physical abnormalities and a specific pre-ribosomal RNA processing defect in diamond-blackfan anemia. Hum. Mutat. 33, 1037–1044.10.1002/humu.22081Search in Google Scholar

Gerovac, M. and Tampé, R. (2019). Control of mRNA translation by versatile ATP-driven machines. Trends Biochem. Sci. 44, 167–180.10.1016/j.tibs.2018.11.003Search in Google Scholar

Gilbert, W.V. (2011). Functional specialization of ribosomes? Trends Biochem. Sci. 36, 127–132.10.1016/j.tibs.2010.12.002Search in Google Scholar

Gohara, D.W. and Yap, M.F. (2018). Survival of the drowsiest: the hibernating 100S ribosome in bacterial stress management. Curr. Genet. 64, 753–760.10.1007/s00294-017-0796-2Search in Google Scholar

Gouridis, G., Hetzert, B., Kiosze-Becker, K., de Boer, M., Heinemann, H., Nürenberg-Goloub, E., Cordes, T., and Tampe, R. (2019). ABCE1 controls ribosome recycling by an asymmetric dynamic conformational equilibrium. Cell Rep. 28, 723–734 e726.10.1016/j.celrep.2019.06.052Search in Google Scholar

Graber, T.E., Hebert-Seropian, S., Khoutorsky, A., David, A., Yewdell, J.W., Lacaille, J.C., and Sossin, W.S. (2013). Reactivation of stalled polyribosomes in synaptic plasticity. Proc. Natl. Acad. Sci. USA 110, 16205–16210.10.1073/pnas.1307747110Search in Google Scholar

Hayes, C.S. and Keiler, K.C. (2010). Beyond ribosome rescue: tmRNA and co-translational processes. FEBS Lett. 584, 413–419.10.1016/j.febslet.2009.11.023Search in Google Scholar

He, F., Li, X., Spatrick, P., Casillo, R., Dong, S., and Jacobson, A. (2003). Genome-wide analysis of mRNAs regulated by the nonsense-mediated and 5′ to 3′ mRNA decay pathways in yeast. Mol. Cell 12, 1439–1452.10.1016/S1097-2765(03)00446-5Search in Google Scholar

Hellen, C.U.T. (2018). Translation termination and ribosome recycling in eukaryotes. Cold Spring Harb. Perspect. Biol. 10, a032656.10.1101/cshperspect.a032656Search in Google Scholar PubMed PubMed Central

Hershey, J.W., Sonenberg, N., and Mathews, M.B. (2012). Principles of translational control: an overview. Cold Spring Harb. Perspect. Biol. 4, a011528.10.1101/cshperspect.a011528Search in Google Scholar PubMed PubMed Central

Heuer, A., Gerovac, M., Schmidt, C., Trowitzsch, S., Preis, A., Kotter, P., Berninghausen, O., Becker, T., Beckmann, R., and Tampé, R. (2017). Structure of the 40S-ABCE1 post-splitting complex in ribosome recycling and translation initiation. Nat. Struct. Mol. Biol. 24, 453–460.10.1038/nsmb.3396Search in Google Scholar PubMed

Hinnebusch, A.G. (2017). Structural insights into the mechanism of scanning and start codon recognition in eukaryotic translation initiation. Trends Biochem. Sci. 42, 589–611.10.1016/j.tibs.2017.03.004Search in Google Scholar PubMed

Holcik, M. and Sonenberg, N. (2005). Translational control in stress and apoptosis. Nat. Rev. Mol. Cell Biol. 6, 318–327.10.1038/nrm1618Search in Google Scholar PubMed

Hopfner, K.P. (2016). Invited review: architectures and mechanisms of ATP binding cassette proteins. Biopolymers 105, 492–504.10.1002/bip.22843Search in Google Scholar PubMed

Hronova, V., Mohammad, M.P., Wagner, S., Panek, J., Gunisova, S., Zeman, J., Poncova, K., and Valasek, L.S. (2017). Does eIF3 promote reinitiation after translation of short upstream ORFs also in mammalian cells? RNA Biol. 14, 1660–1667.10.1080/15476286.2017.1353863Search in Google Scholar PubMed PubMed Central

Huang, B., Gao, Y., Tian, D., and Zheng, M. (2010). A small interfering ABCE1-targeting RNA inhibits the proliferation and invasiveness of small cell lung cancer. Int. J. Mol. Med. 25, 687–693.10.3892/ijmm_00000392Search in Google Scholar

Huter, P., Muller, C., Arenz, S., Beckert, B., and Wilson, D.N. (2017). Structural basis for ribosome rescue in bacteria. Trends Biochem. Sci. 42, 669–680.10.1016/j.tibs.2017.05.009Search in Google Scholar PubMed

Ikeuchi, K. and Inada, T. (2016). Ribosome-associated Asc1/RACK1 is required for endonucleolytic cleavage induced by stalled ribosome at the 3′ end of nonstop mRNA. Sci. Rep. 6, 28234.10.1038/srep28234Search in Google Scholar PubMed PubMed Central

Ikeuchi, K., Tesina, P., Matsuo, Y., Sugiyama, T., Cheng, J., Saeki, Y., Tanaka, K., Becker, T., Beckmann, R., and Inada, T. (2019). Collided ribosomes form a unique structural interface to induce Hel2-driven quality control pathways. EMBO J. 38, e100276.10.15252/embj.2018100276Search in Google Scholar PubMed PubMed Central

Imai, H., Abe, T., Miyoshi, T., Nishikawa, S.I., Ito, K., and Uchiumi, T. (2018). The ribosomal stalk protein is crucial for the action of the conserved ATPase ABCE1. Nucleic Acids Res. 46, 7820–7830.10.1093/nar/gky619Search in Google Scholar PubMed PubMed Central

Imami, K., Milek, M., Bogdanow, B., Yasuda, T., Kastelic, N., Zauber, H., Ishihama, Y., Landthaler, M., and Selbach, M. (2018). Phosphorylation of the ribosomal protein RPL12/uL11 affects translation during mitosis. Mol. Cell. 72, 84–98 e89.10.1016/j.molcel.2018.08.019Search in Google Scholar PubMed

Iwawaki, T., Hosoda, A., Okuda, T., Kamigori, Y., Nomura-Furuwatari, C., Kimata, Y., Tsuru, A., and Kohno, K. (2001). Translational control by the ER transmembrane kinase/ribonuclease IRE1 under ER stress. Nat. Cell Biol. 3, 158–164.10.1038/35055065Search in Google Scholar PubMed

Jaffrey, S.R. and Wilkinson, M.F. (2018). Nonsense-mediated RNA decay in the brain: emerging modulator of neural development and disease. Nat. Rev. Neurosci. 19, 715–728.10.1038/s41583-018-0079-zSearch in Google Scholar PubMed PubMed Central

Jamar, N.H., Kritsiligkou, P., and Grant, C.M. (2017). The non-stop decay mRNA surveillance pathway is required for oxidative stress tolerance. Nucleic Acids Res. 45, 6881–6893.10.1093/nar/gkx306Search in Google Scholar PubMed PubMed Central

Jenner, L., Demeshkina, N., Yusupova, G., and Yusupov, M. (2010). Structural rearrangements of the ribosome at the tRNA proofreading step. Nat. Struct. Mol. Biol. 17, 1072–1078.10.1038/nsmb.1880Search in Google Scholar PubMed

Joazeiro, C.A.P. (2019). Mechanisms and functions of ribosome-associated protein quality control. Nat. Rev. Mol. Cell Biol. 20, 368–383.10.1038/s41580-019-0118-2Search in Google Scholar PubMed PubMed Central

Jobe, A., Liu, Z., Gutierrez-Vargas, C., and Frank, J. (2019). New insights into ribosome structure and function. Cold Spring Harb. Perspect. Biol. 11, a032615.10.1101/cshperspect.a032615Search in Google Scholar PubMed PubMed Central

Juszkiewicz, S., Chandrasekaran, V., Lin, Z., Kraatz, S., Ramakrishnan, V., and Hegde, R.S. (2018). ZNF598 is a quality control sensor of collided ribosomes. Mol. Cell 72, 469–481 e467.10.1016/j.molcel.2018.08.037Search in Google Scholar PubMed PubMed Central

Karcher, A., Buttner, K., Martens, B., Jansen, R.P., and Hopfner, K.P. (2005). X-ray structure of RLI, an essential twin cassette ABC ATPase involved in ribosome biogenesis and HIV capsid assembly. Structure 13, 649–659.10.1016/j.str.2005.02.008Search in Google Scholar PubMed

Karcher, A., Schele, A., and Hopfner, K.P. (2008). X-ray structure of the complete ABC enzyme ABCE1 from Pyrococcus abyssi. J. Biol. Chem. 283, 7962–7971.10.1074/jbc.M707347200Search in Google Scholar PubMed

Karousis, E.D. and Muhlemann, O. (2019). Nonsense-mediated mRNA decay begins where translation ends. Cold Spring Harb. Perspect. Biol. 11, a032862.10.1101/cshperspect.a032862Search in Google Scholar PubMed PubMed Central

Keiler, K.C. (2008). Biology of trans-translation. Annu. Rev. Microbiol. 62, 133–151.10.1146/annurev.micro.62.081307.162948Search in Google Scholar PubMed

Keiler, K.C. (2015). Mechanisms of ribosome rescue in bacteria. Nat. Rev. Microbiol. 13, 285–297.10.1038/nrmicro3438Search in Google Scholar PubMed

Keiler, K.C., Waller, P.R., and Sauer, R.T. (1996). Role of a peptide tagging system in degradation of proteins synthesized from damaged messenger RNA. Science 271, 990–993.10.1126/science.271.5251.990Search in Google Scholar PubMed

Khajuria, R.K., Munschauer, M., Ulirsch, J.C., Fiorini, C., Ludwig, L.S., McFarland, S.K., Abdulhay, N.J., Specht, H., Keshishian, H., Mani, D.R., et al. (2018). Ribosome levels selectively regulate translation and lineage commitment in human hematopoiesis. Cell 173, 90–103 e119.10.1016/j.cell.2018.02.036Search in Google Scholar PubMed PubMed Central

Kiosze-Becker, K., Ori, A., Gerovac, M., Heuer, A., Nürenberg-Goloub, E., Rashid, U.J., Becker, T., Beckmann, R., Beck, M., and Tampé, R. (2016). Structure of the ribosome post-recycling complex probed by chemical cross-linking and mass spectrometry. Nat. Commun. 7, 13248.10.1038/ncomms13248Search in Google Scholar PubMed PubMed Central

Knorr, A.G., Schmidt, C., Tesina, P., Berninghausen, O., Becker, T., Beatrix, B., and Beckmann, R. (2019). Ribosome-NatA architecture reveals that rRNA expansion segments coordinate N-terminal acetylation. Nat. Struct. Mol. Biol. 26, 35–39.10.1038/s41594-018-0165-ySearch in Google Scholar PubMed

Kobayashi, K., Kikuno, I., Kuroha, K., Saito, K., Ito, K., Ishitani, R., Inada, T., and Nureki, O. (2010). Structural basis for mRNA surveillance by archaeal Pelota and GTP-bound EF1alpha complex. Proc. Natl. Acad. Sci. USA 107, 17575–17579.10.1073/pnas.1009598107Search in Google Scholar PubMed PubMed Central

Kobayashi, K., Saito, K., Ishitani, R., Ito, K., and Nureki, O. (2012). Structural basis for translation termination by archaeal RF1 and GTP-bound EF1α complex. Nucleic Acids Res. 40, 9319–9328.10.1093/nar/gks660Search in Google Scholar PubMed PubMed Central

Komar, A.A. (2018). Unraveling co-translational protein folding: concepts and methods. Methods 137, 71–81.10.1016/j.ymeth.2017.11.007Search in Google Scholar PubMed PubMed Central

Kondrashov, N., Pusic, A., Stumpf, C.R., Shimizu, K., Hsieh, A.C., Ishijima, J., Shiroishi, T., and Barna, M. (2011). Ribosome-mediated specificity in Hox mRNA translation and vertebrate tissue patterning. Cell 145, 383–397.10.1016/j.cell.2011.03.028Search in Google Scholar PubMed PubMed Central

Kopeina, G.S., Afonina, Z.A., Gromova, K.V., Shirokov, V.A., Vasiliev, V.D., and Spirin, A.S. (2008). Step-wise formation of eukaryotic double-row polyribosomes and circular translation of polysomal mRNA. Nucleic Acids Res. 36, 2476–2488.10.1093/nar/gkm1177Search in Google Scholar PubMed PubMed Central

Kozak, M. (2005). Regulation of translation via mRNA structure in prokaryotes and eukaryotes. Gene 361, 13–37.10.1016/j.gene.2005.06.037Search in Google Scholar PubMed

Kramer, G., Shiber, A., and Bukau, B. (2019). Mechanisms of cotranslational maturation of newly synthesized proteins. Annu. Rev. Biochem. 88, 337–364.10.1146/annurev-biochem-013118-111717Search in Google Scholar PubMed

Kressler, D., Hurt, E., and Bassler, J. (2017). A puzzle of life: crafting ribosomal subunits. Trends Biochem. Sci. 42, 640–654.10.1016/j.tibs.2017.05.005Search in Google Scholar PubMed

Kurosaki, T., Popp, M.W., and Maquat, L.E. (2019). Quality and quantity control of gene expression by nonsense-mediated mRNA decay. Nat. Rev. Mol. Cell Biol. 20, 406–420.10.1038/s41580-019-0126-2Search in Google Scholar PubMed PubMed Central

Leppek, K., Das, R., and Barna, M. (2018). Functional 5′ UTR mRNA structures in eukaryotic translation regulation and how to find them. Nat. Rev. Mol. Cell Biol. 19, 158–174.10.1038/nrm.2017.103Search in Google Scholar PubMed PubMed Central

Lingappa, U.F., Wu, X., Macieik, A., Yu, S.F., Atuegbu, A., Corpuz, M., Francis, J., Nichols, C., Calayag, A., Shi, H., et al. (2013). Host-rabies virus protein-protein interactions as druggable antiviral targets. Proc. Natl. Acad. Sci. USA 110, E861–E868.10.1073/pnas.1210198110Search in Google Scholar PubMed PubMed Central

Liu, Q. and Fredrick, K. (2016). Intersubunit bridges of the bacterial ribosome. J. Mol. Biol. 428, 2146–2164.10.1016/j.jmb.2016.02.009Search in Google Scholar PubMed PubMed Central

Lodish, H.F. (1974). Model for the regulation of mRNA translation applied to haemoglobin synthesis. Nature 251, 385–388.10.1038/251385a0Search in Google Scholar PubMed

Lomakin, I.B., Stolboushkina, E.A., Vaidya, A.T., Zhao, C., Garber, M.B., Dmitriev, S.E., and Steitz, T.A. (2017). Crystal structure of the human ribosome in complex with DENR-MCT-1. Cell Rep. 20, 521–528.10.1016/j.celrep.2017.06.025Search in Google Scholar PubMed PubMed Central

Lytvynenko, I., Paternoga, H., Thrun, A., Balke, A., Muller, T.A., Chiang, C.H., Nagler, K., Tsaprailis, G., Anders, S., Bischofs, I., et al. (2019). Alanine tails signal proteolysis in bacterial ribosome-associated quality control. Cell 178, 76–90 e22.10.1016/j.cell.2019.05.002Search in Google Scholar PubMed PubMed Central

Mancera-Martinez, E., Brito Querido, J., Valasek, L.S., Simonetti, A., and Hashem, Y. (2017). ABCE1: a special factor that orchestrates translation at the crossroad between recycling and initiation. RNA Biol. 14, 1279–1285.10.1080/15476286.2016.1269993Search in Google Scholar PubMed PubMed Central

Martin, I., Kim, J.W., Lee, B.D., Kang, H.C., Xu, J.C., Jia, H., Stankowski, J., Kim, M.S., Zhong, J., Kumar, M., et al. (2014). Ribosomal protein s15 phosphorylation mediates LRRK2 neurodegeneration in Parkinson’s disease. Cell 157, 472–485.10.1016/j.cell.2014.01.064Search in Google Scholar PubMed PubMed Central

Marygold, S.J., Roote, J., Reuter, G., Lambertsson, A., Ashburner, M., Millburn, G.H., Harrison, P.M., Yu, Z., Kenmochi, N., Kaufman, T.C., et al. (2007). The ribosomal protein genes and minute loci of Drosophila melanogaster. Genome Biol. 8, R216.10.1186/gb-2007-8-10-r216Search in Google Scholar PubMed PubMed Central

Mills, E.W. and Green, R. (2017). Ribosomopathies: there’s strength in numbers. Science 358, eaan2755.10.1126/science.aan2755Search in Google Scholar PubMed

Mills, E.W., Wangen, J., Green, R., and Ingolia, N.T. (2016). Dynamic regulation of a ribosome rescue pathway in erythroid cells and platelets. Cell Rep. 17, 1–10.10.1016/j.celrep.2016.08.088Search in Google Scholar PubMed PubMed Central

Muhs, M., Hilal, T., Mielke, T., Skabkin, M.A., Sanbonmatsu, K.Y., Pestova, T.V., and Spahn, C.M. (2015). Cryo-EM of ribosomal 80S complexes with termination factors reveals the translocated cricket paralysis virus IRES. Mol. Cell 57, 422–432.10.1016/j.molcel.2014.12.016Search in Google Scholar PubMed PubMed Central

Munzarova, V., Panek, J., Gunisova, S., Danyi, I., Szamecz, B., and Valasek, L.S. (2011). Translation reinitiation relies on the interaction between eIF3a/TIF32 and progressively folded cis-acting mRNA elements preceding short uORFs. PLoS Genet. 7, e1002137.10.1371/journal.pgen.1002137Search in Google Scholar PubMed PubMed Central

Myasnikov, A.G., Afonina, Z.A., Menetret, J.F., Shirokov, V.A., Spirin, A.S., and Klaholz, B.P. (2014). The molecular structure of the left-handed supra-molecular helix of eukaryotic polyribosomes. Nat. Commun. 5, 5294.10.1038/ncomms6294Search in Google Scholar PubMed

Narla, A. and Ebert, B.L. (2010). Ribosomopathies: human disorders of ribosome dysfunction. Blood 115, 3196–3205.10.1182/blood-2009-10-178129Search in Google Scholar PubMed PubMed Central

Natchiar, S.K., Myasnikov, A.G., Kratzat, H., Hazemann, I., andKlaholz, B.P. (2017). Visualization of chemical modifications in the human 80S ribosome structure. Nature 551, 472–477.10.1038/nature24482Search in Google Scholar

Nilsson, O.B., Hedman, R., Marino, J., Wickles, S., Bischoff, L., Johansson, M., Muller-Lucks, A., Trovato, F., Puglisi, J.D., O’Brien, E.P., et al. (2015). Cotranslational protein folding inside the ribosome exit tunnel. Cell Rep. 12, 1533–1540.10.1016/j.celrep.2015.07.065Search in Google Scholar

Nürenberg, E. and Tampé, R. (2013). Tying up loose ends: ribosome recycling in eukaryotes and archaea. Trends Biochem. Sci. 38, 64–74.10.1016/j.tibs.2012.11.003Search in Google Scholar

Nürenberg-Goloub, E., Heinemann, H., Gerovac, M., and Tampé, R. (2018). Ribosome recycling is coordinated by processive events in two asymmetric ATP sites of ABCE1. Life Sci. Alliance 1, e201800095.10.26508/lsa.201800095Search in Google Scholar

Ogle, J.M., Murphy, F.V., Tarry, M.J., and Ramakrishnan, V. (2002). Selection of tRNA by the ribosome requires a transition from an open to a closed form. Cell 111, 721–732.10.1016/S0092-8674(02)01086-3Search in Google Scholar

Parks, M.M., Kurylo, C.M., Dass, R.A., Bojmar, L., Lyden, D., Vincent, C.T., and Blanchard, S.C. (2018). Variant ribosomal RNA alleles are conserved and exhibit tissue-specific expression. Sci. Adv. 4, eaao0665.10.1126/sciadv.aao0665Search in Google Scholar PubMed PubMed Central

Pena, C., Hurt, E., and Panse, V.G. (2017). Eukaryotic ribosome assembly, transport and quality control. Nat. Struct. Mol. Biol. 24, 689–699.10.1038/nsmb.3454Search in Google Scholar PubMed

Pisarev, A.V., Skabkin, M.A., Pisareva, V.P., Skabkina, O.V., Rakotondrafara, A.M., Hentze, M.W., Hellen, C.U.T., and Pestova, T.V. (2010). The role of ABCE1 in eukaryotic posttermination ribosomal recycling. Mol. Cell 37, 196–210.10.1016/j.molcel.2009.12.034Search in Google Scholar PubMed PubMed Central

Polikanov, Y.S., Melnikov, S.V., Soll, D., and Steitz, T.A. (2015). Structural insights into the role of rRNA modifications in protein synthesis and ribosome assembly. Nat. Struct. Mol. Biol. 22, 342–344.10.1038/nsmb.2992Search in Google Scholar PubMed PubMed Central

Prabhakar, A., Choi, J., Wang, J., Petrov, A., and Puglisi, J.D. (2017). Dynamic basis of fidelity and speed in translation: coordinated multistep mechanisms of elongation and termination. Protein Sci. 26, 1352–1362.10.1002/pro.3190Search in Google Scholar PubMed PubMed Central

Protter, D.S.W. and Parker, R. (2016). Principles and properties of stress granules. Trends Cell Biol. 26, 668–679.10.1016/j.tcb.2016.05.004Search in Google Scholar PubMed PubMed Central

Qin, D., Liu, Q., Devaraj, A., and Fredrick, K. (2012). Role of helix 44 of 16S rRNA in the fidelity of translation initiation. RNA 18, 485–495.10.1261/rna.031203.111Search in Google Scholar PubMed PubMed Central

Rheinberger, H.J., Sternbach, H., and Nierhaus, K.H. (1981). Three tRNA binding sites on Escherichia coli ribosomes. Proc. Natl. Acad. Sci. USA 78, 5310–5314.10.1073/pnas.78.9.5310Search in Google Scholar PubMed PubMed Central

Robichaud, N., Sonenberg, N., Ruggero, D., and Schneider, R.J. (2018). Translational Control in Cancer. Cold Spring Harb. Perspect. Biol. 11, a032896.10.1101/cshperspect.a032896Search in Google Scholar PubMed PubMed Central

Rodnina, M.V. (2018). Translation in prokaryotes. Cold Spring Harb. Perspect. Biol. 10, a032664.10.1101/cshperspect.a032664Search in Google Scholar PubMed PubMed Central

Roundtree, I.A., Evans, M.E., Pan, T., and He, C. (2017). Dynamic RNA modifications in gene expression regulation. Cell 169, 1187–1200.10.1016/j.cell.2017.05.045Search in Google Scholar PubMed PubMed Central

Roy, B., Vaughn, J.N., Kim, B.H., Zhou, F., Gilchrist, M.A., and Von Arnim, A.G. (2010). The h subunit of eIF3 promotes reinitiation competence during translation of mRNAs harboring upstream open reading frames. RNA 16, 748–761.10.1261/rna.2056010Search in Google Scholar PubMed PubMed Central

Saito, S., Hosoda, N., and Hoshino, S. (2013). The Hbs1-Dom34 protein complex functions in non-stop mRNA decay in mammalian cells. J. Biol. Chem. 288, 17832–17843.10.1074/jbc.M112.448977Search in Google Scholar PubMed PubMed Central

Schosserer, M., Minois, N., Angerer, T.B., Amring, M., Dellago, H., Harreither, E., Calle-Perez, A., Pircher, A., Gerstl, M.P., Pfeifenberger, S., et al. (2015). Methylation of ribosomal RNA by NSUN5 is a conserved mechanism modulating organismal lifespan. Nat. Commun. 6, 6158.10.1038/ncomms7158Search in Google Scholar PubMed PubMed Central

Shao, S., Murray, J., Brown, A., Taunton, J., Ramakrishnan, V., and Hegde, R.S. (2016). Decoding mammalian ribosome-mRNA states by translational GTPase complexes. Cell 167, 1229–1240 e1215.10.1016/j.cell.2016.10.046Search in Google Scholar PubMed PubMed Central

Sharma, S. and Lafontaine, D.L.J. (2015). ‘View from a bridge’: a new perspective on eukaryotic rRNA base modification. Trends Biochem. Sci. 40, 560–575.10.1016/j.tibs.2015.07.008Search in Google Scholar PubMed

Sharma, S., Hartmann, J.D., Watzinger, P., Klepper, A., Peifer, C., Kotter, P., Lafontaine, D.L.J., and Entian, K.D. (2018). A single N1-methyladenosine on the large ribosomal subunit rRNA impacts locally its structure and the translation of key metabolic enzymes. Sci. Rep. 8, 11904.10.1038/s41598-018-30383-zSearch in Google Scholar PubMed PubMed Central

Shi, Z. and Barna, M. (2015). Translating the genome in time and space: specialized ribosomes, RNA regulons, and RNA-binding proteins. Annu. Rev. Cell Dev. Biol. 31, 31–54.10.1146/annurev-cellbio-100814-125346Search in Google Scholar PubMed

Shi, Z., Fujii, K., Kovary, K.M., Genuth, N.R., Rost, H.L., Teruel, M.N., and Barna, M. (2017). Heterogeneous ribosomes preferentially translate distinct subpools of mRNAs genome-wide. Mol. Cell 67, 71–83 e77.10.1016/j.molcel.2017.05.021Search in Google Scholar PubMed PubMed Central

Shirokikh, N.E. and Preiss, T. (2018). Translation initiation by cap-dependent ribosome recruitment: recent insights and open questions. Wiley Interdiscip. Rev. RNA 9, e1473.10.1002/wrna.1473Search in Google Scholar PubMed

Shoemaker, C.J. and Green, R. (2011). Kinetic analysis reveals the ordered coupling of translation termination and ribosome recycling in yeast. Proc. Natl. Acad. Sci. USA 108, E1392–E1398.10.1073/pnas.1113956108Search in Google Scholar PubMed PubMed Central

Simms, C.L., Yan, L.L., and Zaher, H.S. (2017). Ribosome collision is critical for quality control during no-go decay. Mol. Cell 68, 361–373 e365.10.1016/j.molcel.2017.08.019Search in Google Scholar PubMed PubMed Central

Simsek, D., Tiu, G.C., Flynn, R.A., Byeon, G.W., Leppek, K., Xu, A.F., Chang, H.Y., and Barna, M. (2017). The mammalian ribo-interactome reveals ribosome functional diversity and heterogeneity. Cell 169, 1051–1065 e1018.10.1016/j.cell.2017.05.022Search in Google Scholar PubMed PubMed Central

Sitron, C.S., Park, J.H., and Brandman, O. (2017). Asc1, Hel2, and Slh1 couple translation arrest to nascent chain degradation. RNA 23, 798–810.10.1261/rna.060897.117Search in Google Scholar PubMed PubMed Central

Skabkin, M.A., Skabkina, O.V., Dhote, V., Komar, A.A., Hellen, C.U.T., and Pestova, T.V. (2010). Activities of ligatin and MCT-1/DENR in eukaryotic translation initiation and ribosomal recycling. Genes Dev. 24, 1787–1801.10.1101/gad.1957510Search in Google Scholar PubMed PubMed Central

Skabkin, M.A., Skabkina, O.V., Hellen, C.U.T., and Pestova, T.V. (2013). Reinitiation and other unconventional posttermination events during eukaryotic translation. Mol. Cell 51, 249–264.10.1016/j.molcel.2013.05.026Search in Google Scholar

Slavov, N., Semrau, S., Airoldi, E., Budnik, B., and van Oudenaarden, A. (2015). Differential stoichiometry among core ribosomal proteins. Cell Rep. 13, 865–873.10.1016/j.celrep.2015.09.056Search in Google Scholar

Sloan, K.E., Warda, A.S., Sharma, S., Entian, K.D., Lafontaine, D.L.J., and Bohnsack, M.T. (2017). Tuning the ribosome: the influence of rRNA modification on eukaryotic ribosome biogenesis and function. RNA Biol. 14, 1138–1152.10.1080/15476286.2016.1259781Search in Google Scholar

Slobodin, B., Han, R., Calderone, V., Vrielink, J., Loayza-Puch, F., Elkon, R., and Agami, R. (2017). Transcription impacts the efficiency of mRNA translation via co-transcriptional N6-adenosine methylation. Cell 169, 326–337 e312.10.1016/j.cell.2017.03.031Search in Google Scholar

Sonenberg, N. and Hinnebusch, A.G. (2009). Regulation of translation initiation in eukaryotes: mechanisms and biological targets. Cell 136, 731–745.10.1016/j.cell.2009.01.042Search in Google Scholar

Song, H., Mugnier, P., Das, A.K., Webb, H.M., Evans, D.R., Tuite, M.F., Hemmings, B.A., and Barford, D. (2000). The crystal structure of human eukaryotic release factor eRF1 – mechanism of stop codon recognition and peptidyl-tRNA hydrolysis. Cell 100, 311–321.10.1016/S0092-8674(00)80667-4Search in Google Scholar

Sossin, W.S. and Costa-Mattioli, M. (2019). Translational control in the brain in health and disease. Cold Spring Harb. Perspect. Biol. 11, a032912.10.1101/cshperspect.a032912Search in Google Scholar PubMed PubMed Central

Sriram, A., Bohlen, J., and Teleman, A.A. (2018). Translation acrobatics: how cancer cells exploit alternate modes of translational initiation. EMBO Rep. 19, e45947.10.15252/embr.201845947Search in Google Scholar PubMed PubMed Central

Stark, H., Rodnina, M.V., Rinke-Appel, J., Brimacombe, R., Wintermeyer, W., and van Heel, M. (1997). Visualization of elongation factor Tu on the Escherichia coli ribosome. Nature 389, 403–406.10.1038/38770Search in Google Scholar PubMed

Stein, K.C. and Frydman, J. (2019). The stop-and-go traffic regulating protein biogenesis: how translation kinetics controls proteostasis. J. Biol. Chem. 294, 2076–2084.10.1074/jbc.REV118.002814Search in Google Scholar PubMed PubMed Central

Sundaramoorthy, E., Leonard, M., Mak, R., Liao, J., Fulzele, A., and Bennett, E.J. (2017). ZNF598 and RACK1 regulate mammalian ribosome-associated quality control function by mediating regulatory 40S ribosomal ubiquitylation. Mol. Cell 65, 751–760 e754.10.1016/j.molcel.2016.12.026Search in Google Scholar PubMed PubMed Central

Tahmasebi, S., Khoutorsky, A., Mathews, M.B., and Sonenberg, N. (2018). Translation deregulation in human disease. Nat. Rev. Mol. Cell Biol. 19, 791–807.10.1038/s41580-018-0034-xSearch in Google Scholar PubMed

Thomas, C. and Tampé, R. (2018). Multifaceted structures and mechanisms of ABC transport systems in health and disease. Curr. Opin. Struct. Biol. 51, 116–128.10.1016/j.sbi.2018.03.016Search in Google Scholar PubMed

Truitt, M.L. and Ruggero, D. (2017). New frontiers in translational control of the cancer genome. Nat. Rev. Cancer 17, 332.10.1038/nrc.2017.30Search in Google Scholar PubMed

Turoverov, K.K., Kuznetsova, I.M., Fonin, A.V., Darling, A.L., Zaslavsky, B.Y., and Uversky, V.N. (2019). Stochasticity of biological soft matter: emerging concepts in intrinsically disordered proteins and biological phase separation. Trends Biochem. Sci. 44, 716–728.10.1016/j.tibs.2019.03.005Search in Google Scholar PubMed

Van Treeck, B. and Parker, R. (2018). Emerging roles for intermolecular RNA-RNA interactions in RNP assemblies. Cell 174, 791–802.10.1016/j.cell.2018.07.023Search in Google Scholar PubMed PubMed Central

Vogel, C. and Marcotte, E.M. (2012). Insights into the regulation of protein abundance from proteomic and transcriptomic analyses. Nat. Rev. Genet. 13, 227–232.10.1038/nrg3185Search in Google Scholar PubMed PubMed Central

Voorhees, R.M. and Ramakrishnan, V. (2013). Structural basis of the translational elongation cycle. Annu. Rev. Biochem. 82, 203–236.10.1146/annurev-biochem-113009-092313Search in Google Scholar PubMed

Waudby, C.A., Dobson, C.M., and Christodoulou, J. (2019). Nature and regulation of protein folding on the ribosome. Trends Biochem. Sci. 44, 914–926.10.1016/j.tibs.2019.06.008Search in Google Scholar PubMed PubMed Central

Weisser, M., Schafer, T., Leibundgut, M., Bohringer, D., Aylett, C.H.S., and Ban, N. (2017). Structural and functional insights into human re-initiation complexes. Mol. Cell 67, 447–456 e447.10.1016/j.molcel.2017.06.032Search in Google Scholar PubMed

Wethmar, K. (2014). The regulatory potential of upstream open reading frames in eukaryotic gene expression. Wiley Interdiscip. Rev. RNA 5, 765–778.10.1002/wrna.1245Search in Google Scholar PubMed

Whittle, C.A. and Krochko, J.E. (2009). Transcript profiling provides evidence of functional divergence and expression networks among ribosomal protein gene paralogs in Brassica napus. Plant Cell 21, 2203–2219.10.1105/tpc.109.068411Search in Google Scholar PubMed PubMed Central

Williams, M.E. and Sussex, I.M. (1995). Developmental regulation of ribosomal protein L16 genes in Arabidopsis thaliana. Plant J. 8, 65–76.10.1046/j.1365-313X.1995.08010065.xSearch in Google Scholar

Wolozin, B. and Ivanov, P. (2019). Stress granules and neurodegeneration. Nat. Rev. Neurosci. 20, 649–666.10.1038/s41583-019-0222-5Search in Google Scholar PubMed PubMed Central

Wreschner, D.H., James, T.C., Silverman, R.H., and Kerr, I.M. (1981). Ribosomal RNA cleavage, nuclease activation and 2-5A(ppp(A2′p)nA) in interferon-treated cells. Nucleic Acids Res. 9, 1571–1581.10.1093/nar/9.7.1571Search in Google Scholar PubMed PubMed Central

Wu, Z., Wang, Y., Lim, J., Liu, B., Li, Y., Vartak, R., Stankiewicz, T., Montgomery, S., and Lu, B. (2018). Ubiquitination of ABCE1 by NOT4 in response to mitochondrial damage links co-translational quality control to PINK1-directed mitophagy. Cell Metab. 28, 130–144 e137.10.1016/j.cmet.2018.05.007Search in Google Scholar PubMed PubMed Central

Xue, S. and Barna, M. (2012). Specialized ribosomes: a new frontier in gene regulation and organismal biology. Nat. Rev. Mol. Cell Biol. 13, 355–369.10.1038/nrm3359Search in Google Scholar PubMed PubMed Central

Yamada, S.B., Gendron, T.F., Niccoli, T., Genuth, N.R., Grosely, R., Shi, Y., Glaria, I., Kramer, N.J., Nakayama, L., Fang, S., et al. (2019). RPS25 is required for efficient RAN translation of C9orf72 and other neurodegenerative disease-associated nucleotide repeats. Nat. Neurosci. 22, 1383–1388.10.1038/s41593-019-0455-7Search in Google Scholar PubMed PubMed Central

Yap, K. and Makeyev, E.V. (2013). Regulation of gene expression in mammalian nervous system through alternative pre-mRNA splicing coupled with RNA quality control mechanisms. Mol. Cell. Neurosci. 56, 420–428.10.1016/j.mcn.2013.01.003Search in Google Scholar PubMed

Yarunin, A., Panse, V.G., Petfalski, E., Dez, C., Tollervey, D., and Hurt, E.C. (2005). Functional link between ribosome formation and biogenesis of iron-sulfur proteins. EMBO J. 24, 580–588.10.1038/sj.emboj.7600540Search in Google Scholar PubMed PubMed Central

Young, D.J. and Guydosh, N.R. (2019). Hcr1/eIF3j is a 60S ribosomal subunit recycling accessory factor in vivo. Cell Rep. 28, 39–50 e34.10.1016/j.celrep.2019.05.111Search in Google Scholar PubMed PubMed Central

Young, D.J., Guydosh, N.R., Zhang, F., Hinnebusch, A.G., and Green, R. (2015). Rli1/ABCE1 recycles terminating ribosomes and controls translation reinitiation in 3′UTRs in vivo. Cell 162, 872–884.10.1016/j.cell.2015.07.041Search in Google Scholar PubMed PubMed Central

Zacharogianni, M., Aguilera-Gomez, A., Veenendaal, T., Smout, J., and Rabouille, C. (2014). A stress assembly that confers cell viability by preserving ERES components during amino-acid starvation. eLife 3, e04132.10.7554/eLife.04132.030Search in Google Scholar

Zhang, Y. and Lu, H. (2009). Signaling to p53: ribosomal proteins find their way. Cancer Cell 16, 369–377.10.1016/j.ccr.2009.09.024Search in Google Scholar PubMed PubMed Central

Zhouravleva, G., Frolova, L., Le Goff, X., Le Guellec, R., Inge-Vechtomov, S., Kisselev, L., and Philippe, M. (1995). Termination of translation in eukaryotes is governed by two interacting polypeptide chain release factors, eRF1 and eRF3. EMBO J. 14, 4065–4072.10.1002/j.1460-2075.1995.tb00078.xSearch in Google Scholar PubMed PubMed Central

Zinoviev, A., Hellen, C.U.T., and Pestova, T.V. (2015). Multiple mechanisms of reinitiation on bicistronic calicivirus mRNAs. Mol. Cell 57, 1059–1073.10.1016/j.molcel.2015.01.039Search in Google Scholar PubMed PubMed Central

Received: 2019-06-04
Accepted: 2019-10-22
Published Online: 2019-10-29
Published in Print: 2019-12-18

©2020 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 26.4.2024 from https://www.degruyter.com/document/doi/10.1515/hsz-2019-0279/html
Scroll to top button