Next Article in Journal
Rational Design of Chimeric Antisense Oligonucleotides on a Mixed PO–PS Backbone for Splice-Switching Applications
Next Article in Special Issue
Putting It All Together: The Roles of Ribosomal Proteins in Nucleolar Stages of 60S Ribosomal Assembly in the Yeast Saccharomyces cerevisiae
Previous Article in Journal
Readers of RNA Modification in Cancer and Their Anticancer Inhibitors
Previous Article in Special Issue
Loss of the DYRK1A Protein Kinase Results in the Reduction in Ribosomal Protein Gene Expression, Ribosome Mass and Reduced Translation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Beak of Eukaryotic Ribosomes: Life, Work and Miracles

by
Sara Martín-Villanueva
1,2,†,
Carla V. Galmozzi
1,2,†,
Carmen Ruger-Herreros
1,2,
Dieter Kressler
3 and
Jesús de la Cruz
1,2,*
1
Instituto de Biomedicina de Sevilla, Hospital Universitario Virgen del Rocío/CSIC/Universidad de Sevilla, E-41013 Seville, Spain
2
Departamento de Genética, Facultad de Biología, Universidad de Sevilla, E-41012 Seville, Spain
3
Department of Biology, University of Fribourg, CH-1700 Fribourg, Switzerland
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 19 June 2024 / Revised: 19 July 2024 / Accepted: 21 July 2024 / Published: 22 July 2024
(This article belongs to the Special Issue Ribosomal Proteins in Ribosome Assembly)

Abstract

:
Ribosomes are not totally globular machines. Instead, they comprise prominent structural protrusions and a myriad of tentacle-like projections, which are frequently made up of ribosomal RNA expansion segments and N- or C-terminal extensions of ribosomal proteins. This is more evident in higher eukaryotic ribosomes. One of the most characteristic protrusions, present in small ribosomal subunits in all three domains of life, is the so-called beak, which is relevant for the function and regulation of the ribosome’s activities. During evolution, the beak has transitioned from an all ribosomal RNA structure (helix h33 in 16S rRNA) in bacteria, to an arrangement formed by three ribosomal proteins, eS10, eS12 and eS31, and a smaller h33 ribosomal RNA in eukaryotes. In this review, we describe the different structural and functional properties of the eukaryotic beak. We discuss the state-of-the-art concerning its composition and functional significance, including other processes apparently not related to translation, and the dynamics of its assembly in yeast and human cells. Moreover, we outline the current view about the relevance of the beak’s components in human diseases, especially in ribosomopathies and cancer.

1. Introduction: Protuberances of the Ribosomal Subunits

Ribosomes are intricate molecular machinery found in the cytoplasm of all organisms. They play crucial roles during the translation of cellular mRNAs by efficiently and accurately decoding their genetic information. In addition, ribosomes can be found inside two types of organelles, chloroplasts and mitochondria [1,2,3,4].
All ribosomes are universally composed of two subunits, the small and large ribosomal subunits (r-subunits), which comprise ribosomal RNAs (rRNAs) and ribosomal proteins (r-proteins). The large r-subunit is about twice the size of the small r-subunit [2,5]. In a model eukaryote, such as the yeast Saccharomyces cerevisiae, the small r-subunit or 40S is composed of a single 18S rRNA and 33 different r-proteins; in turn, the large r-subunit or 60S contains three rRNAs (5S, 5.8S and 25S rRNAs) and 46 (47 in humans) r-proteins [6,7]. The general shape of the ribosomes has been known since the early 1970s; however, structures at high resolution have only been obtained since the beginning of the twenty-first century, following significant advances in structural technologies, such as X-ray crystallography and cryo-electron microscopy (cryo-EM) (e.g., [7,8,9,10]). In agreement with the functional conservation of ribosomes in all organisms [2,5,11,12], ribosomes display considerable structural similarity in prokaryotes, organelles and eukaryotes, although organellar ribosomes have substantially diverged from bacterial ones, with which they share a common ancestor (e.g., [3,13]). Moreover, structural studies have clearly shown that eukaryotic ribosomes (25–30 nm in diameter) are larger and more complex than their prokaryotic (20 nm in diameter) and organellar counterparts [2,3,5,6,7]. When observed at low resolution, ribosomes appear flattened and spherical in shape; however, this observation is far from being correct. In fact, both r-subunits are extremely irregular complexes containing different domains and exhibiting different discernible protuberances. The small r-subunit is formed by four different structural domains, the head, the platform, the body and the beak; the beak domain is the main protrusion found in small r-subunits (Figure 1A). The large r-subunit, which is a more compact unit, displays three characteristic protuberances, the L1- and the P-stalk and the central protuberance (Figure 1B). Importantly, the structure of several r-subunit protuberances and protrusions has only been determined at high resolution in a few organisms, a fact that is most likely due to the high mobility of these regions, which is clearly connected to their functions (e.g., [6]). Thus, the movement of the beak as part of the head of 40S r-subunits is important to allow the loading of mRNA on the ribosome and the interaction with translation factors, as we will discuss in detail later on. In turn, the primary role of the P-stalk is to provide a flexible hook outside of the core of the ribosome to recruit and then stimulate the activity of different translational GTPases during various stages of translation (e.g., [14,15,16]). The P-stalk also enables the activation of the Gcn2 kinase and mediates the interaction with distinct ribosome-inactivating proteins (e.g., [17,18,19,20]). The L1-stalk acts as a flexible protuberance at the antipodes of the P-stalk, facilitating the binding, movement and release of the deacylated tRNAs [21,22].
In addition to being part of the prominent, above-described protuberances, many r-proteins have long non-globular extensions, typically serpentine ones that become deeply embedded inside the rRNA core of the r-subunits and stabilize their overall structure. There are also r-protein extensions (see Figure 1B for examples) that project out of the ribosome and may serve as additional regions for interactions of the ribosome with factors or cellular structures [23,24,25]. Frequently, these extensions, which are located at the N- and/or C-terminal ends of r-proteins, are not well ordered and are highly mobile due to the lack of stable interactions with the core of the ribosome; consequently, it has not been possible to model all of them from the current cryo-EM density maps.
Protuberances are not only flexible components of ribosomes, but also dynamic structures. The eukaryotic P-stalk is a pentameric complex, composed of the essential r-protein uL10 (previously named P0) and two non-essential heterodimers made up of another two r-proteins (named P1 and P2), that is connected to the body of the 60S r-subunit by the uL11 r-protein [15,26]. P1 and P2, when phosphorylated, can exchange during translation with their equivalent cytoplasmic pool of non-phosphorylated P1 and P2 variants [27,28,29]. Interestingly, in yeast, P1 and P2 r-proteins are not essential for cell viability, but their absence affects differently the translation of specific mRNAs [15,30]. Furthermore, it has been described that the amount of P1/P2 r-proteins bound to the ribosome changes under different physiological conditions. For instance, when yeast cells enter the stationary phase, the P1/P2 r-proteins are practically absent from ribosomes [31]. This and other observations made in different eukaryotes (e.g., [32,33]) have allowed different authors to propose that the eukaryotic P-stalk acts as a regulator of translation and that changes in its composition could selectively control the translation of specific groups of mRNAs [34]. In agreement with a regulatory role of the P-stalk during translation, mutation of the phosphorylation sites of the yeast acidic P1/P2 r-proteins does not alter their interaction with the ribosome but influences the translation of specific mRNAs, among them, those related to osmotic stress [35]. Further experiments, using high-resolution techniques (e.g., ribosome profiling), are required to provide details on such a regulatory process and on the specific proteome translated upon changes in P-stalk abundance and phosphorylation status.
To the best of our knowledge, aside from the acidic P1/P2 r-proteins, there is no evidence of exchangeability of r-proteins from the other ribosomal protuberances, including the beak; however, there are interesting reports about several other r-proteins whose nascent forms can replace an older copy of themselves on a mature ribosome. (i) Thus, uL16 has been described as an r-protein potentially able to cycle on and off large r-subunits [36,37,38]. uL16 is strategically positioned on the surface of the evolutionarily conserved core of the 60S r-subunit, near the corridor through which aminoacyl-tRNAs move during accommodation and also near other functional centers, such as the GTPase-associated center (GAC) and the peptidyl transferase center (PTC). uL16 is required for the joining of r-subunits during translation initiation and the rotation status of the ribosome [39,40,41]; importantly, the availability to assemble uL16 onto large r-subunits could also be used as a translational regulatory mechanism to limit global translation under unfavorable circumstances [37,41]. (ii) Another interesting r-protein is RACK1 (Asc1 in yeast). RACK1 is a WD40-domain protein located at the head region of the 40S r-subunit near the mRNA exit site, where it interacts with other r-proteins, among them uS3, several kinases, and translation initiation factors [42,43]. RACK1 has an important role in different aspects of the translation process: it is required for efficient translation of mRNAs with short open reading frames (e.g., those of r-protein genes), it is critical for translation during heat stress, and it facilitates ribosome-associated quality control (RQC) mechanisms such as those that are involved in the rescue of stalled collided ribosomes at consecutive rare codons, such as CGA (Arg) in yeast [44,45,46,47,48,49]. Although there is no experimental demonstration for RACK1 to cycle on and off ribosomes [50], it is clear that RACK1 is a non-essential r-protein whose loss does not disrupt ribosome integrity and translation [46]. Moreover, as RACK1 protein levels can be modulated by a variety of environmental insults, such as hypoxic stress, glucose deprivation or amino acid starvation or by the physiological cellular status (exponential versus stationary growth phase) (e.g., [51]), it is possible that the RACK1’s ribosome association could be regulated, thereby, promoting differential translation. (iii) Another important inductor of exchangeability is chemical damage. Originally reported in prokaryotes [52], yeast ribosomes containing chemically damaged r-proteins can also be repaired by exchanging these with undamaged r-proteins, as convincingly demonstrated by the Karbstein laboratory [38,53,54]. The repair of damaged ribosomes might represent an important mechanism for maintaining the translational activity of cells following different insults as, for example, those produced by oxidative stress [53,55]. An analogous repair process via protein replacement has also been suggested to occur in neurons; thus, implying that this mechanism has been evolutionarily conserved (e.g., [56]). In yeast, the molecular details for a ribosome repair mechanism have been provided upon oxidation of eS26 and uL16, which are released from damaged ribosomes by their respective dedicated chaperones, Tsr2 and Sqt1, generating transiently eS26- and uL16-deficient ribosomes that are subsequently repaired with newly made r-proteins [38]. Ribosomes lacking eS26 can also be generated in a Tsr2-dependent manner upon the exposure of yeast cells to high salt or high pH conditions. Ribosome repair is extremely relevant from the physiological point of view as eS26-lacking ribosomes preferentially translate specific transcripts bearing Kozak sequence variations, including mRNAs enabling the biological response to high salt and high pH insults [55]. The recovery from stress is concomitant to the reincorporation of eS26 into ribosomes, again in a Tsr2-dependent manner [38,53]. This sophisticated system of autoregulation resembles that previously reported for the translation circuit of leaderless mRNAs (lmRNAs) in bacteria. These lmRNAs can be generated in response to adverse environmental conditions, some of them (e.g., the presence of antibiotics such as kasugamycin in the culture media) also being able to reprogram ribosomes to translate preferentially lmRNAs. Interestingly, this reprogramming involves the formation of stable r-particles (referred to as 61S particles) deficient in almost a dozen r-proteins from the small r-subunit, among them bS1 and other r-proteins associated along the path of the mRNA through this r-subunit [57,58].
Of special attention is the central protuberance (CP), where the 5S ribonucleoprotein particle (RNP), which is composed of 5S rRNA and r-proteins uL5 and uL18, plays an important regulatory role (Figure 1B). The whole 5S RNP, rather than its individual components, is incorporated as a prefabricated complex into early pre-60S r-particles during the nucleolar ribosome biogenesis phase, and it temporally adopts a conformation that is different from the one in the mature 60S r-subunit (e.g., [59,60,61]). From yeast to humans, ribosome biogenesis is tightly coupled to cell growth and proliferation, with the assembly of the 5S RNP playing a central regulatory role. Thus, in yeast, an imbalanced production of rRNAs and r-proteins generates defects in ribosome biogenesis leading to the accumulation of ribosome-unbound uL18, likely as part of the 5S RNP, which induces a delay in the G1/S phase of the cell cycle [62]. This behavior has been interpreted as part of a protective mechanism that prevents cell cycle progression when ribosome biogenesis is impaired, i.e., when not all necessary components are sufficiently available to ensure a complete and satisfactory assembly of ribosomes. In metazoans, the 5S RNP clearly accumulates when ribosome biogenesis is impaired [63,64,65,66]. The free 5S RNP binds to MDM2 (HDM2 in humans), which is an E3 ubiquitin ligase that ubiquitinates p53 and thereby channels it to degradation via the proteasome. Upon binding of the 5S RNP to MDM2, which occurs mutually exclusively to its binding to pre-60S r-particles, p53 escapes from MDM2-mediated degradation and accumulates [67,68]. Concomitantly, p53 is activated and, thus, exerts its different anti-proliferative functions, ranging from temporary cell cycle arrest to apoptosis [69]. The implications of the involvement of ribosome biogenesis in the regulation of p53 in human health and disease have been extensively discussed in other reviews (e.g., [68,70,71,72]) and will also be examined later in this work.
This review is aimed at giving insights into the composition, structure, role, biogenesis and dysfunction of the components of the beak, which is the most prominent protuberance found in all small r-subunits and so-called because of its resemblance to a bird’s beak. We discuss all these features of the eukaryotic beak by specially focusing on the current knowledge about the beak of 40S r-subunits from the yeast S. cerevisiae and by highlighting similarities and differences compared to the beak of human ribosomes.

2. Composition of the Beak of the Eukaryotic Ribosome

Despite the fact that the beak is overall an easily recognizable structure in the small r-subunits of all ribosomes, the composition of the beak is quite different within the three domains of life. In bacteria, the beak is composed exclusively of rRNA, specifically, the helix h33 of 16S rRNA [9,73,74]. In contrast, the beak of eukaryotic ribosomes has been transformed into a mixture of rRNA and specific r-proteins not found in bacteria, with the biological reasons for this transformation remaining unsolved. The r-proteins bound to helix h33 of eukaryotic 18S rRNA, which itself is shorter than the bacterial h33, are eS10, eS12 and eS31 [1,2,6]. It is accepted that the structural core components of the archaeal ribosomes are of prokaryotic origin, to which specific elements, some shared with eukaryotes, have been added; therefore, archaeal ribosomes represent intermediate steps towards the evolution of eukaryotic ribosomes [75]. In this sense, it is interesting to mention that the beak of archaeal small r-subunits has a transitional complexity from an all-rRNA to an rRNA/r-protein protrusion; accordingly, many archaeal genomes encode clear homologs of the eukaryotic eS31 r-protein, which in all cases, however, lack the eukaryote-specific N-terminal extension [76,77]. In addition, cryo-EM has shown that ribosomes of distinct archaea contain at least two copies of eL8, one at the canonical location on the large r-subunit and another one bound at a position on h33 that is equivalent to the one occupied by eS12 on the eukaryotic beak [78]; further predictions suggest that this feature, i.e., the presence of eL8 in the beak, occurs in all archaeal ribosomes [78]. Moreover, this observation suggests that eS12 evolved from eL8, as both proteins share conserved regions and belong to the same family (InterPro entry IPR004038). Finally, it is clear that eukaryotic eS10 has no counterpart in archaeal ribosomes, as evidenced by different database searches, including BLAST [77,79]. A comparison of the beak composition and structure of ribosomes from prototypical bacteria, archaea and eukaryotes is shown in Figure 2.
Ribosomes present in organelles also contain all the structural landmarks that are characteristic of cytoplasmic ribosomes of prokaryotes and eukaryotes, including the beak. However, mito- and chlororibosomes have been found to be extremely diverse in terms of their composition, including the acquisition of organelle-specific r-proteins that has an impact on their overall structures. In general, chlororibosomes resemble bacterial ribosomes [80,81,82]; hence, the beak of these ribosomes is formed by the h33 rRNA protrusion and is devoid of r-proteins (Figure 3). In marked contrast to chlororibosomes, the characterization of mitoribosomes from diverse species representing the different major groups of eukaryotes has revealed that these have diverged considerably from each other and from prokaryotic and eukaryotic ribosomes [3,83,84]. In mitoribosomes, the beak can vary from the all-RNA prototype found in yeast to the massive protein-based beak found in the kinetoplastid Trypanosoma brucei (Figure 3).
Cytoplasmic ribosomes of the yeast S. cerevisiae contain a standard beak, composed of a helix h33 of 52 nucleotides and a single copy of three r-proteins, eS10, eS12 and eS31 (Figure 4) [6]. Yeast eS10 is encoded by two paralogous genes, RPS10A (YOR293W) and RPS10B (YMR230W). The two genes code for the virtually identical eS10A and eS10B r-proteins of 105 amino acids and ca. 12.7 kDa that only differ in three solvent-exposed amino acids (E6, D7 and T98 in eS10A versus Q6, E7 and S98 in eS10B). Mutants harboring individual deletions of the RPS10A and RPS10B genes are viable in different yeast backgrounds; moreover, while the rps10B∆ null mutant grows practically identical to the wild-type strain, the rps10A∆ null mutant exhibits only a mild increase in the doubling time [85]. In all genetic backgrounds, eS10 is an essential protein as the rps10A∆ rps10B∆ double mutant is inviable [85,86]. Yeast eS10 is a mostly globular protein; however, it contains an unstructured C-terminal extension of about 20 amino acids, which interacts with uS3 at the base of the beak in the mRNA entry channel [87]. In contrast to eS10 and most yeast r-proteins, eS12 and eS31 are non-essential r-proteins that are encoded by single-copy genes (RPS12 or YOR369C, and RPS31 or UBI3 or YLR167W, respectively). However, in most genetic backgrounds, both the rps12∆ and the ubi3∆ mutant display a severe growth impairment [85,88,89]. Yeast eS12 is a small globular protein of 143 amino acids and ca. 15.5 kDa, containing an unstructured N-terminal extension of around 25 amino acids whose deletion causes a slow growth phenotype of still uncertain significance (our unpublished results). On the other hand, yeast eS31 is a small r-protein of 76 amino acids consisting of a globular domain, which is well conserved from archaea to eukaryotes, and a eukaryote-specific N-terminal extension of about 25 amino acids, which extends toward the ribosomal A-site and has relevant functions in translation and small r-subunit assembly ([1,76,90]; see later). More interestingly, it has been well reported that in most eukaryotes eS31 as well as eL40 are produced as C-terminal parts of ubiquitin-fused precursor proteins, which are rapidly processed to individual ubiquitin and r-protein moieties before assembly of the corresponding r-protein into the small and large r-subunit, respectively [91]. The biological relevance of maintaining these fusions during evolution for the correct production and assembly of these r-proteins and for the possible co-regulation of two related cellular functions, protein synthesis and protein degradation, has been previously covered and will not be further discussed in this review [91,92].
In consonance with the critical importance of yeast eS10, eS12 and eS31 r-proteins for cell growth, it has also been reported that loss-of-function mutations in the genes coding for these proteins in other model eukaryotes (Caenorhabditis elegans, Drosophila melanogaster, Danio rerio, Mus musculus, Homo sapiens) lead to a myriad of adverse phenotypes, including lethality, increased cell death, cell cycle arrest, reduced fertility, organ development defects and tumorigenesis [93].

3. Roles of the Beak during Translation

The head of the small r-subunit is a flexible and dynamic structure involved in the engagement of the mRNAs and tRNAs during translation. Taking into consideration the strategic position of the beak at the entrance of the mRNA channel in the small r-subunit, it is not surprising that the beak has been linked to diverse functions during the translation process:
(i) During translation initiation, the beak is an important site for the interaction of trans-acting factors both in prokaryotic and eukaryotic ribosomes. For example, cryo-EM has revealed that the bacterial aldehyde-alcohol dehydrogenase E (AdhE) enzyme interacts with ribosomes in the beak region [94]. This enzyme provides a further RNA helicase activity, in addition to the intrinsic one of the ribosome [95], in order to ensure the linear configuration of structured mRNAs at the mRNA entrance to facilitate their translation [94]. In eukaryotes, several RNA helicases play roles during translation initiation. The canonical initiation factor eIF4A and the Ded1 (DDX3 in mammals) RNA helicase assist in the unwinding of the 5′-UTR secondary structure of most mRNAs [96,97,98]. Interestingly, in mammals, the translation initiation of endogenous and viral mRNAs with highly structured 5′-UTRs requires an additional RNA helicase, named DHX29 [99,100]. As revealed by cryo-EM, DHX29 contacts the beak and adjacent regions by interacting with at least uS3, eS10 and eS12 [101]. In other examples, the beak has been described as being important for the recognition of specific mRNAs. Accordingly, mammalian eS10 has been found to specifically interact with a class of cellular mRNAs containing the so-called TISU-element in their short 5′-UTRs [102].
(ii) The loading of the mRNA itself into the mRNA channel of the small r-subunit during translation initiation is regulated by the opening and closing of an mRNA latch situated below the beak that connects the body and the head of the small r-subunit [74]. The open conformation of this structure is promoted by the binding of distinct initiation factors (eIF1 and eIF1A in eukaryotes, IF1 in prokaryotes) to the small r-subunit during the formation of the eukaryotic 43S pre-initiation complex [103]. eIF1A is a globular protein harboring unstructured N- and C- terminal extensions of ca. 25 amino acids. During translation initiation, the globular domain of eIF1A is positioned at the A-site, while its extensions seem to project out of this site of the ribosome; thus, preventing tRNA binding to this site [103]. From X-ray crystallography, it can be inferred that the N-terminus of eIF1A directly contacts the eukaryote-specific N-terminal extension of eS31 and approaches extensions of other r-proteins, such as eS10, uS3 and uS19, that are adjacent to each other in the A-site [104]. These interactions seem to be crucial for translation as mutations in the N-terminal tail of eIF1A, which are frequently observed in several types of cancers [105], result in reduced binding of eIF1A to its r-protein partners and a hyperaccurate recognition of AUG codons that are embedded in an optimal sequence context [106,107]. Importantly, the phenotypic analysis of yeast ubi3 mutations (e.g., ubi3G75,76A) that interfere with the cleavage of the ubiquitin-eS31 fusion protein indicates that the non-cleaved protein can still assemble into mature 40S r-subunits, which are active in translation but mildly defective at the translation initiation stage. This defect is likely due to the interference of the ubiquitin moiety with the binding and proper activity of the initiator tRNA and the eIF1A factor [89,104]. Moreover, interactions between several subunits of the eIF3 complex and beak components, including eS10 (e.g., [108]), have been described to occur within the yeast 43S pre-initiation complex; these are expected to be of functional relevance during translation initiation (e.g., [109]).
(iii) The beak also participates in the formation of the binding surface for the internal ribosome entry site (IRES) of some viral mRNAs on human 40S r-subunits. For instance, among other 40S r-proteins, eS10 contributes to the binding of the hepatitis C virus (HCV) IRES [110]. Viral proteins also interact with the beak to hijack the host’s translation machinery. One interesting example concerns the SARS-CoV-2 coronavirus, whose non-structural protein NSP1 contains a globular N-terminal domain that binds the base of the beak, while its C-terminal extension blocks the mRNA channel entry site and thereby prevents any mRNA accommodation; thus, inhibiting translation of host mRNAs [111,112,113]. However, NSP1 does not impede the translation of viral mRNAs, which is promoted by the presence of a cis-acting RNA hairpin in the 5′-UTR of these mRNAs [114,115].
(iv) Translation is reversibly shutdown upon nutrient starvation in a variety of ways, including the accumulation of inactive or hibernating vacant 80S ribosomes. These dormant ribosomes contain eEF2·GTP in the A-site and the hibernation factor Stm1 (in yeast) or SERBP1 (in mammals) in the mRNA channel, thereby impeding mRNA binding [87,116]. This mechanism of blocking the mRNA entry tunnel resembles that mentioned above for the coronavirus NSP1 protein. It has been described that the C-terminal region of Stm1/SERBP1 also stably associates with the head of 40S r-subunits, likely via binding to eS10, eS12 and eS31 [87,116].
(v) In all ribosomes, the interaction of the different translation elongation factors with the ribosome leads to specific movements of the head domain of the small r-subunit, including that of its associated beak towards the shoulder of the body of the same r-subunit (e.g., [117,118]). In eukaryotes, the beak components themselves interact with distinct domains of the translation elongation factors, as exemplified by the interaction of eS12 and eS31 with domains II and IV of eEF2 [7]. In agreement with the important role of beak r-proteins in the fidelity of translation elongation, the depletion of the essential yeast eS10 as well as the mutation or deletion of the genes encoding yeast eS12 and eS31 result in translation defects, including misreading [76,88,119]. Moreover, due to the specific position of the N-terminal extension of eS31 in the A-site of the ribosome, the assembly of non-cleaved yeast Ubi3 is expected to sterically interfere with the binding of the translation elongation factors to the ribosomal GTPase-associated center [89,91].
(vi) The beak is also expected to be functionally relevant during translation termination. Cryo-EM structures have revealed how eukaryotic translation termination factor 1 (eRF1), whose overall shape resembles a tRNA molecule, interacts with a stop codon in the A-site of the ribosome via its N-terminal lobe (e.g., [120] and references therein). Notably, in the structures of pre-termination complexes, a short segment of the N-terminal lobe of eRF1 is in close proximity to the initial residues of the N-terminal extension of eS31 [120,121]. Moreover, the mini-domain of eRF1, which is an insertion within the C-terminal domain, also interacts with the N-terminal extension of eS31 and protrudes toward the beak where it contacts helix h33 [120,121,122].
(vii) Another example of the role of the beak components in translation comes from studies of the cellular responses to elongation stalls induced by different stresses (including oxidative stress, heat shock or starvation) as well as by particular sequences, strong secondary structures and chemical damage within mRNAs. Normally, when exposed to stressful conditions, cells adapt by halting or decreasing the global synthesis of new proteins, while, concomitantly, inducing the selective translation of mRNAs encoding proteins that are necessary for cell survival and stress recovery [123]. This translational reprogramming can be mediated by multiple parallel and independent signaling pathways that converge on the modulation of the function of a few key translation factors [124]. Relatively recent studies from the Silva laboratory showed that following oxidative stress, induced by an exposure of yeast cells to hydrogen peroxide, a set of r-proteins were K63-specifically polyubiquitinated at different residues by the ubiquitin-conjugating enzyme Rad6 and the ubiquitin-protein ligase Bre1, with the extent of this modification declining very rapidly during stress recovery [125,126]. Most of these r-proteins are located within the head of the small r-subunit of the ribosome and include uS3, the beak components eS10, eS12 and eS31 and the P-stalk proteins uL10, uL11 and P2 [127]. Although oxidative stress induces a rapid inhibition of translation initiation via activation of the Gcn2 kinase (see below), K63-linked ubiquitination of r-proteins leads to an additional response, which results in the stalling of translation at the elongation stage [127]. Using cryo-EM and cryo-electron tomography, the Silva laboratory was also able to demonstrate that K63-linked ubiquitination of ribosomes alters the conformation of distinct r-proteins, including eS31 and eS12, that are located at the interface of the two r-subunits where eEF2 binds, thereby interfering with its efficient binding and/or GTPase activity and promoting the translational halt at the elongation stage, specifically at the rotated pre-translocation stage 2 [128].
If a ribosome persistently stalls on an mRNA, collisions with the trailing ribosomes will eventually occur; this phenomenon triggers different ribosome-associated quality control (RQC) mechanisms. It is thought that these mechanisms have evolved in order to relieve stalled ribosomes, thus avoiding the depletion of active ribosome and tRNA pools, which would prevent their participation in new rounds of protein synthesis and could reduce cellular fitness or survival [129,130]. RQC mechanisms additionally target damaged mRNAs and incomplete polypeptide chains for degradation [129]. In prokaryotes, the rescue of ribosomes stalled at the 3′ end of mRNAs lacking a stop codon, thus containing an empty A-site, often involves the action of the long transfer-messenger RNA (tmRNA), whose interaction with the ribosome occurs through the formation of a ring of its large loop around the beak of the 30S r-subunit (for further information, see [131]). In eukaryotes, different rescue pathways center around the recognition of the empty A-site in the ribosome (e.g., [130,132]). The rescue of eukaryotic ribosomes stalled on truncated and aberrant mRNAs lacking stop codons (NSD, non-stop decay) relies on several factors, such as Dom34 (Pelota in mammals), Hbs1 (HBS1L in mammals), Rli1 (ABCE1 in mammals) and Ski7 (HBS1L3 in mammals). These factors interact or have the potential to interact with ribosomes in a similar manner as translation elongation and termination factors [133]; therefore, it is expected that the binding and function of these factors, and, thus, the fate of NSD, could be altered by mutations affecting beak components. Dom34/Pelota is structurally related to tRNAs and eRF1 and binds the ribosome in a similar way to these two; in turn, Hbs1, which interacts with Dom34/Pelota, is a member of the family of translational GTPases that includes eEF1, which delivers aminoacyl-tRNAs to the A-site, eRF3, which interacts with eRF1 in a similar manner to Hbs1 with Dom34, and Ski7, which is a paralog of Hbs1 [133]. The N-terminal domain of Ski7 mediates the recruitment of the exosome and the Ski2-Ski3-Ski8 complex (SKIV2L-TTC37-WDR61 in humans) [134,135,136], while its C-terminal part contains the GTPase-like domain that it is assumed to interact, similar to other translation GTPases, with the GAC site of the ribosome, but whose exact role is still unknown [137,138]. Cryo-EM structures of different ribosomes, Ski2/3/8 complex and exosome intermediates suggest a scenario where a stalled ribosome bound to the Ski2/3/8 complex recruits a pre-assembled exosome-Ski7 complex [136,139,140]. Through this triple (ribosome–Ski2/3/8 complex–exosome) interaction, aberrant mRNA substrates are unwound and guided into the exosome. The ribosome-bound Ski2/3/8 complex specifically recognizes the 40S r-subunit by binding near the entry of the mRNA channel and connecting the head and beak regions [140]. Concerning the beak, the Ski2 helicase interacts with several r-proteins, among them eS10 and uS3, while the N-terminal part of Ski3 contacts eS12 [139,140].
Prolonged ribosome stalling leads to a ribosome collision of the trailing ribosomes with the stalled ribosome [141,142]. Under these circumstances, the E3 ligase Hel2 (ZNF598 in mammals) recognizes the collided ribosomes and adds ubiquitin to a number of 40S r-proteins at precise lysine residues, among them eS10 and uS3 ([143] and references therein). This ubiquitination is assumed to serve as the starting signal for the progression of the RQC response in order to dissociate the stalled ribosomes into r-subunits and degrade their associated mRNAs and nascent peptides [144]. In addition, beside many other responses [142], ribosome collisions also activate the kinase Gcn2, and evidence suggests that this activation can occur independently of the presence of deacetylated tRNAs [145,146], which constitutes its classical activation pathway. This activation is dependent on Gcn1, and cryo-EM has nicely revealed how this long, tube-like HEAT repeat protein spans across a collided disome by forming an extensive network of interactions both with the leading and trailing ribosome [147]. Interestingly, along its interaction path, the region preceding the central eEF3-like HEAT repeats engages in contacts with eS10, eS12 and eS31 within the beak of the 40S r-subunit of the colliding ribosome [147,148]. Perturbations of the beak, elicited by absent or mutated beak r-proteins, could influence Gcn1 such that its ribosome association or Gcn2-binding capacity, both of which are required for Gcn2 activation, is affected. In turn, activated Gcn2 can then phosphorylate eIF2α to downregulate general translation initiation and to enable the translation of specific mRNAs, such as those encoding Gcn4 in yeast or ATF4 in mammals, in order to adequately respond to the stress that causes ribosomes to collide. In line with the structural integrity of the beak being necessary for an efficient Gcn1-mediated Gcn2 activation, lower levels of eS10 (individual deletion of RPS10A or RPS10B) or the absence of eS31 were shown to reduce the extent of eIF2α phosphorylation or to impair derepression of GCN4 mRNA translation in response to amino acid starvation, respectively (e.g., [148,149]). Recently, the ubiquitination of eS31 has also been reported to occur in circumstances of translation elongation inhibition where the A-site is occluded by a trapped eEF1A factor bound to an aminoacyl-tRNA [150]. In this case, the reaction is dependent on an E3 ligase called RNF25. Ubiquitination of eS31 is required for the degradation of the trapped eEF1A, which itself is ubiquitinated both by RNF25 and an additional E3 ligase RNF14, with the latter directly interacting with GCN1, which is also essential for eEF1A degradation [150].

4. Other Cellular Functions of the Beak Components

The beak r-proteins, in addition to their clear role in translation, participate in other cellular processes, including ribosome biogenesis (see Section 5), activation of the p53-dependent pathway in response to nucleolar stress as well as oncogenesis (see Section 6), and cell competition (see below). Whether or not the effects that mutations in the beak r-proteins have on these processes are translation-dependent or independent is still unclear in some cases.
As mentioned above, a systematic study has analyzed the contribution of loss-of-function mutations for most r-proteins, including eS10, eS12 and eS31, to multiple phenotypic features in six relevant eukaryotic model organisms (S. cerevisiae, C. elegans, D. melanogaster, D. rerio, M. musculus, and H. sapiens) [93]. Several reports on the characterization of specific features of eS10 have highlighted the important role of this r-protein. These include the description of a hypo-proliferative phenotype, known as the Minute phenotype, associated with loss-of-function mutations in one of the two copies of the RPS10 gene during development in Drosophila. The Minute phenotype is characterized by a prolongation of the developmental time, the presence of short and thin bristles, and reduced fertility [151]. In Drosophila, eS10 is encoded by duplicated genes, and, interestingly, the expression of one of the two genes is enriched in the germline cells of embryonic gonads, suggesting a germline-specific role [151]. In Arabidopsis, loss-of-function mutations in one of the genes encoding eS10 lead to a reduction in stamen number, shoot and floral meristem defects, and a leaf polarity deficiency [152].
The r-protein eS12 also plays interesting roles not directly related to ribosome biogenesis or translation. In specific neurons, RPS12 mRNA levels, among other r-protein transcripts, seem to be reduced by an acute period of sleep deprivation (hippocampus) or injury of the sciatic nerve (dorsal root ganglion), suggesting dynamic changes in ribosome composition following these insults (reviewed in [153]). In S. cerevisiae, a specific mutation in the RPS12 gene leads to the suppression of phenotypes elicited by rDNA instability upon Fob1 overexpression [154]. Whether this phenomenon is the result of an extra-ribosomal function of yeast eS12 remains to be explored. Undoubtedly, the most interesting function of eS12 besides its orthodox roles in ribosome biogenesis and translation is that related to cell competition in D. melanogaster. In the classical form of cell competition, wild-type cells (homozygotic cells for r-protein genes; hereafter Rp+/+ cells) in genetic mosaic flies are able to actively eliminate their adjacent heterozygotic cells (heterozygotic cells for r-protein genes; hereafter Rp+/− cells) from imaginal discs via apoptosis [155,156,157]. In this process, eS12 plays a specific role as a sensor of an imbalance of r-proteins to allow the elimination of Rp+/− cells [158]. Thus, the viable missense rps12[G97D] mutant allele of RPS12 in homozygosis prevents cell competition of Rp+/− cells by wild-type Rp+/+ cells [158,159]. In other words, rps12[G97D]+/+ Rp+/− cells are not eliminated by wild-type Rp+/+ cells. Moreover, the relative copy number of the wild-type RPS12 allele in Rp+/− cells is apparently what determines the competitiveness [158]. It has been shown that the eS12[G97D] variant efficiently assembles into 40S r-subunits [158]. Moreover, the yeast rps12[G102D] allele (equivalent to Drosophila rps12[G97D] allele), when it is the sole cellular source of eS12 r-protein, neither confers a growth defect nor a global impairment of translation (S. M.-V., unpublished results). Interestingly, it has been shown that Drosophila eS12 is required to increase the transcription of the gene encoding the transcription factor Xrp1 [159,160], which itself also directly regulates cell competition [156,159]. Consistently, loss-of-function mutations in Xrp1 also prevent competition of Rp+/− by wild-type Rp+/+ cells [159]. Whether the function of eS12 in promoting Xrp1 expression is extra-ribosomal still needs confirmation.
As for eS10 and eS12, there are reports indicating that eS31 could also have ribosome-independent functions (for a recent review, see [161]). First, eS31 has been identified as a regulator of the LMP1 protein encoded by the Epstein-Barr virus (EBV); thus, eS31 binds directly to LMP1 and increases its stability by reducing its proteasome-mediated degradation [162]. Moreover, overexpression of eS31 leads to increased cell growth and survival as the result of LMP1-mediated oncogenic events (e.g., epithelial to mesenchymal transition, motility, migration and invasion). In addition, as RPS31 (also known as RPS27A) mRNA levels are reduced in sperm with low motility, eS31 might be necessary for optimal sperm functionality in humans [163]. In plants, eS31 seems to be highly expressed in meristematic tissues, pollen and ovules [164], and flowers of RPS31-silenced plants exhibit abnormal development [165]. Finally, it has been observed that double-strand break DNA damage results in the MDM2-independent proteasomal degradation of eS31 in HEK293 human cells, and as a consequence, these cells contain ribosomes that specifically lack eS31 and exhibit lower global translation activity [166]. Whether this phenomenon is part of an adaptive response to deal with DNA damage remains to be determined [166].

5. Assembly and Maturation of the Beak Structure

The assembly of the beak has been analyzed in the context of the general maturation of the 40S r-subunit, which begins in the nucleolus and ends in the cytoplasm, both in yeast and in human cell lines (for a review, see [167]). Moreover, given the fact that the beak is a pronounced protrusion in the structure of the 40S r-subunit, the assembly of the beak represents a challenge for the nucleocytoplasmic transport of this r-subunit. Using genetics in yeast and siRNA technology in human cell lines, the role of the beak r-proteins in pre-rRNA processing and r-subunit assembly has also been well examined. In both cases, it has been described that the three r-proteins that form the eukaryotic beak are required for the production and the stability of mature 40S r-subunits.
In yeast, eS10 is an essential r-protein whose contribution to ribosome biogenesis has been assessed by the use of a yeast strain conditionally expressing this r-protein [86,168]. These studies showed that eS10 is required for the efficient maturation of the 20S pre-rRNA, which accumulates to high levels upon eS10 depletion both within nucleoplasmic pre-40S r-particles, as a consequence of a delay in the export of pre-40S r-particles and cytoplasmic pre-40S r-particles, and as a consequence of inefficient 20S pre-rRNA processing at site D [86,168]. Interestingly, knocking down the expression of human RPS10 leads to cytoplasmic accumulation of the 18S-E pre-rRNA, which is the equivalent human form of the yeast 20S pre-rRNA [169,170]. In yeast, we and others have demonstrated that the quasi-essential r-proteins eS12 and eS31 are also crucial for the efficient cytoplasmic processing of the 20S pre-rRNA into mature 18S rRNA [88,89,92]. Similarly, the cytoplasmic accumulation of 18S-E pre-rRNA has also been reported upon siRNA-mediated knockdown of human RPS12 or RPS27A expression [170,171,172].
The precise timing of the assembly of the beak rRNA and r-proteins has also been analyzed in both yeast and humans at a reasonable resolution. Assembly of this structure involves compositional and structural changes of both the beak rRNA and r-proteins. In general terms, while the timing of eS31 assembly (nucle(ol)ar or cytoplasmic) is still controversial, it is likely that eS12 is incorporated early during the formation of 90S pre-ribosomal particles and that eS10 assembly occurs within late cytoplasmic pre-40S r-particles (see below). In yeast and the fungus Chaetomium thermophilum, the structural analysis of 90S pre-ribosomal particles suggests that the four subdomains of the 18S rRNA (5′, central, 3′ major, and 3′ minor) fold independently and associate co-transcriptionally with a set of r-proteins and ribosome assembly factors (RAFs), before being compacted into a defined pre-ribosomal particle. Analysis of the first reported structures of 90S pre-ribosomal particles indicated that the folding of the 3′ major domain of the 18S rRNA, the helix h33 included, requires the prior co-transcriptional structuring of the 5′ domain [173,174,175]. However, a more recent structural determination of a series of 90S assembly intermediates from C. thermophilum provides evidence that the formation of the 90S does not follow a strict 5′ to 3′ co-transcriptional direction; instead, the 3′ major and 3′ minor domains seem to assemble first with the 5′-ETS domain of 35S pre-rRNA, preceding the incorporation of the 5′ and central domains of pre-18S rRNA into 90S pre-ribosomal particles [176]. In any case, from the diverse collection of structurally stable 90S r-particles available in the literature, it is clear that these particles contain a clearly identifiable, immature beak structure. In most of these particles, the nascent beak structure comprises eS12, while only a few of them also contain eS31. As an example of this, in the 90S structure reported by Sun et al. [173], the beak forms a protrusion and it is composed of helices h32-34 and the r-proteins eS12 and eS31, connected to the body of the particle by the RAF Emg1 (see Figure 5). At this level, it is also clear that the presence of Enp1, which stabilizes the beak by binding to helices h32-34, impedes the incorporation of eS10, which can only occur after the release of Enp1 from late pre-40S r-particles in the cytoplasm [177,178]. Moreover, the fact that eS31 (and to a lesser extent eS12) is not present in many of the structural maps of 90S r-particles available in the literature, as well as in a variety of further pre-40S r-particles (see below), clearly indicates that the association of these r-proteins with early precursors of 40S r-subunits might be highly labile and should only become stable during late and cytoplasmic steps of 40S r-subunit maturation, concomitant with the formation of a more rigid beak structure. In this regard, another RAF, Tsr1, apparently blocks the correct binding of eS31 until its repositioning at a late maturation step occurring on cytoplasmic pre-40S r-particles [179]. Alternatively, as favored by other authors, eS12 and especially eS31 are only incorporated into cytoplasmic pre-40S r-particles [179,180]. However, at least in the case of eS31, we have identified a functional nuclear localization signal (NLS) within the first 25 amino acids of the N-terminal extension of yeast eS31, which is conserved in other eukaryotes [76]. This sequence is sufficient to target a triple GFP reporter to the nucleus, and most importantly, a functional GFP-tagged eS31 protein notably accumulates in the nucleus upon depletion of different 90S RAFs, among them Emg1, which leads to the nuclear retention of pre-40S r-particles ([76] and S. M.-V., unpublished results). Whatever the case may be, the N-terminal ubiquitin moiety present in the linear precursor of eS31 in many eukaryotes, including yeast and humans, is very rapidly and efficiently processed. Accordingly, under wild-type conditions, the Ubi3 precursor has so far never been detected; hence, it must be processed prior to the incorporation of eS31 into pre-40S r-subunits [89,92,181]. Consequently, it is unlikely that the ubiquitin moiety fused to eS31 directly participates in the ribosomal assembly of eS31. Moreover, when a wild-type and a cleavage-deficient Ubi3 variant are co-expressed in the same cells, eS31 derived from wild-type Ubi3 is preferentially incorporated into pre-40S r-particles compared to the non-cleaved ubiquitin-eS31 fusion protein, which in turn is rapidly degraded [89]. Forcing the assembly of non-cleaved Ubi3 into nascent 40S r-subunits only mildly impairs their biogenesis, but, as mentioned above, may lead to translation initiation defects [89].
Following the sequential cleavages at sites A0–A2 within the 35S pre-rRNA, the yeast 90S r-particle is dismantled and converted into an early nuclear pre-40S r-particle, which is rapidly exported to the cytoplasm. At this step, and before export, most 90S RAFs have disassembled and only a few others have been recruited, among them Rio2, Tsr1, Ltv1 and Rrp12 [182]. Perhaps the characteristic that defines best the nucleoplasmic pre-40S intermediates is the high flexibility of their head domain, which becomes more structured as the particles transition through their maturation [183,184,185]. The incorporation of Ltv1 is relevant for beak formation as it interacts with Enp1 and the r-protein uS3, which binds at the base of the beak structure [119,186,187]. The recruitment of uS3 is initiated just before or concomitant with that of Ltv1 [188,189]. The r-protein uS3 consists of two distinct N- and C-terminal domains and is delivered to nuclear pre-40S r-particles by its dedicated chaperone Yar1 [190]. Yar1 binds only the N-terminal domain of uS3; thus, initial interaction of uS3 with pre-40S r-particles likely occurs through its C-terminal domain [188,191,192]. The release of Yar1 is concomitant with the interaction of the N-terminal domain of delivered uS3 with Ltv1. This interaction also contributes to preventing uS3 from prematurely acquiring its final and stable position within cytoplasmic pre-40S r-particles, which is only achieved upon the global structural changes occurring in these r-particles after Ltv1 release [119,179,190,191]. Indeed, a subcomplex formed by Ltv1, Enp1 and uS3 can be untethered from purified yeast pre-40S r-particles at high salt concentrations, while uS3 cannot be extracted from mature 40S r-subunits by the same treatment [193], indicating that uS3 is less stably integrated into pre-40S than mature 40S r-subunits. It has been suggested that a certain degree of flexibility in the beak is required at this nucleoplasmic stage because a rigid beak structure close to the head of the pre-40S r-subunit might hinder export through the nuclear pore complex (NPC) [179,193,194,195].
Once exported, the early cytoplasmic pre-40S r-particles undergo a cascade of maturation events, the first ones being essential for beak formation. The precise chronology of these events remains to be elucidated at high resolution, but it seems that it first involves the recruitment of the casein kinase Hrr25, also favored by the previous binding of the uS3 r-protein [190,192]. Moreover, the direct interaction between Hrr25 and Ltv1 appears to weaken the association between Ltv1 and Enp1 [192]. Hrr25, which is an essential protein, then phosphorylates Ltv1 on specific conserved serine residues, leading to the release of Ltv1 from pre-40S r-particles [188,189,192]. Strikingly, Hrr25 is no longer essential in the absence of Ltv1 or upon phosphomimetic substitutions of the specific Ltv1 serine residues [188,189], indicating that the essential function of Hrr25 is linked to Ltv1 in ribosome biogenesis. Interestingly, the release of Ltv1 is coordinated with that of Rio2 on the intersubunit side of the head domain of the pre-40S r-particle; a process that is mediated by the correct assembly of the uS10 r-protein [192]. The release of Ltv1 now provokes the dissociation of Enp1, which is also phosphorylated by the Hrr25 orthologue (CK1δ/ε) in humans [196]; phosphorylation of yeast Enp1 by Hrr25 is still controversial [189,193]. The dissociation of Enp1 and Ltv1 is absolutely required for nascent 40S r-subunits to become translationally competent, as their interaction with the beak environment would hinder the opening of the mRNA channel [177]. Another consequence of the dissociation of these factors is that eS10 gains access to its binding position and is integrated into the beak structure [180,186,197]. Concomitantly, uS19 and the two domains of uS3 are fitted into their mature position [119,179,192]. All these events promote the structural organization of the beak, which then enables the progression of the maturation events in other regions of the pre-40S r-particles [167,180].
Formation of the beak in human 40S r-subunits seems to occur in a similar way to that described in yeast, albeit with certain peculiarities [167,197,198]. Orthologs for all key factors mentioned above have been described in humans, including Enp1, Ltv1 and Hrr25 (Bystin, LTV1 and CK1δ/ε, respectively) [199]. Most importantly, despite differences in pre-rRNA processing between yeast and humans [200], the positioning, timing of interaction and dissociation, and function of all these RAFs have been well conserved [199]. Moreover, the structures of several nuclear and cytoplasmic pre-40S r-particles have been described, providing detailed insights into the maturation steps [184,197,198,201]. As an example, the cryo-EM structures of apparently nucleoplasmic human pre-40S r-particles contain, in addition to the orthologs of Enp1 and Ltv1, the r-proteins eS12, eS31 and uS3, despite the fact that, as in yeast, the assembly time point of eS31 and eS12 is again controversial [184,197].
Finally, although beyond of the scope of this review, the assembly of the prokaryotic beak is also an important late event during the maturation of 30S r-subunits; in this regard, it is interesting to mention that some authors have proposed that the bacterial assembly factor RimM, which is involved in the maturation of the 3′ domain of the head of 30S r-subunits (see [202] and references therein), works as a functional analog of Ltv1 in bacteria [119,189]. Strikingly, in the absence of RimM, the 30S beak (h33) is not correctly folded, and several r-proteins, including uS19 as well as the tertiary binders uS3 and uS10, do not efficiently assemble into pre-30S r-particles [203,204]. Moreover, the assembly of the small subunit of the trypanosomal mitoribosome represents a major challenge for those researchers who are studying this process (e.g., [205]).

6. Beak Components and Human Diseases

It is evident that mutations and dysregulation of the majority of r-protein genes are linked to a range of human genetic diseases, such as ribosomopathies and cancer [206,207]; in this regard, beak r-proteins are not an exception.
Ribosomopathies are a group of rare inherited or acquired genetic diseases linked to defects in r-proteins or ribosome biogenesis factors [207,208,209]. Despite the importance of ribosomes in all cell types, these diseases result mainly in tissue-specific manifestations, especially in the hematopoietic system [207]. Intriguingly, inherited ribosomopathies are congenital and normally exhibit a paradoxical transition from early symptoms related to cellular hypo-proliferation to a hyper-proliferative oncogenic state later in life [210]. The best studied ribosomopathy, which is also one of the most prevalent ones (10 individuals per million live births) is the so-called Diamond–Blackfan Anemia (DBA) [211]. DBA is mostly a dominant genetic disorder (autosomal or X-linked) that is characterized by the reduced formation of red blood cells and is also associated with a series of other congenital anomalies, such as skeletal abnormalities, heart and genitourinary malformations, and an increased cancer susceptibility [212]. Most patients diagnosed with DBA harbor heterozygous loss-of-function mutations in particular genes encoding r-proteins, either of the small or the large r-subunit [212]. About 3% of all DBA patients have been reported to carry mutations in the RPS10 gene, most likely leading in all cases to the production of non-functional eS10 variant proteins [212,213]; these mutations mostly consist of (i) insertion mutations that cause a frameshift and the appearance of a premature termination codon, (ii) nonsense mutations, namely, changes of particular codons to a stop codon (often, the R113Stop mutation), (iii) missense mutations that transform the RPS10 start codon into an isoleucine or a threonine codon (M1I or M1T), and (iv) different other missense mutations (e.g., L14F, P30L) of so far unknown biological significance ([213,214]; for more details, check the UniProt entry P46783 and the OMIM entry 603632). Mutations in the human RPS27A gene, which codes for human eS31, have also been identified in patients with DBA (e.g., S57P); however, whether these mutations are indeed pathogenic genetic variants remains to be determined [215]. To our knowledge, no DBA-linked mutations have so far been reported in the RPS12 gene. However, as the underlying mutations in at least 20% of patients with DBA syndromes have not yet been identified [212], it is still possible that RPS12 alleles could be responsible for DBA manifestations; in line with this possibility, RPS12 haploinsufficiency in mice leads to an erythropoiesis defect that recapitulates the one found in DBA patients (discussed in [216]). From a very simplistic point of view, as the DBA disease is mostly caused by loss-of-function mutations in several r-protein and a few RAF genes, all associated DBA symptoms and manifestations must come from common dysfunctions of the same molecular process that, in this case, can be no other than ribosome biogenesis [217], ultimately leading to an impairment or a limitation of translation. Thus, in a non-exclusive manner, it has been proposed that (i) the hypo-proliferative, pro-apoptotic anemia associated with DBA could be the consequence of a global reduction in translation, limiting below a critical threshold the synthesis of critical proteins, such as the globins and the transcription factor GATA1, with the latter being essential for normal erythropoiesis [218]. In agreement with this possibility, GATA1 translation is reduced in erythroid precursor cells of DBA patients with mutations in different r-protein genes (e.g., [219,220,221]), and loss-of-function mutations in GATA1 result in a DBA-like phenotype (e.g., [222,223] and references therein). (ii) DBA cells display elevated levels of reactive oxygen species (ROS), which, by generating a high oxidative stress, inhibit cell proliferation [206,207]. In this regard, lowering cellular ROS levels by antioxidants can rescue the proliferation defects in cells subjected to r-protein haploinsufficiency or carrying selected r-protein mutations [207,224]. (iii) As a consequence of the ribosome biogenesis deficiency occurring in DBA cells, the so-called nucleolar stress response is triggered in these cells, which induces p53 stabilization and enables p53 to transactivate its target genes, leading to cell cycle arrest, apoptosis, autophagy, and senescence [225,226]. In normal growth conditions, cellular levels of p53 are maintained low due to its efficient recognition and ubiquitination by the E3 ubiquitin ligase MDM2 (HDM2 in humans) and the subsequent degradation of ubiquitinated p53 by the proteasome. However, when ribosome biogenesis is impaired, non-assembled r-proteins tend to accumulate and can be released from the nucleolus to the nucleoplasm. Several different free r-proteins, but primarily uL5 and uL18 as part of the 5S RNP, can bind and sequester MDM2, thereby preventing the degradation of p53; thus, the upregulation of p53 explains many of the hypo-proliferative phenotypes displayed by DBA patients, including bone marrow erythroid hypoplasia [206,227,228]. In consonance with the relevant role of nucleolar stress in DBA, genetic or pharmacological inactivation of p53 can rescue disease-associated phenotypes [206,209,225,229]. Interestingly, it has been shown that eS31 is able to regulate the MDM2-p53 loop in response to nucleolar stress [230,231]. Moreover, eS31 apparently interacts with the central acidic domain of MDM2 through its eukaryote-specific N-terminal extension [230]. Therefore, this interaction seems not to be mutually exclusive from the ones of MDM2 with p53, which used a short, N-terminal segment to bind to the N-terminal domain of MDM2 [232], and with the r-proteins uL5 and uL18, which mostly involve the Zn-finger and RING domains of MDM2 [59,233]. Importantly, the overexpression of eS31 reduces MDM2-mediated ubiquitination of p53, thereby leading to its stabilization and activation [230,234]. The induction of p53 by eS31 may likely be additionally fueled by the observation that the RPS27A gene is apparently also transcriptionally activated by p53 [234]. Moreover, it has also been shown that MDM2 mediates the ubiquitination and proteasomal degradation of eS31 in response to nucleolar stress, indicating that free eS31 could be a physiological substrate of MDM2 [230]. It has been proposed that this mutual inhibitory regulation between MDM2 and eS31 may contribute to cellular recovery after the experienced stress [230]. Another report has suggested that the RAF PICT1 seems to regulate the interaction between eS31 and MDM2, as low levels of PICT1 are apparently required for the efficient translocation of eS31 from the nucleolus to the nucleoplasm, so that it can bind MDM2 [235]. It has also been described in several human lung cancer cell lines that eS31 could interact with uL5 in a way that might weaken the strength of the interaction between uL5 and MDM2; thus, knockdown of RPS27A stabilizes p53 in a uL5-dependent manner, promoting the p53 tumor suppressor functions [231]. Altogether, the above-mentioned data highlight that eS31 is connected to the MDM2-p53 axis and suggest that eS31 may be relevant for fine-tuning the cellular response to and recovery from nucleolar stress. However, its importance is apparently cell-type dependent, as recently shown by the Schneider group, who demonstrated that knockdown of RPS27A robustly induced p53 in certain cell lines but not in others [181].
Cancer cells require a high production of ribosomes to sustain boundless growth and cell division (e.g., [236]). Moreover, many r-proteins, including the beak ones, have been implicated in cancer development (e.g., [206,237,238,239]). Mutations and altered expression of beak r-proteins have been described in many cancer types, in some cases likely displaying an extra-ribosomal function: (i) high expression of RPS10 has been found in colorectal, renal and prostate cancer [240], whereas it has been reported that ribosomes purified from MDA-MB-231 breast cancer cells contain substoichiometric levels of eS10, among other r-proteins [119]. (ii) Overexpression of RPS12 has been observed in colon adenomatous polyps and carcinomas as well as in gastric cancers [241]. Deletions of RPS12 are frequently observed in diffuse large B cell lymphomas [242], and eS12 has been reported to play a role as a stimulator of WNT secretion in cancer cells, which is particularly important in the context of triple-negative breast cancer initiation and progression [243]. (iii) It has been reported that eS31 is overexpressed in renal, colon, cervical, and breast cancers, chronic myeloid leukemia and lung adenocarcinoma [161,231,244,245]. In most of these cases, the overexpression of eS31 correlates with poor prognosis for the patients. A recent review has highlighted the expression and role of eS31 in cancer cells and tumor tissues [161].

7. Concluding Remarks and Future Perspectives

Herein, we have discussed the relevance of the beak, a structurally conserved region of the small r-subunit of cytoplasmic ribosomes in all three domains of life. Notably, the beak has transitioned from a structure composed exclusively of rRNA in bacteria to a protuberance comprising three specific r-proteins, eS10, eS12 and eS31, in eukaryotes, while an intermediate situation prevails in archaea as the beaks of many species only contain two r-proteins, eS31 and eL8, with the latter being clearly the ancestor of the eukaryotic eS12 r-protein. As also discussed, the beak has diverged considerably in the mitoribosomes of some organisms, such as in trypanosomatids, perhaps owing to the particular translation requirements inside these organelles [246,247]. As outlined in this review, the beak has important roles in the three major phases of translation (initiation, elongation and termination) and is involved in many other events of the translation process, including ribosome stalling and collisions, as well as other seemingly translation-unrelated processes, such as cell competition in Drosophila. In many of these processes, the biological significance of post-translational modifications, such as ubiquitination, is still poorly understood. We have also highlighted the implication of the beak r-proteins in the stepwise assembly of nascent 40S r-subunits. Regarding the maturation of the eukaryotic beak, it is evident that further research is required to precisely define the assembly timing of the beak r-proteins; for instance, it is still controversial whether eS31 associates with 40S r-subunit maturation intermediates in the nucleus or the cytoplasm. More studies are also required to elucidate whether the beak r-proteins play active or passive roles during the assembly and nuclear export of pre-40S r-subunits. Finally, we discussed the relevance of the beak r-proteins in human diseases, especially ribosomopathies and cancer. A deeper understanding of the connection between these diseases and the ribosome biogenesis process is expected to offer new perspectives for therapeutic approaches.

Author Contributions

S.M.-V., C.V.G., C.R.-H., D.K. and J.d.l.C. wrote the manuscript. D.K. and J.d.l.C. edited the draft and the final version of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This publication is part of the project R+D+i PID2022-136564NB-I00 funded by MCIN/AEI/10.13039/501100011033/ERDF, EU to J.d.l.C. We also acknowledge the Andalusian Platform for Biomodels and Resources in Genomic Edition, the Fortalece Program (FORT 2023) from MCIN, and the Translacore (CA21154) and ProteoCure (CA20113) COST Actions from the EU for support. D. K. is supported by the Swiss National Science Foundation (project grant 310030_204801). S.M.-V. was an academic research staff member of the Andalusian Research, Development, and Innovation Plan by the Andalusian Regional Government (PAIDI 2020). C.V.G. was supported by a Marie Sklodowska-Curie Individual Fellowship (H2020-MSCA-IF-2020, Grant 101024158 from the EU). C.R.-H. is supported by a María Zambrano Grant from the Spanish Ministry of Universities.

Acknowledgments

We thank members of the de la Cruz laboratory for valuable discussions, critical reading, and suggestions.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Wilson, D.N.; Doudna Cate, J.H. The structure and function of the eukaryotic ribosome. Cold Spring Harb. Perspect. Biol. 2012, 4, a011536. [Google Scholar] [CrossRef] [PubMed]
  2. Melnikov, S.; Ben-Shem, A.; Garreau de Loubresse, N.; Jenner, L.; Yusupova, G.; Yusupov, M. One core, two shells: Bacterial and eukaryotic ribosomes. Nat. Struct. Mol. Biol. 2012, 19, 560–567. [Google Scholar] [CrossRef] [PubMed]
  3. Greber, B.J.; Ban, N. Structure and function of the mitochondrial ribosome. Annu. Rev. Biochem. 2016, 85, 103–132. [Google Scholar] [CrossRef] [PubMed]
  4. Robles, P.; Quesada, V. Unveiling the functions of plastid ribosomal proteins in plant development and abiotic stress tolerance. Plant Physiol. Biochem. 2022, 189, 35–45. [Google Scholar] [CrossRef] [PubMed]
  5. Sulima, S.O.; Dinman, J.D. The expanding riboverse. Cells 2019, 8, 1205. [Google Scholar] [CrossRef] [PubMed]
  6. Yusupova, G.; Yusupov, M. High-resolution structure of the eukaryotic 80S ribosome. Annu. Rev. Biochem. 2014, 83, 467–486. [Google Scholar] [CrossRef] [PubMed]
  7. Anger, A.M.; Armache, J.P.; Berninghausen, O.; Habeck, M.; Subklewe, M.; Wilson, D.N.; Beckmann, R. Structures of the human and Drosophila 80S ribosome. Nature 2013, 497, 80–85. [Google Scholar] [CrossRef] [PubMed]
  8. Klinge, S.; Voigts-Hoffmann, F.; Leibundgut, M.; Ban, N. Atomic structures of the eukaryotic ribosome. Trends Biochem. Sci. 2012, 37, 189–198. [Google Scholar] [CrossRef]
  9. Noeske, J.; Wasserman, M.R.; Terry, D.S.; Altman, R.B.; Blanchard, S.C.; Cate, J.H. High-resolution structure of the Escherichia coli ribosome. Nat. Struct. Mol. Biol. 2015, 22, 336–341. [Google Scholar] [CrossRef]
  10. Amunts, A.; Brown, A.; Toots, J.; Scheres, S.H.W.; Ramakrishnan, V. Ribosome. The structure of the human mitochondrial ribosome. Science 2015, 348, 95–98. [Google Scholar] [CrossRef]
  11. Lafontaine, D.L.; Tollervey, D. The function and synthesis of ribosomes. Nat. Rev. Mol. Cell Biol. 2001, 2, 514–520. [Google Scholar] [CrossRef] [PubMed]
  12. Bernier, C.R.; Petrov, A.S.; Kovacs, N.A.; Penev, P.I.; Williams, L.D. Translation: The universal structural core of life. Mol. Biol. Evol. 2018, 35, 2065–2076. [Google Scholar] [CrossRef] [PubMed]
  13. Agrawal, R.K.; Majumdar, S. Evolution: Mitochondrial ribosomes across species. Methods Mol. Biol. 2023, 2661, 7–21. [Google Scholar] [PubMed]
  14. Gonzalo, P.; Reboud, J.P. The puzzling lateral flexible stalk of the ribosome. Biol. Cell 2003, 95, 179–193. [Google Scholar] [CrossRef] [PubMed]
  15. Ballesta, J.P.G.; Remacha, M. The large ribosomal subunit stalk as a regulatory element of the eukaryotic translational machinery. Prog. Nucleic Acid Res. Mol. Biol. 1996, 55, 157–193. [Google Scholar] [PubMed]
  16. Diaconu, M.; Kothe, U.; Schlunzen, F.; Fischer, N.; Harms, J.M.; Tonevitsky, A.G.; Stark, H.; Rodnina, M.V.; Wahl, M.C. Structural basis for the function of the ribosomal L7/12 stalk in factor binding and GTPase activation. Cell 2005, 121, 991–1004. [Google Scholar] [CrossRef]
  17. Kulczyk, A.W.; Sorzano, C.O.S.; Grela, P.; Tchorzewski, M.; Tumer, N.E.; Li, X.P. Cryo-EM structure of Shiga toxin 2 in complex with the native ribosomal P-stalk reveals residues involved in the binding interaction. J. Biol. Chem. 2023, 299, 102795. [Google Scholar] [CrossRef]
  18. Fan, X.; Zhu, Y.; Wang, C.; Niu, L.; Teng, M.; Li, X. Structural insights into the interaction of the ribosomal P stalk protein P2 with a type II ribosome-inactivating protein ricin. Sci. Rep. 2016, 6, 37803. [Google Scholar] [CrossRef]
  19. Gupta, R.; Hinnebusch, A.G. Differential requirements for P stalk components in activating yeast protein kinase Gcn2 by stalled ribosomes during stress. Proc. Natl. Acad. Sci. USA 2023, 120, e2300521120. [Google Scholar] [CrossRef]
  20. Inglis, A.J.; Masson, G.R.; Shao, S.; Perisic, O.; McLaughlin, S.H.; Hegde, R.S.; Williams, R.L. Activation of GCN2 by the ribosomal P-stalk. Proc. Natl. Acad. Sci. USA 2019, 116, 4946–4954. [Google Scholar] [CrossRef]
  21. Trabuco, L.G.; Schreiner, E.; Eargle, J.; Cornish, P.; Ha, T.; Luthey-Schulten, Z.; Schulten, K. The role of L1 stalk-tRNA interaction in the ribosome elongation cycle. J. Mol. Biol. 2010, 402, 741–760. [Google Scholar] [CrossRef] [PubMed]
  22. Mohan, S.; Noller, H.F. Recurring RNA structural motifs underlie the mechanics of L1 stalk movement. Nat. Commun. 2017, 8, 14285. [Google Scholar] [CrossRef] [PubMed]
  23. Ghosh, A.; Komar, A.A. Eukaryote-specific extensions in ribosomal proteins of the small subunit: Structure and function. Translation 2015, 3, e999576. [Google Scholar] [CrossRef] [PubMed]
  24. Timsit, Y.; Sergeant-Perthuis, G.; Bennequin, D. Evolution of ribosomal protein network architectures. Sci. Rep. 2021, 11, 625. [Google Scholar] [CrossRef] [PubMed]
  25. Kisly, I.; Tamm, T. Archaea/eukaryote-specific ribosomal proteins-guardians of a complex structure. Comput. Struct. Biotechnol. J. 2023, 21, 1249–1261. [Google Scholar] [CrossRef] [PubMed]
  26. Liljas, A.; Sanyal, S. The enigmatic ribosomal stalk. Q. Rev. Biophys. 2018, 51, e12. [Google Scholar] [CrossRef] [PubMed]
  27. Zinker, S.; Warner, J.R. The ribosomal proteins of Saccharomyces cerevisiae. Phosphorylated and exchangeable proteins. J. Biol. Chem. 1976, 251, 1799–1807. [Google Scholar] [CrossRef] [PubMed]
  28. Tsurugi, K.; Ogata, K. Evidence for the exchangeability of acidic ribosomal proteins on cytoplasmic ribosomes in regenerating rat liver. J. Biochem. 1985, 98, 1427–1431. [Google Scholar] [CrossRef] [PubMed]
  29. Bautista-Santos, A.; Zinker, S. The P1/P2 protein heterodimers assemble to the ribosomal stalk at the moment when the ribosome Is committed to translation but not to the native 60S ribosomal subunit in Saccharomyces cerevisiae. Biochemistry 2014, 53, 4105–4112. [Google Scholar] [CrossRef]
  30. Remacha, M.; Jiménez-Díaz, A.; Bermejo, B.; Rodríguez-Gabriel, M.A.; Guarinos, E.; Ballesta, J.P.G. Ribosomal acidic phosphoproteins P1 and P2 are not required for cell viability but regulate the pattern of protein expression in Saccharomyces cerevisiae. Mol. Cell. Biol. 1995, 15, 4754–4762. [Google Scholar] [CrossRef]
  31. Sáenz-Robles, M.T.; Remacha, M.; Vilella, M.D.; Zinker, S.; Ballesta, J.P. The acidic ribosomal proteins as regulators of the eukaryotic ribosomal activity. Biochim. Biophys. Acta 1990, 1050, 51–55. [Google Scholar] [CrossRef] [PubMed]
  32. Derylo, K.; Michalec-Wawiorka, B.; Krokowski, D.; Wawiorka, L.; Hatzoglou, M.; Tchorzewski, M. The uL10 protein, a component of the ribosomal P-stalk, is released from the ribosome in nucleolar stress. Biochim. Biophys. Acta Mol. Cell Res. 2018, 1865, 34–47. [Google Scholar] [CrossRef] [PubMed]
  33. Siodmak, A.; Martínez-Seidel, F.; Rayapuram, N.; Bazin, J.; Alhoraibi, H.; Gentry-Torfer, D.; Tabassum, N.; Sheikh, A.H.; Kise, J.K.G.; Blilou, I.; et al. Dynamics of ribosome composition and ribosomal protein phosphorylation in immune signaling in Arabidopsis thaliana. Nucleic Acids Res. 2023, 51, 11876–11892. [Google Scholar] [CrossRef]
  34. Ballesta, J.P.G.; Rodriguez-Gabriel, M.A.; Bou, G.; Briones, E.; Zambrano, R.; Remacha, M. Phosphorylation of the yeast ribosomal stalk. Functional affects and enzymes involved in the process. FEMS Microbiol. Rev. 1999, 23, 537–550. [Google Scholar] [CrossRef] [PubMed]
  35. Zambrano, R.; Briones, E.; Remacha, M.; Ballesta, J.P. Phosphorylation of the acidic ribosomal P proteins in Saccharomyces cerevisiae: A reappraisal. Biochemistry 1997, 36, 14439–14446. [Google Scholar] [CrossRef] [PubMed]
  36. Subramanian, A.R.; van Duin, J. Exchange of individual ribosomal proteins between ribosomes as studied by heavy isotope-transfer experiments. Mol. Gen. Genet. 1977, 158, 1–9. [Google Scholar] [CrossRef] [PubMed]
  37. Dick, F.A.; Eisinger, D.P.; Trumpower, B.L. Exchangeability of Qsr1p, a large ribosomal subunit protein required for subunit joining, suggests a novel translational regulatory mechanism. FEBS Lett. 1997, 419, 1–3. [Google Scholar] [CrossRef] [PubMed]
  38. Yang, Y.M.; Jung, Y.; Abegg, D.; Adibekian, A.; Carroll, K.S.; Karbstein, K. Chaperone-directed ribosome repair after oxidative damage. Mol. Cell 2023, 83, 1527–1537.e1525. [Google Scholar] [CrossRef] [PubMed]
  39. Eisinger, D.P.; Dick, F.A.; Trumpower, B.L. Qsr1p, a 60S ribosomal subunit protein, is required for joining of 40S and 60S subunits. Mol. Cell. Biol. 1997, 17, 5136–5145. [Google Scholar] [CrossRef]
  40. Sulima, S.O.; Gulay, S.P.; Anjos, M.; Patchett, S.; Meskauskas, A.; Johnson, A.W.; Dinman, J.D. Eukaryotic rpL10 drives ribosomal rotation. Nucleic Acids Res. 2014, 42, 2049–2063. [Google Scholar] [CrossRef]
  41. Pollutri, D.; Penzo, M. Ribosomal protein L10: From function to dysfunction. Cells 2020, 9, 2503. [Google Scholar] [CrossRef]
  42. Adams, D.R.; Ron, D.; Kiely, P.A. RACK1, A multifaceted scaffolding protein: Structure and function. Cell Commun. Signal. 2011, 9, 22. [Google Scholar] [CrossRef] [PubMed]
  43. Singh, N.; Jindal, S.; Ghosh, A.; Komar, A.A. Communication between RACK1/Asc1 and uS3 (Rps3) is essential for RACK1/Asc1 function in yeast Saccharomyces cerevisiae. Gene 2019, 706, 69–76. [Google Scholar] [CrossRef]
  44. Nilsson, J.; Sengupta, J.; Frank, J.; Nissen, P. Regulation of eukaryotic translation by the RACK1 protein: A platform for signalling molecules on the ribosome. EMBO Rep. 2004, 5, 1137–1141. [Google Scholar] [CrossRef]
  45. Thompson, M.K.; Rojas-Duran, M.F.; Gangaramani, P.; Gilbert, W.V. The ribosomal protein Asc1/RACK1 is required for efficient translation of short mRNAs. eLife 2016, 5, e11154. [Google Scholar] [CrossRef] [PubMed]
  46. Gerbasi, V.R.; Weaver, C.M.; Hill, S.; Friedman, D.B.; Link, A.J. Yeast Asc1p and mammalian RACK1 are functionally orthologous core 40S ribosomal proteins that repress gene expression. Mol. Cell. Biol. 2004, 24, 8276–8287. [Google Scholar] [CrossRef]
  47. Gerbasi, V.R.; Browne, C.M.; Samir, P.; Shen, B.; Sun, M.; Hazelbaker, D.Z.; Galassie, A.C.; Frank, J.; Link, A.J. Critical role for Saccharomyces cerevisiae Asc1p in translational initiation at elevated temperatures. Proteomics 2018, 18, e1800208. [Google Scholar] [CrossRef]
  48. Ikeuchi, K.; Inada, T. Ribosome-associated Asc1/RACK1 is required for endonucleolytic cleavage induced by stalled ribosome at the 3′ end of nonstop mRNA. Sci. Rep. 2016, 6, 28234. [Google Scholar] [CrossRef]
  49. Wolf, A.S.; Grayhack, E.J. Asc1, homolog of human RACK1, prevents frameshifting in yeast by ribosomes stalled at CGA codon repeats. RNA 2015, 21, 935–945. [Google Scholar] [CrossRef] [PubMed]
  50. Johnson, A.G.; Lapointe, C.P.; Wang, J.; Corsepius, N.C.; Choi, J.; Fuchs, G.; Puglisi, J.D. RACK1 on and off the ribosome. RNA 2019, 25, 881–895. [Google Scholar] [CrossRef]
  51. Rachfall, N.; Schmitt, K.; Bandau, S.; Smolinski, N.; Ehrenreich, A.; Valerius, O.; Braus, G.H. RACK1/Asc1p, a ribosomal node in cellular signaling. Mol. Cell. Proteom. 2013, 12, 87–105. [Google Scholar] [CrossRef] [PubMed]
  52. Pulk, A.; Liiv, A.; Peil, L.; Maivali, U.; Nierhaus, K.; Remme, J. Ribosome reactivation by replacement of damaged proteins. Mol. Microbiol. 2010, 75, 801–814. [Google Scholar] [CrossRef] [PubMed]
  53. Yang, Y.M.; Karbstein, K. The chaperone Tsr2 regulates Rps26 release and reincorporation from mature ribosomes to enable a reversible, ribosome-mediated response to stress. Sci. Adv. 2022, 8, eabl4386. [Google Scholar] [CrossRef] [PubMed]
  54. Yang, Y.M.; Karbstein, K. Ribosome assembly and repair. Annu. Rev. Cell Dev. Biol. 2024. [Google Scholar] [CrossRef] [PubMed]
  55. Ferretti, M.B.; Ghalei, H.; Ward, E.A.; Potts, E.L.; Karbstein, K. Rps26 directs mRNA-specific translation by recognition of Kozak sequence elements. Nat. Struct. Mol. Biol. 2017, 24, 700–707. [Google Scholar] [CrossRef]
  56. Fusco, C.M.; Desch, K.; Dorrbaum, A.R.; Wang, M.; Staab, A.; Chan, I.C.W.; Vail, E.; Villeri, V.; Langer, J.D.; Schuman, E.M. Neuronal ribosomes exhibit dynamic and context-dependent exchange of ribosomal proteins. Nat. Commun. 2021, 12, 6127. [Google Scholar] [CrossRef] [PubMed]
  57. Leiva, L.E.; Katz, A. Regulation of leaderless mRNA translation in bacteria. Microorganisms 2022, 10, 723. [Google Scholar] [CrossRef] [PubMed]
  58. Kaberdina, A.C.; Szaflarski, W.; Nierhaus, K.H.; Moll, I. An unexpected type of ribosomes induced by kasugamycin: A look into ancestral times of protein synthesis? Mol. Cell 2009, 33, 227–236. [Google Scholar] [CrossRef] [PubMed]
  59. Castillo Duque de Estrada, N.M.; Thoms, M.; Flemming, D.; Hammaren, H.M.; Buschauer, R.; Ameismeier, M.; Bassler, J.; Beck, M.; Beckmann, R.; Hurt, E. Structure of nascent 5S RNPs at the crossroad between ribosome assembly and MDM2-p53 pathways. Nat. Struct. Mol. Biol. 2023, 30, 1119–1131. [Google Scholar] [CrossRef] [PubMed]
  60. Lau, B.; Huang, Z.; Kellner, N.; Niu, S.; Berninghausen, O.; Beckmann, R.; Hurt, E.; Cheng, J. Mechanism of 5S RNP recruitment and helicase-surveilled rRNA maturation during pre-60S biogenesis. EMBO Rep. 2023, 24, e56910. [Google Scholar] [CrossRef]
  61. Leidig, C.; Thoms, M.; Holdermann, I.; Bradatsch, B.; Berninghausen, O.; Bange, G.; Sinning, I.; Hurt, E.; Beckmann, R. 60S ribosome biogenesis requires rotation of the 5S ribonucleoprotein particle. Nat. Commun. 2014, 5, 3491. [Google Scholar] [CrossRef]
  62. Gómez-Herreros, F.; Rodríguez-Galán, O.; Morillo-Huesca, M.; Maya, D.; Arista-Romero, M.; de la Cruz, J.; Chávez, S.; Muñoz-Centeno, M.C. Balanced production of ribosome components is required for proper G1/S transition in Saccharomyces cerevisiae. J. Biol. Chem. 2013, 288, 31689–31700. [Google Scholar] [CrossRef] [PubMed]
  63. Sloan, K.E.; Bohnsack, M.T.; Watkins, N.J. The 5S RNP couples p53 homeostasis to ribosome biogenesis and nucleolar stress. Cell Rep. 2013, 5, 237–247. [Google Scholar] [CrossRef] [PubMed]
  64. Donati, G.; Peddigari, S.; Mercer, C.A.; Thomas, G. 5S ribosomal RNA is an essential component of a nascent ribosomal precursor complex that regulates the Hdm2-p53 checkpoint. Cell Rep. 2013, 4, 87–98. [Google Scholar] [CrossRef] [PubMed]
  65. Gentilella, A.; Morón-Durán, F.D.; Fuentes, P.; Zweig-Rocha, G.; Riano-Canalias, F.; Pelletier, J.; Ruiz, M.; Turón, G.; Castano, J.; Tauler, A.; et al. Autogenous control of 5′TOP mRNA stability by 40S ribosomes. Mol. Cell 2017, 67, 55–70.e54. [Google Scholar] [CrossRef] [PubMed]
  66. Eastham, M.J.; Pelava, A.; Wells, G.R.; Lee, J.K.; Lawrence, I.R.; Stewart, J.; Deichner, M.; Hertle, R.; Watkins, N.J.; Schneider, C. The induction of p53 correlates with defects in the production, but not the levels, of the small ribosomal subunit and stalled large ribosomal subunit biogenesis. Nucleic Acids Res. 2023, 51, 9397–9414. [Google Scholar] [CrossRef] [PubMed]
  67. Zheng, J.; Lang, Y.; Zhang, Q.; Cui, D.; Sun, H.; Jiang, L.; Chen, Z.; Zhang, R.; Gao, Y.; Tian, W.; et al. Structure of human MDM2 complexed with RPL11 reveals the molecular basis of p53 activation. Genes Dev. 2015, 29, 1524–1534. [Google Scholar] [CrossRef] [PubMed]
  68. Pelava, A.; Schneider, C.; Watkins, N.J. The importance of ribosome production, and the 5S RNP-MDM2 pathway, in health and disease. Biochem. Soc. Trans. 2016, 44, 1086–1090. [Google Scholar] [CrossRef] [PubMed]
  69. Chen, J. The cell-cycle arrest and apoptotic functions of p53 in tumor initiation and progression. Cold Spring Harb. Perspect. Med. 2016, 6, a026104. [Google Scholar] [CrossRef] [PubMed]
  70. Bustelo, X.R.; Dosil, M. Ribosome biogenesis and cancer: Basic and translational challenges. Curr. Opin. Genet. Dev. 2018, 48, 22–29. [Google Scholar] [CrossRef]
  71. Penzo, M.; Montanaro, L.; Trere, D.; Derenzini, M. The ribosome biogenesis-cancer connection. Cells 2019, 8, 55. [Google Scholar] [CrossRef] [PubMed]
  72. Lindström, M.S.; Bartek, J.; Maya-Mendoza, A. p53 at the crossroad of DNA replication and ribosome biogenesis stress pathways. Cell Death Differ. 2022, 29, 972–982. [Google Scholar] [CrossRef] [PubMed]
  73. Wimberly, B.T.; Brodersen, D.E.; Clemons, W.M.; Morgan-Warren, R.J.; Carter, A.P.; Vonrhein, C.; Hartsch, T.; Ramakrishnan, V. Structure of the 30S ribosomal subunit. Nature 2000, 407, 327–339. [Google Scholar] [CrossRef] [PubMed]
  74. Schluenzen, F.; Tocilj, A.; Zarivach, R.; Harms, J.; Bashan, A.; Bartels, H.; Agmon, I.; Franceschi, F.; Yonath, A. Structure of functional activated small ribosomal subunit at 3.3 Å resolution. Cell 2000, 102, 615–623. [Google Scholar] [CrossRef]
  75. Londei, P.; Ferreira-Cerca, S. Ribosome biogenesis in archaea. Front Microbiol. 2021, 12, 686977. [Google Scholar] [CrossRef] [PubMed]
  76. Fernández-Pevida, A.; Martín-Villanueva, S.; Murat, G.; Lacombe, T.; Kressler, D.; de la Cruz, J. The eukaryote-specific N-terminal extension of ribosomal protein S31 contributes to the assembly and function of 40S ribosomal subunits. Nucleic Acids Res. 2016, 44, 7777–7791. [Google Scholar] [CrossRef] [PubMed]
  77. Lecompte, O.; Ripp, R.; Thierry, J.C.; Moras, D.; Poch, O. Comparative analysis of ribosomal proteins in complete genomes: An example of reductive evolution at the domain scale. Nucleic Acids Res. 2002, 30, 5382–5390. [Google Scholar] [CrossRef] [PubMed]
  78. Armache, J.P.; Anger, A.M.; Marquez, V.; Franckenberg, S.; Frohlich, T.; Villa, E.; Berninghausen, O.; Thomm, M.; Arnold, G.J.; Beckmann, R.; et al. Promiscuous behaviour of archaeal ribosomal proteins: Implications for eukaryotic ribosome evolution. Nucleic Acids Res. 2013, 41, 1284–1293. [Google Scholar] [CrossRef] [PubMed]
  79. Mendler, K.; Chen, H.; Parks, D.H.; Lobb, B.; Hug, L.A.; Doxey, A.C. AnnoTree: Visualization and exploration of a functionally annotated microbial tree of life. Nucleic Acids Res. 2019, 47, 4442–4448. [Google Scholar] [CrossRef]
  80. Manuell, A.L.; Quispe, J.; Mayfield, S.P. Structure of the chloroplast ribosome: Novel domains for translation regulation. PLoS Biol. 2007, 5, e209. [Google Scholar] [CrossRef]
  81. Bieri, P.; Leibundgut, M.; Saurer, M.; Boehringer, D.; Ban, N. The complete structure of the chloroplast 70S ribosome in complex with translation factor pY. EMBO J. 2017, 36, 475–486. [Google Scholar] [CrossRef] [PubMed]
  82. Perez-Boerema, A.; Aibara, S.; Paul, B.; Tobiasson, V.; Kimanius, D.; Forsberg, B.O.; Wallden, K.; Lindahl, E.; Amunts, A. Structure of the chloroplast ribosome with chl-RRF and hibernation-promoting factor. Nat. Plants 2018, 4, 212–217. [Google Scholar] [CrossRef]
  83. Desai, N.; Brown, A.; Amunts, A.; Ramakrishnan, V. The structure of the yeast mitochondrial ribosome. Science 2017, 355, 528–531. [Google Scholar] [CrossRef]
  84. Scaltsoyiannes, V.; Corre, N.; Waltz, F.; Giegé, P. Types and functions of mitoribosome-specific ribosomal proteins across eukaryotes. Int. J. Mol. Sci. 2022, 23, 3474. [Google Scholar] [CrossRef] [PubMed]
  85. Steffen, K.K.; McCormick, M.A.; Pham, K.M.; Mackay, V.L.; Delaney, J.R.; Murakami, C.J.; Kaeberlein, M.; Kennedy, B.K. Ribosome deficiency protects against ER stress in Saccharomyces cerevisiae. Genetics 2012, 191, 107–118. [Google Scholar] [CrossRef] [PubMed]
  86. Ferreira-Cerca, S.; Pöll, G.; Gleizes, P.E.; Tschochner, H.; Milkereit, P. Roles of eukaryotic ribosomal proteins in maturation and transport of pre-18S rRNA and ribosome function. Mol. Cell 2005, 20, 263–275. [Google Scholar] [CrossRef] [PubMed]
  87. Ben-Shem, A.; Garreau de Loubresse, N.; Melnikov, S.; Jenner, L.; Yusupova, G.; Yusupov, M. The structure of the eukaryotic ribosome at 3.0 Å resolution. Science 2011, 334, 1524–1529. [Google Scholar] [CrossRef]
  88. Martín-Villanueva, S.; Fernández-Fernández, J.; Rodríguez-Galán, O.; Fernández-Boraita, J.; Villalobo, E.; de la Cruz, J. Role of the 40S beak ribosomal protein eS12 in ribosome biogenesis and function in Saccharomyces cerevisiae. RNA Biol. 2020, 17, 1261–1276. [Google Scholar] [CrossRef]
  89. Lacombe, T.; García-Gómez, J.J.; de la Cruz, J.; Roser, D.; Hurt, E.; Linder, P.; Kressler, D. Linear ubiquitin fusion to Rps31 and its subsequent cleavage are required for the efficient production and functional integrity of 40S ribosomal subunits. Mol. Microbiol. 2009, 72, 69–84. [Google Scholar] [CrossRef]
  90. Rössler, I.; Weigl, S.; Fernández-Fernández, J.; Martín-Villanueva, S.; Strauss, D.; Hurt, E.; de la Cruz, J.; Pertschy, B. The C-terminal tail of ribosomal protein Rps15 is engaged in cytoplasmic pre-40S maturation. RNA Biol. 2022, 19, 560–574. [Google Scholar] [CrossRef]
  91. Martín-Villanueva, S.; Gutiérrez, G.; Kressler, D.; de la Cruz, J. Ubiquitin and Ubiquitin-Like proteins and domains in ribosome production and function: Chance or necessity? Int. J. Mol. Sci. 2021, 22, 4359. [Google Scholar] [CrossRef] [PubMed]
  92. Finley, D.; Bartel, B.; Varshavsky, A. The tails of ubiquitin precursors are ribosomal proteins whose fusion to ubiquitin facilitates ribosome biogenesis. Nature 1989, 338, 394–401. [Google Scholar] [CrossRef] [PubMed]
  93. Polymenis, M. Ribosomal proteins: Mutant phenotypes by the numbers and associated gene expression changes. Open Biol. 2020, 10, 200114. [Google Scholar] [CrossRef] [PubMed]
  94. Shasmal, M.; Dey, S.; Shaikh, T.R.; Bhakta, S.; Sengupta, J. E. coli metabolic protein aldehyde-alcohol dehydrogenase-E binds to the ribosome: A unique moonlighting action revealed. Sci. Rep. 2016, 6, 19936. [Google Scholar] [CrossRef] [PubMed]
  95. Qu, X.; Wen, J.D.; Lancaster, L.; Noller, H.F.; Bustamante, C.; Tinoco, I., Jr. The ribosome uses two active mechanisms to unwind messenger RNA during translation. Nature 2011, 475, 118–121. [Google Scholar] [CrossRef] [PubMed]
  96. Valentini, M.; Linder, P. Happy birthday: 30 years of RNA helicases. Methods Mol. Biol. 2021, 2209, 17–34. [Google Scholar] [PubMed]
  97. Yourik, P.; Aitken, C.E.; Zhou, F.; Gupta, N.; Hinnebusch, A.G.; Lorsch, J.R. Yeast eIF4A enhances recruitment of mRNAs regardless of their structural complexity. eLife 2017, 6, e31476. [Google Scholar] [CrossRef] [PubMed]
  98. Parsyan, A.; Svitkin, Y.; Shahbazian, D.; Gkogkas, C.; Lasko, P.; Merrick, W.C.; Sonenberg, N. mRNA helicases: The tacticians of translational control. Nat. Rev. Mol. Cell Biol. 2011, 12, 235–245. [Google Scholar] [CrossRef] [PubMed]
  99. Pisareva, V.P.; Pisarev, A.V. DHX29 and eIF3 cooperate in ribosomal scanning on structured mRNAs during translation initiation. RNA 2016, 22, 1859–1870. [Google Scholar] [CrossRef]
  100. Pisareva, V.P.; Pisarev, A.V.; Komar, A.A.; Hellen, C.U.; Pestova, T.V. Translation initiation on mammalian mRNAs with structured 5′UTRs requires DExH-box protein DHX29. Cell 2008, 135, 1237–1250. [Google Scholar] [CrossRef]
  101. Hashem, Y.; des Georges, A.; Dhote, V.; Langlois, R.; Liao, H.Y.; Grassucci, R.A.; Hellen, C.U.; Pestova, T.V.; Frank, J. Structure of the mammalian ribosomal 43S preinitiation complex bound to the scanning factor DHX29. Cell 2013, 153, 1108–1119. [Google Scholar] [CrossRef]
  102. Haimov, O.; Sinvani, H.; Martin, F.; Ulitsky, I.; Emmanuel, R.; Tamarkin-Ben-Harush, A.; Vardy, A.; Dikstein, R. Efficient and accurate translation initiation directed by TISU involves RPS3 and RPS10e binding and differential eukaryotic initiation factor 1A regulation. Mol. Cell. Biol. 2017, 37, e00150-00117. [Google Scholar] [CrossRef]
  103. Passmore, L.A.; Schmeing, T.M.; Maag, D.; Applefield, D.J.; Acker, M.G.; Algire, M.A.; Lorsch, J.R.; Ramakrishnan, V. The eukaryotic translation initiation factors eIF1 and eIF1A induce an open conformation of the 40S ribosome. Mol. Cell 2007, 26, 41–50. [Google Scholar] [CrossRef]
  104. Weisser, M.; Voigts-Hoffmann, F.; Rabl, J.; Leibundgut, M.; Ban, N. The crystal structure of the eukaryotic 40S ribosomal subunit in complex with eIF1 and eIF1A. Nat. Struct. Mol. Biol. 2013, 20, 1015–1017. [Google Scholar] [CrossRef] [PubMed]
  105. Sehrawat, U.; Koning, F.; Ashkenazi, S.; Stelzer, G.; Leshkowitz, D.; Dikstein, R. Cancer-associated eukaryotic translation initiation factor 1A mutants impair Rps3 and Rps10 binding and enhance scanning of cell cycle genes. Mol. Cell. Biol. 2019, 39, e00441-00418. [Google Scholar] [CrossRef] [PubMed]
  106. Martin-Marcos, P.; Zhou, F.; Karunasiri, C.; Zhang, F.; Dong, J.; Nanda, J.; Kulkarni, S.D.; Sen, N.D.; Tamame, M.; Zeschnigk, M.; et al. eIF1A residues implicated in cancer stabilize translation preinitiation complexes and favor suboptimal initiation sites in yeast. eLife 2017, 6, e31250. [Google Scholar] [CrossRef]
  107. Fekete, C.A.; Mitchell, S.F.; Cherkasova, V.A.; Applefield, D.; Algire, M.A.; Maag, D.; Saini, A.K.; Lorsch, J.R.; Hinnebusch, A.G. N- and C-terminal residues of eIF1A have opposing effects on the fidelity of start codon selection. EMBO J. 2007, 26, 1602–1614. [Google Scholar] [CrossRef]
  108. Zeman, J.; Itoh, Y.; Kukačka, Z.; Rosůlek, M.; Kavan, D.; Kouba, T.; Jansen, M.E.; Mohammad, M.P.; Novák, P.; Valášek, L.S. Binding of eIF3 in complex with eIF5 and eIF1 to the 40S ribosomal subunit is accompanied by dramatic structural changes. Nucleic Acids Res. 2019, 47, 8282–8300. [Google Scholar] [CrossRef] [PubMed]
  109. Poncová, K.; Wagner, S.; Jansen, M.E.; Beznosková, P.; Gunisova, S.; Herrmannova, A.; Zeman, J.; Dong, J.; Valášek, L.S. uS3/Rps3 controls fidelity of translation termination and programmed stop codon readthrough in co-operation with eIF3. Nucleic Acids Res. 2019, 47, 11326–11343. [Google Scholar] [CrossRef] [PubMed]
  110. Malygin, A.A.; Shatsky, I.N.; Karpova, G.G. Proteins of the human 40S ribosomal subunit involved in hepatitis C IRES binding as revealed from fluorescent labeling. Biochemistry 2013, 78, 53–59. [Google Scholar] [CrossRef]
  111. Tidu, A.; Janvier, A.; Schaeffer, L.; Sosnowski, P.; Kuhn, L.; Hammann, P.; Westhof, E.; Eriani, G.; Martin, F. The viral protein NSP1 acts as a ribosome gatekeeper for shutting down host translation and fostering SARS-CoV-2 translation. RNA 2020, 27, 253–264. [Google Scholar] [CrossRef] [PubMed]
  112. Thoms, M.; Buschauer, R.; Ameismeier, M.; Koepke, L.; Denk, T.; Hirschenberger, M.; Kratzat, H.; Hayn, M.; Mackens-Kiani, T.; Cheng, J.; et al. Structural basis for translational shutdown and immune evasion by the Nsp1 protein of SARS-CoV-2. Science 2020, 369, 1249–1255. [Google Scholar] [CrossRef] [PubMed]
  113. Schubert, K.; Karousis, E.D.; Jomaa, A.; Scaiola, A.; Echeverria, B.; Gurzeler, L.A.; Leibundgut, M.; Thiel, V.; Muhlemann, O.; Ban, N. SARS-CoV-2 Nsp1 binds the ribosomal mRNA channel to inhibit translation. Nat. Struct. Mol. Biol. 2020, 27, 959–966. [Google Scholar] [CrossRef] [PubMed]
  114. Schubert, K.; Karousis, E.D.; Ban, I.; Lapointe, C.P.; Leibundgut, M.; Baumlin, E.; Kummerant, E.; Scaiola, A.; Schonhut, T.; Ziegelmuller, J.; et al. Universal features of Nsp1-mediated translational shutdown by coronaviruses. Mol. Cell 2023, 83, 3546–3557.e3548. [Google Scholar] [CrossRef] [PubMed]
  115. Bujanic, L.; Shevchuk, O.; von Kugelgen, N.; Kalinina, A.; Ludwik, K.; Koppstein, D.; Zerna, N.; Sickmann, A.; Chekulaeva, M. The key features of SARS-CoV-2 leader and NSP1 required for viral escape of NSP1-mediated repression. RNA 2022, 28, 766–779. [Google Scholar] [CrossRef] [PubMed]
  116. Brown, A.; Baird, M.R.; Yip, M.C.; Murray, J.; Shao, S. Structures of translationally inactive mammalian ribosomes. eLife 2018, 7, e40486. [Google Scholar] [CrossRef] [PubMed]
  117. Spahn, C.M.; Gomez-Lorenzo, M.G.; Grassucci, R.A.; Jorgensen, R.; Andersen, G.R.; Beckmann, R.; Penczek, P.A.; Ballesta, J.P.; Frank, J. Domain movements of elongation factor eEF2 and the eukaryotic 80S ribosome facilitate tRNA translocation. EMBO J. 2004, 23, 1008–1119. [Google Scholar] [CrossRef] [PubMed]
  118. Hassan, A.; Byju, S.; Freitas, F.C.; Roc, C.; Pender, N.; Nguyen, K.; Kimbrough, E.M.; Mattingly, J.M.; Gonzalez, R.L., Jr.; de Oliveira, R.J.; et al. Ratchet, swivel, tilt and roll: A complete description of subunit rotation in the ribosome. Nucleic Acids Res. 2023, 51, 919–934. [Google Scholar] [CrossRef] [PubMed]
  119. Collins, J.C.; Ghalei, H.; Doherty, J.R.; Huang, H.; Culver, R.N.; Karbstein, K. Ribosome biogenesis factor Ltv1 chaperones the assembly of the small subunit head. J. Cell Biol. 2018, 217, 4141–4154. [Google Scholar] [CrossRef]
  120. Shao, S.; Murray, J.; Brown, A.; Taunton, J.; Ramakrishnan, V.; Hegde, R.S. Decoding mammalian ribosome-mRNA states by translational GTPase complexes. Cell 2016, 167, 1229–1240.e1215. [Google Scholar] [CrossRef]
  121. Taylor, D.; Unbehaun, A.; Li, W.; Das, S.; Lei, J.; Liao, H.Y.; Grassucci, R.A.; Pestova, T.V.; Frank, J. Cryo-EM structure of the mammalian eukaryotic release factor eRF1-eRF3-associated termination complex. Proc. Natl. Acad. Sci. USA 2012, 109, 18413–18418. [Google Scholar] [CrossRef]
  122. Preis, A.; Heuer, A.; Barrio-Garcia, C.; Hauser, A.; Eyler, D.E.; Berninghausen, O.; Green, R.; Becker, T.; Beckmann, R. Cryoelectron microscopic structures of eukaryotic translation termination complexes containing eRF1-eRF3 or eRF1-ABCE1. Cell Rep. 2014, 8, 59–65. [Google Scholar] [CrossRef]
  123. Liu, B.; Qian, S.B. Translational reprogramming in cellular stress response. Wiley Interdiscip. Rev. RNA 2014, 5, 301–315. [Google Scholar] [CrossRef]
  124. Roux, P.P.; Topisirovic, I. Signaling pathways involved in the regulation of mRNA translation. Mol. Cell. Biol. 2018, 38, e00070-00018. [Google Scholar] [CrossRef]
  125. Silva, G.M.; Finley, D.; Vogel, C. K63 polyubiquitination is a new modulator of the oxidative stress response. Nat. Struct. Mol. Biol. 2015, 22, 116–123. [Google Scholar] [CrossRef]
  126. Simoes, V.; Cizubu, B.K.; Harley, L.; Zhou, Y.; Pajak, J.; Snyder, N.A.; Bouvette, J.; Borgnia, M.J.; Arya, G.; Bartesaghi, A.; et al. Redox-sensitive E2 Rad6 controls cellular response to oxidative stress via K63-linked ubiquitination of ribosomes. Cell Rep. 2022, 39, 110860. [Google Scholar] [CrossRef] [PubMed]
  127. Back, S.; Gorman, A.W.; Vogel, C.; Silva, G.M. Site-specific K63 ubiquitinomics provides insights into translation regulation under stress. J. Proteome Res. 2019, 18, 309–318. [Google Scholar] [CrossRef] [PubMed]
  128. Zhou, Y.; Kastritis, P.L.; Dougherty, S.E.; Bouvette, J.; Hsu, A.L.; Burbaum, L.; Mosalaganti, S.; Pfeffer, S.; Hagen, W.J.H.; Forster, F.; et al. Structural impact of K63 ubiquitin on yeast translocating ribosomes under oxidative stress. Proc. Natl. Acad. Sci. USA 2020, 117, 22157–22166. [Google Scholar] [CrossRef] [PubMed]
  129. Inada, T.; Beckmann, R. Mechanisms of translation-coupled quality control. J. Mol. Biol. 2024, 436, 168496. [Google Scholar] [CrossRef] [PubMed]
  130. Filbeck, S.; Cerullo, F.; Pfeffer, S.; Joazeiro, C.A.P. Ribosome-associated quality-control mechanisms from bacteria to humans. Mol. Cell 2022, 82, 1451–1466. [Google Scholar] [CrossRef]
  131. Guyomar, C.; D’Urso, G.; Chat, S.; Giudice, E.; Gillet, R. Structures of tmRNA and SmpB as they transit through the ribosome. Nat. Commun. 2021, 12, 4909. [Google Scholar] [CrossRef]
  132. Joazeiro, C.A.P. Ribosomal stalling during translation: Providing substrates for ribosome-associated protein quality control. Annu. Rev. Cell Dev. Biol. 2017, 33, 343–368. [Google Scholar] [CrossRef] [PubMed]
  133. Powers, K.T.; Szeto, J.A.; Schaffitzel, C. New insights into no-go, non-stop and nonsense-mediated mRNA decay complexes. Curr. Opin. Struct. Biol. 2020, 65, 110–118. [Google Scholar] [CrossRef]
  134. Araki, Y.; Takahashi, S.; Kobayashi, T.; Kajiho, H.; Hoshino, S.; Katada, T. Ski7p G protein interacts with the exosome and the Ski complex for 3′-to-5′ mRNA decay in yeast. EMBO J. 2001, 20, 4684–4693. [Google Scholar] [CrossRef] [PubMed]
  135. Kalisiak, K.; Kulinski, T.M.; Tomecki, R.; Cysewski, D.; Pietras, Z.; Chlebowski, A.; Kowalska, K.; Dziembowski, A. A short splicing isoform of HBS1L links the cytoplasmic exosome and SKI complexes in humans. Nucleic Acids Res. 2017, 45, 2068–2080. [Google Scholar] [CrossRef]
  136. Keidel, A.; Kögel, A.; Reichelt, P.; Kowalinski, E.; Schafer, I.B.; Conti, E. Concerted structural rearrangements enable RNA channeling into the cytoplasmic Ski238-Ski7-exosome assembly. Mol. Cell 2023, 83, 4093–4105.e4097. [Google Scholar] [CrossRef]
  137. Zinder, J.C.; Lima, C.D. Targeting RNA for processing or destruction by the eukaryotic RNA exosome and its cofactors. Genes Dev. 2017, 31, 88–100. [Google Scholar] [CrossRef] [PubMed]
  138. Kowalinski, E.; Schuller, A.; Green, R.; Conti, E. Saccharomyces cerevisiae Ski7 Is a GTP-Binding protein adopting the characteristic conformation of active translational GTPases. Structure 2015, 23, 1336–1343. [Google Scholar] [CrossRef]
  139. Kögel, A.; Keidel, A.; Bonneau, F.; Schafer, I.B.; Conti, E. The human SKI complex regulates channeling of ribosome-bound RNA to the exosome via an intrinsic gatekeeping mechanism. Mol. Cell 2022, 82, 756–769.e758. [Google Scholar] [CrossRef]
  140. Schmidt, C.; Kowalinski, E.; Shanmuganathan, V.; Defenouillère, Q.; Braunger, K.; Heuer, A.; Pech, M.; Namane, A.; Berninghausen, O.; Fromont-Racine, M.; et al. The cryo-EM structure of a ribosome-Ski2-Ski3-Ski8 helicase complex. Science 2016, 354, 1431–1433. [Google Scholar] [CrossRef]
  141. De, S.; Mühlemann, O. A comprehensive coverage insurance for cells: Revealing links between ribosome collisions, stress responses and mRNA surveillance. RNA Biol. 2022, 19, 609–621. [Google Scholar] [CrossRef]
  142. Iyer, K.V.; Müller, M.; Tittel, L.S.; Winz, M.L. Molecular Highway Patrol for Ribosome Collisions. Chembiochem 2023, 24, e202300264. [Google Scholar] [CrossRef]
  143. Matsuo, Y.; Ikeuchi, K.; Saeki, Y.; Iwasaki, S.; Schmidt, C.; Udagawa, T.; Sato, F.; Tsuchiya, H.; Becker, T.; Tanaka, K.; et al. Ubiquitination of stalled ribosome triggers ribosome-associated quality control. Nat. Commun. 2017, 8, 159. [Google Scholar] [CrossRef] [PubMed]
  144. Dougherty, S.E.; Maduka, A.O.; Inada, T.; Silva, G.M. Expanding role of ubiquitin in translational control. Int. J. Mol. Sci. 2020, 21, 1151. [Google Scholar] [CrossRef] [PubMed]
  145. Wu, C.C.; Peterson, A.; Zinshteyn, B.; Regot, S.; Green, R. Ribosome collisions trigger general stress responses to regulate cell fate. Cell 2020, 182, 404–416.e414. [Google Scholar] [CrossRef] [PubMed]
  146. Yan, L.L.; Zaher, H.S. Ribosome quality control antagonizes the activation of the integrated stress response on colliding ribosomes. Mol. Cell 2021, 81, 614–628.e614. [Google Scholar] [CrossRef]
  147. Pochopien, A.A.; Beckert, B.; Kasvandik, S.; Berninghausen, O.; Beckmann, R.; Tenson, T.; Wilson, D.N. Structure of Gcn1 bound to stalled and colliding 80S ribosomes. Proc. Natl. Acad. Sci. USA 2021, 118, e2022756118. [Google Scholar] [CrossRef]
  148. Lee, S.J.; Swanson, M.J.; Sattlegger, E. Gcn1 contacts the small ribosomal protein Rps10, which is required for full activation of the protein kinase Gcn2. Biochem. J. 2015, 466, 547–559. [Google Scholar] [CrossRef]
  149. Mueller, P.P.; Grueter, P.; Hinnebusch, A.G.; Trachsel, H. A ribosomal protein is required for translational regulation of GCN4 mRNA. Evidence for involvement of the ribosome in eIF2 recycling. J. Biol. Chem. 1998, 273, 32870–32877. [Google Scholar] [CrossRef]
  150. Oltion, K.; Carelli, J.D.; Yang, T.; See, S.K.; Wang, H.Y.; Kampmann, M.; Taunton, J. An E3 ligase network engages GCN1 to promote the degradation of translation factors on stalled ribosomes. Cell 2023, 186, 346–362.e317. [Google Scholar] [CrossRef]
  151. 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. The ribosomal protein genes and Minute loci of Drosophila melanogaster. Genome Biol. 2007, 8, R216. [Google Scholar] [CrossRef] [PubMed]
  152. Stirnberg, P.; Liu, J.P.; Ward, S.; Kendall, S.L.; Leyser, O. Mutation of the cytosolic ribosomal protein-encoding RPS10B gene affects shoot meristematic function in Arabidopsis. BMC Plant Biol. 2012, 12, 160. [Google Scholar] [CrossRef] [PubMed]
  153. Islam, R.A.; Rallis, C. Ribosomal biogenesis and heterogeneity in development, disease, and aging. Epigenomes 2023, 7, 26. [Google Scholar] [CrossRef] [PubMed]
  154. Yanagi, S.; Iida, T.; Kobayashi, T. RPS12 and UBC4 are related to senescence signal production in the ribosomal RNA gene cluster. Mol. Cell. Biol. 2022, 42, e0002822. [Google Scholar] [CrossRef] [PubMed]
  155. Morata, G. Cell competition: A historical perspective. Dev. Biol. 2021, 476, 33–40. [Google Scholar] [CrossRef] [PubMed]
  156. Kiparaki, M.; Baker, N.E. Ribosomal protein mutations and cell competition: Autonomous and nonautonomous effects on a stress response. Genetics 2023, 224, iyad080. [Google Scholar] [CrossRef] [PubMed]
  157. Baker, N.E. Emerging mechanisms of cell competition. Nat. Rev. Genet. 2020, 21, 683–697. [Google Scholar] [CrossRef] [PubMed]
  158. Kale, A.; Ji, Z.; Kiparaki, M.; Blanco, J.; Rimesso, G.; Flibotte, S.; Baker, N.E. Ribosomal protein S12e has a distinct function in cell competition. Dev. Cell 2018, 44, 42–55.e44. [Google Scholar] [CrossRef] [PubMed]
  159. Lee, C.H.; Kiparaki, M.; Blanco, J.; Folgado, V.; Ji, Z.; Kumar, A.; Rimesso, G.; Baker, N.E. A regulatory response to ribosomal protein mutations controls translation, growth, and cell competition. Dev. Cell 2018, 46, 456–469.e454. [Google Scholar] [CrossRef]
  160. Ji, Z.; Kiparaki, M.; Folgado, V.; Kumar, A.; Blanco, J.; Rimesso, G.; Chuen, J.; Liu, Y.; Zheng, D.; Baker, N.E. Drosophila RpS12 controls translation, growth, and cell competition through Xrp1. PLoS Genet. 2019, 15, e1008513. [Google Scholar] [CrossRef]
  161. Luo, J.; Zhao, H.; Chen, L.; Liu, M. Multifaceted functions of RPS27a: An unconventional ribosomal protein. J. Cell. Physiol. 2023, 238, 485–497. [Google Scholar] [CrossRef] [PubMed]
  162. Hong, S.W.; Kim, S.M.; Jin, D.H.; Kim, Y.S.; Hur, D.Y. RPS27a enhances EBV-encoded LMP1-mediated proliferation and invasion by stabilizing of LMP1. Biochem. Biophys. Res. Commun. 2017, 491, 303–309. [Google Scholar] [CrossRef] [PubMed]
  163. Bansal, S.K.; Gupta, N.; Sankhwar, S.N.; Rajender, S. Differential genes expression between fertile and infertile spermatozoa revealed by transcriptome analysis. PLoS ONE 2015, 10, e0127007. [Google Scholar] [CrossRef] [PubMed]
  164. Bäurle, I.; Laux, T. Apical meristems: The plant’s fountain of youth. Bioessays 2003, 25, 961–970. [Google Scholar] [CrossRef] [PubMed]
  165. Hanania, U.; Velcheva, M.; Sahar, N.; Flaishman, M.; Or, E.; Degani, O.; Perl, A. The ubiquitin extension protein S27a is differentially expressed in developing flower organs of Thompson seedless versus Thompson seeded grape isogenic clones. Plant Cell Rep. 2009, 28, 1033–1042. [Google Scholar] [CrossRef] [PubMed]
  166. Riepe, C.; Zelin, E.; Frankino, P.A.; Meacham, Z.A.; Fernandez, S.G.; Ingolia, N.T.; Corn, J.E. Double stranded DNA breaks and genome editing trigger loss of ribosomal protein RPS27A. FEBS J. 2022, 289, 3101–3114. [Google Scholar] [CrossRef] [PubMed]
  167. Cerezo, E.; Plisson-Chastang, C.; Henras, A.K.; Lebaron, S.; Gleizes, P.E.; O’Donohue, M.F.; Romeo, Y.; Henry, Y. Maturation of pre-40S particles in yeast and humans. Wiley Interdiscip. Rev. RNA 2019, 10, e1516. [Google Scholar] [CrossRef] [PubMed]
  168. Ferreira-Cerca, S.; Pöll, G.; Kuhn, H.; Neueder, A.; Jakob, S.; Tschochner, H.; Milkereit, P. Analysis of the in vivo assembly pathway of eukaryotic 40S ribosomal proteins. Mol. Cell 2007, 28, 446–457. [Google Scholar] [CrossRef] [PubMed]
  169. Rouquette, J.; Choesmel, V.; Gleizes, P.E. Nuclear export and cytoplasmic processing of precursors to the 40S ribosomal subunits in mammalian cells. EMBO J. 2005, 24, 2862–2872. [Google Scholar] [CrossRef] [PubMed]
  170. O’Donohue, M.F.; Choesmel, V.; Faubladier, M.; Fichant, G.; Gleizes, P.E. Functional dichotomy of ribosomal proteins during the synthesis of mammalian 40S ribosomal subunits. J. Cell Biol. 2010, 190, 853–866. [Google Scholar] [CrossRef]
  171. Nicolas, E.; Parisot, P.; Pinto-Monteiro, C.; de Walque, R.; De Vleeschouwer, C.; Lafontaine, D.L. Involvement of human ribosomal proteins in nucleolar structure and p53-dependent nucleolar stress. Nat. Commun. 2016, 7, 11390. [Google Scholar] [CrossRef] [PubMed]
  172. Wild, T.; Horvath, P.; Wyler, E.; Widmann, B.; Badertscher, L.; Zemp, I.; Kozak, K.; Csucs, G.; Lund, E.; Kutay, U. A protein inventory of human ribosome biogenesis reveals an essential function of exportin 5 in 60S subunit export. PLoS Biol. 2010, 8, e1000522. [Google Scholar] [CrossRef]
  173. Sun, Q.; Zhu, X.; Qi, J.; An, W.; Lan, P.; Tan, D.; Chen, R.; Wang, B.; Zheng, S.; Zhang, C.; et al. Molecular architecture of the 90S small subunit pre-ribosome. eLife 2017, 6, e22086. [Google Scholar] [CrossRef] [PubMed]
  174. Kornprobst, M.; Turk, M.; Kellner, N.; Cheng, J.; Flemming, D.; Kos-Braun, I.; Kos, M.; Thoms, M.; Berninghausen, O.; Beckmann, R.; et al. Architecture of the 90S pre-ribosome: A structural view on the birth of the eukaryotic ribosome. Cell 2016, 166, 380–393. [Google Scholar] [CrossRef] [PubMed]
  175. Cheng, J.; Kellner, N.; Berninghausen, O.; Hurt, E.; Beckmann, R. 3.2-Å-resolution structure of the 90S preribosome before A1 pre-rRNA cleavage. Nat. Struct. Mol. Biol. 2017, 24, 954–964. [Google Scholar] [CrossRef] [PubMed]
  176. Cheng, J.; Bassler, J.; Fischer, P.; Lau, B.; Kellner, N.; Kunze, R.; Griesel, S.; Kallas, M.; Berninghausen, O.; Strauss, D.; et al. Thermophile 90S pre-ribosome structures reveal the reverse order of co-transcriptional 18S rRNA subdomain integration. Mol. Cell 2019, 75, 1256–1269.e1257. [Google Scholar] [CrossRef] [PubMed]
  177. Strunk, B.S.; Loucks, C.R.; Su, M.; Vashisth, H.; Cheng, S.; Schilling, J.; Brooks, C.L., III; Karbstein, K.; Skiniotis, G. Ribosome assembly factors prevent premature translation initiation by 40S assembly intermediates. Science 2011, 333, 1449–1453. [Google Scholar] [CrossRef] [PubMed]
  178. Heuer, A.; Thomson, E.; Schmidt, C.; Berninghausen, O.; Becker, T.; Hurt, E.; Beckmann, R. Cryo-EM structure of a late pre-40S ribosomal subunit from Saccharomyces cerevisiae. eLife 2017, 6, e30189. [Google Scholar] [CrossRef] [PubMed]
  179. Blomqvist, E.K.; Huang, H.; Karbstein, K. A disease associated mutant reveals how Ltv1 orchestrates RP assembly and rRNA folding of the small ribosomal subunit head. PLoS Genet. 2023, 19, e1010862. [Google Scholar] [CrossRef] [PubMed]
  180. Plassart, L.; Shayan, R.; Montellese, C.; Rinaldi, D.; Larburu, N.; Pichereaux, C.; Froment, C.; Lebaron, S.; O’Donohue, M.F.; Kutay, U.; et al. The final step of 40S ribosomal subunit maturation is controlled by a dual key lock. eLife 2021, 10, e61254. [Google Scholar] [CrossRef]
  181. Eastham, M.J.; Pelava, A.; Wells, G.R.; Watkins, N.J.; Schneider, C. RPS27a and RPL40, which are produced as ubiquitin fusion proteins, are not essential for p53 signalling. Biomolecules 2023, 13, 898. [Google Scholar] [CrossRef] [PubMed]
  182. Schäfer, T.; Strauss, D.; Petfalski, E.; Tollervey, D.; Hurt, E. The path from nucleolar 90S to cytoplasmic 40S pre-ribosomes. EMBO J. 2003, 22, 1370–1380. [Google Scholar] [CrossRef] [PubMed]
  183. Cheng, J.; Lau, B.; La Venuta, G.; Ameismeier, M.; Berninghausen, O.; Hurt, E.; Beckmann, R. 90S pre-ribosome transformation into the primordial 40S subunit. Science 2020, 369, 1470–1476. [Google Scholar] [CrossRef] [PubMed]
  184. Cheng, J.; Lau, B.; Thoms, M.; Ameismeier, M.; Berninghausen, O.; Hurt, E.; Beckmann, R. The nucleoplasmic phase of pre-40S formation prior to nuclear export. Nucleic Acids Res. 2022, 50, 11924–11937. [Google Scholar] [CrossRef] [PubMed]
  185. Du, Y.; An, W.; Zhu, X.; Sun, Q.; Qi, J.; Ye, K. Cryo-EM structure of 90S small ribosomal subunit precursors in transition states. Science 2020, 369, 1477–1481. [Google Scholar] [CrossRef] [PubMed]
  186. Scaiola, A.; Pena, C.; Weisser, M.; Bohringer, D.; Leibundgut, M.; Klingauf-Nerurkar, P.; Gerhardy, S.; Panse, V.G.; Ban, N. Structure of a eukaryotic cytoplasmic pre-40S ribosomal subunit. EMBO J. 2018, 37, e98499. [Google Scholar] [CrossRef] [PubMed]
  187. Johnson, M.C.; Ghalei, H.; Doxtader, K.A.; Karbstein, K.; Stroupe, M.E. Structural heterogeneity in pre-40S ribosomes. Structure 2017, 25, 329–340. [Google Scholar] [CrossRef] [PubMed]
  188. Mitterer, V.; Gantenbein, N.; Birner-Gruenberger, R.; Murat, G.; Bergler, H.; Kressler, D.; Pertschy, B. Nuclear import of dimerized ribosomal protein Rps3 in complex with its chaperone Yar1. Sci. Rep. 2016, 6, 36714. [Google Scholar] [CrossRef] [PubMed]
  189. Ghalei, H.; Schaub, F.X.; Doherty, J.R.; Noguchi, Y.; Roush, W.R.; Cleveland, J.L.; Stroupe, M.E.; Karbstein, K. Hrr25/CK1δ-directed release of Ltv1 from pre-40S ribosomes is necessary for ribosome assembly and cell growth. J. Cell Biol. 2015, 208, 745–759. [Google Scholar] [CrossRef] [PubMed]
  190. Mitterer, V.; Murat, G.; Rety, S.; Blaud, M.; Delbos, L.; Stanborough, T.; Bergler, H.; Leulliot, N.; Kressler, D.; Pertschy, B. Sequential domain assembly of ribosomal protein S3 drives 40S subunit maturation. Nat. Commun. 2016, 7, 10336. [Google Scholar] [CrossRef]
  191. Holzer, S.; Ban, N.; Klinge, S. Crystal structure of the yeast ribosomal protein rpS3 in complex with its chaperone Yar1. J. Mol. Biol. 2013, 425, 4154–4160. [Google Scholar] [CrossRef] [PubMed]
  192. Mitterer, V.; Shayan, R.; Ferreira-Cerca, S.; Murat, G.; Enne, T.; Rinaldi, D.; Weigl, S.; Omanic, H.; Gleizes, P.E.; Kressler, D.; et al. Conformational proofreading of distant 40S ribosomal subunit maturation events by a long-range communication mechanism. Nat. Commun. 2019, 10, 2754. [Google Scholar] [CrossRef] [PubMed]
  193. Schäfer, T.; Maco, B.; Petfalski, E.; Tollervey, D.; Bottcher, B.; Aebi, U.; Hurt, E. Hrr25-dependent phosphorylation state regulates organization of the pre-40S subunit. Nature 2006, 441, 651–655. [Google Scholar] [CrossRef] [PubMed]
  194. Hector, R.D.; Burlacu, E.; Aitken, S.; Bihan, T.L.; Tuijtel, M.; Zaplatina, A.; Cook, A.G.; Granneman, S. Snapshots of pre-rRNA structural flexibility reveal eukaryotic 40S assembly dynamics at nucleotide resolution. Nucleic Acids Res. 2014, 42, 12138–12154. [Google Scholar] [CrossRef] [PubMed]
  195. Seiser, R.M.; Sundberg, A.E.; Wollam, B.J.; Zobel-Thropp, P.; Baldwin, K.; Spector, M.D.; Lycan, D.E. Ltv1 is required for efficient nuclear export of the ribosomal small subunit in Saccharomyces cerevisiae. Genetics 2006, 174, 679–691. [Google Scholar] [CrossRef] [PubMed]
  196. Zemp, I.; Wandrey, F.; Rao, S.; Ashiono, C.; Wyler, E.; Montellese, C.; Kutay, U. CK1δ and CK1ɛ are components of human 40S subunit precursors required for cytoplasmic 40S maturation. J. Cell Sci. 2014, 127, 1242–1253. [Google Scholar] [PubMed]
  197. Ameismeier, M.; Cheng, J.; Berninghausen, O.; Beckmann, R. Visualizing late states of human 40S ribosomal subunit maturation. Nature 2018, 558, 249–253. [Google Scholar] [CrossRef] [PubMed]
  198. Larburu, N.; Montellese, C.; O’Donohue, M.F.; Kutay, U.; Gleizes, P.E.; Plisson-Chastang, C. Structure of a human pre-40S particle points to a role for RACK1 in the final steps of 18S rRNA processing. Nucleic Acids Res. 2016, 44, 8465–8478. [Google Scholar] [CrossRef] [PubMed]
  199. Dörner, K.; Ruggeri, C.; Zemp, I.; Kutay, U. Ribosome biogenesis factors-from names to functions. EMBO J. 2023, 42, e112699. [Google Scholar] [CrossRef] [PubMed]
  200. Henras, A.K.; Plisson-Chastang, C.; O’Donohue, M.F.; Chakraborty, A.; Gleizes, P.E. An overview of pre-ribosomal RNA processing in eukaryotes. Wiley Interdiscip. Rev. RNA 2015, 6, 225–242. [Google Scholar] [CrossRef]
  201. Ameismeier, M.; Zemp, I.; van den Heuvel, J.; Thoms, M.; Berninghausen, O.; Kutay, U.; Beckmann, R. Structural basis for the final steps of human 40S ribosome maturation. Nature 2020, 587, 683–687. [Google Scholar] [CrossRef]
  202. Clatterbuck Soper, S.F.; Dator, R.P.; Limbach, P.A.; Woodson, S.A. In vivo X-ray footprinting of pre-30S ribosomes reveals chaperone-dependent remodeling of late assembly intermediates. Mol. Cell 2013, 52, 506–516. [Google Scholar] [CrossRef] [PubMed]
  203. Guo, Q.; Goto, S.; Chen, Y.; Feng, B.; Xu, Y.; Muto, A.; Himeno, H.; Deng, H.; Lei, J.; Gao, N. Dissecting the in vivo assembly of the 30S ribosomal subunit reveals the role of RimM and general features of the assembly process. Nucleic Acids Res. 2013, 41, 2609–2620. [Google Scholar] [CrossRef]
  204. Leong, V.; Kent, M.; Jomaa, A.; Ortega, J. Escherichia coli rimM and yjeQ null strains accumulate immature 30S subunits of similar structure and protein complement. RNA 2013, 19, 789–802. [Google Scholar] [CrossRef]
  205. Saurer, M.; Ramrath, D.J.F.; Niemann, M.; Calderaro, S.; Prange, C.; Mattei, S.; Scaiola, A.; Leitner, A.; Bieri, P.; Horn, E.K.; et al. Mitoribosomal small subunit biogenesis in trypanosomes involves an extensive assembly machinery. Science 2019, 365, 1144–1149. [Google Scholar] [CrossRef]
  206. Kang, J.; Brajanovski, N.; Chan, K.T.; Xuan, J.; Pearson, R.B.; Sanij, E. Ribosomal proteins and human diseases: Molecular mechanisms and targeted therapy. Signal Transduct. Target Ther. 2021, 6, 323. [Google Scholar] [CrossRef]
  207. Kampen, K.R.; Sulima, S.O.; Vereecke, S.; De Keersmaecker, K. Hallmarks of ribosomopathies. Nucleic Acids Res. 2020, 48, 1013–1028. [Google Scholar] [CrossRef] [PubMed]
  208. Mills, E.W.; Green, R. Ribosomopathies: There’s strength in numbers. Science 2017, 358, eaan2755. [Google Scholar] [CrossRef] [PubMed]
  209. Farley-Barnes, K.I.; Ogawa, L.M.; Baserga, S.J. Ribosomopathies: Old concepts, new controversies. Trends Genet. 2019, 35, 754–767. [Google Scholar] [CrossRef]
  210. De Keersmaecker, K.; Sulima, S.O.; Dinman, J.D. Ribosomopathies and the paradox of cellular hypo- to hyperproliferation. Blood 2015, 125, 1377–1382. [Google Scholar] [CrossRef]
  211. Danilova, N.; Gazda, H.T. Ribosomopathies: How a common root can cause a tree of pathologies. Dis. Models Mech. 2015, 8, 1013–1026. [Google Scholar] [CrossRef] [PubMed]
  212. Da Costa, L.; Mohandas, N.; David, N.L.; Platon, J.; Marie, I.; O’Donohue, M.F.; Leblanc, T.; Gleizes, P.E. Diamond-Blackfan anemia, the archetype of ribosomopathy: How distinct is it from the other constitutional ribosomopathies? Blood Cells Mol. Dis. 2024, 106, 102838. [Google Scholar] [CrossRef] [PubMed]
  213. Doherty, L.; Sheen, M.R.; Vlachos, A.; Choesmel, V.; O’Donohue, M.F.; Clinton, C.; Schneider, H.E.; Sieff, C.A.; Newburger, P.E.; Ball, S.E.; et al. Ribosomal protein genes RPS10 and RPS26 are commonly mutated in Diamond-Blackfan anemia. Am. J. Hum. Genet. 2010, 86, 222–228. [Google Scholar] [CrossRef] [PubMed]
  214. 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. The ribosomal basis of Diamond-Blackfan Anemia: Mutation and database update. Hum. Mutat. 2010, 31, 1269–1279. [Google Scholar] [CrossRef] [PubMed]
  215. 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. Ribosomal protein L5 and L11 mutations are associated with cleft palate and abnormal thumbs in Diamond-Blackfan anemia patients. Am. J. Hum. Genet. 2008, 83, 769–780. [Google Scholar] [CrossRef] [PubMed]
  216. Folgado-Marco, V.; Ames, K.; Chuen, J.; Gritsman, K.; Baker, N.E. Haploinsufficiency of the essential gene Rps12 causes defects in erythropoiesis and hematopoietic stem cell maintenance. eLife 2023, 12, e69322. [Google Scholar] [CrossRef] [PubMed]
  217. Choesmel, V.; Bacqueville, D.; Rouquette, J.; Noaillac-Depeyre, J.; Fribourg, S.; Cretien, A.; Leblanc, T.; Tchernia, G.; Da Costa, L.; Gleizes, P.E. Impaired ribosome biogenesis in Diamond-Blackfan anemia. Blood 2007, 109, 1275–1283. [Google Scholar] [CrossRef] [PubMed]
  218. Ferreira, R.; Ohneda, K.; Yamamoto, M.; Philipsen, S. GATA1 function, a paradigm for transcription factors in hematopoiesis. Mol. Cell. Biol. 2005, 25, 1215–1227. [Google Scholar] [CrossRef] [PubMed]
  219. Ludwig, L.S.; Gazda, H.T.; Eng, J.C.; Eichhorn, S.W.; Thiru, P.; Ghazvinian, R.; George, T.I.; Gotlib, J.R.; Beggs, A.H.; Sieff, C.A.; et al. Altered translation of GATA1 in Diamond-Blackfan anemia. Nat. Med. 2014, 20, 748–753. [Google Scholar] [CrossRef] [PubMed]
  220. 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. Ribosome levels selectively regulate translation and lineage commitment in human hematopoiesis. Cell 2018, 173, 90–103.e119. [Google Scholar] [CrossRef]
  221. Rio, S.; Gastou, M.; Karboul, N.; Derman, R.; Suriyun, T.; Manceau, H.; Leblanc, T.; El Benna, J.; Schmitt, C.; Azouzi, S.; et al. Regulation of globin-heme balance in Diamond-Blackfan anemia by HSP70/GATA1. Blood 2019, 133, 1358–1370. [Google Scholar] [CrossRef]
  222. van Dooijeweert, B.; Kia, S.K.; Dahl, N.; Fenneteau, O.; Leguit, R.; Nieuwenhuis, E.; van Solinge, W.; van Wijk, R.; Da Costa, L.; Bartels, M. GATA-1 defects in Diamond-Blackfan anemia: Phenotypic characterization points to a specific subset of disease. Genes 2022, 13, 447. [Google Scholar] [CrossRef]
  223. Ling, T.; Crispino, J.D. GATA1 mutations in red cell disorders. IUBMB Life 2020, 72, 106–118. [Google Scholar] [CrossRef]
  224. Sulima, S.O.; Kampen, K.R.; Vereecke, S.; Pepe, D.; Fancello, L.; Verbeeck, J.; Dinman, J.D.; De Keersmaecker, K. Ribosomal lesions promote oncogenic mutagenesis. Cancer Res. 2019, 79, 320–327. [Google Scholar] [CrossRef] [PubMed]
  225. Golomb, L.; Volarevic, S.; Oren, M. p53 and ribosome biogenesis stress: The essentials. FEBS Lett. 2014, 588, 2571–2579. [Google Scholar] [CrossRef] [PubMed]
  226. Bursac, S.; Brdovcak, M.C.; Donati, G.; Volarevic, S. Activation of the tumor suppressor p53 upon impairment of ribosome biogenesis. Biochim. Biophys. Acta 2014, 1842, 817–830. [Google Scholar] [CrossRef]
  227. Liu, Y.; Deisenroth, C.; Zhang, Y. RP-MDM2-p53 pathway: Linking ribosomal biogenesis and tumor surveillance. Trends Cancer 2016, 2, 191–204. [Google Scholar] [CrossRef]
  228. Dutt, S.; Narla, A.; Lin, K.; Mullally, A.; Abayasekara, N.; Megerdichian, C.; Wilson, F.H.; Currie, T.; Khanna-Gupta, A.; Berliner, N.; et al. Haploinsufficiency for ribosomal protein genes causes selective activation of p53 in human erythroid progenitor cells. Blood 2011, 117, 2567–2576. [Google Scholar] [CrossRef]
  229. Barlow, J.L.; Drynan, L.F.; Trim, N.L.; Erber, W.N.; Warren, A.J.; McKenzie, A.N. New insights into 5q- syndrome as a ribosomopathy. Cell Cycle 2010, 9, 4286–4293. [Google Scholar] [CrossRef] [PubMed]
  230. Sun, X.X.; DeVine, T.; Challagundla, K.B.; Dai, M.S. Interplay between ribosomal protein S27a and MDM2 protein in p53 activation in response to ribosomal stress. J. Biol. Chem. 2011, 286, 22730–22741. [Google Scholar] [CrossRef]
  231. Li, H.; Zhang, H.; Huang, G.; Bing, Z.; Xu, D.; Liu, J.; Luo, H.; An, X. Loss of RPS27a expression regulates the cell cycle, apoptosis, and proliferation via the RPL11-MDM2-p53 pathway in lung adenocarcinoma cells. J. Exp. Clin. Cancer Res. 2022, 41, 33. [Google Scholar] [CrossRef] [PubMed]
  232. Kussie, P.H.; Gorina, S.; Marechal, V.; Elenbaas, B.; Moreau, J.; Levine, A.J.; Pavletich, N.P. Structure of the MDM2 oncoprotein bound to the p53 tumor suppressor transactivation domain. Science 1996, 274, 948–953. [Google Scholar] [CrossRef] [PubMed]
  233. Lindström, M.S.; Jin, A.; Deisenroth, C.; White Wolf, G.; Zhang, Y. Cancer-associated mutations in the MDM2 zinc finger domain disrupt ribosomal protein interaction and attenuate MDM2-induced p53 degradation. Mol. Cell. Biol. 2007, 27, 1056–1068. [Google Scholar] [CrossRef] [PubMed]
  234. Nosrati, N.; Kapoor, N.R.; Kumar, V. DNA damage stress induces the expression of ribosomal protein S27a gene in a p53-dependent manner. Gene 2015, 559, 44–51. [Google Scholar] [CrossRef] [PubMed]
  235. Wang, H.; Zhao, J.; Yang, J.; Wan, S.; Fu, Y.; Wang, X.; Zhou, T.; Zhang, Z.; Shen, J. PICT1 is critical for regulating the Rps27a-Mdm2-p53 pathway by microtubule polymerization inhibitor against cervical cancer. Biochim. Biophys. Acta Mol. Cell Res. 2021, 1868, 119084. [Google Scholar] [CrossRef] [PubMed]
  236. Pelletier, J.; Thomas, G.; Volarevic, S. Ribosome biogenesis in cancer: New players and therapeutic avenues. Nat. Rev. Cancer 2018, 18, 51–63. [Google Scholar] [CrossRef] [PubMed]
  237. de Las Heras-Rubio, A.; Perucho, L.; Paciucci, R.; Vilardell, J.; Lleonart, M.E. Ribosomal proteins as novel players in tumorigenesis. Cancer Metastasis Rev. 2014, 33, 115–141. [Google Scholar] [CrossRef]
  238. Pecoraro, A.; Pagano, M.; Russo, G.; Russo, A. Ribosome Biogenesis and Cancer: Overview on Ribosomal Proteins. Int. J. Mol. Sci. 2021, 22, 5496. [Google Scholar] [CrossRef] [PubMed]
  239. Ajore, R.; Raiser, D.; McConkey, M.; Joud, M.; Boidol, B.; Mar, B.; Saksena, G.; Weinstock, D.M.; Armstrong, S.; Ellis, S.R.; et al. Deletion of ribosomal protein genes is a common vulnerability in human cancer, especially in concert with TP53 mutations. EMBO Mol. Med. 2017, 9, 498–507. [Google Scholar] [CrossRef] [PubMed]
  240. Ponten, F.; Jirström, K.; Uhlen, M. The Human Protein Atlas-a tool for pathology. J. Pathol. 2008, 216, 387–393. [Google Scholar] [CrossRef]
  241. Chen, D.; Zhang, R.; Shen, W.; Fu, H.; Liu, S.; Sun, K.; Sun, X. RPS12-specific shRNA inhibits the proliferation, migration of BGC823 gastric cancer cells with S100A4 as a downstream effector. Int. J. Oncol. 2013, 42, 1763–1769. [Google Scholar] [CrossRef] [PubMed]
  242. Derenzini, E.; Agostinelli, C.; Rossi, A.; Rossi, M.; Scellato, F.; Melle, F.; Motta, G.; Fabbri, M.; Diop, F.; Kodipad, A.A.; et al. Genomic alterations of ribosomal protein genes in diffuse large B cell lymphoma. Br. J. Haematol. 2019, 185, 330–334. [Google Scholar] [CrossRef] [PubMed]
  243. Katanaev, V.L.; Kryuchkov, M.; Averkov, V.; Savitsky, M.; Nikolaeva, K.; Klimova, N.; Khaustov, S.; Solis, G.P. HumanaFly: High-throughput transgenesis and expression of breast cancer transcripts in Drosophila eye discovers the RPS12-Wingless signaling axis. Sci. Rep. 2020, 10, 21013. [Google Scholar] [CrossRef] [PubMed]
  244. Wang, Q.; Cai, Y.; Fu, X.; Chen, L. High RPS27A expression predicts poor prognosis in patients with HPV type 16 cervical cancer. Front. Oncol. 2021, 11, 752974. [Google Scholar] [CrossRef] [PubMed]
  245. Wang, H.; Yu, J.; Zhang, L.; Xiong, Y.; Chen, S.; Xing, H.; Tian, Z.; Tang, K.; Wei, H.; Rao, Q.; et al. RPS27a promotes proliferation, regulates cell cycle progression and inhibits apoptosis of leukemia cells. Biochem. Biophys. Res. Commun. 2014, 446, 1204–1210. [Google Scholar] [CrossRef] [PubMed]
  246. Ramrath, D.J.F.; Niemann, M.; Leibundgut, M.; Bieri, P.; Prange, C.; Horn, E.K.; Leitner, A.; Boehringer, D.; Schneider, A.; Ban, N. Evolutionary shift toward protein-based architecture in trypanosomal mitochondrial ribosomes. Science 2018, 362, 422. [Google Scholar] [CrossRef]
  247. Bochler, A.; Querido, J.B.; Prilepskaja, T.; Soufari, H.; Simonetti, A.; Del Cistia, M.L.; Kuhn, L.; Ribeiro, A.R.; Valášek, L.S.; Hashem, Y. Structural differences in translation initiation between pathogenic trypanosomatids and their mammalian hosts. Cell Rep. 2020, 33, 108534. [Google Scholar] [CrossRef]
Figure 1. Protuberances of the r-subunits. (A) Structure of the mature 40S r-subunit of Saccharomyces cerevisiae showing its different structural domains (head, body, platform and beak) and highlighting the beak elements that include helix h33 of the 18S rRNA and the r-proteins eS10, eS12 and eS31. The 40S r-subunit is shown from the interface view; 18S rRNA is colored in gray (except h33 in yellow) and r-proteins not belonging to the beak structure are colored in green. Protein Data Bank (PDB) code: 4V88. (B) Structure of the mature 60S r-subunit of S. cerevisiae showing its different substructures (L1 and P0 stalks, and central protuberance) and highlighting the different protuberances: L1-stalk formed by helices H76-78 of 25S rRNA and the r-protein uL1; central protuberance formed by 5S rRNA and r-proteins uL5 and uL18; P0-stalk formed by helices H43-44 of 25S rRNA and the r-proteins uL11, uL10 (P0), P1 and P2 (two copies each). As other examples of protrusions found in the 60S r-subunits, we highlight helices H38 and H69 (yellow) and the long extensions of the r-proteins eL19 and eL24 (purple). The 60S r-subunit is shown from the interface view; 25S and 5.8S rRNAs are colored in gray and the rest of the r-proteins are colored in light blue. PDB code: 6OIG. Cartoons were generated using UCSF Chimera (https://www.rbvi.ucsf.edu/chimera; accessed on 1 June 2024).
Figure 1. Protuberances of the r-subunits. (A) Structure of the mature 40S r-subunit of Saccharomyces cerevisiae showing its different structural domains (head, body, platform and beak) and highlighting the beak elements that include helix h33 of the 18S rRNA and the r-proteins eS10, eS12 and eS31. The 40S r-subunit is shown from the interface view; 18S rRNA is colored in gray (except h33 in yellow) and r-proteins not belonging to the beak structure are colored in green. Protein Data Bank (PDB) code: 4V88. (B) Structure of the mature 60S r-subunit of S. cerevisiae showing its different substructures (L1 and P0 stalks, and central protuberance) and highlighting the different protuberances: L1-stalk formed by helices H76-78 of 25S rRNA and the r-protein uL1; central protuberance formed by 5S rRNA and r-proteins uL5 and uL18; P0-stalk formed by helices H43-44 of 25S rRNA and the r-proteins uL11, uL10 (P0), P1 and P2 (two copies each). As other examples of protrusions found in the 60S r-subunits, we highlight helices H38 and H69 (yellow) and the long extensions of the r-proteins eL19 and eL24 (purple). The 60S r-subunit is shown from the interface view; 25S and 5.8S rRNAs are colored in gray and the rest of the r-proteins are colored in light blue. PDB code: 6OIG. Cartoons were generated using UCSF Chimera (https://www.rbvi.ucsf.edu/chimera; accessed on 1 June 2024).
Biomolecules 14 00882 g001
Figure 2. Comparison of small r-subunits of bacteria, archaea and eukaryotes. (A) Small r-subunit of Escherichia coli; PDB code: 7OE1; (B) small r-subunit of Thermus thermophilus; PDB code: 1J5E; (C) small r-subunit of Pyrococcus abyssi; PDB code: 7ZHG; (D) small r-subunit of S. cerevisiae; PDB code: 4V88; (E) small r-subunit of Homo sapiens; PDB code: 7R4X. In all cases, the interface view of the individual r-subunits is shown. The rRNA is colored in gray and the r-proteins in green. The beak (b) of all r-subunits is highlighted; helix h33 of rRNA is colored in yellow, and the r-proteins eS31, eL8, eS10 and eS12 in the indicated colors. The r-protein uS3, which is located at the base of the beak, is colored in pink. Cartoons were generated using UCSF Chimera (https://www.rbvi.ucsf.edu/chimera; accessed on 1 June 2024).
Figure 2. Comparison of small r-subunits of bacteria, archaea and eukaryotes. (A) Small r-subunit of Escherichia coli; PDB code: 7OE1; (B) small r-subunit of Thermus thermophilus; PDB code: 1J5E; (C) small r-subunit of Pyrococcus abyssi; PDB code: 7ZHG; (D) small r-subunit of S. cerevisiae; PDB code: 4V88; (E) small r-subunit of Homo sapiens; PDB code: 7R4X. In all cases, the interface view of the individual r-subunits is shown. The rRNA is colored in gray and the r-proteins in green. The beak (b) of all r-subunits is highlighted; helix h33 of rRNA is colored in yellow, and the r-proteins eS31, eL8, eS10 and eS12 in the indicated colors. The r-protein uS3, which is located at the base of the beak, is colored in pink. Cartoons were generated using UCSF Chimera (https://www.rbvi.ucsf.edu/chimera; accessed on 1 June 2024).
Biomolecules 14 00882 g002
Figure 3. Comparison of the structural features of the small subunit of the mitochondrial ribosome from S. cerevisiae (A) and Trypanosoma brucei (B), and of the chloroplast ribosome from Spinacia oleracea (C). The PDB codes are 5MRC, 6HIW, and 5MMJ, respectively. In all cases, the interface view of the individual r-subunits is shown. The rRNA is colored in gray and the r-proteins in green. The beak (b) of all r-subunits is highlighted; helix h33 of rRNA is colored in yellow, and the set of specific r-proteins present in the T. brucei small r-subunit are shown in different colors. Cartoons were generated using UCSF Chimera (https://www.rbvi.ucsf.edu/chimera; accessed on 1 June 2024).
Figure 3. Comparison of the structural features of the small subunit of the mitochondrial ribosome from S. cerevisiae (A) and Trypanosoma brucei (B), and of the chloroplast ribosome from Spinacia oleracea (C). The PDB codes are 5MRC, 6HIW, and 5MMJ, respectively. In all cases, the interface view of the individual r-subunits is shown. The rRNA is colored in gray and the r-proteins in green. The beak (b) of all r-subunits is highlighted; helix h33 of rRNA is colored in yellow, and the set of specific r-proteins present in the T. brucei small r-subunit are shown in different colors. Cartoons were generated using UCSF Chimera (https://www.rbvi.ucsf.edu/chimera; accessed on 1 June 2024).
Biomolecules 14 00882 g003
Figure 4. The r-proteins from the beak (b) of the yeast 40S r-subunit and the equivalent ones from the beak of the Pyrococcus abyssi 30S r-subunit. (A) Structure of the yeast 40S r-subunit (PDB code 4V88). (B) Structure of yeast eS12, eS31 and eS10 r-proteins are shown as found in the 40S r-subunit before or after 90° rotation on the Y-axis. (C) Structure of the equivalent r-proteins eL8 and eS31 from P. abyssi are also shown as found in the 30S r-subunit (PDB code 7ZHG). Note that the first 19 N-terminal residues of yeast eS12, the six first residues of yeast eS31, the last nine C-terminal residues of yeast eS10, the first residue of P. abyssi eL8 and the first and the last residue of P. abyssi eS31 are not present in the structures. Cartoons were generated using UCSF Chimera (https://www.rbvi.ucsf.edu/chimera; accessed on 1 June 2024).
Figure 4. The r-proteins from the beak (b) of the yeast 40S r-subunit and the equivalent ones from the beak of the Pyrococcus abyssi 30S r-subunit. (A) Structure of the yeast 40S r-subunit (PDB code 4V88). (B) Structure of yeast eS12, eS31 and eS10 r-proteins are shown as found in the 40S r-subunit before or after 90° rotation on the Y-axis. (C) Structure of the equivalent r-proteins eL8 and eS31 from P. abyssi are also shown as found in the 30S r-subunit (PDB code 7ZHG). Note that the first 19 N-terminal residues of yeast eS12, the six first residues of yeast eS31, the last nine C-terminal residues of yeast eS10, the first residue of P. abyssi eL8 and the first and the last residue of P. abyssi eS31 are not present in the structures. Cartoons were generated using UCSF Chimera (https://www.rbvi.ucsf.edu/chimera; accessed on 1 June 2024).
Biomolecules 14 00882 g004
Figure 5. Formation of the early beak structure. Structural model of the yeast 90S pre-ribosomal particle (PDB code 5WYJ). The pre-rRNA is colored in gray, all r-proteins except eS12 and eS31 are colored in green and all ribosome assembly factors in light blue. Helix h33 is highlighted in yellow, eS12 in red and eS31 in blue. The interaction of Enp1 with the beak is shown. The beak is connected to the rest of the 90S particle by its interaction with Emg1/Nep1 (colored in pink). Left, close-up view of the beak (b) region. Cartoons were generated using UCSF Chimera (https://www.rbvi.ucsf.edu/chimera; accessed on 1 June 2024).
Figure 5. Formation of the early beak structure. Structural model of the yeast 90S pre-ribosomal particle (PDB code 5WYJ). The pre-rRNA is colored in gray, all r-proteins except eS12 and eS31 are colored in green and all ribosome assembly factors in light blue. Helix h33 is highlighted in yellow, eS12 in red and eS31 in blue. The interaction of Enp1 with the beak is shown. The beak is connected to the rest of the 90S particle by its interaction with Emg1/Nep1 (colored in pink). Left, close-up view of the beak (b) region. Cartoons were generated using UCSF Chimera (https://www.rbvi.ucsf.edu/chimera; accessed on 1 June 2024).
Biomolecules 14 00882 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Martín-Villanueva, S.; Galmozzi, C.V.; Ruger-Herreros, C.; Kressler, D.; de la Cruz, J. The Beak of Eukaryotic Ribosomes: Life, Work and Miracles. Biomolecules 2024, 14, 882. https://doi.org/10.3390/biom14070882

AMA Style

Martín-Villanueva S, Galmozzi CV, Ruger-Herreros C, Kressler D, de la Cruz J. The Beak of Eukaryotic Ribosomes: Life, Work and Miracles. Biomolecules. 2024; 14(7):882. https://doi.org/10.3390/biom14070882

Chicago/Turabian Style

Martín-Villanueva, Sara, Carla V. Galmozzi, Carmen Ruger-Herreros, Dieter Kressler, and Jesús de la Cruz. 2024. "The Beak of Eukaryotic Ribosomes: Life, Work and Miracles" Biomolecules 14, no. 7: 882. https://doi.org/10.3390/biom14070882

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