Abstract
Ribosomes translate the mRNA code into protein, and this process can be controlled by metabolites that bind to the translating ribosome in interaction with the nascent protein.
The regulation of gene expression is intensely investigated in diverse biological systems. Gene expression involves RNA transcription, RNA splicing, RNA stability, translation, posttranslational modification, and protein stability. Particular attention has been given to mRNA levels due to advances in microarray analysis and RNA-sequencing techniques. However, transcript levels do not necessarily correlate with protein levels or functionality (Conrads et al., 2005; Gibon et al., 2006; Bianchini et al., 2008), and complex layers of posttranscriptional regulation have been uncovered, foremost mRNA translation.
Translation can be regulated both globally and in a transcript-specific manner. Examples of global mRNA translational regulation include availability of ribosomes and translation initiation, elongation, and termination factors. In transcript-specific translational regulation, individual mRNA species or mRNA groups are selectively translated. For example, mRNAs can be sequestered in stress granules, removing them from the translatable mRNA pool (Chantarachot and Bailey-Serres, 2018). mRNA sequence or structural features can affect translatability directly or indirectly, the latter via small RNAs or mRNA-binding proteins (for review, see Merchante et al., 2017). Upstream open reading frames (uORFs) have been shown to participate in both global and transcript-specific regulation (von Arnim et al., 2014). Here, recent advances in translation regulation by uORFs are discussed, focusing on uORFs encoding sequence-conserved peptides (CPuORFs).
THE PROCESS OF PROTEIN TRANSLATION
The general concept of protein translation has been reviewed extensively (Browning and Bailey-Serres, 2015; Merchante et al., 2017). Eukaryotic mRNAs consist of a 5′ leader (also referred to as the 5′ untranslated region) followed by the protein encoding the main open reading frame (mORF) and the 3′ trailer sequence (also referred to as the 3′ untranslated region). mRNA translation initiates with the 43S preinitiation complex (PIC) recognizing the 5′ cap structure present on most eukaryotic mRNAs. Additional eukaryotic initiation factors (eIFs) are then bound to form the 48S PIC (for review, see Browning and Bailey-Serres, 2015). The 48S PIC consists of the 40S ribosomal subunit, multiple eIFs, and the initiator Met-tRNAimet. Once formed, the 48S PIC starts scanning mRNA for a start codon. The initiation factors dissociate upon start codon recognition and the 60S ribosomal subunit joins the small subunit to form the 80S ribosome, which is followed by commencement of protein translation.
During translation elongation, an amino acid-charged tRNA will enter the aminoacyl (A) site of the ribosome. Next, the nascent peptide in the peptidyl site will be transferred to the charged tRNA in the A site in the peptidyl transferase reaction. The tRNA present in the peptidyl site will then transfer to the exit site and leave the ribosome, concomitant with the transfer of the peptidyl-tRNA to the peptidyl site, freeing up the A site for entry of the next charged tRNA (for review, see Browning and Bailey-Serres, 2015).
Elongation ends when a stop codon is perceived in the A site. The stop codon recruits eukaryotic release factors to the A site, resulting in termination of translation. Translation termination is usually followed by disassembly of the 80S ribosome and release from the mRNA. However, ribosomes can reinitiate translation at downstream start codons (von Arnim et al., 2014; Browning and Bailey-Serres, 2015).
Novel technologies enable detailed investigation of the genome-wide translational landscape. Ribosome protection (Ribo-seq) experiments (Liu et al., 2013; Juntawong et al., 2014; Merchante et al., 2015; Hsu et al., 2016; Bazin et al., 2017) and analysis of truncated RNA ends (Hou et al., 2016; Yu et al., 2016) can be used to map the positions of ribosomes on mRNAs with nucleotide precision. Such analyses show the remarkable breadth of translation regulation, which includes both regular mRNAs and other RNA species. Progression of the translating ribosome on RNA occurs in steps of three nucleotides. This triplex repeat is the hallmark of protein translation and can be readily studied using Ribo-seq.
uORFS
Twenty percent to 50% of eukaryotic mRNAs contain uORFs in the 5′ leader sequence preceding the main functional protein-encoding ORF (Kochetov, 2008). Generally, these uORFs repress translation of the mORF as the ribosomes terminate and are released from the mRNA after translation of the uORF. However, the inhibitory effect of uORFs on mORF translation can be bypassed by leaky scanning or translational reinitiation (Hellens et al., 2016; Merchante et al., 2017; Zhang et al., 2019; Box 1).
During leaky scanning, the 48S PIC fails to recognize the uORF start codon usually due to a suboptimal context and proceeds scanning to initiate translation at the mORF start codon. An optimal start codon context contains an AUG start codon, an adenine at the −3 position, and a guanine at the +4 position relative to the A+1UG position (Browning and Bailey-Serres, 2015; Diaz de Arce et al., 2018). Non-AUG initiation codons also present a suboptimal initiation context.
Translation reinitiation involves translation of the uORF but, following termination, the 40S ribosomal subunit remains associated with the mRNA and acquires initiation factors needed for continued mRNA scanning and reinitiation of translation at a downstream start codon. Whether reinitiation occurs depends on multiple factors, including uORF length, whereby short uORF mRNAs are more likely to retain eIF3h to the ribosome that is needed for translation reinitiation (von Arnim et al., 2014). Reinitiation is enhanced by TOR kinase activity, which activates eIF3h via S6 kinase1 (S6K1). TOR kinase in turn can be regulated by multiple factors, including hormones and energy, and nutrient availability (Schepetilnikov et al., 2013; Schepetilnikov and Ryabova, 2017).
uORF-containing genes are prone to translation reinitiation, depending on conditions. For example, in AUXIN RESPONSE FACTOR (ARF) transcripts, uORFs regulate the translation of the mORF depending on auxin levels. Such reinitiation is stimulated by auxin activation of TOR activity (Schepetilnikov et al., 2013; Schepetilnikov and Ryabova, 2017).
CPuORFS
Bioinformatics analyses uncovered a significant number of genes with uORF-encoded peptides that are highly conserved among multiple plant species, named CPuORFs. In the plant kingdom, such observed CPuORF homology groups can be evolutionarily ancient, pointing to their functional significance. Different bioinformatics approaches (Box 2) have been used to identify homology groups (Hayden and Jorgensen, 2007; Tran et al., 2008; Takahashi et al., 2012, 2019; Vaughn et al., 2012; Lorenzo-Orts et al., 2019; van der Horst et al., 2019). Ribo-seq analysis shows that such CPuORFs can be translated (Hsu et al., 2016). Relatively few of the Arabidopsis (Arabidopsis thaliana) CPuORFs have a known biological function; those characterized CPuORFs are involved in translational regulation of the mORF, mostly in a metabolite-dependent manner (Table 1). In the following examples, metabolite levels control mRNA translation in a CPuORF-dependent manner.
Table 1. List of CPuORF genes with known interacting metabolites.
| Gene Containing CPuORF + Homology Group | Brief Description of mORF Function | Biological Role of mORF | CPuORF-Interacting Metabolite | References |
|---|---|---|---|---|
| S1-group bZIPs (HG1) | bZIP transcription factor | Controls amino acid and sugar metabolism | Suc | Rook et al., 1998; Wiese et al., 2004; Rahmani et al., 2009; Peviani et al., 2016; Yamashita et al., 2017 |
| SAC51 (HG15) | bHLH transcription factor | Involved in xylem differentiation | Thermospermine | Imai et al., 2006, 2008; Cai et al., 2016; Ishitsuka et al., 2019 |
| HsfB1/TBF1 (HG18) | HSF transcription factor | Involved in heat tolerance and growth-to-defense transition | Galactinol | Pajerowska-Mukhtar et al., 2012; Zhu et al., 2012, 2018; Xu et al., 2017 |
| AdoMetDCs (HG3) | Polyamine biosynthesis | AdoMetDC decarboxylates AdoMet to produce dcAdoMet, a precursor of spermidine and spermine | Spermine and spermidine | Hanfrey et al., 2002, 2005; Uchiyama-Kadokura et al., 2014 |
| PAO (HG6) | Polyamine degradation | Involved in degradation of PAs | Spermine | Guerrero-González et al., 2014, 2016 |
| XIPOTL1/PEAMT (HG13) | Phosphocholine biosynthesis | Encodes a phosphoethanolamine N-methyltransferase involved in phosphocholine biosynthesis | Phosphocholine | Tabuchi et al., 2006; Alatorre-Cobos et al., 2012 |
| GGP (HG131) | Ascorbate biosynthesis | Involved in the first dedicated step in ascorbate production | Ascorbate | Laing et al., 2015; Zhang et al., 2018 |
Suc
The CPuORF homology group presented by mRNAs of the five Arabidopsis S1-group bZIP transcription factors (bZIP1, bZIP2, bZIP11, bZIP44, and bZIP53) is a well-studied example of metabolite-controlled translation by CPuORFs (Rook et al., 1998; Wiese et al., 2004; Hummel et al., 2009; Weltmeier et al., 2009). These bZIPs are regulators of metabolism and control amino acid and sugar metabolism as well as resource allocation (Hanson et al., 2008; Ma et al., 2011; Thalor et al., 2012; Dröge-Laser and Weiste, 2018). Translation of the mORF is regulated by the CPuORF in a Suc-dependent manner (Rook et al., 1998; Rahmani et al., 2009). Increasing Suc levels promote ribosome stalling on the CPuORF and, consequently, due to steric hindrance on the 5′ leader, ribosomes translating or scanning the uORF are unable to reach and translate the mORF (Fig. 1; Juntawong et al., 2014; Hou et al., 2016; Yamashita et al., 2017). Single amino acid mutations in the CPuORF abolish Suc-dependent ribosome stalling (Rahmani et al., 2009; Yamashita et al., 2017). Thus, the CPuORF peptide and not the nucleotide sequence plays a role in ribosome stalling. Moreover, RNA sequencing and mRNA degradation experiments revealed the stalling site of the bZIP CPuORFs to be close to the stop codon (Juntawong et al., 2014; Hou et al., 2016). Recent in vitro translation studies using wheat germ extract revealed that the peptidyl-tRNA of the last amino acid of the CPuORF is present on the ribosome during Suc-induced stalling (Yamashita et al., 2017). Overexpression of S1 bZIPs lacking the CPuORF in plants results in generally altered metabolism, including the accumulation of Suc and several amino acids (Hanson et al., 2008; Ma et al., 2011; Thalor et al., 2012; Sagor et al., 2016).
Figure 1.
Schematic representation of Suc-induced repression of translation of the bZIP11 mRNA. When Suc levels are low, ribosomes translating the CPuORF terminate at the stop codon. Translation of the mORF involves translational reinitiation or ribosome leaky scanning. Increasing Suc levels will induce ribosome stalling near the stop codon of the CPuORF, thereby blocking upstream translating or scanning ribosomes and inhibiting bZIP11 protein production. Lines and bars indicate RNA and ORFs, respectively.
Polyamines
Polyamines (PAs) provide essential functions in diverse biological processes, including the stability of nucleic acids and membranes, and general protein translation (Poidevin et al., 2019). In plants, PAs are essential for growth and stress tolerance. Translational control is important for PA homeostasis through the involvement of CPuORFs in key metabolic steps. PA biosynthesis initiates with the conversion of Orn or Arg to putrescine. Aminopropyl units derived from decarboxylated S-adenosylmethionine (dcAdoMet) are added to putrescine to produce spermidine and the higher order PAs spermine and thermospermine. The enzyme S-adenosylmethionine decarboxylase (AdoMetDC) decarboxylates adenosylmethionine (AdoMet) to produce dcAdoMet. CPuORF-mediated translational control of AdoMetDC has been well documented in different systems (Ivanov et al., 2010; Uchiyama-Kadokura et al., 2014) and was recently reviewed in detail (Poidevin et al., 2019).
In the presence of spermidine or spermine, ribosomes stall on the CPuORF of the AdoMetDC mRNA. Similar to the S1-bZIP CPuORFs, it was shown that the peptide was necessary to induce the PA-dependent ribosome stalling at the stop codon of the CPuORF (Uchiyama-Kadokura et al., 2014). Interestingly, the AdoMetDC CPuORF is accompanied by an additional smaller uORF, also known as the tiny uORF (Hanfrey et al., 2002, 2005). In Arabidopsis, the tiny uORF is three amino acids long and overlaps with the start codon of the CPuORF in a different frame. When the PA level is low, the ribosome initiates translation on the tiny uORF, thereby skipping the CPuORF start codon, and then successfully reinitiates on the mORF due to the short length of the tiny uORF (Hanfrey et al., 2005). When PA levels rise, the tiny uORF start codon is skipped and the CPuORF is translated, inducing ribosome stalling that then prevents AdoMetDC mORF translation and dcAdoMet production. Exactly how translation initiation on the tiny and CPuORF AUGs is regulated by PAs is unresolved.
CPuORFs in conjunction with their specific metabolites usually promote ribosome stalling, often close to or at the stop codon of the CPuORF. However, metabolites can also promote the release of ribosomes stalled by CPuORF peptides, thereby promoting mORF translation. Polyamine degradation involves Polyamine Oxidase2 (PAO2) to PAO4 that catalyze the oxidative deamination of spermidine and spermine, contributing to PA homeostasis (Guerrero-González et al., 2014, 2016). Particularly, the CPuORF-regulated translation of AtPAO2 mRNA has been well investigated. The AtPAO2 CPuORF constitutively inhibits translation of the mORF, allowing PAs to accumulate. Accumulating PAs then release CPuORF-mediated repression of mORF translation, resulting in AtPAO2 production and PA degradation. Mutational analysis showed that the AtPAO2 CPuORF must be translated in order to repress mORF translation (Guerrero-González et al., 2014, 2016). Again, the ribosome in combination with the CPuORF somehow presents a sensor function to fine-tune cellular PA levels. However, the molecular mechanism of how PAs promote PAO2 translation remains unresolved.
Thermospermine
In Arabidopsis, thermospermine is produced by the thermospermine synthase ACAULIS5 (ACL5) that converts spermidine to thermospermine (Knott et al., 2007). ACL5 mutants display severe defects in stem elongation due to the overproliferation of xylem vessels. In an acl5 suppressor mutagenesis screen, multiple dominant or semidominant suppressor of acaulis (sac) mutants were identified (Imai et al., 2006). One of these mutants, sac51-d, has a premature stop codon mutation in the SAC51 CPuORF that fully restores the wild-type phenotype (Imai et al., 2006). Interestingly, in the sac51-d mutant, the SAC51 mORF is translated, resulting in restoration of function. Thus, the presence of thermospermine promotes SAC51 mORF translation, whereas low thermospermine levels result in poor SAC51 translation. In the sac51-d mutant, a dysfunctional CPuORF peptide is produced that is incapable of inhibiting SAC51 mORF translation (Imai et al., 2006; Ishitsuka et al., 2019). Similar to the AtPAO2 situation, the SAC51 CPuORF peptide inhibits mORF translation that is released by increasing thermospermine levels.
Phosphocholine
Phosphatidylcholine is an essential lipid that functions as a membrane structural lipid and as a signaling molecule. The Arabidopsis XIPOTL1 gene (AT3G18000) encodes a phosphoethanolamine N-methyltransferase involved in phosphocholine biosynthesis. Phosphocholine is considered as a rate-limiting substrate for phosphatidylcholine biosynthesis. Translation of the XIPOTL1 mRNA is regulated by a CPuORF that encodes a peptide of 26 amino acids. Physiological phosphocholine levels repress mORF translation in a concentration-dependent manner without significantly affecting mRNA levels. This translational repression depends on translation of the CPuORF peptide (Alatorre-Cobos et al., 2012).
Ascorbate
Enzyme-encoding mRNAs seem ideal targets for metabolite-regulated translation via CPuORFs, since feedback loops are thus readily established for maintaining metabolite homeostasis. The first dedicated step in ascorbate biosynthesis is catalyzed by GDP-l-Gal phosphorylase (GGP). Ascorbate levels control GGP expression independently of transcription or GGP mRNA levels (Laing et al., 2015). Phylogenetic analyses of the 5′ leader of the mRNA revealed a CPuORF that initiates with a highly conserved ACG codon. Increasing cellular ascorbate levels greatly reduce the translation of GGP mORFs in a CPuORF-dependent manner (Laing et al., 2015). Mutation of a conserved His codon in the CPuORF abolished the ascorbate-mediated regulation of translation. Moreover, ribosome footprinting revealed ribosome stalling on the CPuORF (Laing et al., 2015). Interestingly, in a recent study, CPuORFs of two lettuce (Lactuca sativa) GGP orthologs were mutated using CRISPR-Cas9, which caused cellular ascorbate levels to increase (Zhang et al., 2018). This study demonstrates the potential of CPuORF modification for crop improvement.
Galactinol
Plants induce the secretion of antimicrobial proteins when challenged with microbe-associated molecular patterns. The heat shock factor-like transcription factor HSFB1/TBF1 (AT4G36990) specifically binds to a promoter region (the TL1 element, GAAGAAGAA) present in the genes encoding such antimicrobial proteins (Pajerowska-Mukhtar et al., 2012). HSFB1/TBF1 is itself transcriptionally induced by microbial pathogens or the elf18 microbe-associated molecular pattern. Importantly, HSFB1/TBF1 protein accumulation is highly dependent on the presence of two uORFs in its mRNA, the second of which is a CPuORF. Particularly, this uORF2 inhibits protein translation of the HSFB1/TBF1 mORF (Zhu et al., 2012); however, this inhibition is released following pathogen infection or elf18 treatment (Xu et al., 2017). HSFB1/TBF1 protein promotes the all-important growth-to-defense transition. Recently, it was shown that the HSFB1/TBF1 CPuORF is responsive to galactinol (Zhu et al., 2018), a Gal-inositol disaccharide that is the substrate for the biosynthesis of the raffinose series oligosaccharides involved in adaptation to stressful growth conditions. Galactinol in the micromolar range suppresses HSFB1/TBF1 mORF translation, probably through CPuORF peptide-dependent ribosome stalling. Possibly, pathogens reduce galactinol levels, thereby releasing stalled ribosomes and allowing HSFB1/TBF1 mORF translation and the induction of defense mechanisms. However, galactinol has been implicated previously as an inducer of plant pathogen defense (Kim et al., 2008). Recently, the HSFB1/TBF1 mRNA leader sequence harboring the CPuORF (named the TBF1 cassette) was used to induce pathogen resistance upon infection only by the immediate translation of downstream mORF resistance genes. This powerful translational control system circumvents deleterious effects of the constitutive expression of resistance genes (Bailey-Serres and Ma, 2017; Xu et al., 2017).
Boron
Boron (B) is an essential element for plants but is toxic at high concentrations. A remarkable uORF-dependent translation regulatory system was reported for the B diffusion facilitator NODULIN26-LIKE INTRINSIC PROTEIN5;1 (NIP5;1), which is required for efficient B uptake in roots. In the NIP5;1 mRNA leader, an AUG-stop uORF mediates ribosome stalling under high-B conditions, preventing mORF translation and thereby the production of NIP5;1 and B uptake (Tanaka et al., 2016). Here, ribosome stalling promotes mRNA degradation at a conserved nucleotide sequence immediately upstream of the stalled ribosome. Transfer of the NIP5;1 AUG-stop region (ATGTAA) to an unrelated mRNA imposed B-dependent ribosome stalling. Two additional B-responsive genes (SKU5 and ABS/NGAL1) were identified that are regulated by ribosome stalling at the AUG-stop sequence. The B-dependent AUG-stop-mediated regulation system also functioned in cell-free and animal translation systems. When B is low, translation reinitiation appears to promote NIP5;1 mORF translation. mRNA translation of another B transporter, BOR1, similarly depends on translation reinitiation (Aibara et al., 2018).
mORF RIBOSOME STALLING
AdoMet
Metabolites and their corresponding CPuORF peptides control mORF translation through a mechanism involving ribosome stalling in the mRNA leader region. Such metabolite-induced stalling can also occur in the mORF itself. The CGS1 gene (AT3G01120) encodes the enzyme cystathionine γ-synthase, which catalyzes the first committed step in Met biosynthesis. In the N-terminal region of the Arabidopsis CGS1 gene, a short sequence-conserved peptide (MTO1, 77RRNCSNIGVAQIVA90) is located that promotes ribosome stalling at the Ser-94 position, immediately downstream of the MTO1 region, in response to AdoMet, a direct metabolite of Met. Mutations in the MTO1 region of the CGS1 gene no longer respond to AdoMet, resulting in unrestricted CGS1 enzyme production and Met overaccumulation (Chiba et al., 1999). The MTO1 region of the nascent peptide is present in the ribosomal exit tunnel when translation elongation is halted by AdoMet, following formation of the peptidyl tRNASer located in the A site of the ribosome (Onoue et al., 2011). As a result of ribosome arrest, the CGS1 mRNA is degraded specifically at the MTO1-encoding region (Chiba et al., 2003). AdoMet-induced stalling imposes a compact conformation of the nascent peptide in the exit channel (Onoue et al., 2011).
PF-06446846
An interesting example of mORF-associated, small molecule-induced ribosome stalling is presented by animal proprotein convertase subtilisin/kexin type 9 (PCSK9), a key enzyme involved in regulating levels of plasma low-density lipoprotein cholesterol (Lintner et al., 2017). Chemical library screening for cholesterol-lowering drugs affecting PCSK9 identified an inhibitory compound that was functionally optimized to the active molecule PF-06446846, operating in the low micromolar range. The mechanism of PCSK9 inhibition was investigated and, remarkably, it was observed that the compound inhibits PCSK9 mRNA translation though ribosome stalling near amino acid position 34 of the mORF. Recently, details of the stalling mechanism as obtained by cryo-electron microscopy (cryo-EM) were reported (Li et al., 2019). Ribo-seq experiments showed that PF-06446846 is highly selective for translation inhibition of the PCSK9 mRNA. This finding suggests that, in principle, translation of any stretch of amino acids can mediate ribosome stalling, provided that the matching chemical compound is provided. The challenge is to devise high-throughput chemical compound screening systems whereby translation stalling is readily displayed.
The above examples of small molecule-induced mORF stalling demonstrate the adaptability of the ribosome and suggest a dynamic and immediate way to control the expression of specific genes by ribosome stalling.
THE RIBOSOMAL EXIT TUNNEL PRESENTS A REGULATORY ENVIRONMENT
The ribosomal exit channel is part of the large ribosomal subunit and is structured by the 28S rRNA and several ribosomal proteins. The tunnel accommodates a stretch of approximately 40 amino acids. Peptides in the tunnel can already assume a partially folded state, particularly near the tunnel exit site. The peptidyl transferase center (PTC) at the entry site of the tunnel transfers amino acids to the nascent chain, as the ribosome progresses toward the 3′ end of the mRNA. Studies have shown that the passage of peptides through the tunnel can be highly regulated and involves interactions between the peptide and the tunnel wall. Important in this respect is a constriction at approximately one-third of the tunnel length counting from the PTC that involves r-proteins uL4 and uL10 (Ito and Chiba, 2013; Dao Duc et al., 2019). Interestingly, CPuORFs often are best conserved at the C-terminal 15 to 20 amino acids, spanning the PTC-to-constriction region. Translated peptides can have the intrinsic capacity to stall translation, whereas other peptides require an effector molecule such as a metabolite or drug to induce ribosome stalling (for review, see Ito and Chiba, 2013).
Technological advances in 3D single-particle cryo-EM allow for the near-atomic resolution of biological structures as complex as the eukaryotic ribosome (for review, see Nogales and Scheres, 2015). Particularly, the Beckmann laboratory in Munich has used cryo-EM to study the structural basis of nascent polypeptide-mediated ribosome stalling (for review, see Wilson et al., 2016), and the structural basis for cofactor-induced translational stalling was resolved for several bacterial systems (Arenz et al., 2016; Wilson et al., 2016). Of particular interest is the bacterial tryptophanase (tnaCAB) operon that controls bacterial Trp (l-Trp) levels by l-Trp-induced production of tryptophanase and Trp permease. tnaCAB regulation involves l-Trp-induced ribosome stalling of the tnaC nascent peptide. Such stalling allows for the production of l-Trp-removing enzymes. Cryo-EM structural resolution of the l-Trp-stalled tnaC ribosomal complex showed the presence of two l-Trp-binding pockets in the exit tunnel, created by the tnaC peptide and the tunnel wall. l-Trp binding causes a conformational change that blocks release factor entry (Bischoff et al., 2014; Wilson et al., 2016). Thus, the tnaC peptide together with the exit tunnel functions as an l-Trp receptor, regulating the expression of downstream enzymes. The presence of a particular peptide in the exit channel structural environment together with its cognate cofactor inhibits peptidyl transfer or release by inducing conformational changes in proximity to the PTC, thereby stalling the ribosome. Cryo-EM resolution of erythromycin-induced stalling of the ErmCL peptide in the exit channel presents another example of cofactor-induced ribosome stalling (Wilson et al., 2016).
METABOLITE SENSING BY THE RIBOSOME
The CPuORFs discussed in detail above either promote or release stalling, depending on the concentration of their cognate cofactors. How do these different CPuORF homology groups, which bear no apparent amino acid or structural similarity, promote ribosome stalling or release from stalling in the presence of their cognate cofactors? Similar to the tnaC bacterial ribosome stalling system, the eukaryotic ribosome also functions as a metabolite sensor, a prime example being the fungal Arg attenuator peptide (Bhushan et al., 2010).
Bioinformatics analysis using stringent parameters (van der Horst et al., 2019) shows that 81 highly conserved CPuORF homology groups, and likely many more, are present in plants (Table 2; Box 2). Conceivably, many of these groups respond to a cofactor. Such a cell-autonomous ribosome stalling-based translational regulatory mechanism is a most powerful one, as it operates as a dynamic and immediate metabolite sensor, comparable to allosteric regulation of enzymes (Fig. 2). Cryo-EM information of metabolite-induced stalling complexes in plants is currently lacking, but accumulated evidence clearly points to the ribosome as a metabolite sensor, operating in a similar way as discussed above for the tnaC peptide.
Table 2. Arabidopsis CPuORF homology groups, classified according to Gene Ontology.
| Homology Group | Gene ID | Gene Name |
|---|---|---|
| Transcription regulation | ||
| HG1 | AT1G75390 | bZIP44 |
| AT2G18160 | bZIP2 | |
| AT3G62420 | bZIP53 | |
| AT4G34590 | bZIP11 | |
| AT5G49450 | bZIP1 | |
| HG2 | AT1G06150 | EMB1444 |
| AT2G27230 | LHW | |
| AT2G31280 | ||
| HG4 | AT4G25670 | |
| AT4G25690 | ||
| AT5G52550 | ||
| HG14 | AT3G01470 | HB-1 |
| HG15 | AT1G29950 | |
| AT5G09460 | ||
| AT5G50010 | ||
| AT5G64340 | SAC51 | |
| HG18 | AT4G36990 | HSF4 |
| HG21 | AT1G25470 | CRF12 |
| AT1G68550 | CRF10 | |
| AT3G25890 | CRF11 | |
| HG34 | AT2G27350 | OTLD1 |
| HG36 | AT3G15430 | |
| HG41 | AT5G09330 | NAC082 |
| AT5G64060 | NAC103 | |
| HG43 | AT5G46590 | NAC096 |
| HG44 | AT5G60450 | ARF4 |
| HG50 | AT1G01060 | LHY |
| HG51 | AT4G15180 | SDG2 |
| HG58 | AT2G24530 | |
| HG59 | AT2G35940 | BLH1 |
| HG61 | AT5G55600 | |
| HG71 | AT3G61970 | NGA2 |
| HG72 | AT4G34610 | BLH6 |
| HG73 | AT5G53660 | GRF7 |
| HG132 | AT1G07640 | OBP2 |
| HG136 | AT1G68920 | |
| HG145 | AT4G17980 | NAC071 |
| Metabolism | ||
| HG3 | AT3G02470 | SAMDC |
| AT3G25570 | ||
| AT5G15950 | ||
| HG6 | AT2G43020 | PAO2 |
| AT3G59050 | PAO3 | |
| HG11 | AT4G12430 | TPPF |
| AT4G22590 | TPPG | |
| HG13 | AT1G48600 | PMEAMT |
| AT1G73600 | ||
| AT3G18000 | XPL1 | |
| HG17 | AT3G53400 | |
| AT5G01710 | ||
| AT5G03190 | ||
| HG28 | AT1G67480 | |
| HG29 | AT2G22500 | UCP5 |
| HG38 | AT4G10170 | |
| HG39 | AT4G12790 | |
| HG45 | AT5G63640 | |
| HG46 | AT1G14560 | |
| HG47 | AT1G68100 | IAR1 |
| HG48 | AT1G72820 | |
| HG54 | AT1G65320 | CBSX6 |
| HG70 | AT3G52490 | |
| HG131 | AT4G26850 | VTC2 |
| AT5G55120 | VTC5 | |
| HG133 | AT1G11820 | |
| HG134 | AT1G57680 | Cand1 |
| HG135 | AT1G66540 | |
| AT5G35715 | CYP71B8 | |
| HG137 | AT2G23570 | MES19 |
| HG142 | AT3G57170 | |
| HG143 | AT3G62040 | |
| HG146 | AT5G26140 | LOG9 |
| Protein translation or degradation | ||
| HG7 | AT1G36730 | |
| AT1G77840 | ||
| HG22 | AT1G16860 | |
| AT1G78880 | ||
| HG26 | AT3G10910 | |
| AT5G05280 | ||
| HG42 | AT5G27920 | |
| HG141 | AT3G49430 | SR34a |
| HG147 | AT5G55100 | |
| HG148 | AT5G63190 | |
| Kinase or phosphatase | ||
| HG10 | AT4G19110 | |
| AT5G45430 | ||
| HG16 | AT3G51630 | WNK5 |
| HG23 | AT1G64630 | WNK10 |
| AT5G41990 | WNK8 | |
| HG25 | AT3G45240 | SNAK2 |
| AT5G60550 | SNAK1 | |
| HG27 | AT4G30960 | SIP3 |
| HG31 | AT5G01810 | CIPK15 |
| HG32 | AT3G08730 | S6K1 |
| HG33 | AT1G30270 | CIPK23 |
| HG35 | AT2G42880 | MPK20 |
| HG37 | AT3G55050 | |
| HG49 | AT5G14720 | |
| HG52 | AT3G08850 | RAPTOR1 |
| HG53 | AT1G62400 | HT1 |
| HG144 | AT4G03260 | |
| Other/unknown | ||
| HG5 | AT5G07840 | PIA1 |
| AT5G61230 | ANK6 | |
| HG8 | AT3G12010 | |
| HG9 | AT1G64140 | |
| AT5G09670 | ||
| AT5G64550 | ||
| HG12 | AT1G23150 | |
| AT1G70780 | ||
| HG17 | AT1G58120 | |
| HG19 | AT5G53590 | |
| HG20 | AT2G37480 | |
| AT3G53670 | ||
| HG24 | AT3G22970 | |
| AT4G14620 | ||
| HG30 | AT2G11890 | |
| HG40 | AT5G02480 | |
| HG57 | AT1G54095 | |
| AT1G72510 | ||
| HG60 | AT3G12570 | FYD |
| HG138 | AT2G29290 | |
| HG139 | AT2G42490 | |
| HG140 | AT3G23010 | RLP36 |
Figure 2.
Hypothetical model of the stalled plant ribosome. The CPuORF peptide and the ribosomal exit tunnel wall create a specific metabolite-binding pocket. The presence of the metabolite (green color) in the pocket allosterically controls the peptidyl transferase reaction. uL4 and uL22 are ribosomal proteins present in the exit tunnel constriction. uL16 is adjacent to the PTC.
Translational stalling systems often can be reconstituted in in vitro translation systems (Onouchi et al., 2005; Ebina et al., 2015; Hayashi et al., 2017). Such cell-free system reconstitutions point to a sensor function for the ribosome. Moreover, genetic evidence for the involvement of the plant ribosome in metabolite sensing comes from the SAC51 CPuORF peptide discussed above. This CPuORF promotes ribosome stalling that is then released by increasing thermospermine levels. In the SAC51 system, suppressor mutants were identified that promote SAC51 production in the absence of thermospermine. In addition to mutations in the SAC51 CPuORF itself (Imai et al., 2006), mutations in the ribosomal tunnel wall-associated r-proteins uL16, uL4, and RACK1 were also retrieved, named sac52-d, sac53-d, and sac56-d, respectively (Imai et al., 2008; Kakehi et al., 2015). Each of these three semidominant mutants is able to restore the thermospermine-deficient acl5-1 dwarf phenotype at least partly. In these sac ribosomal mutants, repression of the SAC51 mORF translation was alleviated. These results suggest a role for both the ribosome and the CPuORF peptide in SAC51 translational regulation by thermospermine.
In plants, r-proteins are encoded by up to six paralogs. In Arabidopsis, growth conditions can determine the paralog composition of ribosomes (Hummel et al., 2012, 2015). Such different ribosome isoforms might respond with different kinetics to a CPuORF and its cofactor. In Arabidopsis, many mutants affected in plant development turned out to be mutated in r-protein paralogs (Byrne, 2009; Horiguchi et al., 2011). Such mutants in plant development might be affected in receptor or signaling functions of the ribosome.
For optimal cellular function, ribosome stalling complexes should be dynamic (i.e. under certain cellular conditions, stalling must be released). For spermidine-stalled ribosomes translating the AdoMetDC1 CPuORF in a wheat germ system, it was found that stalled complexes were spontaneously released and peptidyl-tRNA hydrolysis was started within a 30-min time frame of inhibition of translation initiation (Uchiyama-Kadokura et al., 2014).
CPuORF CONTROL OF THE PLANT ENERGY SIGNALING NETWORK
Genes involved in the plant energy signaling network seem particularly prone to expression regulation by CPuORFs (Jorgensen and Dorantes-Acosta, 2012; van der Horst et al., 2019). The S1 group of bZIP transcription factors adapt general metabolism to a low-energy state (Ma et al., 2011; Dröge-Laser and Weiste, 2018). As mentioned above, S1 bZIPs are Suc responsive through their CPuORFs, whereby elevated Suc levels promote ribosome stalling, allowing carbon utilization and growth. Other genes in the plant energy signaling network harboring CPuORFs include those encoding the activating kinases (SnAK1 and SnAK2) of the Suc nonfermenting-related protein kinase1 (SnRK1), trehalose 6-phosphate phosphatase (TPP), the regulatory-associated protein of TOR (RAPTOR), S6K1, and translation factor eIF5-1 (Fig. 3).
Figure 3.
CPuORFs in the plant energy signaling network. Genes of the energy signaling network regulated by CPuORFs include the S1 group bZIP transcription factors, TPP, SnAK (SNAK 1/2), the RAPTOR subunit of the TOR kinase, S6K1, and eIF5-1. CPuORFs of the S1 group bZIPs are responsive to Suc. Presumed ligands of the other CPuORFs in the network are unknown.
The SnRK1 and TOR protein complexes are conserved regulators of growth and development in eukaryotes and are responsive to carbon availability. SnRK1 can be activated by SnAK-mediated phosphorylation of the Thr-175 residue in the SnRK1 activation loop (Jossier et al., 2009; Shen et al., 2009; Glab et al., 2017). CPuORFs are present in both SnAK1 and SnAK2 (Jorgensen and Dorantes-Acosta, 2012). Activated SnRK1 regulates the expression of a multitude of genes and metabolic enzymes during adaptation to a low-energy state (Crepin and Rolland, 2019). TOR responds to nutrient abundance by promoting ribosome biogenesis, cell cycle progression, and growth (Shi et al., 2018). The TOR complex consists of the TOR kinase, LST8, and RAPTOR. A CPuORF initiating with a CUG codon was found in the dominantly expressed RAPTOR1 gene (AT3G08850; van der Horst et al., 2019). SnRK1 can phosphorylate RAPTOR, thereby inhibiting TOR activity (Nukarinen et al., 2016). Following uORF translation, activated TOR phosphorylates S6K1 that subsequently phosphorylates eIF3h to promote reinitiation at the mORF (Schepetilnikov et al., 2013). Cereal species have a CPuORF upstream of S6K (Tran et al., 2008); however, cognate start codons were not identified for CPuORFs in S6Ks of Brassicaceae species (van der Horst et al., 2019).
The signaling molecule trehalose-6-phosphate (T6P) inhibits SnRK1 activity (Crepin and Rolland, 2019). T6P is synthesized by T6P synthase from UDP-Glc and Glc-6-P and is metabolized by TPP to trehalose. CPuORFs were found in two of the 11 Arabidopsis TPP genes, TPPF and TPPG (Jorgensen and Dorantes-Acosta, 2012). The S1 group bZIP11 transcription factor induces TPP expression, thereby lowering T6P levels (Ma et al., 2011) and enabling SnRK1 activity.
The essential translation initiation factor eIF5-1 (AT1G36730) is involved in start codon selection. eIF5-1 is conserved in eukaryotes and harbors a CPuORF in plants (Jorgensen and Dorantes-Acosta, 2012). In human cells, eIF5 is autoregulated via negative feedback involving uORFs. Overexpressing eIF5 in mammalian cells increases translation initiation at upstream non-AUGs and weak AUG initiation contexts (Loughran et al., 2012).
From the above, it is evident that the energy signaling network is tightly interwoven and controlled at multiple points by CPuORFs. In this signaling network, the metabolites that control CPuORF-mediated stalling are known only for S1 bZIPs. Uncovering the controlling cofactors that regulate translation of the other CPuORF-containing genes is urgently needed to improve our understanding of this key signaling network.
CONCLUDING REMARKS
In plants, the large number of CPuORFs and their broad evolutionary conservation reflect their importance as a regulatory principle. Where known (Table 1), CPuORFs respond to metabolites within the context of the ribosomal exit tunnel to regulate mORF translation. The ribosome, the CPuORF, and the metabolite in a ménage a trois control mORF translation and in turn biological function.
It is likely that other CPuORFs in addition to those listed in Table 1 will similarly respond to metabolites, and it will be exciting to uncover such metabolites or, perhaps, cofactors such as oligosaccharides or small peptides, and the processes that are under CPuORF control (see Outstanding Questions). Usually, CPuORFs are embedded in complex uORF ensembles with nonconserved uORFs in front, overlapping and downstream of the CPuORF. How such uORF ensembles affect CPuORF function and mORF translation requires further investigation.
Metabolite-regulated CPuORFs present a direct regulatory mechanism for gene expression, dynamically operating within the physiological range of the controlling metabolite. Several of the experimentally confirmed metabolite-controlled CPuORFs concern mRNAs encoding metabolic enzymes involved in their biosynthesis. Such homeostatic feedback loops allow for direct fine-tuning of metabolite status and cellular function. Other CPuORFs appear to operate indirectly by regulating the expression of regulatory proteins such as transcription factors.
ACKNOWLEDGMENTS
S.S. is most grateful to Andrew Millar and Karen Halliday of SynthSys, University of Edinburgh, for their kind hospitality during a sabbatical stay.
Footnotes
This work was supported by The Netherlands Research Council (project ALWOP.2015.115).
Articles can be viewed without a subscription.
References
- Aibara I, Hirai T, Kasai K, Takano J, Onouchi H, Naito S, Fujiwara T, Miwa K (2018) Boron-dependent translational suppression of the borate exporter BOR1 contributes to the avoidance of boron toxicity. Plant Physiol 177: 759–774 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alatorre-Cobos F, Cruz-Ramírez A, Hayden CA, Pérez-Torres CA, Chauvin AL, Ibarra-Laclette E, Alva-Cortés E, Jorgensen RA, Herrera-Estrella L (2012) Translational regulation of Arabidopsis XIPOTL1 is modulated by phosphocholine levels via the phylogenetically conserved upstream open reading frame 30. J Exp Bot 63: 5203–5221 [DOI] [PubMed] [Google Scholar]
- Arenz S, Bock LV, Graf M, Innis CA, Beckmann R, Grubmüller H, Vaiana AC, Wilson DN (2016) A combined cryo-EM and molecular dynamics approach reveals the mechanism of ErmBL-mediated translation arrest. Nat Commun 7: 12026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bailey-Serres J, Ma W (2017) Plant biology: An immunity boost combats crop disease. Nature 545: 420–421 [DOI] [PubMed] [Google Scholar]
- Bazin J, Baerenfaller K, Gosai SJ, Gregory BD, Crespi M, Bailey-Serres J (2017) Global analysis of ribosome-associated noncoding RNAs unveils new modes of translational regulation. Proc Natl Acad Sci USA 114: E10018–E10027 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bhushan S, Meyer H, Starosta AL, Becker T, Mielke T, Berninghausen O, Sattler M, Wilson DN, Beckmann R (2010) Structural basis for translational stalling by human cytomegalovirus and fungal arginine attenuator peptide. Mol Cell 40: 138–146 [DOI] [PubMed] [Google Scholar]
- Bianchini C, Pastore A, Pelucchi S, Torreggiani E, Lambertini E, Marchesi E, Magri E, Frasson C, Querzoli P, Piva R (2008) Sex hormone receptor levels in laryngeal carcinoma: A comparison between protein and RNA evaluations. Eur Arch Otorhinolaryngol 265: 1089–1094 [DOI] [PubMed] [Google Scholar]
- Bischoff L, Berninghausen O, Beckmann R (2014) Molecular basis for the ribosome functioning as an L-tryptophan sensor. Cell Rep 9: 469–475 [DOI] [PubMed] [Google Scholar]
- Browning KS, Bailey-Serres J (2015) Mechanism of cytoplasmic mRNA translation. The Arabidopsis Book 13: e0176. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Byrne ME. (2009) A role for the ribosome in development. Trends Plant Sci 14: 512–519 [DOI] [PubMed] [Google Scholar]
- Cai Q, Fukushima H, Yamamoto M, Ishii N, Sakamoto T, Kurata T, Motose H, Takahashi T (2016) The SAC51 family plays a central role in thermospermine responses in Arabidopsis. Plant Cell Physiol 57: 1583–1592 [DOI] [PubMed] [Google Scholar]
- Chantarachot T, Bailey-Serres J (2018) Polysomes, stress granules, and processing bodies: A dynamic triumvirate controlling cytoplasmic mRNA fate and function. Plant Physiol 176: 254–269 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chiba Y, Ishikawa M, Kijima F, Tyson RH, Kim J, Yamamoto A, Nambara E, Leustek T, Wallsgrove RM, Naito S (1999) Evidence for autoregulation of cystathionine γ-synthase mRNA stability in Arabidopsis. Science 286: 1371–1374 [DOI] [PubMed] [Google Scholar]
- Chiba Y, Sakurai R, Yoshino M, Ominato K, Ishikawa M, Onouchi H, Naito S (2003) S-Adenosyl-L-methionine is an effector in the posttranscriptional autoregulation of the cystathionine gamma-synthase gene in Arabidopsis. Proc Natl Acad Sci USA 100: 10225–10230 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Conrads KA, Yi M, Simpson KA, Lucas DA, Camalier CE, Yu LR, Veenstra TD, Stephens RM, Conrads TP, Beck GR Jr. (2005) A combined proteome and microarray investigation of inorganic phosphate-induced pre-osteoblast cells. Mol Cell Proteomics 4: 1284–1296 [DOI] [PubMed] [Google Scholar]
- Crepin N, Rolland F (2019) SnRK1 activation, signaling, and networking for energy homeostasis. Curr Opin Plant Biol 51: 29–36 [DOI] [PubMed] [Google Scholar]
- Dao Duc K, Batra SS, Bhattacharya N, Cate JHD, Song YS (2019) Differences in the path to exit the ribosome across the three domains of life. Nucleic Acids Res 47: 4198–4210 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Diaz de Arce AJ, Noderer WL, Wang CL (2018) Complete motif analysis of sequence requirements for translation initiation at non-AUG start codons. Nucleic Acids Res 46: 985–994 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dröge-Laser W, Weiste C (2018) The C/S1 bZIP network: A regulatory hub orchestrating plant energy homeostasis. Trends Plant Sci 23: 422–433 [DOI] [PubMed] [Google Scholar]
- Ebina I, Takemoto-Tsutsumi M, Watanabe S, Koyama H, Endo Y, Kimata K, Igarashi T, Murakami K, Kudo R, Ohsumi A, et al. (2015) Identification of novel Arabidopsis thaliana upstream open reading frames that control expression of the main coding sequences in a peptide sequence-dependent manner. Nucleic Acids Res 43: 1562–1576 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gibon Y, Usadel B, Blaesing OE, Kamlage B, Hoehne M, Trethewey R, Stitt M (2006) Integration of metabolite with transcript and enzyme activity profiling during diurnal cycles in Arabidopsis rosettes. Genome Biol 7: R76. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Glab N, Oury C, Guérinier T, Domenichini S, Crozet P, Thomas M, Vidal J, Hodges M (2017) The impact of Arabidopsis thaliana SNF1-related-kinase 1 (SnRK1)-activating kinase 1 (SnAK1) and SnAK2 on SnRK1 phosphorylation status: Characterization of a SnAK double mutant. Plant J 89: 1031–1041 [DOI] [PubMed] [Google Scholar]
- Guerrero-González ML, Ortega-Amaro MA, Juárez-Montiel M, Jiménez-Bremont JF (2016) Arabidopsis polyamine oxidase-2 uORF is required for downstream translational regulation. Plant Physiol Biochem 108: 381–390 [DOI] [PubMed] [Google Scholar]
- Guerrero-González ML, Rodríguez-Kessler M, Jiménez-Bremont JF (2014) uORF, a regulatory mechanism of the Arabidopsis polyamine oxidase 2. Mol Biol Rep 41: 2427–2443 [DOI] [PubMed] [Google Scholar]
- Hanfrey C, Elliott KA, Franceschetti M, Mayer MJ, Illingworth C, Michael AJ (2005) A dual upstream open reading frame-based autoregulatory circuit controlling polyamine-responsive translation. J Biol Chem 280: 39229–39237 [DOI] [PubMed] [Google Scholar]
- Hanfrey C, Franceschetti M, Mayer MJ, Illingworth C, Michael AJ (2002) Abrogation of upstream open reading frame-mediated translational control of a plant S-adenosylmethionine decarboxylase results in polyamine disruption and growth perturbations. J Biol Chem 277: 44131–44139 [DOI] [PubMed] [Google Scholar]
- Hanson J, Hanssen M, Wiese A, Hendriks MMWB, Smeekens S (2008) The sucrose regulated transcription factor bZIP11 affects amino acid metabolism by regulating the expression of ASPARAGINE SYNTHETASE1 and PROLINE DEHYDROGENASE2. Plant J 53: 935–949 [DOI] [PubMed] [Google Scholar]
- Hayashi N, Sasaki S, Takahashi H, Yamashita Y, Naito S, Onouchi H (2017) Identification of Arabidopsis thaliana upstream open reading frames encoding peptide sequences that cause ribosomal arrest. Nucleic Acids Res 45: 8844–8858 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hayden CA, Jorgensen RA (2007) Identification of novel conserved peptide uORF homology groups in Arabidopsis and rice reveals ancient eukaryotic origin of select groups and preferential association with transcription factor-encoding genes. BMC Biol 5: 32. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hellens RP, Brown CM, Chisnall MAW, Waterhouse PM, Macknight RC (2016) The emerging world of small ORFs. Trends Plant Sci 21: 317–328 [DOI] [PubMed] [Google Scholar]
- Horiguchi G, Mollá-Morales A, Pérez-Pérez JM, Kojima K, Robles P, Ponce MR, Micol JL, Tsukaya H (2011) Differential contributions of ribosomal protein genes to Arabidopsis thaliana leaf development. Plant J 65: 724–736 [DOI] [PubMed] [Google Scholar]
- Hou CY, Lee WC, Chou HC, Chen AP, Chou SJ, Chen HM (2016) Global analysis of truncated RNA ends reveals new insights into ribosome stalling in plants. Plant Cell 28: 2398–2416 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hsu PY, Calviello L, Wu HL, Li FW, Rothfels CJ, Ohler U, Benfey PN (2016) Super-resolution ribosome profiling reveals unannotated translation events in Arabidopsis. Proc Natl Acad Sci USA 113: E7126–E7135 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hummel M, Cordewener JHG, de Groot JCM, Smeekens S, America AHP, Hanson J (2012) Dynamic protein composition of Arabidopsis thaliana cytosolic ribosomes in response to sucrose feeding as revealed by label free MSE proteomics. Proteomics 12: 1024–1038 [DOI] [PubMed] [Google Scholar]
- Hummel M, Dobrenel T, Cordewener JJHG, Davanture M, Meyer C, Smeekens SJCM, Bailey-Serres J, America TAHP, Hanson J (2015) Proteomic LC-MS analysis of Arabidopsis cytosolic ribosomes: Identification of ribosomal protein paralogs and re-annotation of the ribosomal protein genes. J Proteomics 128: 436–449 [DOI] [PubMed] [Google Scholar]
- Hummel M, Rahmani F, Smeekens S, Hanson J (2009) Sucrose-mediated translational control. Ann Bot 104: 1–7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Imai A, Hanzawa Y, Komura M, Yamamoto KT, Komeda Y, Takahashi T (2006) The dwarf phenotype of the Arabidopsis acl5 mutant is suppressed by a mutation in an upstream ORF of a bHLH gene. Development 133: 3575–3585 [DOI] [PubMed] [Google Scholar]
- Imai A, Komura M, Kawano E, Kuwashiro Y, Takahashi T (2008) A semi-dominant mutation in the ribosomal protein L10 gene suppresses the dwarf phenotype of the acl5 mutant in Arabidopsis thaliana. Plant J 56: 881–890 [DOI] [PubMed] [Google Scholar]
- Ishitsuka S, Yamamoto M, Miyamoto M, Kuwashiro Y, Imai A, Motose H, Takahashi T (2019) Complexity and conservation of thermospermine-responsive uORFs of SAC51 family genes in angiosperms. Front Plant Sci 10: 564. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ito K, Chiba S (2013) Arrest peptides: Cis-acting modulators of translation. Annu Rev Biochem 82: 171–202 [DOI] [PubMed] [Google Scholar]
- Ivanov IP, Atkins JF, Michael AJ (2010) A profusion of upstream open reading frame mechanisms in polyamine-responsive translational regulation. Nucleic Acids Res 38: 353–359 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jorgensen RA, Dorantes-Acosta AE (2012) Conserved peptide upstream open reading frames are associated with regulatory genes in angiosperms. Front Plant Sci 3: 191. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jossier M, Bouly JP, Meimoun P, Arjmand A, Lessard P, Hawley S, Hardie DG, Thomas M (2009) SnRK1 (SNF1-related kinase 1) has a central role in sugar and ABA signalling in Arabidopsis thaliana. Plant J 59: 316–328 [DOI] [PubMed] [Google Scholar]
- Juntawong P, Girke T, Bazin J, Bailey-Serres J (2014) Translational dynamics revealed by genome-wide profiling of ribosome footprints in Arabidopsis. Proc Natl Acad Sci USA 111: E203–E212 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kakehi J, Kawano E, Yoshimoto K, Cai Q, Imai A, Takahashi T (2015) Mutations in ribosomal proteins, RPL4 and RACK1, suppress the phenotype of a thermospermine-deficient mutant of Arabidopsis thaliana. PLoS ONE 10: e0117309. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim MS, Cho SM, Kang EY, Im YJ, Hwangbo H, Kim YC, Ryu CM, Yang KY, Chung GC, Cho BH (2008) Galactinol is a signaling component of the induced systemic resistance caused by Pseudomonas chlororaphis O6 root colonization. Mol Plant Microbe Interact 21: 1643–1653 [DOI] [PubMed] [Google Scholar]
- Knott JM, Römer P, Sumper M (2007) Putative spermine synthases from Thalassiosira pseudonana and Arabidopsis thaliana synthesize thermospermine rather than spermine. FEBS Lett 581: 3081–3086 [DOI] [PubMed] [Google Scholar]
- Kochetov AV. (2008) Alternative translation start sites and hidden coding potential of eukaryotic mRNAs. BioEssays 30: 683–691 [DOI] [PubMed] [Google Scholar]
- Laing WA, Martínez-Sánchez M, Wright MA, Bulley SM, Brewster D, Dare AP, Rassam M, Wang D, Storey R, Macknight RC, et al. (2015) An upstream open reading frame is essential for feedback regulation of ascorbate biosynthesis in Arabidopsis. Plant Cell 27: 772–786 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li W, Ward FR, McClure KF, Chang ST, Montabana E, Liras S, Dullea RG, Cate JHD (2019) Structural basis for selective stalling of human ribosome nascent chain complexes by a drug-like molecule. Nat Struct Mol Biol 26: 501–509 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lintner NG, McClure KF, Petersen D, Londregan AT, Piotrowski DW, Wei L, Xiao J, Bolt M, Loria PM, Maguire B, et al. (2017) Selective stalling of human translation through small-molecule engagement of the ribosome nascent chain. PLoS Biol 15: e2001882. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu MJ, Wu SH, Wu JF, Lin WD, Wu YC, Tsai TY, Tsai HL, Wu SH (2013) Translational landscape of photomorphogenic Arabidopsis. Plant Cell 25: 3699–3710 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lorenzo-Orts L, Witthoeft J, Deforges J, Martinez J, Loubéry S, Placzek A, Poirier Y, Hothorn LA, Jaillais Y, Hothorn M (2019) Concerted expression of a cell cycle regulator and a metabolic enzyme from a bicistronic transcript in plants. Nat Plants 5: 184–193 [DOI] [PubMed] [Google Scholar]
- Loughran G, Sachs MS, Atkins JF, Ivanov IP (2012) Stringency of start codon selection modulates autoregulation of translation initiation factor eIF5. Nucleic Acids Res 40: 2898–2906 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ma J, Hanssen M, Lundgren K, Hernández L, Delatte T, Ehlert A, Liu CM, Schluepmann H, Dröge-Laser W, Moritz T, et al. (2011) The sucrose-regulated Arabidopsis transcription factor bZIP11 reprograms metabolism and regulates trehalose metabolism. New Phytol 191: 733–745 [DOI] [PubMed] [Google Scholar]
- Merchante C, Brumos J, Yun J, Hu Q, Spencer KR, Enríquez P, Binder BM, Heber S, Stepanova AN, Alonso JM (2015) Gene-specific translation regulation mediated by the hormone-signaling molecule EIN2. Cell 163: 684–697 [DOI] [PubMed] [Google Scholar]
- Merchante C, Stepanova AN, Alonso JM (2017) Translation regulation in plants: An interesting past, an exciting present and a promising future. Plant J 90: 628–653 [DOI] [PubMed] [Google Scholar]
- Nogales E, Scheres SHW (2015) Cryo-EM: A unique tool for the visualization of macromolecular complexity. Mol Cell 58: 677–689 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nukarinen E, Nägele T, Pedrotti L, Wurzinger B, Mair A, Landgraf R, Börnke F, Hanson J, Teige M, Baena-Gonzalez E, et al. (2016) Quantitative phosphoproteomics reveals the role of the AMPK plant ortholog SnRK1 as a metabolic master regulator under energy deprivation. Sci Rep 6: 31697. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Onouchi H, Nagami Y, Haraguchi Y, Nakamoto M, Nishimura Y, Sakurai R, Nagao N, Kawasaki D, Kadokura Y, Naito S (2005) Nascent peptide-mediated translation elongation arrest coupled with mRNA degradation in the CGS1 gene of Arabidopsis. Genes Dev 19: 1799–1810 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Onoue N, Yamashita Y, Nagao N, Goto DB, Onouchi H, Naito S (2011) S-Adenosyl-L-methionine induces compaction of nascent peptide chain inside the ribosomal exit tunnel upon translation arrest in the Arabidopsis CGS1 gene. J Biol Chem 286: 14903–14912 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pajerowska-Mukhtar KM, Wang W, Tada Y, Oka N, Tucker CL, Fonseca JP, Dong X (2012) The HSF-like transcription factor TBF1 is a major molecular switch for plant growth-to-defense transition. Curr Biol 22: 103–112 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Peviani A, Lastdrager J, Hanson J, Snel B (2016) The phylogeny of C/S1 bZIP transcription factors reveals a shared algal ancestry and the pre-angiosperm translational regulation of S1 transcripts. Sci Rep 6: 30444. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Poidevin L, Unal D, Belda-Palazón B, Ferrando A (2019) Polyamines as quality control metabolites operating at the post-transcriptional level. Plants (Basel) 8: 109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rahmani F, Hummel M, Schuurmans J, Wiese-Klinkenberg A, Smeekens S, Hanson J (2009) Sucrose control of translation mediated by an upstream open reading frame-encoded peptide. Plant Physiol 150: 1356–1367 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rook F, Gerrits N, Kortstee A, van Kampen M, Borrias M, Weisbeek P, Smeekens S (1998) Sucrose-specific signalling represses translation of the Arabidopsis ATB2 bZIP transcription factor gene. Plant J 15: 253–263 [DOI] [PubMed] [Google Scholar]
- Sagor GHM, Berberich T, Tanaka S, Nishiyama M, Kanayama Y, Kojima S, Muramoto K, Kusano T (2016) A novel strategy to produce sweeter tomato fruits with high sugar contents by fruit-specific expression of a single bZIP transcription factor gene. Plant Biotechnol J 14: 1116–1126 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schepetilnikov M, Dimitrova M, Mancera-Martínez E, Geldreich A, Keller M, Ryabova LA (2013) TOR and S6K1 promote translation reinitiation of uORF-containing mRNAs via phosphorylation of eIF3h. EMBO J 32: 1087–1102 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schepetilnikov M, Ryabova LA (2017) Auxin signaling in regulation of plant translation reinitiation. Front Plant Sci 8: 1014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shen W, Reyes MI, Hanley-Bowdoin L (2009) Arabidopsis protein kinases GRIK1 and GRIK2 specifically activate SnRK1 by phosphorylating its activation loop. Plant Physiol 150: 996–1005 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shi L, Wu Y, Sheen J (2018) TOR signaling in plants: Conservation and innovation. Development 145: dev160887. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tabuchi T, Okada T, Azuma T, Nanmori T, Yasuda T (2006) Posttranscriptional regulation by the upstream open reading frame of the phosphoethanolamine N-methyltransferase gene. Biosci Biotechnol Biochem 70: 2330–2334 [DOI] [PubMed] [Google Scholar]
- Takahashi H, Hayashi N, Yamashita Y, Naito S, Takahashi A, Fuse K, Endo T, Kojima S, Onouchi H (2019) ESUCA: A pipeline for genome-wide identification of upstream open reading frames with evolutionarily conserved sequences and determination of the taxonomic range of their conservation. bioRxiv 524090 [Google Scholar]
- Takahashi H, Takahashi A, Naito S, Onouchi H (2012) BAIUCAS: A novel BLAST-based algorithm for the identification of upstream open reading frames with conserved amino acid sequences and its application to the Arabidopsis thaliana genome. Bioinformatics 28: 2231–2241 [DOI] [PubMed] [Google Scholar]
- Tanaka M, Sotta N, Yamazumi Y, Yamashita Y, Miwa K, Murota K, Chiba Y, Hirai MY, Akiyama T, Onouchi H, et al. (2016) The minimum open reading frame, AUG-stop, induces boron-dependent ribosome stalling and mRNA degradation. Plant Cell 28: 2830–2849 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Thalor SK, Berberich T, Lee SS, Yang SH, Zhu X, Imai R, Takahashi Y, Kusano T (2012) Deregulation of sucrose-controlled translation of a bZIP-type transcription factor results in sucrose accumulation in leaves. PLoS ONE 7: e33111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tran MK, Schultz CJ, Baumann U (2008) Conserved upstream open reading frames in higher plants. BMC Genomics 9: 361. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Uchiyama-Kadokura N, Murakami K, Takemoto M, Koyanagi N, Murota K, Naito S, Onouchi H (2014) Polyamine-responsive ribosomal arrest at the stop codon of an upstream open reading frame of the AdoMetDC1 gene triggers nonsense-mediated mRNA decay in Arabidopsis thaliana. Plant Cell Physiol 55: 1556–1567 [DOI] [PubMed] [Google Scholar]
- van der Horst S, Snel B, Hanson J, Smeekens S (2019) Novel pipeline identifies new upstream ORFs and non-AUG initiating main ORFs with conserved amino acid sequences in the 5′ leader of mRNAs in Arabidopsis thaliana. RNA 25: 292–304 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vaughn JN, Ellingson SR, Mignone F, von Arnim A (2012) Known and novel post-transcriptional regulatory sequences are conserved across plant families. RNA 18: 368–384 [DOI] [PMC free article] [PubMed] [Google Scholar]
- von Arnim AG, Jia Q, Vaughn JN (2014) Regulation of plant translation by upstream open reading frames. Plant Sci 214: 1–12 [DOI] [PubMed] [Google Scholar]
- Weltmeier F, Rahmani F, Ehlert A, Dietrich K, Schütze K, Wang X, Chaban C, Hanson J, Teige M, Harter K, et al. (2009) Expression patterns within the Arabidopsis C/S1 bZIP transcription factor network: Availability of heterodimerization partners controls gene expression during stress response and development. Plant Mol Biol 69: 107–119 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wiese A, Elzinga N, Wobbes B, Smeekens S (2004) A conserved upstream open reading frame mediates sucrose-induced repression of translation. Plant Cell 16: 1717–1729 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wilson DN, Arenz S, Beckmann R (2016) Translation regulation via nascent polypeptide-mediated ribosome stalling. Curr Opin Struct Biol 37: 123–133 [DOI] [PubMed] [Google Scholar]
- Xu G, Yuan M, Ai C, Liu L, Zhuang E, Karapetyan S, Wang S, Dong X (2017) uORF-mediated translation allows engineered plant disease resistance without fitness costs. Nature 545: 491–494 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yamashita Y, Takamatsu S, Glasbrenner M, Becker T, Naito S, Beckmann R (2017) Sucrose sensing through nascent peptide-meditated ribosome stalling at the stop codon of Arabidopsis bZIP11 uORF2. FEBS Lett 591: 1266–1277 [DOI] [PubMed] [Google Scholar]
- Yu X, Willmann MR, Anderson SJ, Gregory BD (2016) Genome-wide mapping of uncapped and cleaved transcripts reveals a role for the nuclear mRNA cap-binding complex in cotranslational RNA decay in Arabidopsis. Plant Cell 28: 2385–2397 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang H, Si X, Ji X, Fan R, Liu J, Chen K, Wang D, Gao C (2018) Genome editing of upstream open reading frames enables translational control in plants. Nat Biotechnol 36: 894–898 [DOI] [PubMed] [Google Scholar]
- Zhang H, Wang Y, Lu J (2019) Function and evolution of upstream ORFs in eukaryotes. Trends Biochem Sci 44: 782–794 [DOI] [PubMed] [Google Scholar]
- Zhu X, Li Y, Fang W, Kusano T (2018) Galactinol is involved in sequence-conserved upstream open reading frame-mediated repression of Arabidopsis HsfB1 translation. Environ Exp Bot 156: 120–129 [Google Scholar]
- Zhu X, Thalor SK, Takahashi Y, Berberich T, Kusano T (2012) An inhibitory effect of the sequence-conserved upstream open-reading frame on the translation of the main open-reading frame of HsfB1 transcripts in Arabidopsis. Plant Cell Environ 35: 2014–2030 [DOI] [PubMed] [Google Scholar]







