Next Article in Journal
Expression, Polymorphism, and Potential Functional Sites of the BMPR1A Gene in the Sheep Horn
Previous Article in Journal
Long-Term Tissue Preservation at Ambient Temperature for Post-Mass Fatality Incident DNA-Based Victim Identification
Previous Article in Special Issue
yaaJ, the tRNA-Specific Adenosine Deaminase, Is Dispensable in Bacillus subtilis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Beyond the Anticodon: tRNA Core Modifications and Their Impact on Structure, Translation and Stress Adaptation

by
Marcel-Joseph Yared
,
Agathe Marcelot
and
Pierre Barraud
*
Expression Génétique Microbienne, Université Paris Cité, CNRS, Institut de Biologie Physico-Chimique, F-75005 Paris, France
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Genes 2024, 15(3), 374; https://doi.org/10.3390/genes15030374
Submission received: 21 February 2024 / Revised: 15 March 2024 / Accepted: 18 March 2024 / Published: 19 March 2024
(This article belongs to the Special Issue Transfer RNA Modification)

Abstract

:
Transfer RNAs (tRNAs) are heavily decorated with post-transcriptional chemical modifications. Approximately 100 different modifications have been identified in tRNAs, and each tRNA typically contains 5–15 modifications that are incorporated at specific sites along the tRNA sequence. These modifications may be classified into two groups according to their position in the three-dimensional tRNA structure, i.e., modifications in the tRNA core and modifications in the anticodon-loop (ACL) region. Since many modified nucleotides in the tRNA core are involved in the formation of tertiary interactions implicated in tRNA folding, these modifications are key to tRNA stability and resistance to RNA decay pathways. In comparison to the extensively studied ACL modifications, tRNA core modifications have generally received less attention, although they have been shown to play important roles beyond tRNA stability. Here, we review and place in perspective selected data on tRNA core modifications. We present their impact on tRNA structure and stability and report how these changes manifest themselves at the functional level in translation, fitness and stress adaptation.

1. Introduction

Transfer RNAs (tRNAs) serve as adaptor molecules during the translation process, where they deliver amino acids to the ribosome and hold a central role in decoding the genetic information contained in messenger RNAs (mRNAs). To fulfill this fundamental function in cellular protein synthesis, tRNAs are produced following a multi-step biogenesis process that leads to the formation of functional tRNAs [1,2,3,4]. Mature tRNAs are generally ∼75–95 nucleotides in length and adopt a cloverleaf secondary structure. This structure defines several regions: the acceptor stem; the D-stem and D-loop (also referred to as the D-arm); the anticodon-stem and anticodon-loop (ACL); the variable region (which usually contains 4 nucleotides but can contain ∼10–25 nucleotides in certain tRNAs, e.g., tRNASer, tRNALeu and tRNATyr); the T-stem and T-loop (i.e., the T-arm); and the CCA-end (Figure 1a) [5,6]. This secondary structure assembly folds into a uniform three-dimensional structure: the L-shaped architecture. This architectural hallmark, in which the T-stem is stacked on the acceptor stem and the D-stem is stacked on the anticodon-stem, is assembled as a result of multiple tertiary interactions between the T- and D-loops, thereby forming the tRNA elbow, and between the variable region and the D-stem (Figure 1b) [7,8,9,10,11].
A striking feature of tRNAs is their universal incorporation of post-transcriptional chemical modifications, with tRNAs being the RNA family that exhibits both the greatest diversity of modifications and the greatest number of modifications per molecule [12,13]. Although ∼100 different modifications have been identified in tRNAs, each tRNA typically contains from 5 to 15 modifications that are incorporated at specific sites along the tRNA sequence [14,15]. Many modified nucleotides are implicated in the formation of the tertiary interactions that lead to the L-shaped structure of tRNAs (see, for instance, the position of modified nucleotides in E. coli and S. cerevisiaeFigure 1c,d) and are thus key to tRNA stability [16,17] and resistance to RNA decay pathways [18,19,20,21]. Based on their position in the three-dimensional structure, modifications can be divided into two groups, namely modifications in the tRNA core and modifications in the ACL region (Figure 1).
Figure 1. tRNA structure and post-transcriptional modifications: (a) Cloverleaf representation of tRNA secondary structure. tRNAs are composed of five main regions, namely the acceptor stem (in purple) with the CCA-end (in grey), the D-arm consisting of the D-loop and D-stem (in blue), the anticodon-arm (in green), the variable region (in yellow), and the T-arm consisting of the T-loop and T-stem (in red). Residue numbers are according to the standard tRNA nomenclature [22]. (b) Schematic representation of the L-shaped tertiary structure of tRNA with the same color codes as in (a). The acceptor-stem stacks on the T-arm, and the D-arm stacks on the anticodon-arm. The T-loop and D-loop interact and form the tRNA elbow structure. Main tertiary interactions between the T- and D-loops and between the variable region and the D-stem are represented with dashed lines, and the positions of the implicated nucleotides are indicated. (c) Locations of modifications in Escherichia coli tRNAs. (d) Locations of modifications in Saccharomyces cerevisiae tRNAs. Data are compiled from the MODOMICS database [14] and references [23,24] for E. coli tRNAs and [1,2] for S. cerevisiae tRNAs. Modified positions are represented as red circles on the schematic view of the L-shaped tertiary structure of tRNA. Residue numbers and modifications found in E. coli and S. cerevisiae are given for each modified position.
Figure 1. tRNA structure and post-transcriptional modifications: (a) Cloverleaf representation of tRNA secondary structure. tRNAs are composed of five main regions, namely the acceptor stem (in purple) with the CCA-end (in grey), the D-arm consisting of the D-loop and D-stem (in blue), the anticodon-arm (in green), the variable region (in yellow), and the T-arm consisting of the T-loop and T-stem (in red). Residue numbers are according to the standard tRNA nomenclature [22]. (b) Schematic representation of the L-shaped tertiary structure of tRNA with the same color codes as in (a). The acceptor-stem stacks on the T-arm, and the D-arm stacks on the anticodon-arm. The T-loop and D-loop interact and form the tRNA elbow structure. Main tertiary interactions between the T- and D-loops and between the variable region and the D-stem are represented with dashed lines, and the positions of the implicated nucleotides are indicated. (c) Locations of modifications in Escherichia coli tRNAs. (d) Locations of modifications in Saccharomyces cerevisiae tRNAs. Data are compiled from the MODOMICS database [14] and references [23,24] for E. coli tRNAs and [1,2] for S. cerevisiae tRNAs. Modified positions are represented as red circles on the schematic view of the L-shaped tertiary structure of tRNA. Residue numbers and modifications found in E. coli and S. cerevisiae are given for each modified position.
Genes 15 00374 g001
The implication of ACL modifications in the decoding process has been extensively studied, providing many details on the function and dysfunction associated with these modifications (reviewed in [25,26,27,28,29]). The ACL contains complex tRNA modifications that are involved in codon recognition (e.g., at wobble position 34) and reading frame maintenance (e.g., at position 37) [30,31,32,33,34,35,36,37]. Modifications at the wobble position can either restrict or expand the decoding properties of the tRNA, and modifications in the ACL are thus intricately linked to the efficiency and accuracy of translation. Thus, modifications in the ACL can modulate the expression of specific genes that are enriched in codons decoded by tRNAs modified at the wobble position. This notion of modification-tunable expression of codon-biased genes allows for an additional layer of regulation in the gene expression process, which may have general implications for cellular physiology and stress adaptation (reviewed in [38,39,40,41]).
Modifications in the tRNA core, i.e., outside the ACL region, are also implicated in several important cellular functions, such as translation, fitness and stress adaptation. However, the molecular mechanisms involved are generally less well characterized and, therefore, more difficult to comprehend than those involving modifications in the ACL. In addition, in bacteria and yeast, most essential genes, or genes whose deletion leads to marked growth phenotypes, are involved in the introduction of modifications at positions 32, 34 or 37 in the ACL. These positions and their respective modifications have therefore been the focus of many of the studies on tRNA modifications and their associated functions. Overall, modifications of the tRNA core may have been neglected in comparison. In recent years, however, a number of studies have sought to characterize and understand the role of some of these modifications (a few examples are shown in Figure 2). Some tRNA core modifications appear to have roles of their own, with important functions beyond tRNA stability.
The present review aims to bring together and put into perspective selected data on tRNA core modifications. We focus on modifications of cytoplasmic tRNAs in bacteria and yeast. We may take a few detours into higher eukaryotes, but readers interested in these topics are advised to consult more appropriate reviews [42,43,44]. We first present the impact of these modifications on tRNA structure and stability and the hierarchical introduction of other modifications. We then report and discuss how these changes manifest themselves at the functional level and examine their impact on translation, fitness and stress adaptation. Overall, we aim to emphasize the importance of tRNA core modifications, which should not be overlooked in favor of their anticodon-loop counterparts.

2. Influence of tRNA Core Modification on tRNA Structure and Stability

2.1. Modifications and Their Effect on the Physicochemical Properties of the Nucleotides

The majority of the basic knowledge concerning the impact of core modifications on tRNA structure was mainly discovered towards the end of the last century and has been comprehensively reviewed and discussed elsewhere [16,17,45,46,47,48]. We provide here only a concise overview of these fundamentals, which are at the root of all local and global molecular changes associated with the introduction of modifications.
Pseudouridylation, 2′-O-methylation and 2-thiolation enhance tRNA structural stability through stabilization of the C3′-endo conformation of the ribose moiety and improvement of the base stacking properties of these modified nucleotides [49,50,51,52]. This enhances the thermostability of tRNAs containing these modifications. For instance, Ψ 55, Ψ 40 and Gm18 individually increase the melting temperature of E. coli tRNASer [53], and m5s2U54 increases the melting temperature of thermophile tRNAs [54,55]. In addition, 4-thiolation, such as s4U8, also increases tRNA melting temperatures [53], probably by strongly reinforcing the s4U8-A14 reverse Hoogsteen base pair (Figure 1c) [56]. Furthermore, as Gm18 and Ψ 55 are involved in a tertiary interaction between the D- and T-arms of the tRNAs (Figure 1c,d), they are implicated in stabilizing their global tertiary structure and, more particularly, their elbow region [57]. Pseudouridine exhibits a third stabilizing effect by providing an additional hydrogen-bond donor (H-N1) that may form a water-mediated interaction with the phosphate backbone [52,58,59]. This additional link increases the rigidity of the local tRNA structure and may contribute to the proper tertiary folding of tRNAs [60].
Simple methylations such as m5U or m5C also exhibit stabilizing properties due to their increased hydrophobicity, increased base polarizability, increased stacking capabilities and reinforcement of m5U54-m1A58 and G15-m5C48 tertiary interactions (Figure 1d) [60,61,62]. This explains why the presence of m5U54 increases the melting temperature of tRNAs [53,54]. Other methylations such as m1A58 and m7G46 introduce a positive charge at the purine aromatic ring of the base, which may stabilize the tRNA structure by interacting with negatively charged phosphates ions of the tRNA backbone or by stabilizing certain tertiary interactions, such as the C13-G22-m7G46 base triplet (Figure 1c,d) [47,63]. Other methylations, such as m1A9 and m 2 2 G26, disrupt the Watson–Crick base-pairing capacity of these nucleotides and are individually essential for the formation of the proper cloverleaf secondary structure and three-dimensional folding of certain tRNAs [64,65,66].
Only the dihydrouridine (D) modification has the capacity to enhance tRNA structural flexibility. Dihydrouridines effectively promote the C2′-endo sugar conformation and prevent proper stacking interaction with neighboring bases, thereby allowing for greater conformational flexibility and dynamic motion in the tRNA D-loop (reviewed in [67]).

2.2. Modifications in the tRNA Core and Their Impact on tRNA Structure and Dynamics

First, it is important to note that tRNA modifications can have little to no effect on some tRNAs while dramatically affecting others. For instance, in the yeast elongator tRNA Phe , the introduction of m1A58 has no detectable impact on the structure of the tRNA as observed with NMR [68,69]. On the contrary, the introduction of m1A58 on in vitro transcribed unmodified yeast tRNA i Met has a major effect on the tertiary folding of this tRNA and more particularly on the proper assembly of its elbow region [70]. NMR studies showed that unmodified tRNA i Met adopts different heterogeneously folded conformations while m1A58- tRNA i Met adopts one major homogenously folded conformation. This important structural role of m1A58 in tRNA i Met , which has a structural impact reminiscent of that of m1A9 and m 2 2 G26, is likely related to the unique tRNA substructure in the elbow region of eukaryotic initiator tRNAs [71]. It should be emphasized that the tRNA-dependent nature of the impact of modifications is perfectly in line with the idea that certain modifications do not necessarily have the same beneficial effect and are not equally useful for all tRNAs (reviewed in [72]).
Although dihydrouridines are considered to bring flexibility to tRNAs, their effects can be ambivalent. Dihydrouridines indeed allow greater conformational flexibility and dynamic motion in the tRNA D-loop [73] but also simultaneously constrain the D-stem to adopt a stable conformation, as seen in S. pombe tRNA i Met [74]. The stabilization effect of dihydrouridines is also observed in E. coli tRNASer where the absence of D20 decreases the melting temperature of this tRNA [53]. Lastly, it has been suggested that the interplay between the flexibility of D20 and the rigidity of Ψ 55 is needed to stabilize the elbow region of E. coli tRNA f Met [75]. Another example affecting the melting temperature of tRNAs concerns the acp 3 U47 core modification, which increases the thermal stability of E. coli tRNA Met by 3 °C [76]. The exact mechanism of how acp 3 U47 stabilizes the tertiary structure of tRNAs is not well understood. However, it has been suggested that it could be due to its inability to form Watson–Crick base pairing, thus ensuring the correct folding of tRNAs [76].
Although chemical modifications of nucleotides often increase the overall thermal stability of tRNAs, this may be achieved by an increase or a decrease in the local conformational dynamics in solution [45]. For instance, modified E. coli tRNA f Met has a higher overall structural and thermal stability, compared to its unmodified version, but presents an increase in local conformational fluctuations [75]. This behavior is attributed to entropic effects in which increased local conformational dynamics lead to a stabilization of the tertiary structure of the tRNA [75], in line with previous molecular dynamics simulations [77,78]. Interestingly, the authors observed that tRNA core modifications generate a remote effect of stabilization throughout the entire tRNA [75], which correlates well with an earlier report stating that modifications alter the intrinsic correlated motions within the tRNAs [52]. Effects of tRNA core modifications on the conformational dynamics of tRNAs are likely tRNA-dependent, and fully modified human mt- tRNA Leu , for instance, displays lesser folding dynamics than the unmodified transcript and exhibits a more constrained native structure that does not allow intermediate conformations [79].

2.3. Influence of tRNA Core Modification Enzymes on tRNA Structure: More Than Just Their Catalytic Activity

Modification enzymes have been reported to act as chaperones stabilizing and refolding their substrates independently of their enzymatic activity, thereby giving misfolded tRNAs a second chance to refold into native conformations (reviewed in [80]). Two tRNA modification enzymes acting as chaperones have been identified and characterized in E. coli, namely TruB introducing Ψ 55 and TrmA introducing m5U54 [81,82]. The chaperone activity of both of these enzymes comes from their ability to partially unfold and open the tRNA elbow region by disrupting tertiary interactions. It is thus likely that other modification enzymes possessing the capacity to unfold the tRNA elbow region would also act as tRNA-specific chaperones independently of their catalytic activity [83,84]. For instance, eukaryotic fission yeast Trm1 has been recently reported to have a chaperone-like activity similar to prokaryotic TruB and TrmA [85]. Trm1 enhances tRNA functionality in vivo, even in the absence of catalytic activity, and promotes RNA strand annealing and dissociation, as previously reported for other RNA chaperones [85,86].
It is worth mentioning that the RNA chaperone La, implicated in tRNA 3′-end protection and pre-tRNA folding, shares some functional redundancy with tRNA modification enzymes displaying chaperone activity, such as Trm1 and Pus4 [85,87,88]. The La chaperone does not discriminate between misfolded and properly folded tRNAs (as was also observed for the modification enzyme TruB [81,89]), which creates potential challenges for tRNA chaperones to bind and refold all defective substrates.

2.4. tRNA Core Modifications and Their Impact on tRNA Degradation and Cellular Stability

As described above, the absence of certain modifications may alter the folding and tertiary structure of tRNAs. The recognition and elimination of these hypomodified tRNAs with compromised tertiary structures are carried out by tRNA quality control pathways. They prevent hypomodified tRNAs from massively entering the translation process. The characterization of the pathways targeting the degradation of tRNAs lacking specific modifications is much more advanced in eukaryotes, but similar mechanisms seem to also exist in bacteria (see below).
In eukaryotes, two main pathways for tRNA degradation have been identified: the nuclear surveillance pathway, which primarily degrades defective tRNA precursors (pre-tRNAs) in the nucleus, and the Rapid tRNA Decay (RTD) pathway, which degrades hypomodified mature tRNAs in the cytoplasm (reviewed in [1]). Nuclear surveillance and RTD degradation pathways have been extensively studied in the yeasts S. cerevisiae and S. pombe, where several hypomodified tRNAs have been correlated with phenotypes of thermosensitivity and growth defects (see Table 1 for an overview of the tRNAs targeted by degradation pathways and the tRNA core modifications involved).
Early reports showed that yeast pre- tRNA i Met lacking m1A58 is degraded in the nucleus by the nuclear surveillance pathway [19,90]. However, a more recent study revealed that a fraction of this hypomodified initiator tRNA continues the maturation process and is exported to the cytoplasm, where it is given a second chance to be degraded by the RTD pathway (Table 1) [91]. The latter study also showed that S. pombe tRNA i Met lacking m1A58 is only degraded by the RTD pathway and not by the nuclear surveillance pathway (Table 1) [91]. As mentioned earlier, m1A58 has a drastic impact on the tertiary structure of yeast tRNA i Met , so that the tRNA elbow structure is only properly assembled when this modification is present [70]. This provides a structural explanation of why hypomodified tRNA i Met lacking m1A58 is targeted by RNA decay pathways. This is also in line with a genetic screen, which reported that degradation by the RTD pathway in yeast primarily depends on the structural integrity and stability of the acceptor-stem and T-arm region of tRNAs [92].
Recently, a study reported a new tRNA quality control mechanism in S. cerevisiae, in which mature tRNA Tyr lacking m1G9 is eliminated by a Met22-dependent but RTD- and nuclear surveillance-independent degradation pathway (Table 1) [93]. Further investigation is definitely needed to identify the nucleases responsible for this degradation and the molecular mechanisms of this intriguing tRNA quality control pathway. Interestingly, m1G9 prevents the degradation of tRNA Tyr , whereas its absence in tRNA Gly has no effect on its cellular stability [93], adding to the list of known examples of tRNA core modifications that have different effects depending on the tRNA species and are only required for the cellular stability of a particular tRNA [72].
In bacteria, tRNA surveillance and quality control systems monitoring hypomodified and/or thermodynamically destabilized tRNAs have not been studied extensively. These aspects were initially studied in thermophilic bacteria, where tRNA thermal and structural stability are required for optimal cellular thermotolerance and survival at elevated growth temperatures [94]. The tRNA core modifications and their dynamic regulation represent an important player in the structural and thermal adaptation of tRNAs (Figure 3a), and their absence may lead to the targeting of the hypomodified tRNA to quality control pathways. For instance, the lack of m7G46 in T. thermophilus tRNAs leads to the degradation of tRNA Phe and tRNA Ile by a yet unidentified pathway (Table 1) [95]. Another important report revealed a bacterial tRNA quality control system in Vibrio cholerae where hypomodified tRNAs lacking s4U8 are rapidly eliminated by the RNA degradosome (Table 1) [20]. The similarities observed with the well-characterized mechanisms present in eukaryotes definitely call for further studies in other bacterial species, in order to refine our understanding of the role of tRNA core modifications in the cellular stability of tRNAs in bacteria.
Table 1. List of eukaryotic and bacterial hypomodified tRNAs that are degraded by tRNA quality control pathways.
Table 1. List of eukaryotic and bacterial hypomodified tRNAs that are degraded by tRNA quality control pathways.
OrganismtRNALacking ModificationsConditionsDegradation PathwayReferences
S. cerevisiaepre- tRNA i Met m 1 A5836 °CNuclear surveillance[19,90]
S. cerevisiae tRNA i Met m 1 A5827 °C, 34 °CRTD[91]
S. pombe tRNA i Met m 1 A5830 °C, 38.5 °CRTD[91]
S. cerevisiae tRNA Tyr m 1 G930 °C, late log phase, and w/ or w/o 5FUMet22-dependent but Xrn1, Rat1, Dxo1, Trf4 and Rrp6-independent degradation[93]
S. cerevisiae tRNA Val ( AAC ) m 7 G46 and ( m 5 C or Ψ 13 or D47) *37 °CRTD[18,96]
S. cerevisiae tRNA Ser ( CGA , UGA ) Um44 and ac 4 C12 *36.5 °CRTD[96,97]
S. cerevisiae tRNA Ser ( CGA , UGA ) m 2 2 G26 and m 5 C *37 °CRTD[98]
S. cerevisiae tRNA Val ( AAC ) m 7 G4637 °C after thiolutin treatmentRTD[98]
S. cerevisiae tRNA Ser ( CGA , UGA ) ac 4 C1237 °C after thiolutin treatmentRTD[98]
S. cerevisiae tRNA Ser ( CGA , UGA ) m 2 2 G2637 °CRTD[98]
S. pombe tRNA Tyr ( GUA ) and tRNA Pro ( AGG ) m 7 G4636.5 °C, 37.5 °C, 38.5 °CRTD[99]
T. thermophilus tRNA Phe and tRNA Ile m 7 G4670–80 °C-[95]
V. cholerae tRNA Tyr , tRNA 1 Ser , tRNA 1 Cys , tRNA 2 Cys and others s 4 U8stationary phase at 37 °CRNA degradosome[20]
V. cholerae tRNA 1 A Gln s 4 U8log phase at 37 °CRNA degradosome[20]
V. cholerae tRNA Tyr s 4 U8 and ( Ψ 55 or m 5 U54) *log phase at 37 °C and w/o arabinoseRNA degradosome[20]
* Both listed modifications should be absent for the tRNA to be degraded.

2.5. Interdependence Involving tRNA Core Modifications

Although some modifications are incorporated into tRNAs independently, modification circuits have been identified in which the prior presence of one modification influences the occurrence and incorporation of subsequent modifications [3,100,101,102]. Most of the well-characterized modification circuits concern modifications in the ACL region. Few examples involving tRNA core modifications have been reported, both in bacteria and yeast (Figure 3). The mechanisms underlying the altered efficiency of tRNA modification incorporation by initial modifications are not yet well understood, but it is reasonable to propose that alterations in the local structure and/or dynamics of the tRNA would affect the catalytic efficiency of the subsequent enzymes. These alterations could occur at all steps of the catalytic cycle of the modification enzymes, from substrate binding to substrate release.
The first tRNA core modification circuits identified were those found in T. thermophilus tRNA Phe (Figure 3a). They regulate modification levels overall in response to temperature changes (reviewed in [94]). Briefly, at elevated temperatures, the prior presence of m 7 G46 promotes the introduction of Gm18 and m1A58, which subsequently promotes the incorporation of m5s2U54 [95,103], with m1A58 and m5s2U54 being essential for the growth of several thermophiles at high temperatures [104,105,106]. In addition, m7G46 promotes the incorporation of m1G37 in the ACL [95]. However, at low temperatures, the prior presence of Ψ 55 inhibits the introduction of Gm18, m1A58 and m5s2U54, maintaining sufficient tRNA structural flexibility to survive at lower temperatures [107]. In addition, at low temperatures, m5U54 stimulates the introduction of m1A58 [108].
Figure 3. Modification circuits involving tRNA core modifications: (a) modification circuits identified in T. thermophilus tRNA Phe in response to temperature changes represented on a cloverleaf scheme [94,95,107]; (b) T-arm modification circuits identified in yeast tRNA Phe [69,70]; (c,d) modification circuits identified in E. coli tRNA Phe [109,110,111]; and (e) modification circuits identified in E. coli tRNA Asp [111]. Modifications are represented by colored circles (see the legend in the box for the corresponding colors). Arrows indicate stimulatory effects between modifications, and blunted lines indicate inhibitory effects between modifications.
Figure 3. Modification circuits involving tRNA core modifications: (a) modification circuits identified in T. thermophilus tRNA Phe in response to temperature changes represented on a cloverleaf scheme [94,95,107]; (b) T-arm modification circuits identified in yeast tRNA Phe [69,70]; (c,d) modification circuits identified in E. coli tRNA Phe [109,110,111]; and (e) modification circuits identified in E. coli tRNA Asp [111]. Modifications are represented by colored circles (see the legend in the box for the corresponding colors). Arrows indicate stimulatory effects between modifications, and blunted lines indicate inhibitory effects between modifications.
Genes 15 00374 g003
In the yeast S. cerevisiae, the time-resolved monitoring with NMR spectroscopy of the maturation of a tRNA Phe transcript in yeast cell extracts revealed the existence of dependencies between tRNA core modifications [69]. In particular, a strong modification circuit is present in the T-arm, in which the prior presence of Ψ 55 activates the formation of m5U54 by Trm2 and m1A58 by Trm6/Trm61, and m5U54 promotes the formation of m1A58 (Figure 3b). These cross-talks were shown to be also present in other yeast tRNAs, although they were not specifically identified [69]. A recent study conducted nanopore sequencing of all yeast tRNAs in specific deletion strains [112]. This study confirmed the modification circuit in the T-arm of tRNA Phe and further showed that it is also present in at least 15 other tRNA species [112]. Further biochemical studies on this Ψ 55 → m5U54 → m1A58 modification circuit, have characterized and quantified the direct effect of the initial modifications on the enzymatic activity of the subsequent ones [70]. The Ψ 55 and m5U54 modifications have a strong cumulative effect on the introduction of m1A58, together accelerating this process by a factor of ∼15 (discussed in [113]).
Recently, several reports have shed light on modification cross-talks among tRNA core modifications in E. coli. They reported that the introduction of acp 3 U47 is globally enhanced by the prior presence of m7G46 [109] and also stimulated by m5U54 in E. coli tRNA Arg , tRNA Ile ( GAU ) , tRNA Lys , tRNA Phe and tRNA Val [110,111] (Figure 3c). Ψ 55 seems to also participate in this network, in which it may either slightly stimulate or inhibit the introduction of acp 3 U47 in E. coli tRNA Phe and tRNA Lys , respectively [111]. In addition, the formation of ms2 i 6 A37 in the ACL of E. coli tRNA Phe is activated by m5U54 [110,111] (Figure 3d), and the introduction of s4U8 is enhanced in the presence of Ψ 55 and m5U54 in E. coli tRNA Asp [111] (Figure 3e). These interdependencies in E. coli tRNA core modifications were identified using a combination of nucleotide quantification and sequencing techniques, such as HPLC, liquid-chromatography coupled to tandem mass spectrometry (LC-MS/MS), multiplex small RNA-sequencing (MSR-seq) and primer extension analysis. Finally, an order for the introduction of modifications into E. coli tRNA Phe was suggested where Ψ 55 and m5U54 are introduced in the earliest stages of maturation, followed by m7G46 and then acp 3 U47, with s4U8 being introduced after Ψ 55 and m5U54 [114].
Apart from the few examples described above, the characterization of these interdependencies is still relatively limited, but the latest technological developments, in particular in nanopore sequencing technology, are expected to accelerate the discovery of modification circuits involving tRNA core modifications [115].

3. tRNA Core Modifications and Their Impact on Translation and Stress Adaptation

3.1. tRNA Modification Levels Depend on Growth Conditions

Although tRNA modifications may appear to be constitutively introduced as static decoration on every tRNA of the cell, the level and/or nature of certain modifications in a tRNA population may change in response to environmental factors such as growth phase, temperature, cellular stresses or nutrient availability. These variations consist of the appearance, disappearance or change in the nature of the modification at the individual tRNA molecule level. Collectively, these variations have an impact on the level of modifications in tRNA populations [3].
During the exponential growth phase, most microorganisms double their content on an hourly basis. In E. coli, the tRNA pool does not increase homogeneously during exponential growth. Instead, rare tRNAs tend to decrease in concentration while the concentration of other tRNAs remains apparently constant [116,117,118]. In S. cerevisiae, a similar phenomenon exists where tRNA proportions vary significantly upon stress, which may be part of stress response mechanisms [119]. These changes in tRNA proportions are linked to changes in tRNA modification status. Some abundant tRNAs are hypomodified during the exponential phase and complete their maturation during the stationary phase [116,120]. In E. coli, abundant tRNAs tend to be less fully thiolated than tRNAs with rare codons [116]. In B. subtilis, more complete modifications are found in tRNAs from stationary cells in comparison with tRNAs of exponentially growing cells [120].
In addition to the growth phase, growth conditions such as stress exposure, have been shown to affect tRNA modification levels in microorganisms. A pioneering study used HPLC-coupled mass spectrometry to quantitatively analyze modifications in tRNAs. The study showed that exposure to toxic substances such as methylmethane sulfonate (MMS), hydrogen peroxide ( H 2 O 2 ), sodium arsenite (NaAs O 2 ) and sodium hypochlorite (NaOCl) in S. cerevisiae affects the level of several modifications, many of which being tRNA core modifications [121]. Upon treatment with H 2 O 2 , levels of m5C and m 2 2 G were increased, whereas levels of m5U, m2G and m1A were slightly decreased [121]. Various toxicants induced similar yet distinct changes in tRNA modifications, indicating that dynamic changes in tRNA modifications in response to cellular stresses are widespread and stress specific [121,122]. Although the roles of these changes are still mostly unknown, some changes occurring at the wobble position have been interpreted in terms of modification-tunable expression of codon-biased genes (reviewed in [38,39,40,41]). The possible role of the changes affecting tRNA core modifications upon stress exposure remains much less well characterized in comparison.
Several other studies have reported alterations in the levels of tRNA core modifications upon exposure to various stresses or growth conditions. This includes, for instance, the Gm18 modification that is increased in E. coli upon mild antibiotic stress [123]. In S. cerevisiae, additional m5C modifications are specifically introduced on tRNA His at positions 48 and 50 in response to conditions characterized by global cell growth arrest [124]. In T. thermophilus, changes in the growth temperature were shown to affect levels of tRNA core modifications (see text above and references [94,106]). In E. coli, heat stress and/or heat shock were also found to either increase or decrease the levels of several tRNA core modifications, e.g., s4U8/9, Gm18, m7G46 and acp 3 U47 [125]. These examples show that it is important to bear in mind that tRNA core modifications also undergo changes in response to stressful conditions and that this phenomenon is not limited to the modifications found at the wobble position.

3.2. tRNA Core Modifications and Their Impact on Translation and Fitness

The binding of tRNAs to the ribosome and their accommodation into the A-site constitute the first steps of translation elongation. The reaction only becomes irreversible when fitting tRNAs enter the A-site [126], making accommodation a limiting step for translation elongation [126,127]. Inside the A-site, tRNAs interact with the mRNA codon as well as the small ribosomal subunit via their anticodon-arm, leading to the well-characterized proofreading mechanism, and with the large ribosomal subunit mainly via their T-arm and elbow region [11,128,129,130]. After the accommodation of the incoming tRNA into the A-site and the catalysis of the peptide transfer from the P- to the A-site tRNA, the peptidyl A-site tRNA is translocated to the P-site. During the translocation process, tRNAs also interact with the ribosome through their T-arm and elbow regions [11,129,131]. Modifications in the tRNA core that affect the structure and dynamics of tRNAs may therefore affect the interaction with the ribosome and modulate the efficiency of translation elongation by affecting accommodation and/or translocation.
Several recent studies have shed new light on the mechanisms by which modifications in the tRNA core impact fitness and translation [110,111,132]. Effects of certain tRNA core modifications on fitness have been known for many years. For instance, in E. coli, the trmA and truB genes, responsible for the introduction of m5U54 and Ψ 55 in the T-loop, are non-essential under most growth conditions, but subtle growth phenotypes have been reported for the trmA Δ and truB Δ strains [57,133]. In addition, these genes are important for bacterial fitness in co-culture experiments [134,135]. The fitness phenotypes in E. coli are attributed to the chaperone activity of TruB [81], while both the chaperone and modification activities of TrmA are implicated [82]. The effect of m5U54 on translation was initially investigated in vitro with E. coli tRNAs, which pointed towards a role for m5U54 in both translation speed and fidelity [136]. The binding to the A-site of a tRNA Lys ( UUU ) lacking m5U54 was reduced, and the resulting translation speed increased by ∼10-fold, while translation with tRNA Phe ( GAA ) displayed about a ∼10-fold loss of fidelity [136]. In yeast, there is a slight global increase in the translation efficiency in trm2 Δ strains lacking m5U54 [137]. It is also worth mentioning that in humans, the m5U54 modification appears to be involved in translation fidelity [138].
In a recent report that conducted in vitro kinetic studies of translation with E. coli tRNAs prepared from trmA Δ cells, the m5U54 modification was shown to play a role in the translocation process [110]. The use of hygromycin B—an antibiotic known to block the translocation of ribosomes—led to the proposal that m5U54 affects the translocation step and could reduce the translation speed by restraining the translocation of tRNAs [110]. In another report that determined codon-specific changes in translation efficiency in the trmA Δ and truB Δ single- and double-knockout strains, a set of specific codons affecting the expression of a reporter gene was identified [111]. For instance, in the truB Δ strain, a decrease in translation is detected when some arginine codons, i.e., CGU, CGG and AGG, are placed consecutively. Similarly, in the trmA Δ strain, translation is decreased at consecutive cysteine TGT codons but increased at consecutive tyrosine TAT codons [111]. Although the overall translation is unaffected, the codon-specific impact related to the loss of m5U54 and/or Ψ 55 in the trmA Δ and truB Δ strains leads to changes in the abundance of specific proteins, many of which are implicated in metabolism and gene expression [111]. This decreased synthesis of proteins enriched in the identified codons might explain the fitness phenotype of the trmA Δ and truB Δ strains. In these two studies focused on the impact of tRNA core modifications on translation [110,111], the reported changes at the level of the translocation step, or related to codon-specific changes in the translation efficiency, could be attributed to subtle differences in the local dynamics of the tRNA elbow region that could help tune the rates of different translation steps, e.g., accommodation and translocation, to ensure a good balance between the speed and fidelity of translation [139].
These observations are perfectly in line with another recent report that investigated in a more general and broader context the effect of tRNA core modifications on translation in S. cerevisiae [132]. This report is based on the initial hypothesis that tRNA core modifications may have an impact on the intrinsic decoding efficiency of a tRNA by altering its structure and/or flexibility, which could potentially influence the dynamics of its interaction with the ribosome during accommodation and/or translocation. Mistranslation reporters have been used to directly measure amino acid incorporation errors in vivo, to determine whether the loss of specific tRNA core modifications may affect translational misreading errors [132]. The loss of individual modifications globally led to an increase in the misreading frequency. Each modification affected the misreading error of the selected error-prone codons in an intricate manner, and certain combinations did not give rise to significant changes. Strikingly, however, the absence of almost any of the core modifications, altered the accuracy of the tested tRNAs, i.e., tRNA Lys and tRNA Glu , on at least one of the error-prone codons tested [132].
These recent findings on the impact of tRNA core modifications on the fine tuning of the translation process [110,111,132], which appear to be modification-, tRNA- and codon-specific, support the idea that the structure of the tRNA core, including its modifications, have coevolved with the anticodon to tune tRNAs with strong or weak anticodons to achieve a consistent efficiency of decoding [140].

3.3. tRNA Core Modifications and Their Relation with Stress Adaptation Mechanisms

3.3.1. s4U8 Is Implicated in the Response to UV Stress

The s 4 U8 modification is omnipresent in prokaryotic tRNAs [14] and has been reported early on to carry a protective role against UV radiation (Table 2) (reviewed in [141]). This modification is sensitive to UVs [142], and when cytosine is present at position 13, an intramolecular s 4 U8:C13 cyclobutane pyrimidine dimer is formed between the aromatic bases that are stacked on each other in tRNA 3D structures [142]. This UV-triggered cross-linking is a widespread phenomenon, since in E. coli for instance, about two-thirds of the tRNAs harbor the s 4 U8 modification, with half of them also carrying cytosine at position 13. Thus, UV radiation likely targets about one-third of the E. coli tRNAs, including the entire pool of several isoacceptor tRNAs. Some of these photo-crosslinked tRNAs exhibit a reduced ability to be charged by their aminoacyl-tRNA synthetase [143,144]. The accumulation of uncharged tRNAs—which may be directed to the A-site of the ribosome—causes protein synthesis to stall and leads to growth delay in E. coli and Salmonella typhimurium [145,146,147]. In addition, ribosome stalling gives rise to the production of ppGpp, which triggers the stringent response and a massive decrease in stable RNA production. This seemingly allows for a faster recovery of the UV-exposed cells upon antibiotic treatment [148]. Interestingly, E. coli strains selected for their abnormal sensitivity to UVs, were found to carry mutations in the gene responsible for s 4 U8 formation [149,150] and were unable to produce ppGpp after irradiation [147]. These mechanisms—initiated by a photo-cross-linking reaction at the s 4 U8 modification site—eventually ensure better survival after UV stress [147,150].

3.3.2. Gm18 Regulates the Immune-Stimulatory Effect of tRNAs

The Gm18 modification is present in both bacteria and eukaryotes [14]. In E. coli, this modification was recently shown to be involved in a stress adaptation mechanism to antibiotic treatment in relation to the sensing of bacteria by the immune system of the human host (Table 2) [123]. Bacterial tRNAs containing Gm18 exhibit a reduced immunostimulatory effect in humans [151]. Whereas unmodified tRNAs trigger the activation of the human Toll-like receptor 7 (TRL7) [152], Gm18-modified tRNAs act as TRL7 antagonists and inhibit the immune response [153,154]. In E. coli, the gene encoding the G18 methyltransferase, i.e., trmH, is part of an operon activated during the stringent response, and levels of Gm18 are thus expected to increase in certain stress conditions, for instance antibiotic treatment [155]. An increase in the Gm18 level in some tRNAs was indeed observed upon mild antibiotic treatment in E. coli [123]. In addition, and in line with previous reports, the stress-induced increase in tRNA Gm18 levels led to a reduced immune-stimulatory effect of the bacterial tRNAs [123]. Thus, Gm18 is a tRNA core modification that responds to stress with implications outside protein translation and appears as a key modification to escape the immune system of the host.

3.3.3. m7G46 Is Implicated in the Response to Oxidative Stress

The m 7 G46 modification is present in both bacteria and eukaryotes [14]. The genes responsible for introducing m 7 G46 in tRNAs, i.e., trmB in bacteria and trm8/trm82 in yeast, are not essential for cell viability in optimal growth conditions [95,156,157]. However, they are important in certain growth conditions or together with the deletion of other tRNA modification genes (Table 1 and Table 2) [18,95,96,98]. In Pseudomonas aeruginosa, the m 7 G46 modification has been shown to be involved in an adaptation mechanism to H 2 O 2 oxidative stress (Table 2) [158]. In a trmB Δ strain, translation of phenylalanine and aspartate codons by tRNA Phe ( GAA ) and tRNA Asp ( GUC ) are less efficient, which has a negative effect on the translation of mRNAs enriched in Phe and Asp. It is worth mentioning that, in humans, the m 7 G46 modification seems important as well for the efficient translation of certain codons [159]. In P. aeruginosa, the Phe and Asp codons are particularly enriched and grouped in clusters in the mRNAs of the catalases KatA and KatB, two reductases involved in H 2 O 2 detoxification. Interestingly, the transcription of trmB increases following H 2 O 2 treatment, which results in an increase in the m 7 G46 level in tRNAs and, subsequently, in an increase in KatA and KatB production [158]. This complete regulatory loop is one of the best characterized examples of a tRNA core modification involved in stress response, although some of the underlying mechanisms are still not fully understood. As expected, the trmB Δ mutant strain presents an H 2 O 2 -sensitive phenotype, and the survival rate of trmB Δ cells exposed to 20 mM H 2 O 2 was found to be only 10% of that of the wild-type strain [158]. Interestingly, similar H 2 O 2 -sensitive phenotypes have been reported for trmB or trm8 mutants in E. coli [160], in Acinetobacter baumannii [161] and in the fungus Colletotrichum lagenarium [162], demonstrating the widespread implication of the m 7 G46 tRNA core modification in the response to oxidative stress.
Table 2. List of tRNA core modifications linked to stress.
Table 2. List of tRNA core modifications linked to stress.
OrganismModificationType of StressReferences
S. cerevisiaeAm4MMS, NaOCl[121]
E. coli, S. typhimurium s 4 U8UV[147,149,150]
E. coli s 4 U8/ s 4 U9heat[125]
S. cerevisiae m 2 G10 H 2 O 2 , NaOCl, NaAs O 2 [121]
T. thermophilus, E. coliGm18cold/heat[95,107,125,163]
E. coli, V. choleraeGm18antibiotics[123,155,164]
S. cerevisiaeGm18NaOCl[121]
S. cerevisiae m 2 2 G26 H 2 O 2 , MMS, NaAs O 2 [121,122]
S. cerevisiaeUm44 H 2 O 2 , NaAs O 2 , NaOCl[121,165]
C. lagenarium, P. aeruginosa, A. baumannii, E. coli m 7 G46 H 2 O 2 [158,160,161,162]
T. thermophilus, E. coli m 7 G46heat[95,125]
V. cholerae m 7 G46antibiotics[164]
S. cerevisiae m 7 G46MMS, NaAs O 2 [121,165]
E. coli acp 3 U47heat[125]
S. cerevisiae m 5 C48/ m 5 C50amino acid, glucose and uracil starvation[124]
T. thermophilus m 5 U54cold[107,108]
P. furiosus, T. thermophilus m 5 s 2 U54heat[95,104,107,166]
T. thermophilus Ψ 55cold/heat[107]
V. cholerae Ψ 55antibiotics[164]
T. thermophilus m 1 A58heat[95,105,107]
S. cerevisiae m 1 A58 H 2 O 2 [121]
V. cholerae, E. coliDantibiotics[164]
S. cerevisiae m 5 C H 2 O 2 , MMS, NaOCl, NaAs O 2 [121]
S. cerevisiaeCm H 2 O 2 , MMS, NaAs O 2 [121]

3.3.4. tRNA Core Modifications in the Response to Antibiotic Stress

Although tRNA modification genes may be directly involved in bacterial antibiotic resistance—e.g., trmD introducing m 1 G37 in the anticodon loop (reviewed in [167])—, tRNA core modifications have not been shown to confer resistance to antibiotics. However, a recent pioneering study in Vibrio cholerae has uncovered a link between several tRNA modification genes, including those targeting the tRNA core, and the bacterial response to low doses of antibiotics (Table 2) [164]. In this report, initial transposon mutagenesis followed by deep sequencing (TN-seq) [168] in the presence of low-dose antibiotics (sub-minimal inhibitory concentration; sub-MIC) identified several rRNA and tRNA modification genes as important for growth under these antibiotic stress conditions [164]. Several antibiotics were tested, and tRNA modification genes were found to be more largely involved in the response to antibiotics targeting the ribosome, e.g., tobramycin belonging to the aminoglycoside family, than in the response to antibiotics targeting DNA, e.g., ciprofloxacin. These initial genome-wide screens were subsequently complemented with growth competition experiments between the identified tRNA modification deletion strains and the wild-type strain, to determine the effects on the fitness of several antibiotics [164]. Concerning modifications in the tRNA core, under sub-MIC aminoglycoside treatment, deletion of truB decreased fitness, while deletion of dusB and trmH conferred a growth advantage. For other antibiotics, such as ciprofloxacin, trimethoprim and carbenicillin, deletion of dusB, truB, trmB and trmH were either decreasing or slightly increasing fitness, depending on the antibiotic considered [164]. These diverse phenotypes for a given gene suggest that the observed effects are not general but antibiotic-specific. Most importantly, this study showed that the effect on fitness in the presence of sub-MIC antibiotics in tRNA modification deletion strains also impacts tolerance to lethal doses of the antibiotics. Although the molecular mechanisms underlying the effects of the queuosine Q34 modification in the antibiotic translational stress response have subsequently been described in detail [169], the mechanisms by which tRNA core modification affects antibiotic tolerance are still elusive, and their study may certainly shed new light on the subtle functions of this group of modifications.

4. Conclusions and Perspectives

In this review, we focused on tRNA core modifications, starting from their impact on the structure and dynamics of tRNAs and, subsequently, exploring their reported functions in translation and fitness and their implications in stress adaptation mechanisms. Although the phenotypes associated with particular modifications, or a lack thereof, must ultimately be caused by changes in the structural and/or dynamic properties of the tRNA, thereby altering its ability to interact with cellular partners, what emerges from this review is that the mechanisms by which tRNA core modifications affect fitness, translation and response to stress are not yet fully understood. Although it was not the topic of this review, it is worth mentioning that tRNAs are also implicated in several other stress sensing and response mechanisms, which are not related to post-transcriptional modifications, e.g., the stringent response in bacteria or the nuclear accumulation of tRNAs upon amino acid starvation in yeast [170,171]. In addition, mismethionylation of non-methionine tRNAs confers a protective role against oxidative stress [172,173]. We have also not covered the aspects related to tRNA cleavage upon stress, which is known to be affected by modifications in the ACL region (reviewed in [174]). An important role in stress response mechanisms is therefore occupied by tRNAs, and post-transcriptional modifications are closely involved. This is not surprising given that tRNA modifications are at a crossroads connecting translation with metabolism through their biosynthetic pathways [29]. For instance, the cofactors of tRNA methyltransferases, i.e., S-adenosyl-l-methionine (SAM) and methylenetetrahydrofolate (5,10-methylene-THF), occupy key positions in the one-carbon metabolism, and these metabolites thus affect many aspects of translation, in part through their connections with tRNAs [175]. In recent years, major technological advancements have been made in the fields of mass spectrometry, RNA sequencing, and kinetic studies of translation. These methods have provided, and will continue to provide, important insights into the mechanisms by which tRNA modifications shape gene expression and stress response. In these future studies, although it may be tempting to follow the threads involving anticodon modifications on the pretext that the underlying mechanisms are now better understood, we believe it would be wise not to set aside the threads involving tRNA core modifications.

Author Contributions

Conceptualization, M.-J.Y., A.M. and P.B.; Writing—Original Draft Preparation, M.-J.Y., A.M. and P.B.; Writing—Review and Editing, M.-J.Y., A.M. and P.B.; Visualization, M.-J.Y., A.M. and P.B.; Supervision, P.B.; Funding Acquisition, P.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the ANR JCJC CiMoDyMo (ANR-19-CE44-0013) and the Labex DYNAMO (ANR-11-LABX-0011). M.-J.Y. was supported by a Q-Life PhD fellowship (Qlife ANR-17-CONV-0005).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Phizicky, E.M.; Hopper, A.K. The life and times of a tRNA. RNA 2023, 29, 898–957. [Google Scholar] [CrossRef]
  2. Berg, M.D.; Brandl, C.J. Transfer RNAs: Diversity in form and function. RNA Biol. 2021, 18, 316–339. [Google Scholar] [CrossRef]
  3. Barraud, P.; Tisné, C. To be or not to be modified: Miscellaneous aspects influencing nucleotide modifications in tRNAs. IUBMB Life 2019, 71, 1126–1140. [Google Scholar] [CrossRef]
  4. Hopper, A.K. Transfer RNA post-transcriptional processing, turnover, and subcellular dynamics in the yeast Saccharomyces cerevisiae. Genetics 2013, 194, 43–67. [Google Scholar] [CrossRef]
  5. Lin, B.Y.; Chan, P.P.; Lowe, T.M. tRNAviz: Explore and visualize tRNA sequence features. Nucleic Acids Res. 2019, 47, W542–W547. [Google Scholar] [CrossRef]
  6. Chan, P.P.; Lowe, T.M. GtRNAdb 2.0: An expanded database of transfer RNA genes identified in complete and draft genomes. Nucleic Acids Res. 2016, 44, D184–D189. [Google Scholar] [CrossRef] [PubMed]
  7. Robertus, J.D.; Ladner, J.E.; Finch, J.T.; Rhodes, D.; Brown, R.S.; Clark, B.F.; Klug, A. Structure of yeast phenylalanine tRNA at 3 A resolution. Nature 1974, 250, 546–551. [Google Scholar] [CrossRef] [PubMed]
  8. Kim, S.H.; Suddath, F.L.; Quigley, G.J.; McPherson, A.; Sussman, J.L.; Wang, A.H.; Seeman, N.C.; Rich, A. Three-dimensional tertiary structure of yeast phenylalanine transfer RNA. Science 1974, 185, 435–440. [Google Scholar] [CrossRef]
  9. Giegé, R.; Jühling, F.; Pütz, J.; Stadler, P.; Sauter, C.; Florentz, C. Structure of transfer RNAs: Similarity and variability. Wiley Interdiscip. Rev. RNA 2012, 3, 37–61. [Google Scholar] [CrossRef] [PubMed]
  10. Zhang, J.; Ferré-D’Amaré, A.R. The tRNA Elbow in Structure, Recognition and Evolution. Life 2016, 6, 3. [Google Scholar] [CrossRef] [PubMed]
  11. Biela, A.; Hammermeister, A.; Kaczmarczyk, I.; Walczak, M.; Koziej, L.; Lin, T.Y.; Glatt, S. The diverse structural modes of tRNA binding and recognition. J. Biol. Chem. 2023, 299, 104966. [Google Scholar] [CrossRef] [PubMed]
  12. Jackman, J.E.; Alfonzo, J.D. Transfer RNA modifications: Nature’s combinatorial chemistry playground. Wiley Interdiscip. Rev. RNA 2013, 4, 35–48. [Google Scholar] [CrossRef] [PubMed]
  13. Ontiveros, R.J.; Stoute, J.; Liu, K.F. The chemical diversity of RNA modifications. Biochem. J. 2019, 476, 1227–1245. [Google Scholar] [CrossRef] [PubMed]
  14. Cappannini, A.; Ray, A.; Purta, E.; Mukherjee, S.; Boccaletto, P.; Moafinejad, S.N.; Lechner, A.; Barchet, C.; Klaholz, B.P.; Stefaniak, F.; et al. MODOMICS: A database of RNA modifications and related information. 2023 update. Nucleic Acids Res. 2024, 52, D239–D244. [Google Scholar] [CrossRef] [PubMed]
  15. Machnicka, M.A.; Olchowik, A.; Grosjean, H.; Bujnicki, J.M. Distribution and frequencies of post-transcriptional modifications in tRNAs. RNA Biol. 2014, 11, 1619–1629. [Google Scholar] [CrossRef] [PubMed]
  16. Motorin, Y.; Helm, M. tRNA stabilization by modified nucleotides. Biochemistry 2010, 49, 4934–4944. [Google Scholar] [CrossRef]
  17. Lorenz, C.; Lünse, C.E.; Mörl, M. tRNA Modifications: Impact on Structure and Thermal Adaptation. Biomolecules 2017, 7, 35. [Google Scholar] [CrossRef]
  18. Alexandrov, A.; Chernyakov, I.; Gu, W.; Hiley, S.L.; Hughes, T.R.; Grayhack, E.J.; Phizicky, E.M. Rapid tRNA decay can result from lack of nonessential modifications. Mol. Cell 2006, 21, 87–96. [Google Scholar] [CrossRef]
  19. Kadaba, S.; Krueger, A.; Trice, T.; Krecic, A.M.; Hinnebusch, A.G.; Anderson, J. Nuclear surveillance and degradation of hypomodified initiator tRNAMet in S. cerevisiae. Genes Dev. 2004, 18, 1227–1240. [Google Scholar] [CrossRef]
  20. Kimura, S.; Waldor, M.K. The RNA degradosome promotes tRNA quality control through clearance of hypomodified tRNA. Proc. Natl. Acad. Sci. USA 2019, 116, 1394–1403. [Google Scholar] [CrossRef]
  21. Wichtowska, D.; Turowski, T.W.; Boguta, M. An interplay between transcription, processing, and degradation determines tRNA levels in yeast. Wiley Interdiscip. Rev. RNA 2013, 4, 709–722. [Google Scholar] [CrossRef]
  22. Sprinzl, M.; Horn, C.; Brown, M.; Ioudovitch, A.; Steinberg, S. Compilation of tRNA sequences and sequences of tRNA genes. Nucleic Acids Res. 1998, 26, 148–153. [Google Scholar] [CrossRef]
  23. Björk, G.R.; Hagervall, T.G. Transfer RNA Modification: Presence, Synthesis, and Function. EcoSal Plus 2014, 6, 10-1128. [Google Scholar] [CrossRef]
  24. de Crécy-Lagard, V.; Jaroch, M. Functions of Bacterial tRNA Modifications: From Ubiquity to Diversity. Trends Microbiol. 2021, 29, 41–53. [Google Scholar] [CrossRef]
  25. Smith, T.J.; Giles, R.N.; Koutmou, K.S. Anticodon stem-loop tRNA modifications influence codon decoding and frame maintenance during translation. Semin. Cell Dev. Biol. 2024, 154, 105–113. [Google Scholar] [CrossRef] [PubMed]
  26. Suzuki, T. The expanding world of tRNA modifications and their disease relevance. Nat. Rev. Mol. Cell Biol. 2021, 22, 375–392. [Google Scholar] [CrossRef] [PubMed]
  27. Agris, P.F.; Narendran, A.; Sarachan, K.; Väre, V.Y.P.; Eruysal, E. The Importance of Being Modified: The Role of RNA Modifications in Translational Fidelity. Enzymes 2017, 41, 1–50. [Google Scholar] [CrossRef] [PubMed]
  28. Torres, A.G.; Batlle, E.; Ribas de Pouplana, L. Role of tRNA modifications in human diseases. Trends Mol. Med. 2014, 20, 306–314. [Google Scholar] [CrossRef] [PubMed]
  29. El Yacoubi, B.; Bailly, M.; de Crécy-Lagard, V. Biosynthesis and function of posttranscriptional modifications of transfer RNAs. Annu. Rev. Genet. 2012, 46, 69–95. [Google Scholar] [CrossRef] [PubMed]
  30. Nedialkova, D.D.; Leidel, S.A. Optimization of Codon Translation Rates via tRNA Modifications Maintains Proteome Integrity. Cell 2015, 161, 1606–1618. [Google Scholar] [CrossRef]
  31. Urbonavicius, J.; Qian, Q.; Durand, J.M.; Hagervall, T.G.; Björk, G.R. Improvement of reading frame maintenance is a common function for several tRNA modifications. EMBO J. 2001, 20, 4863–4873. [Google Scholar] [CrossRef] [PubMed]
  32. Urbonavicius, J.; Stahl, G.; Durand, J.M.B.; Ben Salem, S.N.; Qian, Q.; Farabaugh, P.J.; Björk, G.R. Transfer RNA modifications that alter +1 frameshifting in general fail to affect -1 frameshifting. RNA 2003, 9, 760–768. [Google Scholar] [CrossRef] [PubMed]
  33. Ranjan, N.; Rodnina, M.V. Thio-Modification of tRNA at the Wobble Position as Regulator of the Kinetics of Decoding and Translocation on the Ribosome. J. Am. Chem. Soc. 2017, 139, 5857–5864. [Google Scholar] [CrossRef] [PubMed]
  34. Tükenmez, H.; Xu, H.; Esberg, A.; Byström, A.S. The role of wobble uridine modifications in +1 translational frameshifting in eukaryotes. Nucleic Acids Res. 2015, 43, 9489–9499. [Google Scholar] [CrossRef]
  35. Klassen, R.; Bruch, A.; Schaffrath, R. Independent suppression of ribosomal +1 frameshifts by different tRNA anticodon loop modifications. RNA Biol. 2017, 14, 1252–1259. [Google Scholar] [CrossRef]
  36. Manickam, N.; Joshi, K.; Bhatt, M.J.; Farabaugh, P.J. Effects of tRNA modification on translational accuracy depend on intrinsic codon-anticodon strength. Nucleic Acids Res. 2016, 44, 1871–1881. [Google Scholar] [CrossRef]
  37. Su, C.; Jin, M.; Zhang, W. Conservation and Diversification of tRNA t6A-Modifying Enzymes across the Three Domains of Life. Int. J. Mol. Sci. 2022, 23, 13600. [Google Scholar] [CrossRef]
  38. Endres, L.; Dedon, P.C.; Begley, T.J. Codon-biased translation can be regulated by wobble-base tRNA modification systems during cellular stress responses. RNA Biol. 2015, 12, 603–614. [Google Scholar] [CrossRef] [PubMed]
  39. Chan, C.; Pham, P.; Dedon, P.C.; Begley, T.J. Lifestyle modifications: Coordinating the tRNA epitranscriptome with codon bias to adapt translation during stress responses. Genome Biol. 2018, 19, 228. [Google Scholar] [CrossRef]
  40. Pollo-Oliveira, L.; de Crécy-Lagard, V. Can Protein Expression Be Regulated by Modulation of tRNA Modification Profiles? Biochemistry 2019, 58, 355–362. [Google Scholar] [CrossRef] [PubMed]
  41. Mitchener, M.M.; Begley, T.J.; Dedon, P.C. Molecular Coping Mechanisms: Reprogramming tRNAs To Regulate Codon-Biased Translation of Stress Response Proteins. Acc. Chem. Res. 2023, 56, 3504–3514. [Google Scholar] [CrossRef]
  42. Pan, T. Modifications and functional genomics of human transfer RNA. Cell Res. 2018, 28, 395–404. [Google Scholar] [CrossRef]
  43. Chujo, T.; Tomizawa, K. Human transfer RNA modopathies: Diseases caused by aberrations in transfer RNA modifications. FEBS J. 2021, 288, 7096–7122. [Google Scholar] [CrossRef]
  44. Zhou, J.B.; Wang, E.D.; Zhou, X.L. Modifications of the human tRNA anticodon loop and their associations with genetic diseases. Cell Mol. Life Sci. 2021, 78, 7087–7105. [Google Scholar] [CrossRef] [PubMed]
  45. Alexander, R.W.; Eargle, J.; Luthey-Schulten, Z. Experimental and computational determination of tRNA dynamics. FEBS Lett. 2010, 584, 376–386. [Google Scholar] [CrossRef] [PubMed]
  46. Väre, V.Y.P.; Eruysal, E.R.; Narendran, A.; Sarachan, K.L.; Agris, P.F. Chemical and Conformational Diversity of Modified Nucleosides Affects tRNA Structure and Function. Biomolecules 2017, 7, 29. [Google Scholar] [CrossRef] [PubMed]
  47. Roovers, M.; Droogmans, L.; Grosjean, H. Post-Transcriptional Modifications of Conserved Nucleotides in the T-Loop of tRNA: A Tale of Functional Convergent Evolution. Genes 2021, 12, 140. [Google Scholar] [CrossRef] [PubMed]
  48. Hori, H. Transfer RNA Modification Enzymes with a Thiouridine Synthetase, Methyltransferase and Pseudouridine Synthase (THUMP) Domain and the Nucleosides They Produce in tRNA. Genes 2023, 14, 382. [Google Scholar] [CrossRef] [PubMed]
  49. Kawai, G.; Yamamoto, Y.; Kamimura, T.; Masegi, T.; Sekine, M.; Hata, T.; Iimori, T.; Watanabe, T.; Miyazawa, T.; Yokoyama, S. Conformational rigidity of specific pyrimidine residues in tRNA arises from posttranscriptional modifications that enhance steric interaction between the base and the 2’-hydroxyl group. Biochemistry 1992, 31, 1040–1046. [Google Scholar] [CrossRef] [PubMed]
  50. Yamamoto, Y.; Yokoyama, S.; Miyazawa, T.; Watanabe, K.; Higuchi, S. NMR analyses on the molecular mechanism of the conformational rigidity of 2-thioribothymidine, a modified nucleoside in extreme thermophile tRNAs. FEBS Lett. 1983, 157, 95–99. [Google Scholar] [CrossRef] [PubMed]
  51. Davis, D.R. Stabilization of RNA stacking by pseudouridine. Nucleic Acids Res. 1995, 23, 5020–5026. [Google Scholar] [CrossRef]
  52. Zhang, X.; Walker, R.C.; Phizicky, E.M.; Mathews, D.H. Influence of Sequence and Covalent Modifications on Yeast tRNA Dynamics. J. Chem. Theory Comput. 2014, 10, 3473–3483. [Google Scholar] [CrossRef]
  53. Nomura, Y.; Ohno, S.; Nishikawa, K.; Yokogawa, T. Correlation between the stability of tRNA tertiary structure and the catalytic efficiency of a tRNA-modifying enzyme, archaeal tRNA-guanine transglycosylase. Genes Cells 2016, 21, 41–52. [Google Scholar] [CrossRef]
  54. Davanloo, P.; Sprinzl, M.; Watanabe, K.; Albani, M.; Kersten, H. Role of ribothymidine in the thermal stability of transfer RNA as monitored by proton magnetic resonance. Nucleic Acids Res. 1979, 6, 1571–1581. [Google Scholar] [CrossRef]
  55. Horie, N.; Hara-Yokoyama, M.; Yokoyama, S.; Watanabe, K.; Kuchino, Y.; Nishimura, S.; Miyazawa, T. Two tRNAIle1 species from an extreme thermophile, Thermus thermophilus HB8: Effect of 2-thiolation of ribothymidine on the thermostability of tRNA. Biochemistry 1985, 24, 5711–5715. [Google Scholar] [CrossRef]
  56. Griffey, R.H.; Davis, D.R.; Yamaizumi, Z.; Nishimura, S.; Hawkins, B.L.; Poulter, C.D. 15N-labeled tRNA. Identification of 4-thiouridine in Escherichia coli tRNASer1 and tRNATyr2 by 1H-15N two-dimensional NMR spectroscopy. J. Biol. Chem. 1986, 261, 12074–12078. [Google Scholar] [CrossRef]
  57. Urbonavicius, J.; Durand, J.M.B.; Björk, G.R. Three modifications in the D and T arms of tRNA influence translation in Escherichia coli and expression of virulence genes in Shigella flexneri. J. Bacteriol. 2002, 184, 5348–5357. [Google Scholar] [CrossRef] [PubMed]
  58. Arnez, J.G.; Steitz, T.A. Crystal structure of unmodified tRNA(Gln) complexed with glutaminyl-tRNA synthetase and ATP suggests a possible role for pseudo-uridines in stabilization of RNA structure. Biochemistry 1994, 33, 7560–7567. [Google Scholar] [CrossRef] [PubMed]
  59. Westhof, E. Pseudouridines or how to draw on weak energy differences. Biochem. Biophys. Res. Commun. 2019, 520, 702–704. [Google Scholar] [CrossRef] [PubMed]
  60. Nobles, K.N.; Yarian, C.S.; Liu, G.; Guenther, R.H.; Agris, P.F. Highly conserved modified nucleosides influence Mg2+-dependent tRNA folding. Nucleic Acids Res. 2002, 30, 4751–4760. [Google Scholar] [CrossRef] [PubMed]
  61. Sowers, L.C.; Shaw, B.R.; Sedwick, W.D. Base stacking and molecular polarizability: Effect of a methyl group in the 5-position of pyrimidines. Biochem. Biophys. Res. Commun. 1987, 148, 790–794. [Google Scholar] [CrossRef]
  62. Kintanar, A.; Yue, D.; Horowitz, J. Effect of nucleoside modifications on the structure and thermal stability of Escherichia coli valine tRNA. Biochimie 1994, 76, 1192–1204. [Google Scholar] [CrossRef] [PubMed]
  63. Agris, P.F.; Sierzputowska-Gracz, H.; Smith, C. Transfer RNA contains sites of localized positive charge: Carbon NMR studies of [13C]methyl-enriched Escherichia coli and yeast tRNAPhe. Biochemistry 1986, 25, 5126–5131. [Google Scholar] [CrossRef]
  64. Steinberg, S.; Cedergren, R. A correlation between N2-dimethylguanosine presence and alternate tRNA conformers. RNA 1995, 1, 886–891. [Google Scholar]
  65. Helm, M.; Brulé, H.; Degoul, F.; Cepanec, C.; Leroux, J.P.; Giegé, R.; Florentz, C. The presence of modified nucleotides is required for cloverleaf folding of a human mitochondrial tRNA. Nucleic Acids Res. 1998, 26, 1636–1643. [Google Scholar] [CrossRef] [PubMed]
  66. Helm, M.; Giegé, R.; Florentz, C. A Watson-Crick base-pair-disrupting methyl group (m1A9) is sufficient for cloverleaf folding of human mitochondrial tRNALys. Biochemistry 1999, 38, 13338–13346. [Google Scholar] [CrossRef]
  67. Finet, O.; Yague-Sanz, C.; Marchand, F.; Hermand, D. The Dihydrouridine landscape from tRNA to mRNA: A perspective on synthesis, structural impact and function. RNA Biol. 2022, 19, 735–750. [Google Scholar] [CrossRef]
  68. Catala, M.; Gato, A.; Tisné, C.; Barraud, P. 1H, 15N chemical shift assignments of the imino groups of yeast tRNAPhe: Influence of the post-transcriptional modifications. Biomol. NMR Assign. 2020, 14, 169–174. [Google Scholar] [CrossRef] [PubMed]
  69. Barraud, P.; Gato, A.; Heiss, M.; Catala, M.; Kellner, S.; Tisné, C. Time-resolved NMR monitoring of tRNA maturation. Nat. Commun. 2019, 10, 3373. [Google Scholar] [CrossRef]
  70. Yared, M.J.; Yoluç, Y.; Catala, M.; Tisné, C.; Kaiser, S.; Barraud, P. Different modification pathways for m1A58 incorporation in yeast elongator and initiator tRNAs. Nucleic Acids Res. 2023, 51, 10653–10667. [Google Scholar] [CrossRef]
  71. Basavappa, R.; Sigler, P.B. The 3 A crystal structure of yeast initiator tRNA: Functional implications in initiator/elongator discrimination. EMBO J. 1991, 10, 3105–3111. [Google Scholar] [CrossRef] [PubMed]
  72. Phizicky, E.M.; Alfonzo, J.D. Do all modifications benefit all tRNAs? FEBS Lett. 2010, 584, 265–271. [Google Scholar] [CrossRef]
  73. Dalluge, J.J.; Hashizume, T.; Sopchik, A.E.; McCloskey, J.A.; Davis, D.R. Conformational flexibility in RNA: The role of dihydrouridine. Nucleic Acids Res. 1996, 24, 1073–1079. [Google Scholar] [CrossRef]
  74. Dyubankova, N.; Sochacka, E.; Kraszewska, K.; Nawrot, B.; Herdewijn, P.; Lescrinier, E. Contribution of dihydrouridine in folding of the D-arm in tRNA. Org. Biomol. Chem. 2015, 13, 4960–4966. [Google Scholar] [CrossRef] [PubMed]
  75. Biedenbänder, T.; de Jesus, V.; Schmidt-Dengler, M.; Helm, M.; Corzilius, B.; Fürtig, B. RNA modifications stabilize the tertiary structure of tRNAfMet by locally increasing conformational dynamics. Nucleic Acids Res. 2022, 50, 2334–2349. [Google Scholar] [CrossRef] [PubMed]
  76. Takakura, M.; Ishiguro, K.; Akichika, S.; Miyauchi, K.; Suzuki, T. Biogenesis and functions of aminocarboxypropyluridine in tRNA. Nat. Commun. 2019, 10, 5542. [Google Scholar] [CrossRef] [PubMed]
  77. Vermeulen, A.; McCallum, S.A.; Pardi, A. Comparison of the global structure and dynamics of native and unmodified tRNAval. Biochemistry 2005, 44, 6024–6033. [Google Scholar] [CrossRef]
  78. Godwin, R.C.; Macnamara, L.M.; Alexander, R.W.; Salsbury, F.R., Jr. Structure and Dynamics of tRNAMet Containing Core Substitutions. ACS Omega 2018, 3, 10668–10678. [Google Scholar] [CrossRef]
  79. Bhaskaran, H.; Taniguchi, T.; Suzuki, T.; Suzuki, T.; Perona, J.J. Structural dynamics of a mitochondrial tRNA possessing weak thermodynamic stability. Biochemistry 2014, 53, 1456–1465. [Google Scholar] [CrossRef]
  80. Porat, J.; Kothe, U.; Bayfield, M.A. Revisiting tRNA chaperones: New players in an ancient game. RNA 2021, 27, 543–559. [Google Scholar] [CrossRef]
  81. Keffer-Wilkes, L.C.; Veerareddygari, G.R.; Kothe, U. RNA modification enzyme TruB is a tRNA chaperone. Proc. Natl. Acad. Sci. USA 2016, 113, 14306–14311. [Google Scholar] [CrossRef] [PubMed]
  82. Keffer-Wilkes, L.C.; Soon, E.F.; Kothe, U. The methyltransferase TrmA facilitates tRNA folding through interaction with its RNA-binding domain. Nucleic Acids Res. 2020, 48, 7981–7990. [Google Scholar] [CrossRef] [PubMed]
  83. Ishitani, R.; Nureki, O.; Nameki, N.; Okada, N.; Nishimura, S.; Yokoyama, S. Alternative tertiary structure of tRNA for recognition by a posttranscriptional modification enzyme. Cell 2003, 113, 383–394. [Google Scholar] [CrossRef] [PubMed]
  84. Finer-Moore, J.; Czudnochowski, N.; O’Connell, J.D., 3rd; Wang, A.L.; Stroud, R.M. Crystal Structure of the Human tRNA m(1)A58 Methyltransferase-tRNA(3)(Lys) Complex: Refolding of Substrate tRNA Allows Access to the Methylation Target. J. Mol. Biol. 2015, 427, 3862–3876. [Google Scholar] [CrossRef] [PubMed]
  85. Porat, J.; Vakiloroayaei, A.; Remnant, B.M.; Talebi, M.; Cargill, T.; Bayfield, M.A. Crosstalk between the tRNA methyltransferase Trm1 and RNA chaperone La influences eukaryotic tRNA maturation. J. Biol. Chem. 2023, 299, 105326. [Google Scholar] [CrossRef]
  86. Rajkowitsch, L.; Schroeder, R. Dissecting RNA chaperone activity. RNA 2007, 13, 2053–2060. [Google Scholar] [CrossRef]
  87. Chakshusmathi, G.; Kim, S.D.; Rubinson, D.A.; Wolin, S.L. A La protein requirement for efficient pre-tRNA folding. EMBO J. 2003, 22, 6562–6572. [Google Scholar] [CrossRef]
  88. Copela, L.A.; Chakshusmathi, G.; Sherrer, R.L.; Wolin, S.L. The La protein functions redundantly with tRNA modification enzymes to ensure tRNA structural stability. RNA 2006, 12, 644–654. [Google Scholar] [CrossRef]
  89. Vakiloroayaei, A.; Shah, N.S.; Oeffinger, M.; Bayfield, M.A. The RNA chaperone La promotes pre-tRNA maturation via indiscriminate binding of both native and misfolded targets. Nucleic Acids Res. 2017, 45, 11341–11355. [Google Scholar] [CrossRef]
  90. Kadaba, S.; Wang, X.; Anderson, J.T. Nuclear RNA surveillance in Saccharomyces cerevisiae: Trf4p-dependent polyadenylation of nascent hypomethylated tRNA and an aberrant form of 5S rRNA. RNA 2006, 12, 508–521. [Google Scholar] [CrossRef]
  91. Tasak, M.; Phizicky, E.M. Initiator tRNA lacking 1-methyladenosine is targeted by the rapid tRNA decay pathway in evolutionarily distant yeast species. PLoS Genet 2022, 18, e1010215. [Google Scholar] [CrossRef]
  92. Whipple, J.M.; Lane, E.A.; Chernyakov, I.; D’Silva, S.; Phizicky, E.M. The yeast rapid tRNA decay pathway primarily monitors the structural integrity of the acceptor and T-stems of mature tRNA. Genes Dev. 2011, 25, 1173–1184. [Google Scholar] [CrossRef]
  93. Bowles, I.E.; Jackman, J.E. A tRNA-specific function for tRNA methyltransferase Trm10 is associated with a new tRNA quality control mechanism in Saccharomyces cerevisiae. RNA 2024, 30, 171–187. [Google Scholar] [CrossRef]
  94. Hori, H. Regulatory Factors for tRNA Modifications in Extreme-Thermophilic Bacterium Thermus thermophilus. Front. Genet. 2019, 10, 204. [Google Scholar] [CrossRef]
  95. Tomikawa, C.; Yokogawa, T.; Kanai, T.; Hori, H. N7-Methylguanine at position 46 (m7G46) in tRNA from Thermus thermophilus is required for cell viability at high temperatures through a tRNA modification network. Nucleic Acids Res. 2010, 38, 942–957. [Google Scholar] [CrossRef] [PubMed]
  96. Chernyakov, I.; Whipple, J.M.; Kotelawala, L.; Grayhack, E.J.; Phizicky, E.M. Degradation of several hypomodified mature tRNA species in Saccharomyces cerevisiae is mediated by Met22 and the 5’-3’exonucleases Rat1 and Xrn1. Genes Dev. 2008, 22, 1369–1380. [Google Scholar] [CrossRef]
  97. Kotelawala, L.; Grayhack, E.J.; Phizicky, E.M. Identification of yeast tRNA Um(44) 2’-O-methyltransferase (Trm44) and demonstration of a Trm44 role in sustaining levels of specific tRNA(Ser) species. RNA 2008, 14, 158–169. [Google Scholar] [CrossRef] [PubMed]
  98. Dewe, J.M.; Whipple, J.M.; Chernyakov, I.; Jaramillo, L.N.; Phizicky, E.M. The yeast rapid tRNA decay pathway competes with elongation factor 1A for substrate tRNAs and acts on tRNAs lacking one or more of several modifications. RNA 2012, 18, 1886–1896. [Google Scholar] [CrossRef]
  99. De Zoysa, T.; Phizicky, E.M. Hypomodified tRNA in evolutionarily distant yeasts can trigger rapid tRNA decay to activate the general amino acid control response, but with different consequences. PLoS Genet. 2020, 16, e1008893. [Google Scholar] [CrossRef]
  100. Han, L.; Phizicky, E.M. A rationale for tRNA modification circuits in the anticodon loop. RNA 2018, 24, 1277–1284. [Google Scholar] [CrossRef] [PubMed]
  101. Sokołowski, M.; Klassen, R.; Bruch, A.; Schaffrath, R.; Glatt, S. Cooperativity between different tRNA modifications and their modification pathways. Biochim. Biophys. Acta Gene Regul. Mech. 2018, 1861, 409–418. [Google Scholar] [CrossRef]
  102. Li, J.; Zhu, W.Y.; Yang, W.Q.; Li, C.T.; Liu, R.J. The occurrence order and cross-talk of different tRNA modifications. Sci. China Life Sci. 2021, 64, 1423–1436. [Google Scholar] [CrossRef] [PubMed]
  103. Tomikawa, C. 7-Methylguanosine Modifications in Transfer RNA (tRNA). Int. J. Mol. Sci. 2018, 19, 4080. [Google Scholar] [CrossRef] [PubMed]
  104. Kowalak, J.A.; Dalluge, J.J.; McCloskey, J.A.; Stetter, K.O. The role of posttranscriptional modification in stabilization of transfer RNA from hyperthermophiles. Biochemistry 1994, 33, 7869–7876. [Google Scholar] [CrossRef] [PubMed]
  105. Droogmans, L.; Roovers, M.; Bujnicki, J.M.; Tricot, C.; Hartsch, T.; Stalon, V.; Grosjean, H. Cloning and characterization of tRNA (m1A58) methyltransferase (TrmI) from Thermus thermophilus HB27, a protein required for cell growth at extreme temperatures. Nucleic Acids Res. 2003, 31, 2148–2156. [Google Scholar] [CrossRef] [PubMed]
  106. Ohira, T.; Suzuki, T. Transfer RNA modifications and cellular thermotolerance. Mol. Cell 2024, 84, 94–106. [Google Scholar] [CrossRef] [PubMed]
  107. Ishida, K.; Kunibayashi, T.; Tomikawa, C.; Ochi, A.; Kanai, T.; Hirata, A.; Iwashita, C.; Hori, H. Pseudouridine at position 55 in tRNA controls the contents of other modified nucleotides for low-temperature adaptation in the extreme-thermophilic eubacterium Thermus thermophilus. Nucleic Acids Res. 2011, 39, 2304–2318. [Google Scholar] [CrossRef]
  108. Yamagami, R.; Tomikawa, C.; Shigi, N.; Kazayama, A.; Asai, S.I.; Takuma, H.; Hirata, A.; Fourmy, D.; Asahara, H.; Watanabe, K.; et al. Folate-/FAD-dependent tRNA methyltransferase from Thermus thermophilus regulates other modifications in tRNA at low temperatures. Genes Cells 2016, 21, 740–754. [Google Scholar] [CrossRef]
  109. Meyer, B.; Immer, C.; Kaiser, S.; Sharma, S.; Yang, J.; Watzinger, P.; Weiß, L.; Kotter, A.; Helm, M.; Seitz, H.M.; et al. Identification of the 3-amino-3-carboxypropyl (acp) transferase enzyme responsible for acp3U formation at position 47 in Escherichia coli tRNAs. Nucleic Acids Res. 2020, 48, 1435–1450. [Google Scholar] [CrossRef]
  110. Jones, J.D.; Franco, M.K.; Tardu, M.; Smith, T.J.; Snyder, L.R.; Eyler, D.E.; Polikanov, Y.; Kennedy, R.T.; Niederer, R.O.; Koutmou, K.S. Conserved 5-methyluridine tRNA modification modulates ribosome translocation. bioRxiv 2023. [Google Scholar] [CrossRef]
  111. Schultz, S.K.; Katanski, C.D.; Halucha, M.; Pena, N.; Fahlman, R.P.; Pan, T.; Kothe, U. Modifications in the T arm of tRNA globally determine tRNA maturation, function and cellular fitness. bioRxiv 2023. [Google Scholar] [CrossRef]
  112. Lucas, M.C.; Pryszcz, L.P.; Medina, R.; Milenkovic, I.; Camacho, N.; Marchand, V.; Motorin, Y.; Ribas de Pouplana, L.; Novoa, E.M. Quantitative analysis of tRNA abundance and modifications by nanopore RNA sequencing. Nat. Biotechnol. 2024, 42, 72–86. [Google Scholar] [CrossRef] [PubMed]
  113. Smoczynski, J.; Yared, M.J.; Meynier, V.; Barraud, P.; Tisné, C. Advances in the Structural and Functional Understanding of m1A RNA Modification. Acc. Chem. Res. 2024, 57, 429–438. [Google Scholar] [CrossRef] [PubMed]
  114. Schultz, S.K.L.; Kothe, U. tRNA elbow modifications affect the tRNA pseudouridine synthase TruB and the methyltransferase TrmA. RNA 2020, 26, 1131–1142. [Google Scholar] [CrossRef] [PubMed]
  115. Thomas, N.K.; Poodari, V.C.; Jain, M.; Olsen, H.E.; Akeson, M.; Abu-Shumays, R.L. Direct Nanopore Sequencing of Individual Full Length tRNA Strands. ACS Nano 2021, 15, 16642–16653. [Google Scholar] [CrossRef] [PubMed]
  116. Emilsson, V.; Näslund, A.K.; Kurland, C.G. Thiolation of transfer RNA in Escherichia coli varies with growth rate. Nucleic Acids Res. 1992, 20, 4499–4505. [Google Scholar] [CrossRef] [PubMed]
  117. Emilsson, V.; Näslund, A.K.; Kurland, C.G. Growth-rate-dependent accumulation of twelve tRNA species in Escherichia coli. J. Mol. Biol. 1993, 230, 483–491. [Google Scholar] [CrossRef]
  118. Dong, H.; Nilsson, L.; Kurland, C.G. Co-variation of tRNA abundance and codon usage in Escherichia coli at different growth rates. J. Mol. Biol. 1996, 260, 649–663. [Google Scholar] [CrossRef]
  119. Torrent, M.; Chalancon, G.; de Groot, N.S.; Wuster, A.; Madan Babu, M. Cells alter their tRNA abundance to selectively regulate protein synthesis during stress conditions. Sci. Signal. 2018, 11. [Google Scholar] [CrossRef]
  120. Singhal, R.P.; Vold, B. Changes in transfer ribonucleic acids of Bacillus subtilis during different growth phases. Nucleic Acids Res. 1976, 3, 1249–1262. [Google Scholar] [CrossRef]
  121. Chan, C.T.Y.; Dyavaiah, M.; DeMott, M.S.; Taghizadeh, K.; Dedon, P.C.; Begley, T.J. A quantitative systems approach reveals dynamic control of tRNA modifications during cellular stress. PLoS Genet. 2010, 6, e1001247. [Google Scholar] [CrossRef]
  122. Pang, Y.L.J.; Abo, R.; Levine, S.S.; Dedon, P.C. Diverse cell stresses induce unique patterns of tRNA up- and down-regulation: TRNA-seq for quantifying changes in tRNA copy number. Nucleic Acids Res. 2014, 42, e170. [Google Scholar] [CrossRef]
  123. Galvanin, A.; Vogt, L.M.; Grober, A.; Freund, I.; Ayadi, L.; Bourguignon-Igel, V.; Bessler, L.; Jacob, D.; Eigenbrod, T.; Marchand, V.; et al. Bacterial tRNA 2’-O-methylation is dynamically regulated under stress conditions and modulates innate immune response. Nucleic Acids Res. 2020, 48, 12833–12844. [Google Scholar] [CrossRef] [PubMed]
  124. Preston, M.A.; D’Silva, S.; Kon, Y.; Phizicky, E.M. tRNAHis 5-methylcytidine levels increase in response to several growth arrest conditions in Saccharomyces cerevisiae. RNA 2013, 19, 243–256. [Google Scholar] [CrossRef] [PubMed]
  125. Yamagami, R.; Sieg, J.P.; Assmann, S.M.; Bevilacqua, P.C. Genome-wide analysis of the in vivo tRNA structurome reveals RNA structural and modification dynamics under heat stress. Proc. Natl. Acad. Sci. USA 2022, 119, e2201237119. [Google Scholar] [CrossRef] [PubMed]
  126. Pape, T.; Wintermeyer, W.; Rodnina, M.V. Complete kinetic mechanism of elongation factor Tu-dependent binding of aminoacyl-tRNA to the A site of the E. coli ribosome. EMBO J. 1998, 17, 7490–7497. [Google Scholar] [CrossRef] [PubMed]
  127. Whitford, P.C.; Geggier, P.; Altman, R.B.; Blanchard, S.C.; Onuchic, J.N.; Sanbonmatsu, K.Y. Accommodation of aminoacyl-tRNA into the ribosome involves reversible excursions along multiple pathways. RNA 2010, 16, 1196–1204. [Google Scholar] [CrossRef] [PubMed]
  128. Zhou, J.; Lancaster, L.; Donohue, J.P.; Noller, H.F. How the ribosome hands the A-site tRNA to the P site during EF-G-catalyzed translocation. Science 2014, 345, 1188–1191. [Google Scholar] [CrossRef] [PubMed]
  129. Rundlet, E.J.; Holm, M.; Schacherl, M.; Natchiar, S.K.; Altman, R.B.; Spahn, C.M.T.; Myasnikov, A.G.; Blanchard, S.C. Structural basis of early translocation events on the ribosome. Nature 2021, 595, 741–745. [Google Scholar] [CrossRef]
  130. Milicevic, N.; Jenner, L.; Myasnikov, A.; Yusupov, M.; Yusupova, G. mRNA reading frame maintenance during eukaryotic ribosome translocation. Nature 2024, 625, 393–400. [Google Scholar] [CrossRef]
  131. Ling, C.; Ermolenko, D.N. Structural insights into ribosome translocation. Wiley Interdiscip. Rev. RNA 2016, 7, 620–636. [Google Scholar] [CrossRef]
  132. Saleh, S.; Farabaugh, P.J. Posttranscriptional modification to the core of tRNAs modulates translational misreading errors. RNA 2023, 30, 37–51. [Google Scholar] [CrossRef] [PubMed]
  133. Kinghorn, S.M.; O’Byrne, C.P.; Booth, I.R.; Stansfield, I. Physiological analysis of the role of truB in Escherichia coli: A role for tRNA modification in extreme temperature resistance. Microbiology 2002, 148, 3511–3520. [Google Scholar] [CrossRef] [PubMed]
  134. Björk, G.R.; Neidhardt, F.C. Physiological and biochemical studies on the function of 5-methyluridine in the transfer ribonucleic acid of Escherichia coli. J. Bacteriol. 1975, 124, 99–111. [Google Scholar] [CrossRef] [PubMed]
  135. Gutgsell, N.; Englund, N.; Niu, L.; Kaya, Y.; Lane, B.G.; Ofengand, J. Deletion of the Escherichia coli pseudouridine synthase gene truB blocks formation of pseudouridine 55 in tRNA in vivo, does not affect exponential growth, but confers a strong selective disadvantage in competition with wild-type cells. RNA 2000, 6, 1870–1881. [Google Scholar] [CrossRef] [PubMed]
  136. Kersten, H.; Albani, M.; Männlein, E.; Praisler, R.; Wurmbach, P.; Nierhaus, K.H. On the role of ribosylthymine in prokaryotic tRNA function. Eur. J. Biochem. 1981, 114, 451–456. [Google Scholar] [CrossRef]
  137. Chou, H.J.; Donnard, E.; Gustafsson, H.T.; Garber, M.; Rando, O.J. Transcriptome-wide Analysis of Roles for tRNA Modifications in Translational Regulation. Mol. Cell 2017, 68, 978–992.e4. [Google Scholar] [CrossRef]
  138. Witzenberger, M.; Burczyk, S.; Settele, D.; Mayer, W.; Welp, L.M.; Heiss, M.; Wagner, M.; Monecke, T.; Janowski, R.; Carell, T.; et al. Human TRMT2A methylates tRNA and contributes to translation fidelity. Nucleic Acids Res. 2023, 51, 8691–8710. [Google Scholar] [CrossRef]
  139. Girodat, D.; Wieden, H.J.; Blanchard, S.C.; Sanbonmatsu, K.Y. Geometric alignment of aminoacyl-tRNA relative to catalytic centers of the ribosome underpins accurate mRNA decoding. Nat. Commun. 2023, 14, 5582. [Google Scholar] [CrossRef] [PubMed]
  140. Uhlenbeck, O.C.; Schrader, J.M. Evolutionary tuning impacts the design of bacterial tRNAs for the incorporation of unnatural amino acids by ribosomes. Curr. Opin. Chem. Biol. 2018, 46, 138–145. [Google Scholar] [CrossRef] [PubMed]
  141. Favre, A.; Hajnsdorf, E.; Thiam, K.; Caldeira de Araujo, A. Mutagenesis and growth delay induced in Escherichia coli by near-ultraviolet radiations. Biochimie 1985, 67, 335–342. [Google Scholar] [CrossRef] [PubMed]
  142. Favre, A.; Michelson, A.M.; Yaniv, M. Photochemistry of 4-thiouridine in Escherichia coli transfer RNA1Val. J. Mol. Biol. 1971, 58, 367–379. [Google Scholar] [CrossRef] [PubMed]
  143. Carré, D.S.; Thomas, G.; Favre, A. Conformation and functioning of tRNAs: Cross-linked tRNAs as substrate for tRNA nucleotidyl-transferase and aminoacyl synthetases. Biochimie 1974, 56, 1089–1101. [Google Scholar] [CrossRef]
  144. Thomas, G.; Thiam, K.; Favre, A. tRNA thiolated pyrimidines as targets for near-ultraviolet-induced synthesis of guanosine tetraphosphate in Escherichia coli. Eur. J. Biochem. 1981, 119, 381–387. [Google Scholar] [CrossRef] [PubMed]
  145. Thomas, G.; Favre, A. 4-Thiouridine as the target for near-ultraviolet light induced growth delay in Escherichia coli. Biochem. Biophys. Res. Commun. 1975, 66, 1454–1461. [Google Scholar] [CrossRef]
  146. Thiam, K.; Favre, A. Role of the stringent response in the expression and mechanism of near-ultraviolet induced growth delay. Eur. J. Biochem. 1984, 145, 137–142. [Google Scholar] [CrossRef]
  147. Kramer, G.F.; Baker, J.C.; Ames, B.N. Near-UV stress in Salmonella typhimurium: 4-thiouridine in tRNA, ppGpp, and ApppGpp as components of an adaptive response. J. Bacteriol. 1988, 170, 2344–2351. [Google Scholar] [CrossRef]
  148. Ramabhadran, T.V.; Jagger, J. Mechanism of growth delay induced in Escherichia coli by near ultraviolet radiation. Proc. Natl. Acad. Sci. USA 1976, 73, 59–63. [Google Scholar] [CrossRef]
  149. Ramabhadran, T.V.; Fossum, T.; Jagger, J. Escherichia coli mutant lacking 4-thiouridine in its transfer ribonucleic acid. J. Bacteriol. 1976, 128, 671–672. [Google Scholar] [CrossRef]
  150. Probst-Rüd, S.; McNeill, K.; Ackermann, M. Thiouridine residues in tRNAs are responsible for a synergistic effect of UVA and UVB light in photoinactivation of Escherichia coli. Environ. Microbiol. 2017, 19, 434–442. [Google Scholar] [CrossRef] [PubMed]
  151. Gehrig, S.; Eberle, M.E.; Botschen, F.; Rimbach, K.; Eberle, F.; Eigenbrod, T.; Kaiser, S.; Holmes, W.M.; Erdmann, V.A.; Sprinzl, M.; et al. Identification of modifications in microbial, native tRNA that suppress immunostimulatory activity. J. Exp. Med. 2012, 209, 225–233. [Google Scholar] [CrossRef]
  152. Jöckel, S.; Nees, G.; Sommer, R.; Zhao, Y.; Cherkasov, D.; Hori, H.; Ehm, G.; Schnare, M.; Nain, M.; Kaufmann, A.; et al. The 2’-O-methylation status of a single guanosine controls transfer RNA-mediated Toll-like receptor 7 activation or inhibition. J. Exp. Med. 2012, 209, 235–241. [Google Scholar] [CrossRef] [PubMed]
  153. Kaiser, S.; Rimbach, K.; Eigenbrod, T.; Dalpke, A.H.; Helm, M. A modified dinucleotide motif specifies tRNA recognition by TLR7. RNA 2014, 20, 1351–1355. [Google Scholar] [CrossRef] [PubMed]
  154. Rimbach, K.; Kaiser, S.; Helm, M.; Dalpke, A.H.; Eigenbrod, T. 2’-O-Methylation within Bacterial RNA Acts as Suppressor of TLR7/TLR8 Activation in Human Innate Immune Cells. J. Innate Immun. 2015, 7, 482–493. [Google Scholar] [CrossRef]
  155. Jain, V.; Kumar, M.; Chatterji, D. ppGpp: Stringent response and survival. J. Microbiol. 2006, 44, 1–10. [Google Scholar] [PubMed]
  156. De Bie, L.G.S.; Roovers, M.; Oudjama, Y.; Wattiez, R.; Tricot, C.; Stalon, V.; Droogmans, L.; Bujnicki, J.M. The yggH gene of Escherichia coli encodes a tRNA (m7G46) methyltransferase. J. Bacteriol. 2003, 185, 3238–3243. [Google Scholar] [CrossRef] [PubMed]
  157. Alexandrov, A.; Martzen, M.R.; Phizicky, E.M. Two proteins that form a complex are required for 7-methylguanosine modification of yeast tRNA. RNA 2002, 8, 1253–1266. [Google Scholar] [CrossRef] [PubMed]
  158. Thongdee, N.; Jaroensuk, J.; Atichartpongkul, S.; Chittrakanwong, J.; Chooyoung, K.; Srimahaeak, T.; Chaiyen, P.; Vattanaviboon, P.; Mongkolsuk, S.; Fuangthong, M. TrmB, a tRNA m7G46 methyltransferase, plays a role in hydrogen peroxide resistance and positively modulates the translation of katA and katB mRNAs in Pseudomonas aeruginosa. Nucleic Acids Res. 2019, 47, 9271–9281. [Google Scholar] [CrossRef] [PubMed]
  159. Lin, S.; Liu, Q.; Lelyveld, V.S.; Choe, J.; Szostak, J.W.; Gregory, R.I. Mettl1/Wdr4-Mediated m7G tRNA Methylome Is Required for Normal mRNA Translation and Embryonic Stem Cell Self-Renewal and Differentiation. Mol. Cell 2018, 71, 244–255.e5. [Google Scholar] [CrossRef]
  160. Schultz, S.K.; Meadows, K.; Kothe, U. Molecular mechanism of tRNA binding by the Escherichia coli N7 guanosine methyltransferase TrmB. J. Biol. Chem. 2023, 299, 104612. [Google Scholar] [CrossRef]
  161. McGuffey, J.C.; Jackson-Litteken, C.D.; Di Venanzio, G.; Zimmer, A.A.; Lewis, J.M.; Distel, J.S.; Kim, K.Q.; Zaher, H.S.; Alfonzo, J.; Scott, N.E.; et al. The tRNA methyltransferase TrmB is critical for Acinetobacter baumannii stress responses and pulmonary infection. mBio 2023, 14, e0141623. [Google Scholar] [CrossRef]
  162. Takano, Y.; Takayanagi, N.; Hori, H.; Ikeuchi, Y.; Suzuki, T.; Kimura, A.; Okuno, T. A gene involved in modifying transfer RNA is required for fungal pathogenicity and stress tolerance of Colletotrichum lagenarium. Mol. Microbiol. 2006, 60, 81–92. [Google Scholar] [CrossRef]
  163. Kumagai, I.; Watanabe, K.; Oshima, T. Thermally induced biosynthesis of 2’-O-methylguanosine in tRNA from an extreme thermophile, Thermus thermophilus HB27. Proc. Natl. Acad. Sci. USA 1980, 77, 1922–1926. [Google Scholar] [CrossRef]
  164. Babosan, A.; Fruchard, L.; Krin, E.; Carvalho, A.; Mazel, D.; Baharoglu, Z. Nonessential tRNA and rRNA modifications impact the bacterial response to sub-MIC antibiotic stress. Microlife 2022, 3, uqac019. [Google Scholar] [CrossRef]
  165. Yoluç, Y.; van de Logt, E.; Kellner-Kaiser, S. The Stress-Dependent Dynamics of Saccharomyces cerevisiae tRNA and rRNA Modification Profiles. Genes 2021, 12, 1344. [Google Scholar] [CrossRef]
  166. Shigi, N.; Suzuki, T.; Terada, T.; Shirouzu, M.; Yokoyama, S.; Watanabe, K. Temperature-dependent biosynthesis of 2-thioribothymidine of Thermus thermophilus tRNA. J. Biol. Chem. 2006, 281, 2104–2113. [Google Scholar] [CrossRef]
  167. Hou, Y.M.; Masuda, I.; Foster, L.J. tRNA methylation: An unexpected link to bacterial resistance and persistence to antibiotics and beyond. Wiley Interdiscip. Rev. RNA 2020, 11, e1609. [Google Scholar] [CrossRef]
  168. Burby, P.E.; Nye, T.M.; Schroeder, J.W.; Simmons, L.A. Implementation and Data Analysis of Tn-seq, Whole-Genome Resequencing, and Single-Molecule Real-Time Sequencing for Bacterial Genetics. J. Bacteriol. 2017, 199, 10-1128. [Google Scholar] [CrossRef]
  169. Fruchard, L.; Babosan, A.; Carvalho, A.; Lang, M.; Li, B.; Duchateau, M.; Giai-Gianetto, Q.; Matondo, M.; Bonhomme, F.; Hatin, I.; et al. Aminoglycoside tolerance in Vibrio cholerae engages translational reprogramming associated to queuosine tRNA modification. bioRxiv 2024. [Google Scholar] [CrossRef]
  170. Irving, S.E.; Choudhury, N.R.; Corrigan, R.M. The stringent response and physiological roles of (pp)pGpp in bacteria. Nat. Rev. Microbiol. 2021, 19, 256–271. [Google Scholar] [CrossRef]
  171. Huang, H.Y.; Hopper, A.K. Multiple Layers of Stress-Induced Regulation in tRNA Biology. Life 2016, 6, 16. [Google Scholar] [CrossRef] [PubMed]
  172. Jones, T.E.; Alexander, R.W.; Pan, T. Misacylation of specific nonmethionyl tRNAs by a bacterial methionyl-tRNA synthetase. Proc. Natl. Acad. Sci. USA 2011, 108, 6933–6938. [Google Scholar] [CrossRef] [PubMed]
  173. Lee, J.Y.; Kim, D.G.; Kim, B.G.; Yang, W.S.; Hong, J.; Kang, T.; Oh, Y.S.; Kim, K.R.; Han, B.W.; Hwang, B.J.; et al. Promiscuous methionyl-tRNA synthetase mediates adaptive mistranslation to protect cells against oxidative stress. J. Cell Sci. 2014, 127, 4234–4245. [Google Scholar] [CrossRef] [PubMed]
  174. Lyons, S.M.; Fay, M.M.; Ivanov, P. The role of RNA modifications in the regulation of tRNA cleavage. FEBS Lett. 2018, 592, 2828–2844. [Google Scholar] [CrossRef]
  175. Shetty, S.; Varshney, U. Regulation of translation by one-carbon metabolism in bacteria and eukaryotic organelles. J. Biol. Chem. 2021, 296, 100088. [Google Scholar] [CrossRef]
Figure 2. Chemical structure of common modified residues discussed in this review: 1-methyladenosine ( m 1 A), 2′-O-methylguanosine (Gm), 7-methylguanosine ( m 7 G), 5-methylcytidine ( m 5 C), 5-methyluridine ( m 5 U), pseudouridine ( Ψ ), dihydrouridine (D) and 4-thiouridine ( s 4 U). Modifications are highlighted in red.
Figure 2. Chemical structure of common modified residues discussed in this review: 1-methyladenosine ( m 1 A), 2′-O-methylguanosine (Gm), 7-methylguanosine ( m 7 G), 5-methylcytidine ( m 5 C), 5-methyluridine ( m 5 U), pseudouridine ( Ψ ), dihydrouridine (D) and 4-thiouridine ( s 4 U). Modifications are highlighted in red.
Genes 15 00374 g002
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

Yared, M.-J.; Marcelot, A.; Barraud, P. Beyond the Anticodon: tRNA Core Modifications and Their Impact on Structure, Translation and Stress Adaptation. Genes 2024, 15, 374. https://doi.org/10.3390/genes15030374

AMA Style

Yared M-J, Marcelot A, Barraud P. Beyond the Anticodon: tRNA Core Modifications and Their Impact on Structure, Translation and Stress Adaptation. Genes. 2024; 15(3):374. https://doi.org/10.3390/genes15030374

Chicago/Turabian Style

Yared, Marcel-Joseph, Agathe Marcelot, and Pierre Barraud. 2024. "Beyond the Anticodon: tRNA Core Modifications and Their Impact on Structure, Translation and Stress Adaptation" Genes 15, no. 3: 374. https://doi.org/10.3390/genes15030374

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