Next Article in Journal
Role of Sigma-1 Receptor in Calcium Modulation: Possible Involvement in Cancer
Next Article in Special Issue
Analysis of RNA Modifications by Second- and Third-Generation Deep Sequencing: 2020 Update
Previous Article in Journal
Mitochondrial Genome Evolution, Genetic Diversity, and Population Structure in British Water Voles (Arvicola amphibius)
Previous Article in Special Issue
Manganese Ions Individually Alter the Reverse Transcription Signature of Modified Ribonucleosides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Post-Transcriptional Modifications of Conserved Nucleotides in the T-Loop of tRNA: A Tale of Functional Convergent Evolution

1
Labiris, Avenue E. Gryson 1, 1070 Bruxelles, Belgium
2
Laboratoire de Chimie Biologique, Université Libre de Bruxelles (ULB), Labiris, Avenue E. Gryson 1, 1070 Bruxelles, Belgium
*
Author to whom correspondence should be addressed.
Genes 2021, 12(2), 140; https://doi.org/10.3390/genes12020140
Submission received: 29 December 2020 / Revised: 18 January 2021 / Accepted: 20 January 2021 / Published: 22 January 2021
(This article belongs to the Special Issue Functions and Dynamics of RNA Modifications)

Abstract

:
The high conservation of nucleotides of the T-loop, including their chemical identity, are hallmarks of tRNAs from organisms belonging to the three Domains of Life. These structural characteristics allow the T-loop to adopt a peculiar intraloop conformation able to interact specifically with other conserved residues of the D-loop, which ultimately folds the mature tRNA in a unique functional canonical L-shaped architecture. Paradoxically, despite the high conservation of modified nucleotides in the T-loop, enzymes catalyzing their formation depend mostly on the considered organism, attesting for an independent but convergent evolution of the post-transcriptional modification processes. The driving force behind this is the preservation of a native conformation of the tRNA elbow that underlies the various interactions of tRNA molecules with different cellular components.

1. Introduction

Transfer RNAs play a central role in deciphering the genetic code during the complex multistep translation process. To fulfil this main cellular function, each isoacceptor species has to be sufficiently distinct to serve as substrate for a large array of specific proteins and enzymes such as modification enzymes and cognate aminoacyl-tRNA synthetase (identity problem), while it also has to be sufficiently identical to serve in a uniform way for interactions with other partners such as elongation/initiation factors and with elements of the ribosome during the various steps of the translation process (conformity problem). To this end, these small RNA molecules have to fold into a uniform 3D-conformation, the so-called L-shaped architecture (reviewed in [1,2]). Conservation of certain nucleotides at key positions combined with chemical alterations of selected nucleotides catalyzed by specific enzymes during the post-transcriptional maturation processes contribute to solve both the identity and conformity problems of the tRNA repertoire within a given cell. These also allow to regulate tRNA plasticity according to the physiological conditions of the cell (especially the growth temperature) and fine-tune their interactions with the various cellular components of the translation machinery (reviewed in [3,4]). Notice, a few metazoan mitochondrial tRNAs lack the D-or T-arm and contain less modified nucleotides as the cytoplasmic counterparts. However, based on computer simulations, the overall L-shaped conformation of these “truncated” tRNAs appears to be maintained through combinatorial networks of alternative tertiary interactions (for details see [5,6]). These atypical tRNAs will not be considered in this review.
Here, we examine the structural parameters that allow fully mature cytoplasmic tRNAs from different origins (Bacteria, Eukarya, and Archaea) to adopt a universal L-shaped three-dimensional structure. Particularly important is the remarkable conservation of nucleotides (modified or not) at each of the 7 positions of the T-loop. A few nucleotides of the T-loop also specifically interact with a few conserved bases of the D-loop. The resulting L-shaped tRNA tertiary structure is further stabilized by additional tertiary base pairs between conserved nucleotides of the D-loop and of the V-loop. While conservation of nucleotides at given positions of the present-day tRNAs results from long coevolution of tRNA population with all elements of translation machinery [7,8], the enzymes that produce identical modified nucleotides are often distinct from one organism to another. This is particularly true when considering organisms belonging to different Domains of Life, attesting for convergent type of enzyme evolution for maintaining a universal functional tertiary tRNA fold.

2. The Intricate T-D-Loop Anatomy

Figure 1A displays the nearly universal secondary structure of mature tRNA sequences in which the conserved and semiconserved nucleotides present in both the T-loop and the D-loop are highlighted. In the T-loop, those nucleotides which are usually post-transcriptionally modified are indicated in red (Figure 1A,B and Figure 2). Their presence is dependent on both the type of isoacceptor tRNA and the group of organisms considered (Bacteria, Eukarya, and Archaea) (Figure 1C). In the D-loop, only 2′-O-ribose methylated G at position 18 (Gm) is mentioned among others not discussed here. The case of initiator tRNA of Eukarya is special because an A instead of the universally conserved U is found at position 54 of T-loop [9,10].
X-ray diffraction data of several tRNAs reveal the presence of a trans Watson–Crick/Hoogsteen pair between the uridine-54 and the adenosine-58 (both in anti-conformation) [11,12,13]. In the case of eukaryotic tRNAiMet, a longer purine–purine Hoogsteen pair (A54-A58) is found [14]. This conserved U54 (A54)-A58 base-pair stacks with the conserved G53-C61 base-pair at the end of the helical T-arm. This forces the conserved pyrimidine-60 (notified as Y) and nucleotide-59 (notified as N for any type of base, but mostly A or U) to flip out of the remaining 3-member mini-T-loop composed of (i) the conserved U55, which stacks with U54 and interacts with G18 of the D-loop, (ii) C56, which pairs with G19, and (iii) purine-57 (notified as R; mostly G in Bacteria and Eukarya but A in Archaea). The residue R57 is sandwiched between A58-U54 and G19-C56. Thus, the series of nucleotides C61, A58, G18, R57, and C56-G19 are forming a continuous stack (see right part of Figure 1A and Figure 3). The flipped dinucleotide, located between the T- and the D-domain is further stabilized out of the rest of T-loop by an H-bond between the 5′-phosphate of pyrimidine 60 and N4 of C61. This explains the strict conservation of the G53-C61 base pair at the end of the T-stem. In this intricate network of interacting nucleotides, key contacts are the U54-A58 Hoogsteen pair (highlighted in red in Figure 3) and the base pairs formed between the conserved U55 and C56 of the T-loop and the conserved G18 and G19 of the D-loop (highlighted in black in Figure 3). The region where the D-loop and the T-loop interact is designated as the tRNA elbow [15]. The final functional L-shaped conformation is further stabilized by a few additional tertiary base pairs between other conserved or semiconserved nucleotides of the D-loop and V-loop, such as the conserved so-called Levitt base-pair R15-Y48 [15,16] and the conserved triple interacting bases R21-A14-U8 [2,17]. Microstructural heterogeneities in the hinge part of the 3D-core results mostly from the variable number of additional bases present in the D-loop (often dihydrouridine) and V-loop of the different isoacceptor tRNAs (indicated by small blue triangles in Figure 1A). This variability in the hinge region allows some tRNA flexibility but does not alter much the global canonical L-shaped tRNA architecture that is mostly dominated by the T-D-loop kissing-type of complex (elbow).
Figure 1. Presence of modified nucleotides in the tRNA elbow structure. (A) Two-dimensional representation of tRNA emphasizing conserved residues involved in D-and T-loop interaction. In red are the residues subject to enzymatic modifications. The little blue box on the right corresponds to the one in Figure 3 (see text for explanation). R stands for purine, Y for pyrimidine, and N for any of the four canonical nucleotides. (B) Residues at G18 and in the T-loop and their observed modifications are shown in red. (C) Distribution of modified residues of the T-loop among organisms of the three Domains of Life. Acronyms for modified nucleotides are those used in Modomics [10]. The 2′-O-methylation of m1I57 into m1Im57 is putative (see text).
Figure 1. Presence of modified nucleotides in the tRNA elbow structure. (A) Two-dimensional representation of tRNA emphasizing conserved residues involved in D-and T-loop interaction. In red are the residues subject to enzymatic modifications. The little blue box on the right corresponds to the one in Figure 3 (see text for explanation). R stands for purine, Y for pyrimidine, and N for any of the four canonical nucleotides. (B) Residues at G18 and in the T-loop and their observed modifications are shown in red. (C) Distribution of modified residues of the T-loop among organisms of the three Domains of Life. Acronyms for modified nucleotides are those used in Modomics [10]. The 2′-O-methylation of m1I57 into m1Im57 is putative (see text).
Genes 12 00140 g001
Figure 2. Chemical structure of the modified nucleotides found in the T-loop and G18.
Figure 2. Chemical structure of the modified nucleotides found in the T-loop and G18.
Genes 12 00140 g002
Figure 3. Scheme recapitulating the organization of the elbow within the general 3D organization of tRNA. (A) Rectangles correspond to nucleotides involved in stacking or base-pairing with another nucleotide of the T- and D-loop. Residues involved in the trans Watson–Crick/Hoogsteen interaction are in red boxes, other residues of the T-loop are in blue boxes, while those of the D-loop are in yellow boxes, except G18 and G19 in black boxes. The possible modifications of each residue are indicated in red. (B) Schematic representation of the 3D architecture of tRNA molecule. (C) The 3D structure of tRNA elbow as found in the crystal structure at 1.93 Å resolution of yeast tRNAPhe harboring m5U54, Ψ55, and m1A58 [13]. The trans Watson–Crick/Hoogsteen interaction is in red, while G18 and G19 are in black.
Figure 3. Scheme recapitulating the organization of the elbow within the general 3D organization of tRNA. (A) Rectangles correspond to nucleotides involved in stacking or base-pairing with another nucleotide of the T- and D-loop. Residues involved in the trans Watson–Crick/Hoogsteen interaction are in red boxes, other residues of the T-loop are in blue boxes, while those of the D-loop are in yellow boxes, except G18 and G19 in black boxes. The possible modifications of each residue are indicated in red. (B) Schematic representation of the 3D architecture of tRNA molecule. (C) The 3D structure of tRNA elbow as found in the crystal structure at 1.93 Å resolution of yeast tRNAPhe harboring m5U54, Ψ55, and m1A58 [13]. The trans Watson–Crick/Hoogsteen interaction is in red, while G18 and G19 are in black.
Genes 12 00140 g003

3. Modified Nucleotides within the T-Loop and G18 Stabilize the Elbow of tRNA

The chemical structures of the various modified nucleotides found in the T-loop are depicted in Figure 2. Isomerization of U55 into the universally conserved Ψ55 favors the 3′-endo sugar pucker as in the A-form of RNA and enhances the stacking interaction with the trans Watson–Crick/Hoogsteen pair U54-A58 and the nearest-neighbor purine 57 [18]. The additional H-bound at N1 of Ψ55 involves the 5′-adjacent phosphate via a coordinated water molecule [19]. Likewise, methylation of the C5 atom of U54 into m5U and isomerization of U54 into Ψ54 followed by the N1-methylation of Ψ54 similarly favors the 3′-endo ribose puckering of the modified nucleotides and reinforces the hydrophobicity of the base. This increases base stacking with the neighboring bases G53 and Ψ55. Methyl groups also increase steric hindrance, block hydrogen bonds at Watson–Crick positions, and increase resistance against hydrolysis by nucleases. In thermophilic Bacteria and Archaea (growing at temperatures above 60 °C), the uridine-54 is almost invariably hypermodified into thiolated derivatives (m5s2U or m1s4Ψ) [20,21]. Remarkably, hypermodified m5s2U and m1s4Ψ at position 54 are isosteric, which implies that in both cases, the bulky highly polarizable thio group reinforces the 3′-endo ribose puckering of the nucleotide. This considerably enhances the stacking interaction with the nearest-neighboring nucleotides G53 and Ψ55 [22] and allows base pairing in an identical way with the trans-Hoogsteen nucleotide 58.
Methylation of the 2′-hydroxyl of the ribose favors the 3′-endo-conformation and changes the hydration sphere around the oxygen of the ribose [23]. Further, 2′-O-ribose methylation of cytidine at position 56 (Cm) was found in tRNAs of the crenarchaeon Pyrobaculum aerophilum [24] and also of N1-methyl-inosine at position 57 (m1Im) in tRNAs of Pyrobaculum islandicum and of Pyrodictium occultum. However, in this latter case information comes from analysis of bulk tRNA with no evidence for the presence of m1Im at position 57 exclusively [25]. 2′-O-ribose methylation of m5U at position 54 (m5Um) exists in certain tRNAs of Eukarya [26]. Residue Gm18 in the D-loop is found in tRNAs of all Domains of Life but are more frequently found in thermophilic Bacteria and hyperthermophilic Archaea [20]. This methylation is diagonally opposite to the 2′-O-methylation of C56 (Cm) that base pairs with the conserved neighboring G19.
Methylation of the N1 atom of adenosine 58 blocks a Watson–Crick position and favors the formation of an intraloop Watson–Crick/Hoogsteen pairing with m5U54 or Ψ/m1Ψ54. The electropositive character of this methylated adenosine also promotes ionic interaction with the negatively neighboring charged phosphates of the backbone. In Bacteria and Eukarya, the conserved purine 57 is more often G57 than A57, while in thermophilic and hyperthermophilic Archaea, a neutral N1-methylinosine (m1I) is often found instead. Such deaminated adenosine can establish H-bonds as a guanosine. Nucleotide in position 57 (G57 or derivatives of A57) is sandwiched between the base pairs G18-U55 and G19-C56 of the tRNA elbow (Figure 3). Notice, the presence of m1A58 and/or m1I57 depends much on the archaeon considered and the type of tRNA analyzed (especially the nature of nucleotide-59, see below). For example, most tRNAs from halophilic Archaea such as Haloferax volcanii display m1I57 and unmodified A58 [27,28], while tRNAs of thermophilic and hyperthermophilic Archaea often display m1A58 and/or m1I57 [29] and references therein.
In sum, the highly conserved nucleotide modifications of the T-loop and of G18 of the D-loop mainly serve to enhance the intrinsic stability of the tRNA elbow and also to enhance the affinity between the T- and D-loop, especially in the thermophilic organisms [30].

4. The Enzymatic Toolbox for Nucleotide Modifications at Positions 18 (D-Loop) and 54–58 (T-Loop) of tRNA

The elbow region of tRNAs of Bacteria and Eukarya contains characteristic conserved noncanonical nucleotides at positions 18, 54, 55, and 58, while in Archaea, additional conserved modifications also occur at positions 56 and 57 (Figure 1 and Figure 2). They are all catalyzed by specific enzymes during the tRNA maturation process (Figure 4 and Figure 5).

4.1. G18 Modification in the D-Loop

In tRNA species of most Gram-negative Bacteria the highly conserved guanosine-18 is converted to a 2′-O-methylated derivative (Gm). Gm18 is absent in Gram-positive Bacteria with a few exceptions such as the thermophilic Geobacillus sp (reviewed in [31]). This modification is accomplished by the dimeric S-adenosyl-L-methionine (SAM)-dependent methyltransferase (MTase) TrmH belonging to the SPOUT superfamily [32] (COG0566–Cluster of Orthologous Genes, see [33]). Two types of TrmH enzymes exist with different tRNA recognition specificities: one methylating all tRNA types, like TrmH of Thermus thermophilus, and the other one methylating only a specific subset of tRNAs like TrmH of Escherichia coli and Aquifex aeolicus [34,35,36], reviewed in [37].
In yeast, Gm18 formation is catalyzed by the dimeric TRM3 MTase also belonging to the SPOUT superfamily (COG0566). Yeast TRM3 exhibits a very long N-terminal extension of unknown function [38]. The human homologous enzyme is TARBP1 [39].
In Archaea, while Gm18 has been experimentally identified in tRNA of different thermophilic organisms [29], no stand-alone homolog of TrmH/TRM3 has yet been identified. Instead, an alternative pathway involving a C/D box guide has been suggested, as in the case of 2′-O-ribose methylation of C56 of the T-loop (see below) and also for other nucleotides in the anticodon loop and position 50 of the T-arm [40,41]. Methylation will then occur via a nucleoprotein complex involving the ad hoc C/D box guide RNA, two protein cofactors Nop5 and L7Ae, and a SAM-dependent fibrillarin-like MTase (FlpA, belonging to COG1889).
Figure 4. Enzymatic systems involved in post-transcriptional modifications of selected nucleotides of the D-and T-loop of tRNAs of various organisms. Acronyms of enzymes are boxed with various backgrounds according to the type of modification together with the COG (Cluster of Orthologous Genes) number. The Greek symbols indicate the oligomerization state of the functional enzymes. Only in a few cases, the enzyme corresponding to a given modification remains unknown (indicated by?). The 2′-O-methylation of m1I57 into m1Im57 is putative (see text). Acronyms for modification enzymes are those used in Modomics [10], except for those in human [39]. gRNA stands for guide RNA. A, B, and E refer, respectively, to Archaea, Bacteria, and Eukarya.
Figure 4. Enzymatic systems involved in post-transcriptional modifications of selected nucleotides of the D-and T-loop of tRNAs of various organisms. Acronyms of enzymes are boxed with various backgrounds according to the type of modification together with the COG (Cluster of Orthologous Genes) number. The Greek symbols indicate the oligomerization state of the functional enzymes. Only in a few cases, the enzyme corresponding to a given modification remains unknown (indicated by?). The 2′-O-methylation of m1I57 into m1Im57 is putative (see text). Acronyms for modification enzymes are those used in Modomics [10], except for those in human [39]. gRNA stands for guide RNA. A, B, and E refer, respectively, to Archaea, Bacteria, and Eukarya.
Genes 12 00140 g004

4.2. Modification of A58 and A57

N1-methyladenosine (m1A) is often present at position 58 of tRNAs of organisms belonging to all three Domains of Life. Its formation, role, and structural implications has been the subject of various reviews [42,43,44,45,46].
In Bacteria, like T. thermophilus, for which m1A58 is present in all sequenced tRNAs, the enzyme catalyzing this N1-adenine methylation is the homotetrameric TrmI [47,48]. This SAM-dependent enzyme belongs to the Rossman fold-like (RFM/class I) family of MTases (COG2519) and is site-specific for the N1-position of A58 [47]. The T-loop in combination with the aminoacyl-stem and the variable region are required for efficient methyl transfer [49].
In Eukarya, like yeast and human, a site-specific A58 TrmI-like ortholog exists but displays a different protein architecture. Here, the enzyme is a heterotetramer composed of two different kinds of subunits: a catalytic SAM-binding protein (TRM6/TRMT6 belonging to RFM/class I) and an evolutionary related RNA binding protein (TRM61/TRMT61A), both belonging to COG2519. As for bacterial TrmI, eukaryal N1-adenine MTase is site-specific for A58 and requires the T-loop in combination with the aminoacyl-stem [50,51,52,53], reviewed in [54].
In hyperthermophilic Archaea, like Pyrococcus furiosus or Pyrococcus abyssi, the TrmI ortholog works as a homotetramer. However, at variance with the bacterial enzyme, it involves intersubunit disulfide bridges that reinforce the tetrameric oligomerization [55,56]. This archaeal TrmI can work in vitro on a minisubstrate composed of only the T-branch [50]. Remarkably, it methylates the N1-atom of both neighboring adenosines 58 and 57 of the T-loop, with a preference for position A57. This region specificity depends on the presence of an adenosine at position 59 of tRNA. In absence of A59, archaeal TrmI behaves as an A57-specific MTase [57]. A model based on fluorescence measurements of 2-aminopurine placed at positions 57 or 58, showed that methylation is sequential, forming m1A57 before m1A58 [58]. In archaeal tRNAs, m1A57 never accumulates, i.e., the 6-amino group of adenine is always very efficiently deaminated by a new type of hydrolytic deaminase (still to be identified) leading to formation of 1-methylinosine (m1I57) [59,60]. The nucleotide 57 is sandwiched between G19-C56 and G18-U55 (see above). The still unknown m1A-dependent deaminase removes the positive electrostatic charge of the N1-atom of m1A (Figure 2), leading to the neutral m1I residue. Although it is difficult to evaluate the molecular effects of the latter modification, it is conceivable that m1I stabilizes stacking interactions better than m1A. On the contrary, the presence of a positively charged m1A58 has been demonstrated to be critical for the folding of the T-loop and proper functioning of certain tRNAs (discussed in [61,62]). In the majority of other bacterial and eukaryal tRNAs, nucleotide 57 is usually G or unmodified A. As for N6-methyladenosine (m6A) in mRNA, m1A in human tRNAs has been found to be occasionally demethylated by a specific enzyme [63], reviewed in [64].

4.3. Modification of C56

The presence of 2′-O-methylcytidine (Cm) at position 56 is a hallmark of archaeal tRNAs. It has never been found in any tRNA of Bacteria nor of Eukarya. As for the formation of Gm18 (see above), two different evolutionary unrelated enzymatic mechanisms exist to catalyze this ribose methylation. In most Archaea, Cm56 formation is achieved by the stand-alone SAM-dependent Trm56 belonging to COG1303 [65]. Little is known about tRNA recognition by this enzyme. Nevertheless it seems that besides the C56, the enzyme recognizes more than just the T-stem-loop, at least in vitro and with unmodified tRNA transcripts as substrates [65,66]. This dimeric MTase belongs to the same SPOUT superfamily as the G18-MTase (TrmH/TRM3/TARBP1) but belongs to a different COG [37,67]. In other Archaea, such as the hyperthermophilic Pyrobaculum aerophilum, Cm56 formation is achieved by a nucleoprotein complex composed of a fibrillarin-like MTase (FlpA, COG1889), two protein cofactors (Nop5 and L7Ae) and a specific small C/D guide RNA that allows the FlpA enzyme to find its target nucleotide [65], reviewed in [24]. This enzymatic pathway is the same as the one catalyzing formation of Gm18, except that distinct guide RNAs are used. Interestingly, a simplified heterodimer composed of P. abyssi FlpA and Nop5 can perform SAM-dependent 2′-O-methylation of RNA in vitro independently of L7Ae and C/D guide RNA. This stand-alone activity of dimeric FlpA-Nop5 may reflect the ancestral activity of the enzyme, before its recruitment into larger complexes [68].

4.4. Modification of U55

Pseudouridine (Ψ) at position 55 is the most universally conserved modified nucleotide in tRNA of all three phylogenetic Domains of Life. Nevertheless, as for several other modified nucleotides discussed in this work, different enzymatic strategies are used to catalyze this simple isomerization reaction in different organisms.
In Bacteria and Eukarya, stand-alone Ψ55-synthase TruB (in E. coli) and its ortholog PUS4/TRUB1 (in Saccharomyces cerevisiae/human) catalyze this modification [69,70,71]. These monomeric enzymes belong to COG0130, which also includes the close homolog Cbf5 (see below). Bacterial TruB in complex with its RNA substrate is probably the best characterized system involving a modification enzyme [72,73,74,75]. This bacterial TruB enzyme, as well as its eukaryal orthologs PUS4/TRUB1 require no cofactors and accept as substrate a simple 17 base oligonucleotide analog to the T-arm and containing the sequence 53-GUUYRANYC-61 [70,76]. This explains why these enzymes efficiently work on unmodified precursor tRNAs (including synthetic molecules) and on many cellular RNAs (other than tRNAs) in vitro and in vivo, as long as the target position in the RNA substrate remains accessible to the enzyme. Besides its catalytic function, bacterial TruB also acts as RNA chaperone [77].
In Eukarya and Archaea, but not in Bacteria, two additional enzymes are implicated in Ψ55 formation. The first usually functions in a ribonucleoprotein complex that utilizes a well-structured box H/ACA guide RNA to target the sequence to be modified, two or three protein cofactors (Nop10, Gar1, and Nhp2 in Eukarya and Nop10, Gar1, and L7Ae in Archaea) and the enzyme Cbf5 belonging to the same COG0130 as PUS4/TRUB1. In Eukarya, such H/ACA guided enzymatic system is used essentially for site-specific catalysis of Ψ formation in various cellular RNA targets, mostly rRNA, and not tRNA (reviewed in [78,79,80]). In Archaea, Cbf5 can function as a stand-alone Ψ-synthase for tRNA Ψ55 formation, at least in vitro [81,82,83,84]. This guide-independent tRNA specific activity of archaeal Cbf5 is enhanced in vitro by the proteins Nop10 and Gar1, probably by stabilizing the active conformation of Cbf5 [82,85,86,87]. It is worth mentioning that the Haloferax volcanii cbf5 deletion strain still contains tRNA carrying Ψ55, attesting that Cbf5 is not the main physiological Ψ55 synthase, at least in this archaeon [88]. This is another example of the probable ancestral activity retained in the present-day Cbf5 enzyme that was recruited into larger complexes later during evolution.
In Archaea, the genuine enzyme catalyzing Ψ55 formation in tRNA is the stand-alone Pus10 belonging to COG1258, phylogenetically unrelated to TruB/PUS4/TRUB1 (COG0130) [81,83]. This monomeric Ψ-synthase needs no protein cofactors and is present in a large variety of Archaea (mostly in Euryarchaeota and a few Crenarchaeota). It is also present in a few Eukarya (including human) but not in Bacteria [89,90]. In archaeal tRNAs the main target is U55, but depending on the archaeon considered, Pus10 behaves as a region-specific enzyme catalyzing Ψ formation at both positions 55 and 54 (see below).
In human, the situation is more complicated due to the coexistence of nuclear/cytoplasmic Ψ55-specific TRUB1 and of mitochondrial TRUB2 (not discussed here). While recombinant human PUS10 can generate Ψ55 in human cytoplasmic tRNAs in vitro [91], its main cellular function is to promote pseudouridylation of U54 in only a small subset of the total cytoplasmic tRNA population (see below), the formation of Ψ55 in the majority of the other human tRNAs being achieved in the nucleus by TRUB1.
Figure 5. Enzymatic systems involved in post-transcriptional modifications of U55 and U54. Same as legend to Figure 4.
Figure 5. Enzymatic systems involved in post-transcriptional modifications of U55 and U54. Same as legend to Figure 4.
Genes 12 00140 g005

4.5. Modification of U54

Methylation of the C5-atom of uridine at position 54 (m5U, also referred as ribothymidine rT), is the second most commonly found modified nucleotide in tRNAs of Bacteria and Eukarya. In Archaea, it is only found in a small subset of organisms belonging to Thermococcales and the sister lineage Nanoarchaeota.
In E. coli, m5U54 formation is catalyzed by the monomeric SAM-dependent TrmA enzyme of the RFM superfamily belonging to GOG2265 [92]. Its presence, but not its enzymatic activity is essential for E. coli viability [93,94]. This is probably explained by the chaperone activity of TrmA during tRNA maturation [95]. In yeast, the orthologous enzyme is TRM2 and in human TRMT2A [96,97].
In Bacteria, the use of SAM as a methyl donor by the TrmA enzyme is common for m5U54 formation in β-γ-and ε-Proteobacteria, while in Firmicutes, α-and δ-Proteobacteria as well as in Cyanobacteria and Deinococci, the enzyme responsible for the methylation of U54 is the flavoprotein TrmFO. This enzyme uses N5, N10-methylenetetrahydrofolate (CH2 = THF) in a more complex mechanism than the rather simple methyl transfer from SAM. Here, a methylene group is transferred first and a subsequent reduction step involving the flavin adenine dinucleotide hydroquinone (FADH-) is required to form the final methylated derivative [98,99], reviewed in [100]. The monomeric folate/FAD-dependent TrmFO enzyme belongs to COG1206. In Mollicutes, a group derived from Firmicutes and displaying extensive genome erosion (retrograde evolution), several SAM-dependent m5U MTases present in their common ancestor have been lost and replaced by folate-dependent paralogs [101,102]. These are clear examples of independent and mutually exclusive evolution of analogous enzymes catalyzing the formation of the same modification (m5U54) using totally unrelated mechanisms in different bacterial lineages.
In the hyperthermophilic Archaea of the phyla Thermococci (such as P. abyssi and P. furiosus) and Nanoarchaeota (as Nanoarchaeum equitans), m5U54 formation is catalyzed by TrmU54, a SAM-dependent MTase belonging to COG2265. Although this monomeric protein belongs to the same COG as TrmA/TRM2/TRMT2A, the protein sequence of archaeal TrmU54, as well as eukaryal TRM2, are more similar to the bacterial paralog RlmD (formerly, RumA/RumT) catalyzing m5U1939 formation in 23S rRNA [103]. One main difference is the presence of an iron–sulfur cluster in the archaeal TrmU54 [104]. In both cases, these two enzymes, TrmU54 and TRM2, arose from an ancestral horizontal transfer (endosymbiosis in the case of TRM2) followed by a shift of function from rRNA to tRNA modification [103,105].
All these MTases are site-specific and methylate exclusively the C5-atom of U54. As for TruB/PUS4 above, the minimal RNA structure for substrate recognition by TrmA, TRM2, and TrmU54 is a 5-base-pair stem, eventually prolonged with the amino acid arm, and a seven-membered loop as in a canonical T-loop of tRNA [50,66,70,75,106,107,108]. The same applies to tRNA recognition by bacterial TrmFO [109].
Depending on the organism, residue m5U54 can be further modified by either 2′-O-methylation or thiolation. In rabbit and human, 2′-O-methylated derivative of m5U54 (m5Um) was reported for tRNALys3 [26,110]. The corresponding MTase remains to be identified. In both thermophilic Bacteria and Archaea, m5U54 is further modified in m5s2U54 by a thiolation enzymatic machinery involving several proteins (TtuA, TtuB, TtuC, and cysteine desulfurases IscS or SufS in Bacteria, not yet known in Archaea) [111], reviewed in [112]. In T. thermophilus, the level of thiolation enzyme and, consequently, the degree of tRNA m5U thiolation increases with increasing cultivation temperatures [113]. The major determinants for 2-thiolation of m5U54 is the presence of the conserved C56 and m1A58 of the T-loop. Tertiary interactions between the T- and D-loops and nonconserved nucleotides in the T-loop are not important for the reaction [114]. The importance of m1A58 for m5s2U54 formation is supported by the fact that a Δ trmI strain of T. thermophilus loses its ability to grow at higher temperature range [47,113], reviewed in [20]. Moreover, the presence of m7G46 in the V-loop allows a better efficient thiolation reaction attesting for a clear interplay between the different modifications in tRNA [115].
In Archaea, other than Thermococci and Nanoarchaeota, and a few Eukarya, U54 is modified to Ψ. This isomerization reaction is catalyzed by the Ψ-synthase Pus10/PUS10 already mentioned above for catalyzing Ψ55 formation. However, depending on the archaeon, Pus10 can be either site-specific for U54 or region-specific for both U55 and U54. When tested in vitro, recombinant Pus10 of Methanocaldococcus jannaschii catalyzes formation of both Ψ55 and Ψ54 with the same efficiency, whereas recombinant Pus10 of P. furiosus is inefficient for catalyzing Ψ54 formation under the same experimental conditions [83]. When tested in vivo, recombinant M. jannaschii Pus10 also efficiently forms both Ψ55 and Ψ54 in tRNAs of a double (truB-trmA)-mutant E. coli strain. In contrast, recombinant P. furiosus Pus10 only forms Ψ55 in the same tRNAs of the E. coli double-mutant strain [116]. Such versatility of archaeal Pus10 is consistent with the fact that in naturally occurring tRNAs of P. furiosus an m5U54 is present, while in M. jannaschii, tRNAs m1Ψ54 is found. Thus, the restricted specificity of P. furiosus Pus10 for U55 avoids a competition with methylation of U54 by TrmA [83]. Recombinant human PUS10 can also catalyze Ψ55 and Ψ54 formation in vitro [117], attesting for adaptation of the enzyme specificity according to the cellular need [116]. Maximal Ψ54 synthase activity is observed for tRNA substrates with the sequence 53-GUUCAm1AAUC-61 along with a stable acceptor stem. As for the archaeal Pus10, the crystal structure of human PUS10 shows the presence of a conserved Ψ-synthase catalytic domain fused to a characteristic THUMP domain [118]. THUMP is an RNA-binding module that is also present in a few other tRNA modification enzymes such as thiouridine synthases and MTases [119,120,121,122,123,124].
In Archaea (mostly Euryarchaeota but also in a few Crenarchaeota), Ψ54 can be further methylated into m1Ψ54 by a SAM-dependent MTase (TrmY) belonging to COG1901 [125,126]. The structure of this dimeric enzyme shows that it belongs to the SPOUT superfamily [127]. In Vitro experiments with purified recombinant region-specific M. jannaschii Pus10 and TrmY demonstrates that TrmY (as Pus10 above) can produce m1Ψ in a 17-base fragment of a tRNA transcript corresponding to the T-arm of which U55 (possibly Ψ55) cannot be mutated [125]. In thermophilic Archaea, residue m1Ψ54 can be further thiolated to m1s4Ψ54 most probably by the same thiolation enzymatic machinery reported above for m5s2U54 formation in thermophilic Bacteria (involving proteins TtuA, TtuB, TtuC, and cysteine desulfurases not yet identified in Archaea). For example, in tRNAs of Ignicoccus hospitalis, a member of the Crenarchaeota, m1s4Ψ54, is found, whereas in tRNAs of Nanoarchaeum equitans, closely related to the Thermococcales, m5s2U54 is present [21]. These two Archaea are growing physically attached as obligate partners exchanging various metabolites under hyperthermophilic conditions [128]. Remarkably, the methyl and thio groups in m1s4Ψ and m5s2U are presented in the same spatial manner for the stabilization of the 3′ endo-anti conformation of nucleotide 54, showing that these Archaea use different but structurally equivalent modifications to stabilize their T-loops. This illustrates how different solutions have been adopted during evolution to fulfil the same function.
Finally, in the case of the hyperthermophilic crenarchaeon Sulfolobus acidocaldarius, s2U has been found at position 54 of several tRNAs. In the particular case of tRNAVal, a methylthio U-derivative was also found, but its chemical structure remains elusive [29]. Examination of the S. acidocaldarius genome reveals the presence of genes coding for TtuA and TtuB responsible for thiolation of U54 but no gene coding for TrmA (catalyzing m5U54), Pus10 (catalyzing Ψ54 in addition of Ψ55), and TrmY (catalyzing m1Ψ54) was found. One possibility is that ribose methylation of U54 occurs in only a few tRNA species, such as the initiator tRNA [129]. Indeed, 2′-O-methylated nucleotides are found at several positions of various S. acidocaldarius tRNAs and are known to be formed by the guide RNA-dependent FlpA machinery [130,131]. However, no guide RNA specific for U54 methylation has been reported up to now. Therefore, in absence of a clear-cut evidence for Um54 existence and of convincing information about the corresponding methylation system, this problem remains unsolved.

5. Temporal Order of Nucleotide Modifications and Their Interdependencies

After transcription, each tRNA primary transcript (pre-tRNA) is engaged in a complex multistep maturation during which 5′-leader and 3′-trailer are removed, introns are spliced, 5′-G (in the case of tRNAHis) and CCA-end are added (if not encoded in the genome), and a bunch of nucleotides become modified. In Eukarya, this process starts in the nucleus and terminates in the cytoplasm. Along this complex sequential maturation, both the structural conformation and the sequence of the pre-tRNAs are progressively transformed in fully mature and functional tRNA molecules with the canonical L-shaped structure.
During this process, enzymatic modifications introduced by a machinery utilizing guide RNAs that base pair with the pre-tRNA substrate and by enzymes requiring limited small substructures are expected to occur first, while enzymes requiring an intact tRNA architecture are expected to act later. Hence, the earliest expected modifications are Gm18 and Cm56 (in Archaea), which both depend on guide RNA, and Ψ55, Ψ54, m5U, m1Ψ54, which all depend on stand-alone enzymes working on minimal T-arm substructures. These are precisely the modifications that were experimentally shown to be the first ones to appear during pre-tRNA processing using either in vivo or in vitro assays [50,132,133,134]. Recently, experiments aiming at monitoring the time-resolved introduction of modifications in yeast tRNAPhe in vitro not only confirm these observations but also demonstrate the existence of a robust modification circuit (interdependency) starting with the introduction of Ψ55, then m5U54, and finally m1A58 [135]. Since the presence of N1-methyl group on A58 is required for efficient enzymatic transfer of sulfur on the C2-atom of m5U54 [113], the biogenesis of m5s2U54 by TtuA/TtuB/TtuC at elevated temperature in thermophilic Bacteria and possibly of m1s4Ψ in hyperthermophilic Archaea should occur later during tRNA maturation. These observations are in complete agreement with the theory of interdependency between the various enzymatic modification systems acting on nucleotides in the tRNA elbow, in which Ψ55-synthase TruB/PUS4/TRUB1 (and probably Pus10) appears to be the conductor, besides acting itself as a chaperone protein [136,137,138,139]. Interdependency between different nucleotide modifications and their modification pathways has been found to be particularly important for modifications in the anticodon-stem loop [60,140,141].

6. tRNA T-Loop Modification Enzymes Can Modify Other T-Loop Containing RNAs

Conserved base-pairings between the T-and D-loop are intrinsic properties of the tRNA itself: they are present in synthetic tRNA transcripts lacking any modified nucleotide, yet in slightly more relaxed conformation [142,143]. Genuine T-loop architecture also exists in local regions of the ribosomal RNAs [144], at the 3′ end of the plant viral TYMV RNA [145] and in bacterial tmRNA [146]. In these two last examples, TrmA/TRM2/TRMT2A was shown to catalyze m5U formation at the position equivalent to U54 [147,148]. The main role of post-transcriptional modifications of the various conserved nucleotides of the T-loop is therefore to fine-tune the dynamics (rigidity/flexibility) of the self-formation of this important part of the tRNA molecule and to facilitate its interaction with the D-loop. As illustrated in Figure 3, the various conserved modified nucleotides (Gm18, m5(s2)U/m1s4Ψ54, Ψ55, Cm56, and m1I57/m1A58) are very close together in the L-shaped conformation of tRNA.

7. Conclusions and Future Prospects: Unraveling the Evolutionary Origin of tRNA Modification Machinery

The importance of a few invariant and semi-invariant nucleotides for self-forming 3D-conformation of the T-loop of tRNAs is now well documented. This unique structural domain is characterized by a trans Watson–Crick/Hoogsteen pair between U54 and A58, which, in turn, allows U54, U55, the bulged pyrimidine 60, and nucleotide 59 to further interact specifically with nucleotides of the D-loop (Figure 3) [149,150]. Since transcription proceeds in the 5′ to 3′ direction, this T-loop/D-loop interaction cannot take place before almost the full sequence of the tRNA gene is transcribed. In addition, among the many modified nucleotides found in fully matured tRNA molecules, Gm18, Ψ55, m5U54/m1Ψ, m1I57/m1A58, and Cm56 are likely the first to occur in the maturation pathway, the most universally conserved Ψ55 being probably the very earliest one. These combinatorial networks of base pairings and nucleotide modifications, together with other tertiary interactions and base modifications within the entire core of the tRNA molecule, permit each individual fully matured isoacceptor tRNA species to adopt a stable still flexible L-shaped conformation [151]. In this final functional conformation, the distance between the anticodon and the CCA-3′-terminus at the opposite is around 80 Å. This property is essential for each aminoacyl-tRNA to accommodate in the decoding center of the small ribosomal subunit and then in the peptidyl-transferase center of the large subunit.
Because of the highly conserved characters of the modified nucleotides found in the T-loop and around G18-G19 among tRNAs of organisms belonging to the three Domains of Life, the simplest ones like Ψ, Cm, m5U, and m1A might have been present in the prebiotic soup [152,153,154]. The corresponding modification enzymes might have appeared later, yet very early in evolution, at least at the root of the common ancestral lineages of Bacteria/Archaea [155,156]. If the present-day enzymes would have been vertically inherited from these earliest forms, their sequences should retain substantial similarities among orthologous enzymes present in majority if not all organisms of the Tree of Life. On the contrary, and this is the main message of the present review, almost all the enzymes involved in the modification of nucleotides at positions 18, 54, 55, 56, and 57/58 differ much from one group of organisms to another, U54 being the most diversely modified one. These nonhomologous enzymes (also named “analogous enzymes” [157]) belong to different COGs, use sometimes different cofactors (such as SAM or CH2 = THF) and mechanisms (standalone or guided-RNA machineries), and recognize different portions of the tRNA molecule. They can be site-, region-, or even dual-specific (for tRNA and other types of RNA). Evidently, emergence and evolution of a given enzymatic activity is more complex than just by gene duplication and neofunctionalization followed by eventual elimination of the ancestral gene. Indeed, a gene present in a given organism can be lost and replaced by a nonorthologous paralog coding for a protein with similar or identical function (see, e.g., [102]). In addition, any time along the cellular evolution, a new enzymatic activity, or a new enzymatic mechanism can be inherited from a lateral gene transfer from an organism living in endosymbiosis or in symbiosis within the same environment. Lateral gene transfer between Bacteria and Archaea or between Prokaryotae and Eukarya are frequent [158,159]. In all cases, the driving force is to optimize an ad hoc function required by the novel needs of the cell and according to the evolving cellular physiology and/or of growth conditions. In the particular case of tRNA modification, the need is to constantly adapt and preserve the stability of the overall architecture and consequently the functionality of the tRNA molecules according to the growing conditions as well as the physiological changes that occur all along the complex evolution process.
To conclude, this review shows the complexity of tRNA modification enzymes evolution, which makes genome annotation a difficult task [160,161] and prompts to start studies on the emergence and evolution of enzymes catalyzing the formation of a given modified nucleotide at a given conserved position of the tRNA molecule.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank Eric Westhof for advice, for helping at generating Figure 3C, and for critical reading of the manuscript; André Feller for helping at generating the figures; and Dany Van Elder for helping at managing the references. We acknowledge the referees for improvements to the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Westhof, E.; Auffinger, P. tRNA stucture. In Citable Reviews in the Life Sciences; John Wiley & Sons Ltd.: Chichester, UK, 2001; pp. 1–11. [Google Scholar] [CrossRef]
  2. 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]
  3. Lorenz, C.; Lünse, C.E.; Mörl, M. tRNA Modifications: Impact on Structure and Thermal Adaptation. Biomolecules 2017, 7, 35. [Google Scholar] [CrossRef] [Green Version]
  4. Väre, V.Y.; 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] [Green Version]
  5. Steinberg, S.; Leclerc, F.; Cedergren, R. Structural rules and conformational compensations in the tRNA L-form. J. Mol. Biol. 1997, 266, 269–282. [Google Scholar] [CrossRef]
  6. Watanabe, Y.; Suematsu, T.; Ohtsuki, T. Losing the stem-loop structure from metazoan mitochondrial tRNAs and co-evolution of interacting factors. Front. Genet. 2014, 5, 109. [Google Scholar] [CrossRef] [Green Version]
  7. Dick, T.P.; Schamel, W.A. Molecular evolution of transfer RNA from two precursor hairpins: Implications for the origin of protein synthesis. J. Mol. Evol. 1995, 41, 1–9. [Google Scholar] [CrossRef]
  8. Ardell, D.H. Computational analysis of tRNA identity. FEBS Lett. 2010, 584, 325–333. [Google Scholar] [CrossRef] [Green Version]
  9. Jühling, F.; Mörl, M.; Hartmann, R.K.; Sprinzl, M.; Stadler, P.F.; Pütz, J. tRNAdb 2009: Compilation of tRNA sequences and tRNA genes. Nucleic Acids Res. 2009, 37, D159–D162. [Google Scholar] [CrossRef] [Green Version]
  10. Boccaletto, P.; Machnicka, M.A.; Purta, E.; Piatkowski, P.; Baginski, B.; Wirecki, T.K.; de Crecy-Lagard, V.; Ross, R.; Limbach, P.A.; Kotter, A.; et al. MODOMICS: A database of RNA modification pathways. 2017 update. Nucleic Acids Res. 2018, 46, D303–D307. [Google Scholar] [CrossRef]
  11. Westhof, E.; Dumas, P.; Moras, D. Crystallographic refinement of yeast aspartic acid transfer RNA. J. Mol. Biol. 1985, 184, 119–145. [Google Scholar] [CrossRef]
  12. Romby, P.; Carbon, P.; Westhof, E.; Ehresmann, C.; Ebel, J.P.; Ehresmann, B.; Giegé, R. Importance of conserved residues for the conformation of the T-loop in tRNAs. J. Biomol. Struct. Dyn. 1987, 5, 669–687. [Google Scholar] [CrossRef]
  13. Shi, H.; Moore, P.B. The crystal structure of yeast phenylalanine tRNA at 1.93 A resolution: A classic structure revisited. RNA 2000, 6, 1091–1105. [Google Scholar] [CrossRef] [Green Version]
  14. 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]
  15. Quigley, G.J.; Rich, A. Structural domains of transfer RNA molecules. Science 1976, 194, 796–806. [Google Scholar] [CrossRef]
  16. Levitt, M. Detailed molecular model for transfer ribonucleic acid. Nature 1969, 224, 759–763. [Google Scholar] [CrossRef]
  17. Oliva, R.; Cavallo, L.; Tramontano, A. Accurate energies of hydrogen bonded nucleic acid base pairs and triplets in tRNA tertiary interactions. Nucleic Acids Res. 2006, 34, 865–879. [Google Scholar] [CrossRef] [Green Version]
  18. Davis, D.R. Stabilization of RNA stacking by pseudouridine. Nucleic Acids Res. 1995, 23, 5020–5026. [Google Scholar] [CrossRef]
  19. Westhof, E.; Dumas, P.; Moras, D. Hydration of transfer RNA molecules: A crystallographic study. Biochimie 1988, 70, 145–165. [Google Scholar] [CrossRef]
  20. Hori, H.; Kawamura, T.; Awai, T.; Ochi, A.; Yamagami, R.; Tomikawa, C.; Hirata, A. Transfer RNA Modification Enzymes from Thermophiles and Their Modified Nucleosides in tRNA. Microorganisms 2018, 6, 110. [Google Scholar] [CrossRef] [Green Version]
  21. Rose, S.; Auxilien, S.; Havelund, J.F.; Kirpekar, F.; Huber, H.; Grosjean, H.; Douthwaite, S. The hyperthermophilic partners Nanoarchaeum and Ignicoccus stabilize their tRNA T-loops via different but structurally equivalent modifications. Nucleic Acids Res. 2020, 48, 6906–6918. [Google Scholar] [CrossRef]
  22. Shigi, N. Biosynthesis and functions of sulfur modifications in tRNA. Front. Genet. 2014, 5, 67. [Google Scholar] [CrossRef] [Green Version]
  23. 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]
  24. Clouet-d’Orval, B.; Gaspin, C.; Mougin, A. Two different mechanisms for tRNA ribose methylation in Archaea: A short survey. Biochimie 2005, 87, 889–895. [Google Scholar] [CrossRef]
  25. Edmonds, C.G.; Crain, P.F.; Gupta, R.; Hashizume, T.; Hocart, C.H.; Kowalak, J.A.; Pomerantz, S.C.; Stetter, K.O.; McCloskey, J.A. Posttranscriptional modification of tRNA in thermophilic archaea (Archaebacteria). J. Bacteriol 1991, 173, 3138–3148. [Google Scholar] [CrossRef] [Green Version]
  26. Keller, P.; Freund, I.; Marchand, V.; Bec, G.; Huang, R.; Motorin, Y.; Eigenbrod, T.; Dalpke, A.; Helm, M. Double methylation of tRNA-U54 to 2′-O-methylthymidine (Tm) synergistically decreases immune response by Toll-like receptor 7. Nucleic Acids Res. 2018, 46, 9764–9775. [Google Scholar] [CrossRef] [Green Version]
  27. Gupta, R. Halobacterium volcanii tRNAs. Identification of 41 tRNAs covering all amino acids, and the sequences of 33 class I tRNAs. J. Biol. Chem. 1984, 259, 9461–9471. [Google Scholar] [CrossRef]
  28. Grosjean, H.; Gaspin, C.; Marck, C.; Decatur, W.A.; de Crécy-Lagard, V. RNomics and Modomics in the halophilic archaea Haloferax volcanii: Identification of RNA modification genes. BMC Genom. 2008, 9, 470. [Google Scholar] [CrossRef] [Green Version]
  29. Wolff, P.; Villette, C.; Zumsteg, J.; Heintz, D.; Antoine, L.; Chane-Woon-Ming, B.; Droogmans, L.; Grosjean, H.; Westhof, E. Comparative patterns of modified nucleotides in individual tRNA species from a mesophilic and two thermophilic archaea. RNA 2020, 26, 1957–1975. [Google Scholar] [CrossRef]
  30. 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] [Green Version]
  31. Ayadi, L.; Galvanin, A.; Pichot, F.; Marchand, V.; Motorin, Y. RNA ribose methylation (2′-O-methylation): Occurrence, biosynthesis and biological functions. Biochim. Biophys. Acta Gene. Regul. Mech. 2019, 1862, 253–269. [Google Scholar] [CrossRef]
  32. Nureki, O.; Watanabe, K.; Fukai, S.; Ishii, R.; Endo, Y.; Hori, H.; Yokoyama, S. Deep knot structure for construction of active site and cofactor binding site of tRNA modification enzyme. Structure 2004, 12, 593–602. [Google Scholar] [CrossRef] [Green Version]
  33. Galperin, M.Y.; Wolf, Y.I.; Makarova, K.S.; Vera Alvarez, R.; Landsman, D.; Koonin, E.V. COG database update: Focus on microbial diversity, model organisms, and widespread pathogens. Nucleic Acids Res. 2020. [Google Scholar] [CrossRef]
  34. Hori, H.; Yamazaki, N.; Matsumoto, T.; Watanabe, Y.; Ueda, T.; Nishikawa, K.; Kumagai, I.; Watanabe, K. Substrate recognition of tRNA (Guanosine-2′-)-methyltransferase from Thermus thermophilus HB27. J. Biol. Chem. 1998, 273, 25721–25727. [Google Scholar] [CrossRef] [Green Version]
  35. Persson, B.C.; Jäger, G.; Gustafsson, C. The spoU gene of Escherichia coli, the fourth gene of the spoT operon, is essential for tRNA (Gm18) 2′-O-methyltransferase activity. Nucleic Acids Res. 1997, 25, 4093–4097. [Google Scholar] [CrossRef] [Green Version]
  36. Hori, H.; Kubota, S.; Watanabe, K.; Kim, J.M.; Ogasawara, T.; Sawasaki, T.; Endo, Y. Aquifex aeolicus tRNA (Gm18) methyltransferase has unique substrate specificity. TRNA recognition mechanism of the enzyme. J. Biol. Chem. 2003, 278, 25081–25090. [Google Scholar] [CrossRef] [Green Version]
  37. Hori, H. Transfer RNA methyltransferases with a SpoU-TrmD (SPOUT) fold and their modified nucleosides in tRNA. Biomolecules 2017, 7, 23. [Google Scholar] [CrossRef] [Green Version]
  38. Cavaillé, J.; Chetouani, F.; Bachellerie, J.P. The yeast Saccharomyces cerevisiae YDL112w ORF encodes the putative 2′-O-ribose methyltransferase catalyzing the formation of Gm18 in tRNAs. RNA 1999, 5, 66–81. [Google Scholar] [CrossRef] [Green Version]
  39. De Crécy-Lagard, V.; Boccaletto, P.; Mangleburg, C.G.; Sharma, P.; Lowe, T.M.; Leidel, S.A.; Bujnicki, J.M. Matching tRNA modifications in humans to their known and predicted enzymes. Nucleic Acids Res. 2019, 47, 2143–2159. [Google Scholar] [CrossRef]
  40. Omer, A.D.; Lowe, T.M.; Russell, A.G.; Ebhardt, H.; Eddy, S.R.; Dennis, P.P. Homologs of small nucleolar RNAs in Archaea. Science 2000, 288, 517–522. [Google Scholar] [CrossRef] [Green Version]
  41. Lui, L.M.; Uzilov, A.V.; Bernick, D.L.; Corredor, A.; Lowe, T.M.; Dennis, P.P. Methylation guide RNA evolution in archaea: Structure, function and genomic organization of 110 C/D box sRNA families across six Pyrobaculum species. Nucleic Acids Res. 2018, 46, 5678–5691. [Google Scholar] [CrossRef]
  42. Anderson, J.; Droogmans, L. Biosynthesis and function of 1-methyladenosine in transfer RNA. In Fine-Tuning of RNA Functions by Modification and Editing; Grosjean, H., Ed.; Springer: Heidelberg/Berlin, Germany, 2005; pp. 121–139. [Google Scholar]
  43. Swinehart, W.E.; Jackman, J.E. Diversity in mechanism and function of tRNA methyltransferases. RNA Biol. 2015, 12, 398–411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Dégut, C.; Ponchon, L.; Folly-Klan, M.; Barraud, P.; Tisné, C. The m1A(58) modification in eubacterial tRNA: An overview of tRNA recognition and mechanism of catalysis by TrmI. Biophys. Chem. 2016, 210, 27–34. [Google Scholar] [CrossRef] [PubMed]
  45. Oerum, S.; Dégut, C.; Barraud, P.; Tisné, C. m1A Post-Transcriptional Modification in tRNAs. Biomolecules 2017, 7, 20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Shima, H.; Igarashi, K. N 1-methyladenosine (m1A) RNA modification: The key to ribosome control. J. Biochem. 2020, 167, 535–539. [Google Scholar] [CrossRef]
  47. 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] [Green Version]
  48. Barraud, P.; Golinelli-Pimpaneau, B.; Atmanene, C.; Sanglier, S.; Van Dorsselaer, A.; Droogmans, L.; Dardel, F.; Tisné, C. Crystal structure of Thermus thermophilus tRNA m1A58 methyltransferase and biophysical characterization of its interaction with tRNA. J. Mol. Biol. 2008, 377, 535–550. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Takuma, H.; Ushio, N.; Minoji, M.; Kazayama, A.; Shigi, N.; Hirata, A.; Tomikawa, C.; Ochi, A.; Hori, H. Substrate tRNA recognition mechanism of eubacterial tRNA (m1A58) methyltransferase (TrmI). J. Biol. Chem. 2015, 290, 5912–5925. [Google Scholar] [CrossRef] [Green Version]
  50. Grosjean, H.; Edqvist, J.; Stråby, K.B.; Giegé, R. Enzymatic formation of modified nucleosides in tRNA: Dependence on tRNA architecture. J. Mol. Biol. 1996, 255, 67–85. [Google Scholar] [CrossRef]
  51. Anderson, J.; Phan, L.; Cuesta, R.; Carlson, B.A.; Pak, M.; Asano, K.; Björk, G.R.; Tamame, M.; Hinnebusch, A.G. The essential Gcd10p-Gcd14p nuclear complex is required for 1-methyladenosine modification and maturation of initiator methionyl-tRNA. Genes Dev. 1998, 12, 3650–3662. [Google Scholar] [CrossRef] [Green Version]
  52. Ozanick, S.; Krecic, A.; Andersland, J.; Anderson, J.T. The bipartite structure of the tRNA m1A58 methyltransferase from S. cerevisiae is conserved in humans. RNA 2005, 11, 1281–1290. [Google Scholar] [CrossRef] [Green Version]
  53. Wang, M.; Zhu, Y.; Wang, C.; Fan, X.; Jiang, X.; Ebrahimi, M.; Qiao, Z.; Niu, L.; Teng, M.; Li, X. Crystal structure of the two-subunit tRNA m(1)A58 methyltransferase TRM6-TRM61 from Saccharomyces cerevisiae. Sci. Rep. 2016, 6, 32562. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Guy, M.P.; Phizicky, E.M. Two-subunit enzymes involved in eukaryotic post-transcriptional tRNA modification. RNA Biol. 2014, 11, 1608–1618. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Roovers, M.; Wouters, J.; Bujnicki, J.M.; Tricot, C.; Stalon, V.; Grosjean, H.; Droogmans, L. A primordial RNA modification enzyme: The case of tRNA (m1A) methyltransferase. Nucleic Acids Res. 2004, 32, 465–476. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Guelorget, A.; Barraud, P.; Tisné, C.; Golinelli-Pimpaneau, B. Structural comparison of tRNA m(1)A58 methyltransferases revealed different molecular strategies to maintain their oligomeric architecture under extreme conditions. BMC Struct. Biol. 2011, 11, 48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Guelorget, A.; Roovers, M.; Guérineau, V.; Barbey, C.; Li, X.; Golinelli-Pimpaneau, B. Insights into the hyperthermostability and unusual region-specificity of archaeal Pyrococcus abyssi tRNA m1A57/58 methyltransferase. Nucleic Acids Res. 2010, 38, 6206–6218. [Google Scholar] [CrossRef] [Green Version]
  58. Hamdane, D.; Guelorget, A.; Guérineau, V.; Golinelli-Pimpaneau, B. Dynamics of RNA modification by a multi-site-specific tRNA methyltransferase. Nucleic Acids Res. 2014, 42, 11697–11706. [Google Scholar] [CrossRef] [Green Version]
  59. Grosjean, H.; Constantinesco, F.; Foiret, D.; Benachenhou, N. A novel enzymatic pathway leading to 1-methylinosine modification in Haloferax volcanii tRNA. Nucleic Acids Res. 1995, 23, 4312–4319. [Google Scholar] [CrossRef] [Green Version]
  60. Dixit, S.; Henderson, J.C.; Alfonzo, J.D. Multi-Substrate Specificity and the Evolutionary Basis for Interdependence in tRNA Editing and Methylation Enzymes. Front. Genet. 2019, 10, 104. [Google Scholar] [CrossRef]
  61. Agris, P.F. The importance of being modified: Roles of modified nucleosides and Mg2+ in RNA structure and function. Prog. Nucleic Acid Res. Mol. Biol. 1996, 53, 79–129. [Google Scholar] [CrossRef]
  62. Auxilien, S.; Keith, G.; Le Grice, S.F.; Darlix, J.L. Role of post-transcriptional modifications of primer tRNALys,3 in the fidelity and efficacy of plus strand DNA transfer during HIV-1 reverse transcription. J. Biol. Chem. 1999, 274, 4412–4420. [Google Scholar] [CrossRef] [Green Version]
  63. Liu, F.; Clark, W.; Luo, G.; Wang, X.; Fu, Y.; Wei, J.; Wang, X.; Hao, Z.; Dai, Q.; Zheng, G.; et al. ALKBH1-Mediated tRNA Demethylation Regulates Translation. Cell 2016, 167, 816–828.e16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Zhang, C.; Jia, G. Reversible RNA Modification N(1)-methyladenosine (m(1)A) in mRNA and tRNA. Genom. Proteom. Bioinform. 2018, 16, 155–161. [Google Scholar] [CrossRef] [PubMed]
  65. Renalier, M.H.; Joseph, N.; Gaspin, C.; Thebault, P.; Mougin, A. The Cm56 tRNA modification in archaea is catalyzed either by a specific 2′-O-methylase, or a C/D sRNP. RNA 2005, 11, 1051–1063. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Constantinesco, F.; Motorin, Y.; Grosjean, H. Transfer RNA modification enzymes from Pyrococcus furiosus: Detection of the enzymatic activities in vitro. Nucleic Acids Res. 1999, 27, 1308–1315. [Google Scholar] [CrossRef]
  67. Tkaczuk, K.L.; Dunin-Horkawicz, S.; Purta, E.; Bujnicki, J.M. Structural and evolutionary bioinformatics of the SPOUT superfamily of methyltransferases. BMC Bioinform. 2007, 8, 73. [Google Scholar] [CrossRef] [Green Version]
  68. Tomkuvienė, M.; Ličytė, J.; Olendraitė, I.; Liutkevičiūtė, Z.; Clouet-d’Orval, B.; Klimašauskas, S. Archaeal fibrillarin-Nop5 heterodimer 2′-O-methylates RNA independently of the C/D guide RNP particle. RNA 2017, 23, 1329–1337. [Google Scholar] [CrossRef] [PubMed]
  69. Nurse, K.; Wrzesinski, J.; Bakin, A.; Lane, B.G.; Ofengand, J. Purification, cloning, and properties of the tRNA psi 55 synthase from Escherichia coli. RNA 1995, 1, 102–112. [Google Scholar]
  70. Becker, H.F.; Motorin, Y.; Planta, R.J.; Grosjean, H. The yeast gene YNL292w encodes a pseudouridine synthase (Pus4) catalyzing the formation of psi55 in both mitochondrial and cytoplasmic tRNAs. Nucleic Acids Res. 1997, 25, 4493–4499. [Google Scholar] [CrossRef]
  71. Zucchini, C.; Strippoli, P.; Biolchi, A.; Solmi, R.; Lenzi, L.; D’Addabbo, P.; Carinci, P.; Valvassori, L. The human TruB family of pseudouridine synthase genes, including the Dyskeratosis Congenita 1 gene and the novel member TRUB1. Int. J. Mol. Med. 2003, 11, 697–704. [Google Scholar] [CrossRef]
  72. Pan, H.; Agarwalla, S.; Moustakas, D.T.; Finer-Moore, J.; Stroud, R.M. Structure of tRNA pseudouridine synthase TruB and its RNA complex: RNA recognition through a combination of rigid docking and induced fit. Proc. Natl. Acad. Sci. USA 2003, 100, 12648–12653. [Google Scholar] [CrossRef] [Green Version]
  73. Phannachet, K.; Elias, Y.; Huang, R.H. Dissecting the roles of a strictly conserved tyrosine in substrate recognition and catalysis by pseudouridine 55 synthase. Biochemistry 2005, 44, 15488–15494. [Google Scholar] [CrossRef] [PubMed]
  74. Hoang, C.; Hamilton, C.S.; Mueller, E.G.; Ferré-D’Amaré, A.R. Precursor complex structure of pseudouridine synthase TruB suggests coupling of active site perturbations to an RNA-sequestering peripheral protein domain. Protein Sci. 2005, 14, 2201–2206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Hur, S.; Stroud, R.M.; Finer-Moore, J. Substrate recognition by RNA 5-methyluridine methyltransferases and pseudouridine synthases: A structural perspective. J. Biol. Chem. 2006, 281, 38969–38973. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Gu, X.; Yu, M.; Ivanetich, K.M.; Santi, D.V. Molecular recognition of tRNA by tRNA pseudouridine 55 synthase. Biochemistry 1998, 37, 339–343. [Google Scholar] [CrossRef] [PubMed]
  77. 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] [Green Version]
  78. Hamma, T.; Ferré-D’Amaré, A.R. The box H/ACA ribonucleoprotein complex: Interplay of RNA and protein structures in post-transcriptional RNA modification. J. Biol. Chem. 2010, 285, 805–809. [Google Scholar] [CrossRef] [Green Version]
  79. Yu, Y.T.; Meier, U.T. RNA-guided isomerization of uridine to pseudouridine—Pseudouridylation. RNA Biol. 2014, 11, 1483–1494. [Google Scholar] [CrossRef] [Green Version]
  80. Spenkuch, F.; Motorin, Y.; Helm, M. Pseudouridine: Still mysterious, but never a fake (uridine)! RNA Biol. 2014, 11, 1540–1554. [Google Scholar] [CrossRef] [Green Version]
  81. Roovers, M.; Hale, C.; Tricot, C.; Terns, M.P.; Terns, R.M.; Grosjean, H.; Droogmans, L. Formation of the conserved pseudouridine at position 55 in archaeal tRNA. Nucleic Acids Res. 2006, 34, 4293–4301. [Google Scholar] [CrossRef]
  82. Gurha, P.; Joardar, A.; Chaurasia, P.; Gupta, R. Differential roles of archaeal box H/ACA proteins in guide RNA-dependent and independent pseudouridine formation. RNA Biol. 2007, 4, 101–109. [Google Scholar] [CrossRef] [Green Version]
  83. Gurha, P.; Gupta, R. Archaeal Pus10 proteins can produce both pseudouridine 54 and 55 in tRNA. RNA 2008, 14, 2521–2527. [Google Scholar] [CrossRef] [Green Version]
  84. Tillault, A.S.; Fourmann, J.B.; Loegler, C.; Wieden, H.J.; Kothe, U.; Charpentier, B. Contribution of two conserved histidines to the dual activity of archaeal RNA guide-dependent and -independent pseudouridine synthase Cbf5. RNA 2015, 21, 1233–1239. [Google Scholar] [CrossRef] [Green Version]
  85. Muller, S.; Fourmann, J.B.; Loegler, C.; Charpentier, B.; Branlant, C. Identification of determinants in the protein partners aCBF5 and aNOP10 necessary for the tRNA:Psi55-synthase and RNA-guided RNA:Psi-synthase activities. Nucleic Acids Res. 2007, 35, 5610–5624. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Fujikane, R.; Behm-Ansmant, I.; Tillault, A.S.; Loegler, C.; Igel-Bourguignon, V.; Marguet, E.; Forterre, P.; Branlant, C.; Motorin, Y.; Charpentier, B. Contribution of protein Gar1 to the RNA-guided and RNA-independent rRNA:Ψ-synthase activities of the archaeal Cbf5 protein. Sci. Rep. 2018, 8, 13815. [Google Scholar] [CrossRef] [PubMed]
  87. Kamalampeta, R.; Kothe, U. Archaeal proteins Nop10 and Gar1 increase the catalytic activity of Cbf5 in pseudouridylating tRNA. Sci. Rep. 2012, 2, 663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Blaby, I.K.; Majumder, M.; Chatterjee, K.; Jana, S.; Grosjean, H.; de Crécy-Lagard, V.; Gupta, R. Pseudouridine formation in archaeal RNAs: The case of Haloferax volcanii. RNA 2011, 17, 1367–1380. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Watanabe, Y.; Gray, M.W. Evolutionary appearance of genes encoding proteins associated with box H/ACA snoRNAs: cbf5p in Euglena gracilis, an early diverging eukaryote, and candidate Gar1p and Nop10p homologs in archaebacteria. Nucleic Acids Res. 2000, 28, 2342–2352. [Google Scholar] [CrossRef] [Green Version]
  90. Fitzek, E.; Joardar, A.; Gupta, R.; Geisler, M. Evolution of Eukaryal and Archaeal Pseudouridine Synthase Pus10. J. Mol. Evol. 2018, 86, 77–89. [Google Scholar] [CrossRef]
  91. Mukhopadhyay, S.; Deogharia, M.; Gupta, R. Mammalian nuclear TRUB1, mitochondrial TRUB2 and cytoplasmic PUS10 produce conserved pseudouridine 55 in different sets of tRNA. RNA 2020. [Google Scholar] [CrossRef]
  92. Ny, T.; Björk, G.R. Cloning and restriction mapping of the trmA gene coding for transfer ribonucleic acid (5-methyluridine)-methyltransferase in Escherichia coli K-12. J. Bacteriol. 1980, 142, 371–379. [Google Scholar] [CrossRef] [Green Version]
  93. Persson, B.C.; Gustafsson, C.; Berg, D.E.; Björk, G.R. The gene for a tRNA modifying enzyme, m5U54-methyltransferase, is essential for viability in Escherichia coli. Proc. Natl. Acad. Sci. USA 1992, 89, 3995–3998. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Kealey, J.T.; Gu, X.; Santi, D.V. Enzymatic mechanism of tRNA (m5U54)methyltransferase. Biochimie 1994, 76, 1133–1142. [Google Scholar] [CrossRef]
  95. 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]
  96. Nordlund, M.E.; Johansson, J.O.; von Pawel-Rammingen, U.; Byström, A.S. Identification of the TRM2 gene encoding the tRNA(m5U54)methyltransferase of Saccharomyces cerevisiae. RNA 2000, 6, 844–860. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Carter, J.M.; Emmett, W.; Mozos, I.R.; Kotter, A.; Helm, M.; Ule, J.; Hussain, S. FICC-Seq: A method for enzyme-specified profiling of methyl-5-uridine in cellular RNA. Nucleic Acids Res. 2019, 47, e113. [Google Scholar] [CrossRef]
  98. Urbonavicius, J.; Brochier-Armanet, C.; Skouloubris, S.; Myllykallio, H.; Grosjean, H. In Vitro detection of the enzymatic activity of folate-dependent tRNA (Uracil-54,-C5)-methyltransferase: Evolutionary implications. Methods Enzymol. 2007, 425, 103–119. [Google Scholar] [CrossRef]
  99. Urbonavicius, J.; Skouloubris, S.; Myllykallio, H.; Grosjean, H. Identification of a novel gene encoding a flavin-dependent tRNA:m5U methyltransferase in bacteria—Evolutionary implications. Nucleic Acids Res. 2005, 33, 3955–3964. [Google Scholar] [CrossRef]
  100. Hamdane, D.; Grosjean, H.; Fontecave, M. Flavin-Dependent Methylation of RNAs: Complex Chemistry for a Simple Modification. J. Mol. Biol. 2016, 428, 4867–4881. [Google Scholar] [CrossRef]
  101. Grosjean, H.; Breton, M.; Sirand-Pugnet, P.; Tardy, F.; Thiaucourt, F.; Citti, C.; Barré, A.; Yoshizawa, S.; Fourmy, D.; de Crécy-Lagard, V.; et al. Predicting the minimal translation apparatus: Lessons from the reductive evolution of mollicutes. PLoS Genet. 2014, 10, e1004363. [Google Scholar] [CrossRef]
  102. Sirand-Pugnet, P.; Brégeon, D.; Béven, L.; Goyenvalle, C.; Blanchard, A.; Rose, S.; Grosjean, H.; Douthwaite, S.; Hamdane, D.; Crécy-Lagard, V. Reductive Evolution and Diversification of C5-Uracil Methylation in the Nucleic Acids of Mollicutes. Biomolecules 2020, 10, 587. [Google Scholar] [CrossRef] [Green Version]
  103. Urbonavicius, J.; Auxilien, S.; Walbott, H.; Trachana, K.; Golinelli-Pimpaneau, B.; Brochier-Armanet, C.; Grosjean, H. Acquisition of a bacterial RumA-type tRNA(uracil-54, C5)-methyltransferase by Archaea through an ancient horizontal gene transfer. Mol. Microbiol. 2008, 67, 323–335. [Google Scholar] [CrossRef] [PubMed]
  104. Lee, T.T.; Agarwalla, S.; Stroud, R.M. Crystal structure of RumA, an iron-sulfur cluster containing E. coli ribosomal RNA 5-methyluridine methyltransferase. Structure 2004, 12, 397–407. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Auxilien, S.; Rasmussen, A.; Rose, S.; Brochier-Armanet, C.; Husson, C.; Fourmy, D.; Grosjean, H.; Douthwaite, S. Specificity shifts in the rRNA and tRNA nucleotide targets of archaeal and bacterial m5U methyltransferases. RNA 2011, 17, 45–53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Sengupta, R.; Vainauskas, S.; Yarian, C.; Sochacka, E.; Malkiewicz, A.; Guenther, R.H.; Koshlap, K.M.; Agris, P.F. Modified constructs of the tRNA TPsiC domain to probe substrate conformational requirements of m(1)A(58) and m(5)U(54) tRNA methyltransferases. Nucleic Acids Res. 2000, 28, 1374–1380. [Google Scholar] [CrossRef] [Green Version]
  107. Alian, A.; Lee, T.T.; Griner, S.L.; Stroud, R.M.; Finer-Moore, J. Structure of a TrmA-RNA complex: A consensus RNA fold contributes to substrate selectivity and catalysis in m5U methyltransferases. Proc. Natl. Acad. Sci. USA 2008, 105, 6876–6881. [Google Scholar] [CrossRef] [Green Version]
  108. Walbott, H.; Leulliot, N.; Grosjean, H.; Golinelli-Pimpaneau, B. The crystal structure of Pyrococcus abyssi tRNA (uracil-54, C5)-methyltransferase provides insights into its tRNA specificity. Nucleic Acids Res. 2008, 36, 4929–4940. [Google Scholar] [CrossRef]
  109. Yamagami, R.; Yamashita, K.; Nishimasu, H.; Tomikawa, C.; Ochi, A.; Iwashita, C.; Hirata, A.; Ishitani, R.; Nureki, O.; Hori, H. The tRNA recognition mechanism of folate/FAD-dependent tRNA methyltransferase (TrmFO). J. Biol. Chem. 2012, 287, 42480–42494. [Google Scholar] [CrossRef] [Green Version]
  110. Gross, H.J.; Simsek, M.; Raba, M.; Limburg, K.; Heckman, J.; Raj Bhandary, U.L. 2′-O-methyl ribothymidine: A component of rabbit liver lysine transfer RNA. Nucleic Acids Res. 1974, 1, 35–43. [Google Scholar] [CrossRef] [Green Version]
  111. Shigi, N.; Sakaguchi, Y.; Asai, S.; Suzuki, T.; Watanabe, K. Common thiolation mechanism in the biosynthesis of tRNA thiouridine and sulphur-containing cofactors. EMBO J. 2008, 27, 3267–3278. [Google Scholar] [CrossRef] [Green Version]
  112. Čavužić, M.; Liu, Y. Biosynthesis of Sulfur-Containing tRNA Modifications: A Comparison of Bacterial, Archaeal, and Eukaryotic Pathways. Biomolecules 2017, 7, 27. [Google Scholar] [CrossRef] [Green Version]
  113. Shigi, N.; Sakaguchi, Y.; Suzuki, T.; Watanabe, K. Identification of two tRNA thiolation genes required for cell growth at extremely high temperatures. J. Biol. Chem. 2006, 281, 14296–14306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Shigi, N.; Suzuki, T.; Tamakoshi, M.; Oshima, T.; Watanabe, K. Conserved bases in the TPsi C loop of tRNA are determinants for thermophile-specific 2-thiouridylation at position 54. J. Biol. Chem. 2002, 277, 39128–39135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. 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] [Green Version]
  116. Joardar, A.; Jana, S.; Fitzek, E.; Gurha, P.; Majumder, M.; Chatterjee, K.; Geisler, M.; Gupta, R. Role of forefinger and thumb loops in production of Ψ54 and Ψ55 in tRNAs by archaeal Pus10. RNA 2013, 19, 1279–1294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Deogharia, M.; Mukhopadhyay, S.; Joardar, A.; Gupta, R. The human ortholog of archaeal Pus10 produces pseudouridine 54 in select tRNAs where its recognition sequence contains a modified residue. RNA 2019, 25, 336–351. [Google Scholar] [CrossRef] [Green Version]
  118. McCleverty, C.J.; Hornsby, M.; Spraggon, G.; Kreusch, A. Crystal structure of human Pus10, a novel pseudouridine synthase. J. Mol. Biol. 2007, 373, 1243–1254. [Google Scholar] [CrossRef]
  119. Aravind, L.; Koonin, E.V. THUMP—A predicted RNA-binding domain shared by 4-thiouridine, pseudouridine synthases and RNA methylases. Trends Biochem. Sci. 2001, 26, 215–217. [Google Scholar] [CrossRef]
  120. Gabant, G.; Auxilien, S.; Tuszynska, I.; Locard, M.; Gajda, M.J.; Chaussinand, G.; Fernandez, B.; Dedieu, A.; Grosjean, H.; Golinelli-Pimpaneau, B.; et al. THUMP from archaeal tRNA:m22G10 methyltransferase, a genuine autonomously folding domain. Nucleic Acids Res. 2006, 34, 2483–2494. [Google Scholar] [CrossRef]
  121. Waterman, D.G.; Ortiz-Lombardía, M.; Fogg, M.J.; Koonin, E.V.; Antson, A.A. Crystal structure of Bacillus anthracis ThiI, a tRNA-modifying enzyme containing the predicted RNA-binding THUMP domain. J. Mol. Biol. 2006, 356, 97–110. [Google Scholar] [CrossRef]
  122. Fislage, M.; Roovers, M.; Tuszynska, I.; Bujnicki, J.M.; Droogmans, L.; Versées, W. Crystal structures of the tRNA:m2G6 methyltransferase Trm14/TrmN from two domains of life. Nucleic Acids Res. 2012, 40, 5149–5161. [Google Scholar] [CrossRef] [Green Version]
  123. Neumann, P.; Lakomek, K.; Naumann, P.T.; Erwin, W.M.; Lauhon, C.T.; Ficner, R. Crystal structure of a 4-thiouridine synthetase-RNA complex reveals specificity of tRNA U8 modification. Nucleic Acids Res. 2014, 42, 6673–6685. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Hirata, A.; Nishiyama, S.; Tamura, T.; Yamauchi, A.; Hori, H. Structural and functional analyses of the archaeal tRNA m2G/m22G10 methyltransferase aTrm11 provide mechanistic insights into site specificity of a tRNA methyltransferase that contains common RNA-binding modules. Nucleic Acids Res. 2016, 44, 6377–6390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Chatterjee, K.; Blaby, I.K.; Thiaville, P.C.; Majumder, M.; Grosjean, H.; Yuan, Y.A.; Gupta, R.; de Crécy-Lagard, V. The archaeal COG1901/DUF358 SPOUT-methyltransferase members, together with pseudouridine synthase Pus10, catalyze the formation of 1-methylpseudouridine at position 54 of tRNA. RNA 2012, 18, 421–433. [Google Scholar] [CrossRef] [Green Version]
  126. Wurm, J.P.; Griese, M.; Bahr, U.; Held, M.; Heckel, A.; Karas, M.; Soppa, J.; Wöhnert, J. Identification of the enzyme responsible for N1-methylation of pseudouridine 54 in archaeal tRNAs. RNA 2012, 18, 412–420. [Google Scholar] [CrossRef] [Green Version]
  127. Chen, H.Y.; Yuan, Y.A. Crystal structure of Mj1640/DUF358 protein reveals a putative SPOUT-class RNA methyltransferase. J. Mol. Cell Biol. 2010, 2, 366–374. [Google Scholar] [CrossRef] [Green Version]
  128. Forterre, P.; Gribaldo, S.; Brochier-Armanet, C. Happy together: Genomic insights into the unique Nanoarchaeum/Ignicoccus association. J. Biol. 2009, 8, 7. [Google Scholar] [CrossRef] [Green Version]
  129. Kuchino, Y.; Ihara, M.; Yabusaki, Y.; Nishimura, S. Initiator tRNAs from archaebacteria show common unique sequence characteristics. Nature 1982, 298, 684–685. [Google Scholar] [CrossRef]
  130. Tang, T.H.; Polacek, N.; Zywicki, M.; Huber, H.; Brugger, K.; Garrett, R.; Bachellerie, J.P.; Hüttenhofer, A. Identification of novel non-coding RNAs as potential antisense regulators in the archaeon Sulfolobus solfataricus. Mol. Microbiol. 2005, 55, 469–481. [Google Scholar] [CrossRef]
  131. Zago, M.A.; Dennis, P.P.; Omer, A.D. The expanding world of small RNAs in the hyperthermophilic archaeon Sulfolobus solfataricus. Mol. Microbiol. 2005, 55, 1812–1828. [Google Scholar] [CrossRef]
  132. Nishikura, K.; De Robertis, E.M. RNA processing in microinjected Xenopus oocytes. Sequential addition of base modifications in the spliced transfer RNA. J. Mol. Biol. 1981, 145, 405–420. [Google Scholar] [CrossRef]
  133. Koski, R.A.; Clarkson, S.G. Synthesis and maturation of Xenopus laevis methionine tRNA gene transcripts in homologous cell-free extracts. J. Biol Chem. 1982, 257, 4514–4521. [Google Scholar] [CrossRef]
  134. Jiang, H.Q.; Motorin, Y.; Jin, Y.X.; Grosjean, H. Pleiotropic effects of intron removal on base modification pattern of yeast tRNAPhe: An in vitro study. Nucleic Acids Res. 1997, 25, 2694–2701. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. 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] [PubMed] [Green Version]
  136. 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] [Green Version]
  137. Schultz, S.K.; Kothe, U. tRNA elbow modifications affect the tRNA pseudouridine synthase TruB and the methyltransferase TrmA. RNA 2020, 26, 1131–1142. [Google Scholar] [CrossRef]
  138. 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]
  139. Hori, H. Regulatory Factors for tRNA Modifications in Extreme- Thermophilic Bacterium Thermus thermophilus. Front. Genet. 2019, 10, 204. [Google Scholar] [CrossRef] [Green Version]
  140. 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]
  141. Han, L.; Phizicky, E.M. A rationale for tRNA modification circuits in the anticodon loop. RNA 2018, 24, 1277–1284. [Google Scholar] [CrossRef] [Green Version]
  142. Perret, V.; Garcia, A.; Puglisi, J.; Grosjean, H.; Ebel, J.P.; Florentz, C.; Giegé, R. Conformation in solution of yeast tRNA(Asp) transcripts deprived of modified nucleotides. Biochimie 1990, 72, 735–743. [Google Scholar] [CrossRef]
  143. Byrne, R.T.; Konevega, A.L.; Rodnina, M.V.; Antson, A.A. The crystal structure of unmodified tRNAPhe from Escherichia coli. Nucleic Acids Res. 2010, 38, 4154–4162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Nagaswamy, U.; Fox, G.E. Frequent occurrence of the T-loop RNA folding motif in ribosomal RNAs. RNA 2002, 8, 1112–1119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Fechter, P.; Rudinger-Thirion, J.; Florentz, C.; Giegé, R. Novel features in the tRNA-like world of plant viral RNAs. Cell Mol. Life Sci. 2001, 58, 1547–1561. [Google Scholar] [CrossRef] [PubMed]
  146. Barends, S.; Björk, K.; Gultyaev, A.P.; de Smit, M.H.; Pleij, C.W.; Kraal, B. Functional evidence for D- and T-loop interactions in tmRNA. FEBS Lett. 2002, 514, 78–83. [Google Scholar] [CrossRef] [Green Version]
  147. Becker, H.F.; Motorin, Y.; Florentz, C.; Giegé, R.; Grosjean, H. Pseudouridine and ribothymidine formation in the tRNA-like domain of turnip yellow mosaic virus RNA. Nucleic Acids Res. 1998, 26, 3991–3997. [Google Scholar] [CrossRef] [Green Version]
  148. Ranaei-Siadat, E.; Fabret, C.; Seijo, B.; Dardel, F.; Grosjean, H.; Nonin-Lecomte, S. RNA-methyltransferase TrmA is a dual-specific enzyme responsible for C5-methylation of uridine in both tmRNA and tRNA. RNA Biol. 2013, 10, 572–578. [Google Scholar] [CrossRef] [Green Version]
  149. Zagryadskaya, E.I.; Kotlova, N.; Steinberg, S.V. Key elements in maintenance of the tRNA L-shape. J. Mol. Biol. 2004, 340, 435–444. [Google Scholar] [CrossRef]
  150. Zagryadskaya, E.I.; Doyon, F.R.; Steinberg, S.V. Importance of the reverse Hoogsteen base pair 54-58 for tRNA function. Nucleic Acids Res. 2003, 31, 3946–3953. [Google Scholar] [CrossRef] [Green Version]
  151. Kuhn, C.D. RNA versatility governs tRNA function: Why tRNA flexibility is essential beyond the translation cycle. Bioessays 2016, 38, 465–473. [Google Scholar] [CrossRef]
  152. Levy, M.; Miller, S.L. The prebiotic synthesis of modified purines and their potential role in the RNA world. J. Mol. Evol. 1999, 48, 631–637. [Google Scholar] [CrossRef]
  153. Cermakian, N.; Cedergren, R. Modified Nucleosides Always Were: An Evolutionary Model. In Modification and Editing of RNA; Grosjean, H., Benne, R., Eds.; ASM Press: Washington, DC, USA, 1998; pp. 535–541. [Google Scholar]
  154. McKenney, K.M.; Alfonzo, J.D. From Prebiotics to Probiotics: The Evolution and Functions of tRNA Modifications. Life Basel 2016, 6, 13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Anantharaman, V.; Koonin, E.V.; Aravind, L. Comparative genomics and evolution of proteins involved in RNA metabolism. Nucleic Acids Res. 2002, 30, 1427–1464. [Google Scholar] [CrossRef] [PubMed]
  156. McKenney, K.M.; Rubio, M.A.T.; Alfonzo, J.D. The Evolution of Substrate Specificity by tRNA Modification Enzymes. Enzymes 2017, 41, 51–88. [Google Scholar] [CrossRef] [PubMed]
  157. Galperin, M.Y.; Walker, D.R.; Koonin, E.V. Analogous enzymes: Independent inventions in enzyme evolution. Genome Res. 1998, 8, 779–790. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Soucy, S.M.; Huang, J.; Gogarten, J.P. Horizontal gene transfer: Building the web of life. Nat. Rev. Genet. 2015, 16, 472–482. [Google Scholar] [CrossRef] [PubMed]
  159. Sieber, K.B.; Bromley, R.E.; Dunning Hotopp, J.C. Lateral gene transfer between prokaryotes and eukaryotes. Exp. Cell Res. 2017, 358, 421–426. [Google Scholar] [CrossRef]
  160. Galperin, M.Y.; Koonin, E.V. Sources of systematic error in functional annotation of genomes: Domain rearrangement, non-orthologous gene displacement and operon disruption. Silico Biol. 1998, 1, 55–67. [Google Scholar]
  161. Zallot, R.; Harrison, K.J.; Kolaczkowski, B.; de Crécy-Lagard, V. Functional Annotations of Paralogs: A Blessing and a Curse. Life 2016, 6, 39. [Google Scholar] [CrossRef] [Green Version]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

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. https://doi.org/10.3390/genes12020140

AMA Style

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(2):140. https://doi.org/10.3390/genes12020140

Chicago/Turabian Style

Roovers, Martine, Louis Droogmans, and Henri Grosjean. 2021. "Post-Transcriptional Modifications of Conserved Nucleotides in the T-Loop of tRNA: A Tale of Functional Convergent Evolution" Genes 12, no. 2: 140. https://doi.org/10.3390/genes12020140

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