Skip to main content
Microbiology and Molecular Biology Reviews : MMBR logoLink to Microbiology and Molecular Biology Reviews : MMBR
. 2024 Feb 29;88(1):e00199-23. doi: 10.1128/mmbr.00199-23

Biosynthesis and function of 7-deazaguanine derivatives in bacteria and phages

Valérie de Crécy-Lagard 1,2,, Geoffrey Hutinet 3, José D D Cediel-Becerra 1, Yifeng Yuan 1, Rémi Zallot 4, Marc G Chevrette 1, R M Madhushi N Ratnayake 5,2, Marshall Jaroch 1,3, Samia Quaiyum 1, Steven Bruner 5
Editor: Corrella S Detweiler6
PMCID: PMC10966956  PMID: 38421302

SUMMARY

Deazaguanine modifications play multifaceted roles in the molecular biology of DNA and tRNA, shaping diverse yet essential biological processes, including the nuanced fine-tuning of translation efficiency and the intricate modulation of codon-anticodon interactions. Beyond their roles in translation, deazaguanine modifications contribute to cellular stress resistance, self-nonself discrimination mechanisms, and host evasion defenses, directly modulating the adaptability of living organisms. Deazaguanine moieties extend beyond nucleic acid modifications, manifesting in the structural diversity of biologically active natural products. Their roles in fundamental cellular processes and their presence in biologically active natural products underscore their versatility and pivotal contributions to the intricate web of molecular interactions within living organisms. Here, we discuss the current understanding of the biosynthesis and multifaceted functions of deazaguanines, shedding light on their diverse and dynamic roles in the molecular landscape of life.

KEYWORDS: queuosine, archaeosine, toyocamycin, bacteriophage, restriction/modification, genetic code

INTRODUCTION

7-Deazapurines are analogs of nucleosides that feature pyrrolo[2,3-d]pyrimidine or pyrrolopyrimidine moieties. They can be found either as independent molecules or integrated into RNA or DNA polymers (1, 2). A distinctive structural characteristic of 7-deazapurines involves substitutions of the nitrogen atom typically located at the purine scaffold’s position 7 (Fig. 1). Deazapurine-containing antibiotics such as tubercidin and sangivamycin (3) (Fig. 2) and deazapurine RNA modifications such as queuosine (Q) (4) and archaeosine (G+) (5) were first described in the 1960s to 1980s. More recently, their large structural diversity has become apparent, propelled by the identification of synthesis pathway genes and the growing availability of diverse genome sequencing data.

Fig 1.

Fig 1

General pathway of 7-deazaguanine modifications in tRNA and DNA. The illustration includes representations of all final and distinctive molecules in the pathway, excluding redundant ones. Key proteins catalyzing reactions are highlighted in bold text adjacent to the corresponding arrows and as described in the text. When two proteins or more are listed, these are alternate enzymes catalyzing the same reactions. Unknown enzymes have question marks associated (?).

Fig 2.

Fig 2

7-Deazapurine natural product representatives with entries in the MIBiG database. Huimycin (A), sangivamycin (B), tubercidin (C), and toyocamycin (D) are the only metabolites with known minimal annotations for the BGCs in the database. The rest of the compounds described in Table 1 remain orphaned. All those compounds have in common the preQ0 scaffold in their chemical structures.

Though the pyrrolopyrimidine core structure is strictly conserved, the deazapurines are a diverse class of molecules regarding their chemical structures, taxonomic distribution, and functional roles. While the diversity of 7-deazapurines has long been appreciated in the context of secondary metabolites, it was initially believed that only a few derivatives were present in nucleic acids, namely G+, Q, and Q derivatives (1). However, recent reports of 7-deazapurines in DNA have altered this perspective, identifying eight distinct pyrrolopyrimidine modifications in bacteriophages so far (6) (Fig. 1).

Depending on the deazapurine, taxonomic distribution may be wide or narrow and the degree of enzymatic conservation of the biosynthetic pathway can vary. For example, Q is predicted to be present in the tRNAs of over 90% of sequenced bacterial and eukaryotic species (7, 8), while many pyrrolopyrimidine secondary metabolites are only found in select bacterial lineages (Table 1). The functions of deazapurines vary greatly depending on their final structures and/or location, as the same molecule can have different roles when located in RNA or DNA. G+ in archaea plays a role in stabilizing tRNA tertiary structure (9), whereas the same modified base in DNA shields bacteriophages from restriction enzymes (10). Q is crucial for tRNA decoding efficiency or accuracy (11, 12) and has been adapted for regulatory functions in some individual species (13). Some deazapurines in bacterial DNA are components of restriction-modification islands (14). The natural functions of deazapurine secondary metabolites are not entirely understood but some have demonstrated anticancer, antiviral, or antibacterial activities (Table 1).

TABLE 1.

7-Deazapurine derived natural products

Natural product Bioactivitya Organism MIBiG BGCb References
Tubercidin AB, AP, AV, AT Streptomyces tubercidis BGC0001937 (1, 15)
Sangivamycin AB, AT, AV Streptomyces rimosus ATCC 14673 BGC0000879 (1, 16, 17)
Toyocamycin AB, AT, AV Streptomyces rimosus ATCC 14673 BGC0000879 (1, 17, 18)
Cadeguomycin AB, AT Streptomyces hygroscopicus U (19)
Kanagawamicin AT Actinoplanes kanagawaensis U (20, 21)
Echiguanine A AT Streptomyces M1698-50F1 U (22)
Echiguanine B AT Streptomyces M1698-50F1 U (22)
Ara-A H Actinoplanes sp. A9222 U (23)
Dapiramicin A AB, AF Micromonospora sp. SF-1917 U (24, 25)
Dapiramicin B AB, AF Micromonospora sp. SF-1917 U (26)
5′-deoxyguanosine H Thermoactinomycete sp. A6019 U (23)
Coaristeromycin H Streptomyces sp. A6308 U (23)
Aristeromycin H Streptomyces sp. A6308 U (23)
5′-deoxytoyocamycin H Streptomyces sp. A14345 U (23)
Coformycin H Unclassified U (23)
5′-deoxy-5-iodotubercidin AKI Hypnea valentiae U (27)
4-amino-5-bromo-pyrrolo[2,3-d] pyrimidine B Echinodictyum U (27)
Huimycin U Kutzneria albida DSM 43870 BGC0002354 (28)
Tubercidin-5′-α-d-glucopyranose CT, AF Plectonema radiosum, Tolypothrix tenuis U (29)
Toyocamycin-5′-α-d-glucopyranose CT, AF Plectonema radiosum, Tolypothrix tenuis U (29)
Mycalisine A AB Mycale sp. U (30)
Mycalisine B AM Mycale sp. U (30, 31)
5′-deoxy-5-iodotubercidin NB Hypnea valendiae, Didemnum voeltzkowi U (32)
5-(methoxycarbonyl) tubercidin CT Jaspis johnstoni U (33)
Toyomycin CT Jaspis johnstoni U (34)
Rigidins B-D CA Cystodytes sp. U (35)
7-deazainosine CT Aplidium pantherinum U (36)
5′-deoxy-3-bromotubercidin CT Didemnum voeltzkowi U (37)
5′-deoxytubercidin CT Didemnum voeltzkowi U (37)
Unamycin B AB Streptomyces . fungicidicus U (38)
Vengicide AB Streptomyces vendargensis U (38)
a

antibiotic (AB); antiparasitic (AP); antiviral (AV); antitumor (AT); antifungal (AF); herbicidal (H); adenosine kinase inhibitor (AKI); bronchodilator (B); unknown (U); antimitotic (AM); cytotoxic (CT); neuromuscular blocking (NB); calmodulin antagonistic (CA).

b

BGCs reported in MIBiG with minimal annotation.

The exploration of various 7-deazapurine biosynthesis pathways has unveiled unprecedented enzyme chemistries and novel structural folds (1, 2). This has both practical applications and evolutionary significance, especially considering that many of these enzymes belong to the tunnel-fold (T-fold) family, which appears to have been recruited to execute different types of reactions on similar substrates (39).

Certain precursors leading to the formation of the ultimate 7-deazapurine molecules can be reclaimed and viewed as micronutrients (40). Consequently, the notion that competition for these deazapurine precursors might influence the ecology of distinct niches, such as the mammalian microbiota or other host-associated environments, is just beginning to take root (41, 42). Thus, deazapurines are widespread metabolites with diverse biological functions, and the complete scope of their roles continues to reveal itself as our understanding advances.

PREQ0 AND PREQ1 ARE THE COMMON PRECURSORS FOR NUMEROUS DEAZAPURINE DERIVATIVES

PreQ0, also known as 7-cyano-7-deazaguanine, serves as the precursor to most natural 7-deazaguanine-containing molecules (1, 2). Both preQ0 and its derivative, 7-aminomethyl-7-deazaguanine (preQ1), can not only be incorporated into DNA or RNA and undergo further modifications but also act as precursors to numerous secondary metabolites. The biosynthesis of preQ0 occurs from guanosine-5'-triphosphate (GTP) in a four-step pathway (Fig. 1; Table 2) that remained enigmatic for over three decades. This pathway was eventually elucidated between 2004 and 2009: the gene candidates were identified by a combination of taxonomic distribution filters and gene neighborhood analyses and were validated by a combination of bioinformatic, genetic, and biochemical studies (4346). These investigations also identified QueF as the enzyme responsible for the biosynthesis of preQ1 from preQ0 (47).

TABLE 2.

Known bacterial and archaeal Q and G+ synthesis and salvage enzymes and transportersb

Protein name Protein function COG KO Cofactors
FolE/GCHI GTP cyclohydrolase IA (EC 3.5.4.16) COG0302 K01495 Zn
FolE2 GTP cyclohydrolase IB (EC 3.5.4.16) COG1469 K09007 Mn
QueD 6-carboxy-5,6,7,8-tetrahydropterin synthase (EC 4.1.2.50) COG0720a K01737 (1) Zn
QueD2 6-carboxy-5,6,7,8-tetrahydropterin synthase (EC 4.1.2.50) COG0720a K01737 (1) Fe, Zn
QueE 7-carboxy-7-deazaguanine synthase (EC 4.3.99.3) COG0602a K10026 Fe
QueC 7-cyano-7-deazaguanine synthase (EC 6.3.4.20) COG0603 K06920 ATP, Zn, NH4+
QueF type I 7-cyano-7-deazaguanine reductase (EC 1.7.1.13) Type II COG0780,
COG2904
K06879 NADPH
QueF type II 7-cyano-7-deazaguanine reductase (EC 1.7.1.13) Type I COG0789 K09457 NADPH
bTgt Queuine tRNA-ribosyltransferase (EC 2.4.2.29) COG0343 K00773 Zn
QueA S-adenosylmethionine:tRNA ribosyltransferase-isomerase (EC 2.4.99.17) COG0809 K07568 SAM
QueG Epoxyqueuosine reductase (EC 1.17.99.6) COG1600 K18979 Fe, Cobalamin
QueH epoxyqueuosine reductase (EC 1.17.99.6) COG1636a K09765 Fe
GluQ Glutamyl-Q tRNA(Asp) synthetase (EC 6.1.1.B3) COG0008a K01894 Zn
aTGT 7-cyano-7-deazaguanine tRNA-ribosyltransferase (EC 2.4.2.48) COG1370 K18779 Zn
ArcS Archaeosine synthase alpha-subunit (EC 2.6.1.97; 2.6.1.-) COG1549 K07557 Lys or Gln
RaSEA Archaeosine synthase beta-subunit [EC 2.6.1.-] COG1244a K06936
QueF-Like QueF-like amidinotransferase (EC 2.6.1.-) Pcal_0221 (2) NH4+
QPTR/YhhQ Queuosine precursor transporter (TC 3.A.1.28) COG1738 K09125
QrtT/QueT Energy-coupling factor transport system substrate-specific component (TC 3.A.1.28) COG4708 K16923
QueK Queuosine hydrolase (EC 3.2.2.1) COG1957 No KO, CD630_16820 (2)
QueL Queuine lyase (EC 4.3.99.M4) COG1244a CD630_16840 (2) SAM
Qng1 Queuosine 5′-phosphate N-glycosylase/hydrolase (EC 3.2.2.-) pfam10343 (3) Sthe_2331 (2)
a

Not isofunctional COG.

b

COG: Clusters of Orthologous Groups; KO: KEGG Orthology Number (1); KO not specific for QueD/PTPS-I (2); No KO number, posted a specific ID with experimental validation. (3); No COG number posted a pfam number instead.

QueD: The first enzyme of preQ0 biosynthesis is shared with the tetrahydrofolate and biopterin pathways

As predicted by the foundational work that identified GTP as the deazapurine precursor (48, 49), the preQ0, tetrahydrofolate (THF), and biopterin (BH4) synthesis pathways share a common first step with the formation of 7,8-dihydroneopterin triphosphate (H2NTP) from GTP (Fig. 3) (46, 50). GTP cyclohydrolase I (EC 3.5.4.16), an enzyme class with various cofactors, mediates this reaction. The Zn2+-dependent variant (FolE; COG0302) is used by most bacteria and mammals, which contrasts with the FolE2 type (COG1469) used by other bacteria and most archaea, which employs different metals like Mn2+ (51). Both FolE and FolE2 are members of the T-fold structural superfamily (39, 52). Shared pathway intermediates might enable shifts between THF and Q pools under conditions of GTP scarcity or elevated Q biosynthesis demand, as seen with riboflavin, another molecule originating from GTP (53). The first committed step in preQ0 synthesis, catalyzed by 6-carboxy-5,6,7,8-tetrahydropterin (6-CPH4) synthase EC (EC4.1.2.50) or QueD (43, 44) (Fig. 1), is a textbook example of the difficulty in annotating paralogous families (54). The Escherichia coli QueD (or PTPS-I) protein was first annotated as 6-pyruvoyltetrahydropterin synthase (PTPS) because of its similarity with the mammalian homolog involved in biopterin synthesis, now named PTPS-II. QueD/PTPS-I and PTPS-II are both members of the T-fold derived COG0720 family (Pfam PF01242). Comparative genomic predictions (43) combined with genetic (43), biochemical (44), and structural studies (50, 55) revealed that slight variations in the active site can shift catalytic activity between PTPS-II and QueD/PTPS-I (Fig. 3). A third subgroup in this family, PTPS-III, catalyzes the conversion of H2NTP into neopterin, bypassing the FolB and FolQ enzymes in the THF pathways of certain bacteria and parasites (Fig. 3) (56). Remarkably, some bacteria have bifunctional QueD/PTPS-III enzymes that must be used both for Q and THF biosynthesis (50). Another QueD variant, QueD2 was identified based on its signature motif and metal binding properties (Fig. 3) (50, 57). QueD is a Zn2+-dependent (44, 55) lyase that catalyzes the elimination of triphosphate and acetaldehyde from H2NTP. A coordinated zinc ion plays a key role, stabilizing oxyanion-containing intermediates leading to the product carboxylate. Comparative genomic analysis suggests QueD2 is regulated by the zinc uptake regulator (Zur) under metal-limiting conditions. The paralogs contain a second zinc-binding site located in an inserted region not present in QueD, and based on its structure, a model has been proposed where the second metal site of QueD2 slows the dissociation of the catalytic metal (50, 57). The different COG0720 subgroups can be separated by strictly conserved signature motifs (50, 57, 58) but many annotation mistakes remain in most databases (59) and several discernable subgroups of this family remain to be functionally characterized [see Fig. S4 in (57)].

Fig 3.

Fig 3

Functional roles of different PTPS subfamilies. Biosynthesis pathways in which PTPS-I, II, and III are involved. Specific reactions catalyzed by PTPS-I (QueD or QueD2), PTPS-II, and PTPS-III, and conserved motifs were identified. Abbreviations: GTP: guanosine triphosphate; FolE: GTP cyclohydrolase I; FolE2: GTP cyclohydrolase II; DHN-P3: dihydroneopterin-triphosphate; CTHP: 6-carboxytetrahydropterin; PTP: 6-pyruvoyl-tetrahydropterin; HMDHP: 6-hydroxymethyldihydropterin; preQ0: 7-cyano-7-deazaguanosine; QueE: 7-carboxy-7-deazaguanine synthase; QueC: 7-cyano-7-deazaguanine synthase; SR: sepiapterin reductase; DHN: dihydroneopterin; [P-ase]: phosphatase; FolB: dihydroneopterin aldolase; FolK: 6-hydroxymethyl-7,8-dihydropterin pyrophosphokinase; FolP: dihydropteroate synthase; FolC: dihydrofolate:folylpolyglutamate synthase; FolA: dihydrofolate reductase.

QueE: an atypical radical SAM enzyme

The next step in preQ0 biosynthesis is the heterocyclic radical-mediated conversion of 6-carboxy-5,6,7,8-tetrahydropterin (CPH4) to 7-carboxy-7-deazaguanine (CDG) catalyzed by 7-carboxy-7-deazaguanine synthase or QueE (EC 4.3.99.3) (60). First identified by in silico and genetics methods (43), QueE is a member of the radical SAM superfamily of enzymes [(61) and see https://radicalsam.org/explore.php?id=cluster-3-1&v=3.0] that perform a wide array of chemical reactions initiated by the highly reactive 5′-deoxyadenosyl radical (62) including C-H activation, atom/group transfer, isomerizations, bond cleavage, and rearrangements. QueE has been extensively studied using biophysical, structural, and biochemical approaches (60, 6365). The overall mechanism involves nitrogen atom migration, resulting in ring contraction, followed by the elimination of ammonia. The role of QueE in catalysis is to stabilize and control the fate of high-energy radical intermediates. Among the larger family of radical SAM enzymes, QueE is atypical because of its dependence on an Mg+2 cation and overall turnover of the radical S-adenosylmethionine cofactor.

QueC: an ATPase that catalyzes two reactions

The enzyme 7-cyano-7-deazaguanine synthase, also known as QueC (EC 6.3.4.20), plays a pivotal role in catalyzing the conversion of CDG into preQ0 (45). This process involves two ATP molecules: one is consumed to generate a 7-amido-7-deazaguanine (ADG) intermediate while the other is used to process ADG into preQ0 (66) (Fig. 1). Mutating two strictly conserved residues located within 7 Å of the phosphate ligand (K163A/R204A) in the Bacillus subtilis QueC protein, effectively halts the reaction at the ADG intermediate, showing the importance of these two residues in processing the second reaction from ADG to preQ0 (67). Each monomeric subunit of the QueC homodimer consists of an N-terminal domain exhibiting a Rossman fold architecture, a characteristic feature shared with many nucleotide-binding proteins. In addition, it contains a helical zinc-binding C-terminal domain. Notably, the active site is predicted to reside at the interface between these two domains (68, 69). QueC-like encoding genes have been found in phage defense clusters, such as QatC in the QueC-like associated with ATPase and TatD DNase system (Qat) (70) or Cap9 in the type IV cyclic oligonucleotide-based anti-phage signaling system (CBASS) (71), yet their functions remain to be elucidated.

QueF: the four-electron reduction of preQ0 to preQ1 from nitrile to primary amine

In contrast to non-specific reductions by nitrogenases, the QueF-mediated reduction of preQ0 to preQ1 is the only known nitrile reduction found in a natural biosynthetic pathway (72). Although the catalytic activity was witnessed in the late 1970s (73), the NADPH-dependent 7-cyano-7-deazaguanine QueF (EC 1.7.1.13) enzyme was only characterized in 2004 (47). QueF was first identified by comparative genomic and genetic studies and found to be involved in Q synthesis in Acinetobacter baylyi and Bacillus subtilis (43). The exact biochemical function was elucidated a few years later despite the initial misannotation that QueF was a GTP cyclohydrolase I based on its membership in the T-fold family (47). Sequence determinants specific to QueF’s reductase activity have been identified and allow for the differentiation of QueF from GTP cyclohydrolases. As with other members of the T-fold family, the QueF active site pocket is located at the interface between subunits (72). Two types of QueF enzyme architectures have been characterized to date. In the QueF type I exemplified by B. subtilis YkvM (K09457), two independent subunits form the catalytic interface while in the QueF type II exemplified by E. coli YqcD (K06879), the interface is formed by two domains of the same subunit. Therefore, type II proteins are predicted to have arisen from a duplication of the type I domain with a catalytic inactive portion at the C-terminus (47). Mechanistic and structural studies identified Cys55 in B. subtilis QueF as a key catalytic residue, forming an α,β-unsaturated thioamide covalent intermediate with preQ0, supporting a covalent catalysis reaction mechanism (72, 7477). To prevent oxidation of the catalytic cysteine, QueF forms a large homodecameric complex with active sites at the inter-monomer interfaces, facilitated by an intermolecular disulfide bridge with another cysteine (78). Because of their unique reductase activity and their potential exploitation for biocatalysis (see in dedicated section), QueF enzymes have been extensively studied (76, 77). However, these characterized QueF enzymes are very specific for preQ0, with only a limited number of other substrates reported (79).

An analysis of the distribution of the preQ1 synthesis proteins (QueC, QueD, QueE, and QueF type I and Type II) and of the Q biosynthesis proteins [(Tgt), QueA see section below)] found 148 out of 7,267 bacterial genomes encoded all proteins but the QueF homologs (data extracted from https://www.kegg.jp/kegg-bin/view_ortholog_table?orthology=K01737+K10026+K06920+K06879+K09457+K00773+K07568 Release 108.0, October 1, 2023). This suggests a non-orthologous displacement for the yet unidentified enzyme catalyzing preQ0 reductase activity, as enzymes with preQ1 substrate specificity exist in those organisms, namely Tgt and QueA. Proposed candidates for this cryptic activity are members of the 2-hydroxyacyl-CoA dehydratase superfamily, but experimental validation remains to be performed (80).

Conservation and regulation of preQ0/preQ1 biosynthetic genes

The dedicated preQ0/preQ1 biosynthetic genes (queD, queC, queE, and queF) are generally physically clustered in different combinations (see https://www.kegg.jp/kegg-bin/view_ortholog_table?orthology=K01737+K10026+K06920+K06879+K09457+K18979+K09765+K00773+K07568). In some organisms, the folE/folE2 genes can also be present in the clusters (Fig. 4). The regulation of preQ0 and preQ1 biosynthetic genes remains poorly understood in many organisms, including the model E. coli K12. The only known dedicated regulatory elements are preQ1 responsive riboswitches, categorized into one of three classes (81). Members belonging to class-I were identified upstream of the B. subtilis preQ1 biosynthesis operon (82) and are mainly found in Proteobacteria and Firmicutes. Class-II preQ1 riboswitches are primarily present in Lactobacillales, while class-III is primarily found in Clostridiales (81). Not all preQ1 riboswitches employ the same mechanisms: the B. subtilis class-I riboswitch uses ligand-triggered transcriptional termination, while the class-II Streptococcus pneumoniae R6 riboswitch utilizes metabolite-mediated sequestration of the Shine-Dalgarno sequence (83). A class-I preQ1 riboswitch in Carnobacterium antarcticus was recently found to bind stacked effector molecules, a unique property in riboswitches studied to date (84). preQ1 riboswitches have a characteristically short length, some as short as 25 nucleotides, making them suitable for structural and biophysical studies (8587). They are used as inducible regulatory elements in synthetic biology (88, 89).

Fig 4.

Fig 4

Physical clustering of genes encoding preQ0/preQ1 synthesis or Q synthesis/salvage enzymes. Different representatives of the gene neighborhood clusters discussed in the text are shown and were drawn using the GeneGraphics App (https://v2.genegraphics.net/) (90). Numeric protein identifiers are given for every example to retrieve specific information. All abbreviations of 7-deazapurine metabolism-related proteins are given in the text or Table 2 with the exception of short-chain dehydrogenase reductase (SDR). DeoB-like and QueA-like reflect that the functions might not be the same as the canonical DeoB (phosphopentomutase) or QueA enzymes.

DIVERSITY OF DEAZAGUANINE MODIFICATIONS IN NUCLEIC ACIDS

Transglycosylases insert free preQ0, preQ1, or Q bases in nucleic acids

preQ0 and preQ1 can be used directly as precursors of natural products (see dedicated section below). Their insertion as bases into nucleic acid molecules requires transglycosylase enzymes that exchange the deazapurine and guanine bases (Fig. 1 and 5). The founding members of this family of enzymes (Interpro ID: IPR002616) modify tRNAs and fall into three subgroups: (i) homodimeric bacterial tRNA-guanine transglycosylases (EC 2.4.2.29, bTGT) that exchange the wobble position guanine in tRNA with GUN anticodons with preQ1; (ii) eukaryotic heterodimeric Queuine tRNA-ribosyltransferases [EC 2.4.2.64, eTGT composed of a catalytic subunit (QTRT1) and an accessory subunit (QTRT2)] that introduce the Q base directly in the same target GUN-anticodon tRNAs; and (iii) archaeal tRNA-guanine transglycosylases (EC 2.4.2.48, arcTGT) that exchange the G at position 15, and sometimes at position 13 (91) with preQ0 in many target tRNAs in most Archaea. The structures and catalytic mechanisms of these three groups of enzymes have been extensively studied and previously reviewed (92). The residues of the substrate binding pocket that allow discrimination between preQ0, preQ1, and Q substrates are well characterized (92, 93) as shown in Fig. 5.

Fig 5.

Fig 5

Substrate specificity of various deazaguanine transglycosylases. Logos on top represent the conserved region for 7-deazaguanine substrate binding and were generated from a relevant set of orthologous sequences of the protein depicted using MAFTT (94) and WebLogo2 (95). Representative structures for some orthologs are depicted on the bottom with their cognate substrate (colored in orange). Bacterial TGT from Z. mobilis complexed with preQ1 (green, PDB ID: 1P0E), AlphaFold (96, 97) modeled C. trachomatis TGT with queuine (cyan, UniProt O84196), human TGT in complex with queuine (magenta, PDB ID: 6H45), and archaeal TGT from P. horikoshii complexed with preQ0 (yellow, PDB ID: 1IT8). All structural illustrations were prepared using PyMOL (https://pymol.org/2/).

In the last 5 years, the understanding of the functional diversity of those different subgroups has greatly expanded. A subgroup of bacterial enzymes was found to have shifted substrate specificity from preQ1 to q in bacteria that live in queuine-rich environments (e.g., intracellular pathogens) (Fig. 5 and 6B) (7). Another subgroup, renamed DpdA for deazapurine in DNA, are homologs of archaeal TGT proteins encoded by bacteria or phages that have shifted their substrate specificities from RNA to DNA (14). In some bacteria, such as Salmonella enterica serovar Montevideo, DpdA incorporates preQ0 into DNA only with the help of the DpdB protein (98, 99). The functional roles of DpdB are not yet clearly defined (and discussed below). Phage DpdA proteins do not require DpdB to insert 7-deazapurines into DNA (6, 10). Phage DpdA can be categorized into four groups that differ by their substrate specificities and potentially their sequence specificities. For example, the E. coli phages 9g and CAjan DpdA1 enzymes are specific for preQ0 (10, 14) while the Haloarcula virus HVTV-1 DpdA4 is specific for preQ1 (6). The DpdA2 enzymes seem to be more promiscuous. For example, the Vibrio natriegens phage nt-1 DpdA2 prefers preQ0 while the Vibrio phage VH7D DpdA2 prefers preQ1 but can insert preQ0 and CDG into DNA, albeit at lower efficiencies, when preQ1 is not available (10). The in vivo substrates of the DpdA3 enzymes are yet to be determined, but the phage genomes that encode them are modified with preQ0 precursors such as CDG or ADG (6).

Fig 6.

Fig 6

Known bacterial Q synthesis or salvage pathways. (Upper feft) De novo Q synthesis and preQ0/preQ1 salvage pathways in E. coli; (Upper right) q salvage pathway in C. trachomatis; (Lower left ) preQ1, q, and Q salvage pathways in C. difficile. The ECF transporters include four subunits: S, the substrate-specific transmembrane component (QueT); T, the energy-coupling module; A and A′, the pair of ABC ATPase. (Lower right) Possible q and Q salvage pathways in B. henselae Houston 1. All abbreviations of 7-deazapurine metabolism-related proteins are given in the text or Table 2.

Target sequence specificity has been determined for two DpdA1 enzymes, both of which utilize preQ0 as a substrate. The DpdA1 enzyme from Enterobacteria phage CAjan recognizes both “GA” and “GGC” sequences (100), while the DpdA enzyme from Pseudomonas phage Iggy specifically recognizes “GA” (101). Both enzymes replace the first guanine of the recognition sequence with preQ0.

A predicted structural analysis of CAjan DpdA1 revealed striking similarities to TGTs (100). Notably, the binding pocket in CAjan DpdA1 resembles that of arcTGT, and both enzymes share two conserved catalytic aspartates. In addition, several residues (including Ser64, Phe67, Asp105, Gly153, Gly154, His132, and Phe189) are predicted to be involved in base binding activity. Asp206 is anticipated to catalyze the first step of the transglycosylation process, while Asp63 is likely responsible for the deprotonation of preQ0 (100).

Diversity of deazapurine derivatives identified in phage DNA

To date, eight derivatives of 7-deazaguanine have been identified in phage genomes (Fig. 1). The first, known as dG+, was initially discovered in Enterobacteria phage 9g (14), where it replaced ~25% of the guanine content. Subsequently, three additional modifications were observed in various phage genomes: 2′-deoxy-7-cyano-7-deazaguanosine (dpreQ0), 2′-deoxy-7-amido-7-deazaguanosine (dADG), and 2′-deoxy-7-aminomethyl-7-deazaguanine (dpreQ1) (10). dpreQ0 was found in Escherichia phage CAjan, with 32% of the guanine undergoing modification, as well as in Mycobacterium phage Rosebush (28% guanine modification) and Vibrio phage nt-1 (0.1% guanine modification). dADG was found to modify 100% of the guanine in Campylobacter phage CP220 (102). It was hypothesized that QueC only performs the initial reaction resulting in ADG, which would then be loaded onto a 2′-deoxyribose and inserted into its genome by DpdA3 (Fig. 1), but this scenario needs to be further validated. The presence of dADG was also detected in small amounts within Halovirus HVTV-1, Vibrio phage nt-1, and Mycobacterium phage Rosebush, while it constituted the primary modification in Salmonella phage 7–11, affecting 0.02% of the guanine (10). In addition, dpreQ1 was observed in Halovirus HVTV-1, where it modified ~30% of the guanine, and in Streptococcus phage Dp-1, affecting 1.7% of the guanine residues (10). In more recent discoveries, methylated and formylated forms of preQ1 were identified, namely 2′-deoxy-7-(methylamino)methyl-7-deazaguanine (mdpreQ1) and 2′-deoxy-7-(formylamino)methyl-7-deazaguanine (fdpreQ1), alongside dCDG and its decarboxylated form, 2′-deoxy-7-deazaguanine (dDG) (6). mdpreQ1 was found in Cellulophaga phage phiSM, affecting 0.1% of the guanine residues, in combination with dpreQ1, which modified 1.1% of the guanines. Similar findings were observed for Cellulophaga phage phi38:2 and phi 47:1. Meanwhile, fdpreQ1 and dDG were both observed at a 100% replacement rate in Flavobacterium phage vB_FspM_immuto_2–6A and Cellulophaga phage phST, respectively. Lastly, dCDG was noted in Sulfolobus virus SVST-2, affecting 0.04% of the guanine content. The process through which dCDG is produced in the genome of Sulfolobus virus SVST-2 remains unclear but for the other seven modifications, the pathways have been nearly fully elucidated.

Phages that undergo modification with dpreQ0 consistently carry the genes encoding the preQ0 synthesis proteins FolE, QueE, QueD, and QueC, in addition to the signature enzyme DpdA (10, 14). Phages modified with dG+ encode different non-orthologous enzymes like ArcS or Gat_QueC, which facilitate the final amidotransferase step akin to G+ synthesis in Archaea (see below) (10). Phages encoding QueF are subject to modification by dpreQ1, involving a change in the substrate of the cognate DpdA to preQ1 (10). A preQ1 methyltransferase, known as DpdM, has been identified and experimentally validated (6). DpdM is likely a metalloprotein with four cysteine residues capable of binding two metals. Furthermore, a proposed preQ1 formyltransferase, DpdN, is a paralog of PurN involved in purine synthesis (103). A candidate CDG decarboxylase DpdL has also been suggested. DpdL, a member of the T-fold superfamily with a known affinity for pterins and purines (39), features an LxxxHRHxF signature motif binding a metal, indicative of an alkaline decarboxylation mechanism.

Synthesis of Q and Q derivatives in RNA

Once preQ1 has been inserted in target tRNAs by the bTGT enzyme, two additional catalytic steps are required to finalize the synthesis of the Q molecule using quite unusual enzymes.

SAM is a ribose donor in the formation of epoxy-Q

The initial enzyme in the transformation of preQ1-tRNA into Q-tRNA, known as S-adenosylmethionine:tRNA ribosyltransferase-isomerase or QueA (EC 2.4.99.17), transfers the ribose moiety from S-adenosylmethionine (SAM) with L-Met and adenine as byproducts. The identification of the queA gene occurred two decades ago in E. coli because it was upstream of the tgt gene (104). This close syntenic association is prevalent, observed in ~50% of the 4,610 bacterial genomes in the KEGG database that encode both tgt and queA (data extracted from https://www.kegg.jp/kegg-bin/view_ortholog_table?orthology=K00773+K07568+K18979+K09765, Release 108.0, October 1, 2023).

Initial kinetic (105) and structural studies (106, 107) have revealed that this enzyme possesses a unique fold and operates through a fully ordered sequential bi-ter kinetic mechanism. In this mechanism, preQ1-tRNA-Tyr binds first, followed by SAM, with product release occurring in the order of adenine, methionine, and epoxyQ-tRNA (or oQ-tRNA). The proposed mechanism involves a unique enzymatic pathway that includes sulfonium ylide and vinyl sulfonium intermediates (108). The five carbons from the ribose moiety of SAM are transformed into the cyclopropyl epoxide of the final product. As of now, no structures bound to substrates have been resolved, leaving the characterization of this unconventional use of SAM as a ribosyl donor incomplete (109).

The QueA protein family generally exhibits iso-functionality, except for members found in numerous Actinomycetes. In most bacteria within this clade, the Q modification is absent (110112). QueA homologs in these Actinomycetes are encoded by genes that cluster with those encoding proteins of the short-chain dehydrogenase superfamily (Fig. 4) and should be renamed QueA-like.

Cobalamin-dependent or independent solutions for epoxy-Q reductase synthesis

In the late 1980s, it was discovered that the last step in Q synthesis, the reduction of epoxy-Q (oQ) to Q, is dependent on cobalamin (Vitamin B12) (113). This insight originated from the observation that Q is present in E. coli tRNAs when grown anaerobically under fermentation conditions with limited iron but not in iron-abundant conditions. This pattern mirrored cobalamin biosynthesis, which is upregulated in anaerobic, iron-limited conditions but downregulated in aerobic or iron-abundant conditions. A genetic approach involving the deletion of hemA, a gene essential for cobalamin production, resulted in Q depletion in tRNA, accompanied by the accumulation of its direct precursor, oQ. This biochemical phenotype could be reversed by supplementing with 5-aminolevulinic acid. The gene responsible for the enzymatic reduction was identified 23 years later in 2011 through a biochemical screen of over 1,700 E. coli Keio collection deletion mutants, and the enzyme named epoxyqueuosine reductase (EC 1.17.99.6) or QueG (114). The QueG protein is homologous to cobalamin-dependent iron-sulfur proteins involved in halorespiration (114, 115). Recombinant QueG from B. subtilis exhibited activity on a synthetic substrate and hypomodified tRNAs from queG-deleted E. coli, requiring a reductant, a redox mediator, and stimulation by cobalamin. Structural studies of the B. subtilis QueG, including one with a bound tRNA-Tyr anticodon stem loop (indicating the positioning of the Q nucleoside in the enzyme’s active site), have led to the proposal of a reaction mechanism involving the formation of a covalent cobalamin-tRNA intermediate (116).

A survey of bacterial genomes encoding the Q synthesis proteins revealed that QueG orthologs were absent in nearly half of these genomes (117), leading to the hypothesis that an alternative to QueG must exist in those genomes. A comparative genomics approach, focused on genomes lacking QueG, identified members of the DUF208 family as candidates for the missing epoxyQ-reductase enzyme and this hypothesis was validated using genetic approaches leading to renaming this family QueH (117). It was initially noted that purified recombinant QueH proteins did not bind cobalamin. Subsequent structural characterization of the Thermotoga maritima homolog confirmed that this protein adopted a novel fold, containing a [4Fe-4S] metallocluster with an intriguing adjacent, coordinated iron metal and an unprecedented mechanism for the reduction of epoxyqueuosine was proposed (118).

An updated analysis of the over 7,000 genomes present in the KEGG database shows that 65% of the genomes encoding both Tgt and QueA also encode the cobalamin-dependent QueG while 25% encode the cobalamin-independent QueH (data extracted from https://www.kegg.jp/kegg-bin/view_ortholog_table?orthology=K07568+K18979+K09765+K00773). A small percentage of genomes (~9%) encoded both QueG and QueH (e.g., Acinetobacter baylyi) suggesting that in these organisms the availability of the cobalamin cofactor could drive the use of one enzyme over the other and that this could be a driving force behind the observed taxonomic distributions of queG and queH. Physical cluster data show that both queG and queH can sometimes be found adjacent to the tgt and queA genes (Fig. 4), reinforcing the strength of gene neighborhood information to link genes and functions (119). Finally, ~10% of the bacteria encoding Tgt and QueA lack homologs of both QueG and QueH, (e.g., in members of the Polaribacter clade). The pathway could stop at oQ in these organisms (e.g., as in E. coli MRE600) (120) or another reductase (specific or nonspecific) is yet to be discovered.

The only Q hypermodification found in bacteria is inserted by a paralog of glutamyl-tRNA synthase

Hypermodification of queuosine, by the addition of sugar or amino acid side chains, has been shown to occur sporadically and only for a subset of tRNAs. For example, galactosyl-Q and mannosyl-Q are only found in mammalian tRNAs (40) and the corresponding enzymes have only been recently identified (121). The only hypermodification identified in bacteria (mainly Proteobacteria) is glutamyl-Q (or GluQ) which is introduced specifically on the Q moiety present on tRNA-Asn by a paralog of glutamyl-tRNA synthase GluQ (YadB) (122124).

Synthesis of archaeosine in DNA and RNA

The archaeosine base has been found in tRNA and DNA. In both molecules, its synthesis starts with the incorporation of preQ0 into the target polymer by a member of the Tgt or DpdA family. preQ0 moiety undergoes subsequent transformation into the archaeosine base through one or two catalytic steps, a process that varies depending on the organism (Fig. 1). The diverse enzymatic systems enabling the conversion of a nitrile to a formamidine moiety showcase instances of both convergent and divergent evolution (125). To date, three non-orthologous enzymatic systems catalyzing this reaction have been identified, all involving enzymes that are paralogs of those involved in queuosine biosynthesis.

Archaeosine synthase, or glutamine: preQ0-tRNA amidinotransferase (ArcS, EC 2.6.1.97) was the first preQ0 aminotransferase discovered nearly 15 years ago in Haloferax volcanii utilizing comparative genomics and genetics (126). ArcS, mainly found in Euryarchaeota, is a paralog of the aTgt enzyme, featuring an additional domain. It was first shown that the Methanocaldococcus janaschii ArcS could catalyze the amidinotransferase reaction in vitro (126). However, follow-up studies revealed a more complex pathway, requiring an additional radical-SAM enzyme, RaSEA (127). In Methanosarcina acetivorans, ArcS was observed to first link the ε-amino group of lysine to the cyano group of preQ0. Subsequently, RaSEA activates the molecule for C-N bond cleavage, resulting in the formation of G+ and 1-piperidine-6-carboxylic acid as a by-product (127). The vast majority (98%) of sequenced Euryarchaeota contain homologs of both ArcS and RaSEA, suggesting that the two-enzyme pathway is the primary route for G+ synthesis in this clade (see data at https://www.kegg.jp/kegg-bin/view_ortholog_table?orthology=K06936+K07557). However, a few Crenoarchaeota, such as Ignicoccus hospitalis KIN4/I and Thermofilum pendens, only encode an ArcS homolog but no RaSEA homolog. In addition, ArcS homologs are found in dG+ insertion clusters from phages lacking any neighboring radical-SAM encoding gene (10). This implies that a direct one-step route may be biologically possible and that further structural and biochemical studies are required for clarification (128).

In Crenearchaeota, the formation of G+ can be catalyzed by distinct enzymes (129). For example, in Pyrobaculum calidifontis, this reaction is catalyzed by the QueF-like (QueF-L) ammonium: preQ0-tRNA aminotransferase (EC 2.6.1.B18). This enzyme is a paralog of QueF that lacks an NADPH-binding site but still makes a thioamide intermediate with the preQ0-modified target tRNA and using NH3 as a donor to make the G+ product (129131). In Sulfolobus solfataricus, the preQ0 amidinotransferase reaction is likely catalyzed by a protein fusion between glutamine amidotransferase (Gat) and QueC (129) but the enzymatic details of this Gat-QueC remain unexplored. Both QueF-L and Gat-QueC homologs have been found in phages involved in the synthesis of the archaeosine base in DNA (10, 14), as discussed above. The molecular determinants that drive the substrate switch from RNA to DNA in these phage enzymes are still unknown.

Biosynthesis of dADG in bacteria: roles of DpdB and DpdC

The paradigm that deazapurine derivatives were found only in tRNA was broken by the detection of dADG and dPreQ0 in bacterial genomic DNA, both originating from the preQ0 precursor (14). In Salmonella enterica serovar Montevideo, a DpdA-DpdB complex integrates preQ0 into DNA through a transglycosylation base exchange reaction, producing dPreQ0, subsequently converted to dADG by DpdC (98, 99) (Fig. 1).

The dpdB gene is detected in 92% of genomes encoding DpdA proteins (132). It belongs to the DNA sulfur modification protein family DndB (IPR017642) which regulates the transcription of phosphorothioate (PT) DNA-modifying genes (133). The ATP hydrolysis function of DndB triggers the disassociation of the DndB-DNA complex, converting DndB-ATP into free DndB. This free DndB can then rebind to promoter DNA, thereby inhibiting transcription. DndB possesses a conserved DGQHR motif in its ATP-binding pocket, which corresponds to the DGQQR motif found in DpdB (14).

Although all DpdABC proteins can bind to DNA, DpdB shows the least DNA binding affinity (98), indicating that its ATPase activity, rather than DNA binding, is key for the base exchange reaction. A recent study on the DpdABC complex revealed that the DpdB ATP hydrolysis activity is essential for the in vitro base exchange reaction of DpdA (99). dpdC is found in 88% of the genomes harboring dpdA (132). DpdC possesses a domain resembling the peroxide stress protein YaaA (PF03883). The X-ray crystal structure analysis of E. coli YaaA revealed a positively charged cleft and a helix-hairpin-helix DNA-binding motif, characteristics shared by DNA repair enzymes (134). This aligns with the observation that DpdC has DNA-binding capabilities (98). In vitro incubation of preQ0-modified DNA with DpdC resulted in the production of dADG-modified DNA, either with or independently of DpdA/B (99). This suggests that DpdC can convert preQ0-modified DNA to ADG-modified DNA without relying on DpdA/B. However, it is important to note the possibility that DpdC first transforms free preQ0 into ADG, which is then inserted into DNA by DpdA to create dADG. This hypothetical pathway has not yet been definitively excluded.

SALVAGE AND RECYCLING OF Q PRECURSORS

The biosynthesis of Q-tRNA imposes a significant demand on cellular energy and resources, involving the utilization of GTP, various metalloenzymes, and cofactors (Table 2). To mitigate this metabolic burden, all eukaryotes and many bacteria opt to salvage precursors rather than synthesize Q de novo. Eukaryotes acquire the queuine base (q), derived from Q-tRNA, through food or microbiota (40). While most bacteria appear to salvage preQ0 and preQ1, some also employ Q salvage routes, particularly among pathogens (7, 135) (Fig. 6). Dedicated transporters facilitating these salvage routes are only beginning to be characterized. Furthermore, given that Q is the only known tRNA modification that can be recycled, it is evident that cellular mechanisms must exist for reusing/recycling Q degradation products such as Q nucleosides and their phosphate derivatives. Their identity and role in Q metabolism are slowly emerging.

Diversity of transporters involved in salvaging Q

Although several strong transporter candidates had been predicted in silico (136, 137), the first experimental evidence for Q precursor salvage was reported in 2017 for the COG1738 aka YhhQ family classified as the vitamin uptake transporter (VUT) family (TC 2.A.88). It has since been reclassified as Queuosine Precursor Transporter or QPTR (135). Genes encoding members of this family are associated strongly with Q pathway genes when analyzed by comparative genomic approaches (Fig. 4). QPTR (UniProt: P37619) from E. coli was shown to transport both preQ0 and preQ1 (Fig. 6A) with a slight preference for the latter (135).

The substrate specificity of Q precursor transporters can be predicted from the presence and absence of Q pathway genes (7, 135): QPTR homologs found in bacteria that harbor queF, tgt, queA, and queG/H are predicted to transport both preQ0 and preQ1, while those in organisms that lack queF but harbor downstream enzymes are predicted to only transport preQ1. Noteworthy, some bacteria only encode TGT and QPTR homologs. The implication, supported by experimental validation of the corresponding Chlamydia trachomatis D/UW-3/CX genes, is that in these organisms, the TGT enzymes and QPTR transporters have switched their substrate specificity for queuine as observed in eukaryotes (7, 135) (Fig. 6B). There are some exceptions, such as Bartonella henselae that encode only TGT and QPTR enzymes that have retained the capacity to use preQ1 as a substrate (138) (Fig. 6D).

The size of QPTR family members ranges from ~19 to 32 kDa and exhibit a predicted six transmembrane helices, and like other transporters, must be located at the inner membrane with a C-terminal inside appendage in the cytosol (135, 139, 140). However, as there are no known structural homologs present in the Protein Data Bank (PDB) for the QPTR family, it has not been possible to determine the residues involved in substrate recognition that could explain the observed shifts in substrate specificity (7, 135).

Energy-coupling factor (ECF)-type transporters are a subfamily of ATP-binding cassette (ABC) transporters (136, 141). While being exclusive to prokaryotes, they consist of two identical copies of cytoplasmic ATPases (A and A′) and two transmembrane units, namely the transmembrane component (T) and substrate-binding component (S). Group I ECF transporters use a dedicated energy coupling module [TAA′; e.g., Rhodobacter capsulatus bioMNY (A, T, S) transports biotin] while group II ECF transporters share the energy coupling module with other S components that transport different molecules (e.g., Bacillus subtilis thiT (S) transports thiamine) (141). Comparative genomic analyses predicted that members of the ECF family transported preQ1: the group I QrtTUVW and the group II ECF-QueT (137). These predictions have been experimentally validated in only one organism, Clostridioides difficile, that encodes three ECF-QueT homologs (7). The heterologous expression of a reconstituted ECF complex in E. coli shows that one of the S components (CD630_16830) can transport preQ1 and Q while another (CD630_2097) could only transport preQ1 (7) (Fig. 6C). The distribution of QPTR/YhhQ and QueT/QrtT homologs in bacteria predicted to be transporting a Q precursor is sporadic (see data at https://www.kegg.jp/kegg-bin/view_ortholog_table?orthology=K00773+K09125+K16787+K16786+K16785+K16923+K01737+K10026+K06920+K06879+K09457+K07566, Release 108.0, October 1, 2023), suggesting that many more bacterial Q precursor transporters are yet to be identified. Candidates identified from gene fusion and physical clustering studies are currently under investigation (de Crécy-Lagard laboratory, unpublished).

Salvage enzymes can regenerate preQ1 and Q from Q/QMP derived from tRNA degradation

Queuosine hydrolases are responsible for catalyzing the hydrolysis of the queuosine ribonucleoside to produce the queuine base and ribose. Two families with Q hydrolase activity have been characterized so far, QueK and Qng1 (7, 142, 143). Initially predicted through the analysis of genes regulated by preQ1 riboswitches and named IunH (82), the first experimentally validated queuosine hydrolase is encoded by a gene under the predicted control of a preQ1 riboswitch in C. difficile (7). Renamed QueK, this enzyme belongs to the Ca++-dependent nucleoside hydrolase family (Fig. 6C). Further analyses, including sequence and structural assessments, identified signature motifs for specifically annotating the QueK subgroup. The queK gene is frequently found in physical clusters, both with ECF-queT genes and yhhQ genes (Fig. 4).

Another family with Q hydrolase activity has recently been biochemically and structurally characterized. Initially designated DUF2419, this family was observed to co-distribute with the eukaryotic TGT enzyme subunit QTRT1 (8). Genetic studies demonstrated the involvement of members from this family in Q salvage in S. pombe and plants, although the precise reaction remained undetermined, despite structure modeling hinting at a potential nucleoside hydrolase role. Recent biochemical and structural characterizations confirm that the homolog from S. thermophilus, S. pombe, and humans, named Qng1, indeed hydrolyzes Q in vitro (142, 143). However, it preferentially targets the Q-5′MP and Q-3′MP substrates (143). The widespread presence of Qng1 homologs in many bacteria, along with their clustering with tgt genes or potential Q hydrolase genes (Fig. 4), suggests potential involvement in Q salvage, recycling, and/or degradation in these organisms. This hypothesis is yet to be experimentally validated and would require a TGT dedicated to queuine incorporation [similar to C. trachomatis (7)] or another enzyme for further breakdown of queuine into preQ1 that could be used as substrate by canonical bacterial TGTs.

While numerous bacteria, including intracellular pathogens like C. trachomatis, directly salvage Q with a TGT with altered specificity from preQ1 to Q (Fig. 6B), a subset of pathogenic bacteria has developed an indirect queuine salvage pathway (7) (Fig. 6C). In those organisms, preQ1 is regenerated from queuine through the action of a recently identified enzyme, queuine lyase or QueL, which belongs to the radical-SAM family (see https://radicalsam.org/explore.php?id=cluster-2-7&v=3.0). The uncommon chemical mechanism involves a radical-mediated cleavage of a C-N bond along with the generation of cyclopentenone compounds.

QueL encoding genes are generally located in an operon with queT and queK but also with yhhQ genes (Fig. 4). These physical clustering associations suggest that queuine is imported from an environment where it is available (e.g., in mammalian blood) and recycled to form preQ1 that, in turn, can be salvaged by most bacterial tgt without necessitating any changes in their sequence for substrate specificity adaptation.

We recently performed a phylogenomic prediction of intracellular organisms that encode the direct q pathway by encoding only a full-length TGT [Fig, 7 of (138)]. We found it was prevalent and predicted in nearly all members of the Dietziaceae, Gordionaceae, and Anasplamataceae families and half of the species in the Borreliaceae and Corynabacteriaceae families that all include major human pathogens. We also did a prediction of organisms that rely on the q indirect pathway (Fig. 7). These are sparse, spread all around the bacterial tree, and are mainly members of the Fusobacteriia, Clostridia, Spirochaetia, and Erysipelotrichia classes. Of note, other yet unidentified Q lyases might exist and current analyses are likely to underestimate the prevalence of the indirect q salvage capabilities.

Fig 7.

Fig 7

Presence and absence of encoded signature Q pathways protein TGT and of indirect q salvage pathways in representative bacterial genomes. A maximum likelihood tree of 10 concatenated ribosomal proteins was created for the species 4,231 complete representative genomes in the BV-BRC database (https://www.bv-brc.org/, version 3.31.12) (144) and the presence (red) or absence (blue) of a full length (>200 aa) TGT encoded protein annotated as Queuine tRNA-ribosyltransferase (EC 2.4.2.29) were noted in the outside circles. Genomes that encode the indirect Q salvage pathway (QueA, QueG/H, QueK, and QueL) are noted in green (outer circle). For better visualization, the branches are grouped and colored by phyla or clade. The branches were colored by bootstrap support value. The tree was visualized using the iTOL platform https://itol.embl.de/ (version 6.8.1) (145). The branch length scale bar indicates the evolutionary distance of 0.5 amino acid substitutions per site.

PHYLOGENETIC DISTRIBUTION OF THE Q PATHWAY IN BACTERIA AND ARCHAEA

bTGT is the signature enzyme of the Q pathway, as it is the enzyme responsible for the base exchange. Hence, when a given genome harbors a tgt gene, it can be inferred with confidence that the corresponding organism salvages or synthesizes Q. However, one should ensure that tgt annotations are correct, as otherwise biological inferences will be erroneous. Common annotation issues arise due to the presence of tgt gene fragments and the miscalling of dpdA genes as tgt genes. The fragmentation of the tgt gene is commonly observed in organisms such as Bartonella quintana that have lost the Q pathway (138). Analysis of 4,245 complete representative genomes in the BV-BRC database (version 3.31.12) revealed they encoded 3,714 proteins annotated as queuine tRNA-ribosyltransferase (EC 2.4.2.29). In all, 50 (~12%) of those were shorter than 260 amino acids in length and further analyses showed these were tgt gene fragments, like in B. quintana (138). bDpdA proteins are currently annotated as “archaeosine tRNA-ribosyltransferase (EC 2.4.2.-) type 5” in the BV-BRC database so they cannot be mistaken for bTGT. In other databases such as Uniprot, DpdA proteins are sometimes incorrectly annotated as queuine tRNA-ribosyltransferase (see Verru16b_03187, Uniprot ID A0A1D8AZ010) though the two families can be readily differentiated based on sequence similarity and gene neighborhoods (14).

The analysis of the tgt gene distribution across bacterial kingdoms shows it is uniformly spread, with independent losses in various clades (Fig. 7). This supports the hypothesis that Q was present in the common ancestor of bacteria. The clades that have lost tgt are the Actinomycetiae class, the Tenericute phylum, and the Lactobacillaceae family. For a few individual organisms in these groups, the absence of Q in tRNA has been experimentally validated (112, 146). Across the broader phylogeny, the loss of tgt is sporadic. It occurs mainly in symbionts or intracellular pathogens with minimal genomes even if the Q synthesis pathways are retained in many of such organisms, as seen in several Buchera or Rickettsia species (147, 148).

It was previously thought that the Q modification was only found in bacteria and not in archaea. However, it was recently reported that the entire Q pathway is encoded in some archaea, notably in Woesearcheaota genomes (149). A current limitation of these observations is the lack of experimental validation. To date, Q has not been observed in any archaeal tRNA.

DIVERSITY OF DEAZAPURINE-DERIVED NATURAL PRODUCTS

Microorganisms are prolific producers of a diverse array of natural products (NPs) (150) also referred to as secondary or specialized metabolites, which often confer fitness advantages in their environments (151). Exploring the biosynthetic capacities of the microbial world has revealed numerous anticancer (152), anti-inflammatory, photoprotectant (153), and antibiotic (154) NPs, including those harboring 7-deazapurine moieties (155). Deazapurine-derived NPs have recently attracted increasing scientific interest due to their diverse chemical structures and biological activities (156158). The distinctive chemistry and biology of these nucleosides and nucleoside-like compounds offer an intriguing path to investigate their range of structures, biosynthetic pathways, and evolutionary histories.

Pyrrolopyrimidine-derived NPs exhibit a remarkable diversity, still far from being fully described. They are exemplified by the nucleosides toyocamycin from Streptomyces toyocaensis and S. rimosus, as well as sangivamycin from S. rimosus, that exhibit diverse bioactive properties including antibiotic, antitumor, and antiviral activities (16, 18). Tubercidin, discovered in S. tubercidis, also demonstrates versatile characteristics, including antimicrobial, antiparasitic, antiviral, and antitumor properties (15). Beyond these exemplars, the broader array of deazapurine-derived compounds reveals a myriad of functionalities, underscoring their potential as promising sources for novel therapeutic agents (Table 1).

The central precursor, preQ0, serves a dual role, being not only a critical participant in DNA and tRNA modifications but also a precursor for putative NPs. In fact, preQ0 itself has shown anticancer properties (159), potentially contributing to the bioactivity seen in its downstream NPs. Although sangivamycin, toyocamycin, tubercidin, huimycin, kanagawamicin, echiguanine, cadeguomycin, and dapiramicin share a common pyrrolopirimidine core, deazapurine-derived NPs have notable chemical diversity with examples of distinct structural modifications. Commonly, deazapurine-derived metabolites attach a ribose moiety to their core (17). The biosynthesis of pyrrolopyrimidines is carried out by a series of reactions, which is initiated with the conversion of GTP to preQ0, orchestrated by pivotal genes including folE, queD, queE, and queC as previously described. For example, the toyocamycin biosynthetic gene cluster harbors dedicated homologs from preQ0 biosynthesis. ToyD catalyzes the reaction of GTP to H2NTP (as in GTP cyclohydrolase I; FolE), ToyB catalyzes H2NTP to CPH4 (as in QueD), ToyC catalyzes CPH4 to CDG (as in QueE), and ToyM catalyzes CDG to preQ0 (as in QueC). From here, further tailoring occurs depending on the gene content of the biosynthetic gene cluster (1, 28). For example, the presence and/or regulation of other biosynthetic genes can drive the conversion of one NP to another, as in Streptomyces rimosus where both toyocamycin and sangivamycin are produced through a common biosynthetic pathway (17). Sangivamycin emerges as a downstream product of toyocamycin under certain regulatory conditions that, when met, modify toyocamycin by a nitrile hydratase (TNHase), introducing a non-heme iron or non-corrin cobalt ion to amide nitrogen and cysteine sulfurs (1). In another example, huimycin is produced by preQ0 methylation by HuiC (a SAM-dependent methyltransferase) before attaching N-acetylglucosamine through the glycosyltransferase HuiG (28). These tailoring variations represent only a subset of what has been described and what is yet to be discovered. As with other NPs, their unique chemistries can influence their interactions with enzymatic, biological, and ecological processes.

While many deazapurine-derived natural compounds have been identified (Table 1), their biosynthetic pathways remain elusive. This renders these compounds “orphans” in a sense, as the genes responsible for their biosynthesis have not yet been characterized. In the MIBiG (Minimum Information about Biosynthetic Gene Cluster) database (160), only tubercidin, toyocamycin, sangivamycin, and huimycin have their biosynthetic gene clusters (BGCs) partially annotated. In part, this is due to the lack of their biosynthetic rules being incorporated in conventional genome mining algorithms, such as antiSMASH (161). This gap underscores the need for orthogonal approaches to unveil the genetic diversity of these intriguing deazapurine-derived NPs, de-orphan known NPs by linking them to their cognate BGCs, and mine genomes for BGCs predicted to be novel examples of the class. As more putative deazapurine BGCs are identified, biosynthetic genes must be carefully annotated to properly predict their enzymatic functions. Comparative genomic analyses have uncovered genetic signatures associated with DNA (dpdAs genes) (14) and tRNA modifications (bacterial tgt genes and their homologs in archaea arcTGT/arcS) (91). If we exclude regions that encode these characteristic DNA and tRNA-modifying enzymes, the remainder may be BGCs responsible for generating deazapurine-derived NPs.

FUNCTION OF DEAZAPURINES

Functions of Q and G+ in RNA

Complex roles of Q in decoding accuracy and efficiency

Q is exclusively found at position 34 of the anticodon stem-loop of the four tRNAs with GUN anticodons that decode the NAC/U codons encoding His/Tyr/Asn/Asp, all located in split codon boxes (Fig. 8). Over the last 20 years, a combination of studies in different organisms using +1 or −1 frameshifts (162, 163), amino acid misincorporations (11, 12, 164), stop-codon readthrough assays (13, 165), and sense codon reassignment analyses (166) have been employed to better understand Q function. Decoding of reporter genes with enrichments of C or U ending Q-dependent codons (13, 167169), structures focusing on codon/anticodon interactions in the ribosome decoding sites (170, 171), evolutionary analyses (172), and ribosome profiling studies (12, 173) can be combined in a model where the presence of Q can stabilize or destabilize the interactions of the Q34U35N36 anticodon with N1A2U/C3 codons in the ribosome A site. This, in turn, homogenizes the translation rates of C or U ending codons and modulates the efficiency of second codon mismatch in both directions near cognate recognition (for Cys and Gly) (12, 164). The specific role of the Q modification in translation speed and accuracy does, however, vary greatly between tRNA isoacceptors and organisms (Fig. 8). For example, RiboSeq provides a genome-wide measurement of translation speed at every codon (174). Applied to mammals, it suggests that Q increases the speed or efficiency of decoding at all NAC/U codons but with a marked difference in ratios: the NAU codons are more dependent on the Q modification than are the NAC codons (173) as predicted from the pioneering studies of Grosjean and Nishimura that found that Q-containing tRNAs bind better to U than to C codons (4, 175). In Schizosaccharomyces pombe however, Q increases the translation speed of the codons G/CAC but not of the G/CAU Asp and His codons, whereas it decreases the translation speed of the A/UAU but not of the A/UAC Asn and Tyr codons (12). These results are consistent with the theory of Grosjean and Westhof (170) where codon-anticodon strength is equilibrated across the genetic code and Q plays different roles for intermediate strength codons (e.g., Asp, His) compared to weak codons (e.g., Asn, Tyr). The only bacterial RiboSeq study of Q-deficient mutants was recently performed in Vibrio cholerae where the absence of Q led to a more efficient translation of UAU (Tyr) and GAU (Asp) (13). These results contrast with a recent analysis of EGFP reporter genes recoded with only C or U ending Q-codons. Here, in an E. coli queF mutant, the U-ending codon reporter gene is translated less efficiently (20%) (168). Of note, the distinct types of Q-dependent U ending codons were not differentiated in this study. Some of the differences observed between codons could be caused by the presence of the hypermodification of Glu-Q on tRNA-Asp (166). It has recently been observed that Shewanella glacialimarina phage 1/4 influences the level of Q in tRNA throughout the phage infection cycle, lowering it in the early stage and gradually increasing it (176). It is proposed to help in the translation of the genes expressed late in the cycle to have a preferred GUA codon for tyrosine decoding.

Fig 8.

Fig 8

Effects of Q on decoding speed in different organisms. Decoding of GUN codons by Q-modified tRNAs in ribosomal A sites. The hydrogen pairing pattern of the wobble base is affected by the presence/absence of Q and decoding speed has been measured by RiboSeq in three organisms to date: H. sapiens (173), S. pombe (12), and V. cholerae (13).

Although Q at position 34 clearly fine-tunes the efficiency and accuracy of translation, its role varies with each of the four Q-modified tRNAs and one must be cautious not to generalize findings from one organism to another without further experimental validations. Indeed, the role of Q is a consequence of an evolutionary adaptation to all other components of the translation machinery, including the presence of other tRNA modifications or the codon usage that is specific to every species. Q is predicted to have been present at the origin of bacteria (177), yet organisms can adapt to life without this modification, as it has been repeatedly lost along the tree of life (Fig. 7). More studies are required to understand the role of Q in translation in different bacteria and in the few Archaea where it is present. This could include RiboSeq analyses of Q+ and Q- in E. coli and B. subtilis, to better link the pleiotropic phenotypes caused by Q deficiency (discussed below) and the underlying molecular mechanisms.

Q is rarely a determinant for other enzymes interacting with tRNAs

Q functions as a determinant for Dnmt2, the enzyme responsible for inserting the m5C38 modification in tRNAs in Eukaryotes (178). However, it is not known to act as a determinant for any modification enzyme in bacteria. In E. coli, tRNA extracted from a Q- strain does not show any variations in the levels of other tRNA modifications (unpublished data from Dedon and de Crécy-Lagard). Although it has been reported that tRNA-Tyr in E. coli is less efficiently charged when Q34 is replaced by C34 (179), the absence of significant growth defects in an E. coli tgt mutant suggests that this defect may not be relevant in vivo (180, 181). Q has been implicated in protecting against ribonuclease cleavage in mammals (182), but this protective role has not been observed in bacteria to date. Ribotoxin E5 specifically cleaves Q-modified tRNAs in vivo, yet Q is not a determinant for recognition (183).

Pleiotropic phenotypes linked to Q or GluQ deficiency vary across organisms

As listed in Table 3, the absence of Q leads to a wide variety of phenotypes across bacteria. Until recently, it was thought that Q was dispensable as several important model organisms such as Saccharomyces cerevisiae, Arabidopsis thaliana, or Mycoplasma genitalium have lost the enzymatic capacity for the modification (184) and the tgt mutant of E. coli does not show any growth defects in most conditions (180). The only notable phenotype was the virulence deficiency of a Shigella flexneri Q- strain caused by a reduced expression of the virF regulator (185). Several studies in the last 3 years have now changed this view with roles in oxidative stress resistance, biofilm formation, and metal homeostasis emerging as common themes.

TABLE 3.

Phenotypes linked to Q genes deficiency or overexpression in bacteria

Organism Phenotype References
Oxidative stress
Escherichia coli tgt mutant is slightly more sensitive to oxidative stress. (181)
Streptococcus thermophilus tgt mutant is more sensitive to oxidative stress (186)
Vibrio cholerae Translation of regulator of oxidative stress rtxA is increased in tgt mutant (13)
Metal homeostasis
Escherichia coli tgt mutant is more resistant to cobalt and nickel and more sensitive to cadmium (181)
Acinetobacter baumanii queD and tgt expression induced by metal sequestration enzyme and metal limitation reduces Q levels in tRNA (57, 187)
Arthrobacter viscosus Overexpression of the queC gene in E. coli confers aluminum resistance (188)
Erwinia amylovira yhhQ and queF overexpressed in high copper (189)
Neisseria meningitidis queC and queF induced by zinc limitation (190)
Pseudomonas putida queF/cinQ expression is induced by copper but mutant does not give any copper sensitivity (191)
Agrobacterium tumefaciens QueF is highly induced by manganese limitation (192)
Virulence
Shigella flexneri Reduced virulence in tgt mutant because of decreased levels of VirF (193)
Rhizobium meliloti Mutants in queC, queF, and tgt are deficient in triggering cytoskeleton modification in uninvaded Hela cells (194)
Escherichia coli Biofilm and cell aggregates diminish in a ΔqueF and iboth with the addition of LPS l levels increase when some of the Q synthesis genes are overexpressed (168)
Miscellaneous
Escherichia coli tgt mutant has fitness defect in the stationary phase
Growth defect with streptomycin but not ampicillin or spectinomycin
(180)
(181)
Streptococcus gordonii queA mutant has fitness cost in the stationary phase (195)
Vibrio cholerae TnSeq data show that mutations in tgt and queADEF genes confer sensitivity to sub-MIC tobramycin (196)
Bacillus subtilis Overexpression of queCDEF genes in VBNC cells and queG mutant more sensitive to Kan in those cells (197)
Pseudomonas simiae TnSeq data show that queA mutants are deficient in deoxyribose catabolism (198)
Staphylococcus epidermidis queF, queH, and tgt induced by pH (199)
Bacillus subtilis Sporulation and biofilm reduced in ΔqueF mutant (168)
Pseudomonas putida Growth inhibition of E. coli by P. putida increased when queF is overexpressed (168)

Two types of oxidative stress phenotypes have been linked to Q deficiency in bacteria to date. The first is a mild sensitivity to oxidative stress as seen in S. thermophilus and E. coli (Table 3) that is not yet understood at the molecular level, but it is possible that this sensitivity may be due to a general response to protein aggregation triggered by conditions that affect translation speed, as previously discussed (181). This mechanism tying Q and oxidative stress could be conserved between kingdoms (200, 201). The second is a Modification Tunable Transcript (or MoTT) regulatory mechanism (202, 203), as seen in V. cholerae where the translation of the rtxA gene is decreased under Q excess because it is enriched in tyrosine encoding TAT codons (13). In V. cholerae, RtxA inactivates the main oxidative stress activator SoxR. High Q levels would lead to an increased oxidative stress response than would low Q levels. This response is required to resist aminoglycosides and would explain why the V. cholerae Q- mutants are more sensitive to tobramycin (196). It was also shown in V. cholerae that the transcription of the tgt gene is regulated by the central regulator CRP and by the stringent response. This leads to an intricate regulatory model where stress increases the levels of Q, leading to decreased levels of TrxA and the induction of oxidative stress response through the activation of SoxR. This second mechanism appears species specific as the regulation observed in V. cholerae is absent in the fellow Enterobacteriaceae E. coli.

The role of metals in Q synthesis was first observed by one of the pioneers in the study of Q synthesis, Helga Kersten, who found that the presence of iron and B12 in the media affected the ratio of oQ/Q in S. typhimurium and E. coli (113) because the last enzyme in the Q pathway in this organism is the B12-dependent iron-sulfur cluster enzyme QueG (114). Additional sporadic observations have linked metal and Q synthesis genes over the years (Table 3). Most of the enzymes in Q synthesis are metal dependent (57, 204) (Table 2). FolE, QueD, QueC, and Tgt are zinc-dependent enzymes and QueE is an iron-dependent radical-SAM enzyme. Comparative genomic analyses of genes in the Zur regulon predicted that Q might be required under zinc limitation in certain organisms (50, 205). In addition, metal limitation was shown to lower Q levels in Acinetobacter baumanii and induce the expression of Q biosynthesis genes (57, 187), but it is yet to be shown that these phenotypes are part of a regulatory circuit. The E. coli tgt mutant is more resistant to nickel and cobalt and more sensitive to cadmium (181). The sensitivity to nickel is possibly caused by a lower expression of the nickel transporter encoding operon nikABCDE when Q is absent, but the underlying mechanism has not been elucidated. One hypothesis that is yet to be experimentally validated is that NikR the repressor is enriched in TAT codon and could be efficiently translated in the absence of Q (13).

A recent study combining proteomic, codon-usage analyses, and phenotypic validations in several model bacteria reported reduced biofilm formation in queF mutants of both E. coli and B. subtilis (168). The study also linked several other virulence- related traits/proteins to Q deficiency and/or to an enrichment in Q-dependent U ending codons. The authors proposed that Q could have a general role in regulating bacterial virulence by modulating the translation of virulence genes. This hypothesis is yet to be validated with RiboSeq data, recoding of target genes, and identification of the signal(s) that would modulate Q levels in conditions where virulence genes would be differentially expressed.

Q biosynthesis is complex, requires several metals (as discussed above), draws on many building blocks from central metabolism (GTP, SAM, ATP, cobalamin) (Table 2), and shares intermediates with essential cofactors such as tetrahydrofolate (Fig. 3). Thus, Q could be used to monitor many aspects of cellular physiology (206). We anticipate that the next few years will reveal more examples of regulatory roles of the Q modification in bacteria as more phenotypes get reported and their molecular mechanisms get fully dissected (as done to date only in the case of the V. cholerae RtxA). This will likely be accelerated as new methods for Q detection become more accessible (207212) and more RiboSeq data sets of Q-deficient cells are generated.

We emphasize, however, that based on the current known cases, MoTT-dependent regulations when identified will be very species-specific. In many organisms, the absence of Q might just lead to a mild increase in amino acid misincorporations and/or aggregation phenotypes (12) that could become problematic under additional proteotoxic stresses, as with many other tRNA modifications deficient cells (213).

The precise role of the hypermodified Glu-Q tRNA-Asp is not fully understood, but it has been associated with stress resistance, observed through its co-transcription with the stringent response-regulated gene dksA in many gammaproteobacteria, including a Shigella flexneri mutant that lacks the Glu-Q modification that demonstrates increased sensitivity to osmotic stress (214).

Q and its precursors are micronutrients

All eukaryotes salvage the queuine (q) base derived from bacterial Q directly from the microbiota or indirectly through the diet (40). The last 10 years have seen a reemergence of Q as a micronutrient important for human health (215, 216), particularly for optimal brain function (217, 218). Even if it is yet to be explored, queuine should also be an important micronutrient for the health of most plants but crucifers that have lost Q biosynthesis genes (8). Another unexplored area is how bacteria compete for Q precursors particularly in specific niches. Different bacteria of the microbiota can make Q de novo, be preQ1/preQ0/Q scavengers, or have lost all genes of the pathway (7, 168). This may encourage competition between sympatric organisms for Q as is observed for B vitamins (219, 220). Indeed, Q supplementation leads to an increased level of α-diversity among intestinal microbiota (42). The amount of Q produced and utilized by the gut microbiome will have health consequences on the host that are just starting to be appreciated. For example, the gut microbiome is enriched in Q-producing bacteria in obese mice (41) or chickens raised outdoors compared to indoors (221).

Structural role of G+ in tRNA stability

Since its discovery in the early 1990s (5), the proposition that G+, primarily located at position 15 and occasionally at 13 in archaeal tRNAs (91), plays a structural role has been substantiated. This hypothesis has been confirmed both in vivo and in vitro across various archaeal models. Experiments involving random and targeted deletion of G+ synthesis genes in thermophilic and mesophilic archaea (9, 91, 222, 223) demonstrated that the absence of G+ in tRNAs resulted in a thermosensitivity (Ts) phenotype in the hyperthermophile Thermococcus kodakarensis but not in the mesophiles Haloferax volcanii and Methanosarcina mazei. A comparison of thermal denaturation profiles between fully modified T. kodakarensis tRNAs and naked transcripts with or without G+ revealed that the presence of G+ protected from melting, particularly in the transcripts. These findings, obtained by two independent laboratories, collectively underscore the role of this modification in adapting to high growth temperatures. Nevertheless, the conservation of this modification in most sequenced Archaea (i.e., not only limited to thermophiles), the absence of a Ts phenotype in G+-deficient M. mazei, and the cold-sensitive phenotype of G+-deficient H. volcanii suggest that the roles of G+ might extend beyond thermotolerance.

Function of deazapurines in DNA

dADG is used by bacteria to discriminate self from non-self

The bacterial dpdABC genes that modify genomic DNA with dADG are located in genomic islands called dpd islands. These islands contain a consistent set of nine genes (DpdABC-DpdEGIJKD) with minor variations, that are sporadically distributed around the bacterial phylogenetic tree and most certainly spread through horizontal gene transfer (14). Classical transformation efficiency experiments revealed that plasmids extracted from cells expressing dpdABC and subsequently modified with dADG were more efficiently transformed in host cells harboring dpdEFGHIJ than plasmids extracted from cells expressing dpdAB and hence modified with preQ0 or unmodified plasmids. These differences in transformation efficiencies disappeared after the disruption of any of the dpdEGIJKD genes (98).

These features bear a resemblance to the well-characterized “self-nonself discrimination” mechanism of methylation-based Restriction-Modification (R-M) systems. R-M systems, typically composed of a methyltransferase (MTase) and a restriction endonuclease (REase), are considered primitive immune systems in bacteria, protecting against bacteriophages or other invading DNA. A similar defensive feature suggests that DpdABC-DpdEGIJKD constitutes a novel dADG-based R-M system, recognizing the dADG status of invading foreign DNA, such as plasmids (14, 98). DpdABC modifies DNA with dADG, while DpdEGIJKD acts as the cognate restriction enzymes that recognize foreign DNA lacking ADG modification and may initiate its cleavage.

Deazapurines are used as anti-restriction strategies by phages

The suggestion that 7-deazaguanine modifications might impede digestion by restriction enzymes was proposed upon their discovery in Enterobacteria phage 9g (14), given the observed resistance of the phage’s DNA to initial digestion (224). Subsequent comprehensive testing by New England Biolabs on this phage DNA revealed that only restriction enzymes interacting with guanines would be inhibited (225). Notably, EcoRV, known to have guanine in its recognition site, was entirely inhibited when the guanine was replaced by a 7-deazaguanine (226). Further exploration involving various bacteriophages with 7-deazaguanine derivatives confirmed that all natural 7-deazaguanines protect against digestion (6, 10). A connection between the recognition sequence of Enterobacteria phage CAjan (“GA” and “GGC”) and the range of inhibited restriction enzymes was established, indicating that only enzymes with recognition sites containing “GA” are affected (100). In addition, dPreQ0 in Pseudomonas phage iggy was found to protect against Cas9 digestion and potentially other DNA-degrading defense systems (101).

Function of deazapurine small molecules

In therapeutic contexts, toyocamycin, while potent against tumors, also demonstrates substantial host toxicity (3, 17). Tubercidin and its analogs are potent antimicrobials, particularly against Candida species and Mycobacterium tuberculosis (227). Sangivamycin and echiguanines A-B exhibit high cytotoxicity and inhibit protein kinase C. The mechanism of action for echiguanines could be involved in phosphatidylinositol turnover and with cell surface tyrosine kinase receptors [members of the platelet-derived growth factor (PDGF) family]. On the other hand, sangivamycin might have two possible mechanisms of action: cell death by apoptosis [i.e., protein kinase C (PKC) and c-Jun NH2-terminal kinase (JNK) activation] and by growth arrest [i.e., cyclin-dependent kinase (CDK) inhibition, DNA damage, and p21 induction) associated with multidrug-resistant breast cancer lines (16, 22). Collectively, these compounds intervene in cellular processes associated with adenine nucleosides, exhibiting diverse effects rather than targeting a singular cellular entity or process (1). As such, these compounds showcase various modes of action, leading to hypotheses that their impacts on cellular metabolism can occur at multiple levels. Despite the paucity of identified self-resistance mechanisms for 7-deazaguanines in the producing organisms, the extracellular release of these molecules in the culture might suggest that it safeguards the producing strains (1). However, there is no empirical evidence for this assumption and further analyses are required to validate this hypothesis.

DETECTION AND BIOTECHNOLOGICAL USES OF DEAZAPURINES

7-deazapurine detection methods

Gel-based assays coupled with Northern blotting that separate and detect Q-modified tRNA through the addition of acryloylaminophenyl boronic (APB) acid have made Q detection accessible to numerous laboratories since the 1980s (228). However, the utility of APB gels is confined to Q detection alone. Traditionally, the detection and quantification of other deazapurines relied on liquid chromatography-mass spectrometry (LC-MS) (229). Although lacking single nucleotide resolution, the heightened sensitivity of these methods now allows the detection of modifications with minimal starting material, making them indispensable when combined with synthetic chemistry for the discovery of new modifications (230, 231).

Nanopore technology was successfully used to detect preQ0 in phage DNA (100, 101) but next-generation methods have also recently been developed to detect Q, preQ1, and preQ0 in RNA. To be detected at the single nucleoside level, chemical treatment (211) or labeling of the tRNA by a non-natural preQ1 derivative (232) must be performed before sequencing, even though some polymerases have been found to make more mistakes in the presence of Q and hence can be used to detect the modification by mapping the errors (208, 212). Direct sequencing by nanopore was recently used to detect Q and preQ1 in tRNA (208, 212). This method does not require prior treatments or labeling but does require comparing modified to unmodified samples, reminiscent of bisulfide sequencing strategies to measure DNA methylation patterns. The toolbox for deazapurine detection is expanding with single base resolution methods and although many are not yet cost-effective for tRNA modifications, they may be critical to survey the locations of deazapurine modifications in DNA.

Biotechnological uses of deazapurines and their biosynthetic enzymes

At the frontiers of unique biochemistry and microbial defense systems, deazapurines and their biosynthetic enzymes have emerged as versatile molecular tools in green chemistry, genetic engineering, and pharmaceutical development. Harnessing the enzymatic reduction of nitrile to primary amine would be of great interest for green chemistry applications (233). This reaction is crucial in synthetic chemistry, traditionally involving harmful reducing agents and complex blocking/deblocking processes (234237). Exploring enzymatic nitrile reduction offers an alternative to synthetic methods. Despite QueF enzymes typically favoring preQ0 as a substrate, advancements in enzyme design raise the possibility of using QueF as a template for creating versatile nitrile reductases that accept a broader range of substrates. A systematic screening of QueF type I, II, and QueF-like enzymes could identify homologs that could accept non-canonical substrates and serve as the basis for directed evolution approaches that could greatly expand our chemical/enzymatic toolkit (238).

Wild-type bacteria encode multiple defense systems against mobile genetic elements (MGEs), such as plasmids, transposons, and bacteriophages (239). In many cases in both bacteria and archaea, the first line of defense is provided by restriction/modification systems. Several of these MGEs are employed in genetic engineering applications, including plasmids in complementation assays, transposons in mutagenesis, and strategies against pathogens (e.g., phage therapy). Thus, it has been proposed to use 7-deazaguanine DNA modifications as a shield to protect against a wide variety of restriction enzymes (225) [and potentially other defense mechanisms (101)] during the initial entry of the MGE (International Patent Application No. PCT/US20/21886). Consequently, the 7-deazaguanine-modified MGE would withstand the first defensive barrier of the target bacteria, avoiding degradation by an organism’s restriction-modification systems.

As we discussed before, pyridopyrimidines derivatives exhibit utility as anticancer, antiviral, and antibiotic agents, which can lead to the discovery of deazapurine-like NPs or serve as a scaffold for synthetic or mimics NPs (240, 241). For instance, some FDA-approved drugs such as ribociclib (242) fododesine, and ruxolitinib contain 7-deazapurine moieties, which are being used to treat breast cancer, leukemia, and pleural mesothelioma, respectively (243). In another study, using 7-deazapurine, as an alternative to conventional purine structures, researchers were able to synthesize compounds that exhibited significantly enhance characteristics. These compounds not only displayed greater potency but also showcased an increased level of selectivity. Moreover, the resultant compounds exhibited favorable pharmacokinetic properties, making them highly promising candidates for applications in the context of cardiac troponin I-interacting kinase (TNNI3K) (244).

7-Deazapurines act as important analogs of biogenic purine nucleosides. Upon replacing the N7 atom with carbon, these compounds gain increased electron density, enabling diverse substituents at the C7 position. This increased electron density makes 7-deazapurines particularly versatile in terms of their chemical functionalization. These modifications may be crucial in the development of compounds with enhanced biological activity, especially in the context of interactions with nucleic acids like DNA and RNA (157).

CONCLUSIONS

The recent surge in 7-deazapurine-related research prompted this review. Following the initial exploration of Q and G+ in the 1980s, the field experienced a downturn by the end of the century, with only a few papers annually. However, the advent of whole-genome sequences, coupled with the rising interest in epigenetics and epitranscriptomics, has sparked a renaissance, resulting in over 20 papers per year. This body of knowledge remains however confined to specialists, as pathway databases inadequately capture it. Comparison among Gene Ontology (245), KEGG (246), and MetaCyc (247) (Table 4) reveals that while Q synthesis is well documented, archaeosine synthesis is incomplete, and DNA modification pathways, even for those like dG+ published seven years ago, are nonexistent. Consequently, recent papers may overlook that preQ0-related clusters in certain phage genomes pertain to DNA modification genes, not tRNA modification genes (248).

TABLE 4.

Comparison of deazapurine-related pathway annotations in knowledge databases

DatabAse objects Pathway names Notes
Metacyc a Good coverage of RNA modifications with a few missing enzymes or intermediates, No DNA modifications, one natural product.
 PWY-6700 Pathway: queuosine biosynthesis I (de novo) Q and GluQ synthesis complete
 PWY-8105 Pathway: queuosine biosynthesis II (queuine salvage) Direct q salvage, no transporter, no Qng1
 PWY-8106 Pathway: queuosine biosynthesis III (queuosine salvage) Indirect pathway from C. difficile, complete with transporter (RXN-21036)
 PWY-6703 Pathway: preQ0 biosynthesis Missing ADG intermediate
 PWY-6720 Pathway: toyocamycin biosynthesis Complete
 PWY-6711 Pathway: archaeosine biosynthesis I ArcS missing raSEA
 PWY-7923 Pathway: archaeosine biosynthesis II QueF-like but missing Gat-QueC
KEGGb All pathways nonexistent except for preQ0
 map00790  Folate biosynthesis Only preQ0 synthesis as part of the folate pathway map, no ADG intermediate
Geneontology.orgc Queuosine synthesis is well described; the other pathways are inexistent or fragmentary
 GO:0046116 Queuosine metabolic process From GTP to Q in Bacteria, missing FolE2, and QueH
 GO:0002927 Archaeosine-tRNA biosynthetic process Very fragmentary in terms of enzyme captures
 GO:1990397 Queuosine salvage Just QPTR no other transporter gene
 CHEBI:134606 Toyocamycin Not linked to genes
 CHEBI:45075 preQ0 Not linked to genes

This review addresses lingering questions that will drive future research. Although pathways for preQ0-derived molecules are better understood than ever, validation or discovery of some pathway enzymes is still required (Fig. 1). Many transporters for Q precursors are unidentified, and comprehension of how Q precursors circulate within microbial communities or between communities and hosts remains inadequate. The newfound associations between Q, considered a quasi-vitamin source (215), and various human diseases (249, 250) have spurred investigations into Q synthesis in the human microbiota (251), but more studies are needed to fully understand the interplay of the gut microbiome and the diet in supplying the Q precursors to the human host. Q degradation is unexplored, and like other modified ribonucleosides (252, 253), Q may serve as a carbon or nitrogen source.

Furthermore, knowledge about the regulation of 7-deazapurine synthesis genes in bacteria without riboswitches is nonexistent. Mechanistically, the observed differences in how Q affects the decoding speed of U or C-ending GUN codons between species (Fig. 8) lack understanding. In addition, the prevalence of regulatory circuits using Q, as described in V. cholerae (13), and whether preQ0 and preQ1 have roles as signaling molecules remain unknown.

Comparative genomics of the enzymatic machinery involved in the production of deazapurine NPs can unveil valuable insights into the evolutionary forces that shape these pathways, including patterns of conservation, duplication, and adaptation (151, 254). Importantly, 7-deazapurines are not confined solely to bacterial sources. Their occurrence in sponges and algae underscores their broader distribution (1), perhaps suggesting an ancient ancestral origin and potential biological interactions mediated by this privileged structural motif that has perpetuated it across diverse evolutionary lineages. The structural diversity seen in these compounds might reflect an evolutionary divergence that has occurred to adapt to distinct niches.

The potential use of 7-deazaguanine derivatives in biotechnology is a promising avenue for exploration. These DNA modifications, protecting against restriction enzymes and possibly other defense systems, offer opportunities for research involving naturally isolated bacteria that are challenging to genetically manipulate. Alternatively, bacteriophages modified with 7-deazaguanine emerge as strong candidates for phage therapy, given their increased likelihood of surviving the initial infection round (International Patent Application No. PCT/US20/21886).

ACKNOWLEDGMENTS

We thank Mythili Merchant for the exploratory analysis of the Q pathway in Archaeal genomes.

This work was funded in part by the National Institutes of Health (awards GM070641 to VdC-L and SB and GM146075 to VdC-L) and by the National Science Foundation (award MCB-1817942 to MC). The funders had no role in study design, data collection and interpretation, or the decision to submit the work for publication.

Biographies

graphic file with name mmbr.00199-23.f009.gif

Valérie de Crécy-Lagard, after obtaining a bachelor’s degree at Ecole Polytechnique in 1987 and a Ph.D. in microbial genetics at the Pasteur Institute (Paris) in 1991, worked in diverse academic and industrial settings using the power of bacterial genetics to study primary and secondary metabolism as well as mechanisms of regulation by proteolysis. In the past 25 years, her work has focused on combining comparative genomic analysis with experimental methods to discover the function of the many "unknowns" found in sequenced genomes, first at the Scripps Research Institute and then, since 2004, in the Microbiology and Cell Science Department at the University of Florida where she is now a Distinguished Professor. This led to solving many long-standing mysteries, particularly in the fields of coenzyme metabolism and transfer RNA (tRNA) modification. In parallel, she collaborates with biotech groups on using long-term cultures to evolve microorganisms with specific traits.

graphic file with name mmbr.00199-23.f010.gif

Geoffrey Hutinet holds a bachelor's degree from the University of Paris XI in France, earned in 2011, and completed his Ph.D. at INRAE of Jouy-en-Josas, France, in 2014. During his Ph.D., he delved into bacteriophages and protein-DNA interactions. In early 2015, Geoffrey joined the University of Florida as a post-doctoral fellow, rising to the position of Biology Scientist III in 2019. His work initially focused on a newly discovered system inserting deazaguanine into bacterial DNA and shifted to a similar system in bacteriophages that he identified. Geoffrey Hutinet played a pivotal role in uncovering various types of deazaguanine DNA modifications and deciphering the pathways leading to them. Currently serving as a Visiting Assistant Professor at Haverford College, he continues his research on deazaguanine modification systems and expands to explore other modification systems, aiming to comprehend their roles in virus-host interactions.

graphic file with name mmbr.00199-23.f011.gif

José D. D. Cediel-Becerra earned his Summa Cum Laude bachelor’s degree in biology from the Industrial University of Santander in Colombia in 2021. His honors research thesis focused on uncovering natural products (NP) with photoprotective potential in bacterial wild-type strains. As an undergraduate, José published his thesis findings as the first author, leading him to receive the laureate thesis award. Fueled by his interests in the NP field, bacterial genomics, and computational biology, José joined the Microbiology and Cell Science Ph.D. Program at the University of Florida in 2022. Currently, in his second year, he is dedicated to developing computational approaches that describe the yet-to-be-discovered biosynthetic gene cluster (BGC) diversity in the microbial world. José is particularly passionate about unraveling the biosynthetic machinery that escapes conventional genome miner algorithms. Throughout his Ph.D. dissertation, José aims to develop a tool for targeted genome mining and to elucidate the potential diversity of 7-deazapurine BGCs.

graphic file with name mmbr.00199-23.f012.gif

Yifeng Yuan obtained a bachelor's degree in biotechnology at Jilin University in China in 2009, a Master’s Degree at California State University in 2012 and a Ph.D. in microbiology at University of Florida, in 2019. Yifeng Yuan worked at the department of biological engineering at MIT as a postdoctoral associate. He worked in diverse academic fields from cancer biology and signaling transduction to bioinformatics and microbiome metagenomics. He discovered gene families involved in deazaguanine and phosphorothioate modification and developed sequencing technology to map phosphorothioate modifications in bacterial genomes. Since 2023, he worked at the department of microbiology and cell science at University of Florida as a biological scientist. He continues interest in combining comparative genomic analysis, machine-learning bioinformatics and multi-omics approaches to connect genes and phenotypes, with an emphasis on bacterial epigenetics and epitranscriptomics and microbiome-host interactions.

graphic file with name mmbr.00199-23.f013.gif

Rémi Zallot received his Doctorate from Université Bordeaux 2 as a plant biochemist. During his work at the University of Florida, while exploring queuosine and B vitamin pathways from plants and microbes, he learned how comparative genomics is leveraged to guide and define the characterization of unknown genes. Subsequently, at the University of Illinois at Urbana-Champaign, he characterized enzymes from the human gut microbiome and contributed to the development and promotion of the EFI web tools, integral to the genomic enzymology approach. With a European-funded MSCA fellowship at Swansea University, he investigated uncharacterized CYPs in Mycobacterium species. At the Manchester Institute of Biotechnology, he characterized specialized metabolism enzymes. Since June 2023, Dr. Zallot is a Lecturer at the Department of Life Sciences of Manchester Metropolitan University. He is focused on characterizing relevant genes from microbial human pathogens and continues to develop and promote bioinformatics approaches.

graphic file with name mmbr.00199-23.f014.gif

Marc G. Chevrette received a B.Sc. in Molecular Biology and Bioinformatics from Rensselaer Polytechnic Institute, master's degrees in Bioengineering and Genetics from Harvard University Extension and the University of Wisconsin-Madison, respectively, a Ph.D. in Genetics from the University of Wisconsin-Madison, and postdoctoral training at the Wisconsin Institute of Discovery. Marc was the Head of Experimental Genomics at Warp Drive Bio and an Associate at the Broad Institute of MIT & Harvard. He is currently an Assistant Professor in the Department of Microbiology and Cell Science at the University of Florida.

graphic file with name mmbr.00199-23.f015.gif

R. M. Madhushi N. Ratnayake earned her B.Sc. (Special) degree in Chemical Biology from the University of Colombo, Sri Lanka in 2016 along with a Diploma in Information Technology in 2013. She worked as a Chemist at the Food and Environment Laboratory, Bureau Veritas, Sri Lanka before commencing her doctoral studies at the University of Florida. Under the mentorship of Professor Steven D. Bruner, her research focused on elucidating the transcriptional regulation of genotoxin colibactin biosynthesis. She explored microviridins as inhibitors to the human transmembrane protease and studied the structure and function of bacterial transporters and nucleoside hydrolases involved in queuosine salvage allowing her to gain expertise in Biochemistry, X-ray crystallography, and Molecular Biology. Madhushi earned her Ph.D. in Chemistry from the University of Florida in 2023. Currently, she continues pursuing research as a postdoctoral fellow at the School of Medicine Basic Sciences, Vanderbilt University with Prof. John Kuriyan.

graphic file with name mmbr.00199-23.f016.gif

Marshall Jaroch completed a bachelor’s degree in Microbiology from the University of South Florida in Tampa, Florida, in 2016 and began his career at Brammer Bio in Alachua, Florida, where he gained a strong foundation in modern analytical techniques while working in the Assay Development and Analytics department. He returned to academia in 2019, joined the Microbiology and Cell Sciences Department at the University of Florida in Gainesville, Florida, and completed his Ph.D. in 2023. After graduating, he began postdoctoral training at the Oral Biology Department at the University of Florida in 2023 and is currently investigating the mechanisms of metal homeostasis in oral pathogens. He has contributed to the understanding of tRNA modifications in Gram-positive organisms, particularly during his Ph.D., by utilizing bioinformatic-guided experimental investigations.

graphic file with name mmbr.00199-23.f017.gif

Samia Quaiyum holds a bachelor's and master's degree in microbiology from Stamford University Bangladesh. Additionally, she pursued a second master's degree in Biodiversity and Molecular Biology at Hokkaido University, Japan. This was supported by a Japanese government scholarship (MEXT). Samia also completed an internship at Algarve University, Portugal, focusing on Marine Biodiversity and Climate Change. In 2020, Samia earned her Ph.D. in Environmental Molecular Microbiology from the Faculty of Agriculture, Hokkaido University. She specializes in bacterial cell degradation in aerobic and anaerobic environments. She served as a Research Assistant at the National Institute of Advanced Industrial Science and Technology Hokkaido Center, Japan, from 2017 to 2020. Currently, Samia is part of the Microbial Genomics and Genetics lab at the University of Florida, acquiring expertise in Molecular Biology, Microbiology, and Bioinformatics, she is set to complete her 3-year postdoctoral training in January 2024.

graphic file with name mmbr.00199-23.f018.gif

Steven Bruner received a bachelor’s degree in Chemistry from Boston College and a Ph.D. in Chemistry from Harvard University in Gregory Verdine’s group. Postdoctoral training was in Christopher Walsh’s group at Harvard Medical School. Steven was an Assistant Professor at Boston College before moving to the University of Florida where he is currently a Professor in the Chemistry Department. The Bruner group uses organic chemistry, enzymology. and structural biology to probe enzyme mechanisms.

Contributor Information

Valérie de Crécy-Lagard, Email: vcrecy@ufl.edu.

Corrella S. Detweiler, University of Colorado Boulder, Boulder, Colorado, USA

REFERENCES

  • 1. McCarty RM, Bandarian V. 2012. Biosynthesis of pyrrolopyrimidines. Bioorg Chem 43:15–25. doi: 10.1016/j.bioorg.2012.01.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2. Hutinet G, Swarjo MA, de Crécy-Lagard V. 2017. Deazaguanine derivatives, examples of crosstalk between RNA and DNA modification pathways. RNA Biol 14:1175–1184. doi: 10.1080/15476286.2016.1265200 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3. Gupta PK, Daunert S, Nassiri MR, Wotring LL, Drach JC, Townsend LB. 1989. Synthesis, cytotoxicity, and antiviral activity of some acyclic analogs of the pyrrolo[2,3-d]pyrimidine nucleoside antibiotics tubercidin, toyocamycin, and sangivamycin. J Med Chem 32:402–408. doi: 10.1021/jm00122a019 [DOI] [PubMed] [Google Scholar]
  • 4. Harada F, Nishimura S. 1972. Possible anticodon sequences of tRNAHis, tRNAAsn, and tRNAAsp from Escherichia coli. Universal presence of nucleoside O in the first position of the anticodons of these transfer ribonucleic acid. Biochemistry 11:301–308. doi: 10.1021/bi00752a024 [DOI] [PubMed] [Google Scholar]
  • 5. Gregson JM, Crain PF, Edmonds CG, Gupta R, Hashizume T, Phillipson DW, McCloskey JA. 1993. Structure of the archaeal transfer RNA nucleoside G*-15 (2-amino-4,7-dihydro- 4-oxo-7-beta-D-ribofuranosyl-1H-pyrrolo[2,3-d]pyrimidine-5-carboximi dam ide (archaeosine). J Biol Chem 268:10076–10086. [PubMed] [Google Scholar]
  • 6. Cui L, Balamkundu S, Liu C-F, Ye H, Hourihan J, Rausch A, Hauß C, Nilsson E, Hoetzinger M, Holmfeldt K, Zhang W, Martinez-Alvarez L, Peng X, Tremblay D, Moinau S, Solonenko N, Sullivan MB, Lee Y-J, Mulholland A, Weigele PR, de Crécy-Lagard V, Dedon PC, Hutinet G. 2023. Four additional natural 7-deazaguanine derivatives in phages and how to make them. Nucleic Acids Res 51:9214–9226. doi: 10.1093/nar/gkad657 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7. Yuan Y, Zallot R, Grove TL, Payan DJ, Martin-Verstraete I, Šepić S, Balamkundu S, Neelakandan R, Gadi VK, Liu C-F, Swairjo MA, Dedon PC, Almo SC, Gerlt JA, de Crécy-Lagard V. 2019. Discovery of novel bacterial queuine salvage enzymes and pathways in human pathogens. Proc Natl Acad Sci U S A 116:19126–19135. doi: 10.1073/pnas.1909604116 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8. Zallot R, Brochier-Armanet C, Gaston KW, Forouhar F, Limbach PA, Hunt JF, de Crécy-Lagard V. 2014. Plant, animal, and fungal micronutrient queuosine is salvaged by members of the DUF2419 protein family. ACS Chem Biol 9:1812–1825. doi: 10.1021/cb500278k [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9. Turner B, Burkhart BW, Weidenbach K, Ross R, Limbach PA, Schmitz RA, de Crécy-Lagard V, Stedman KM, Santangelo TJ, Iwata-Reuyl D. 2020. Archaeosine modification of archaeal tRNA: role in structural stabilization. J Bacteriol 202:00748–19. doi: 10.1128/JB.00748-19 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10. Hutinet G, Kot W, Cui L, Hillebrand R, Balamkundu S, Gnanakalai S, Neelakandan R, Carstens AB, Fa Lui C, Tremblay D, Jacobs-Sera D, Sassanfar M, Lee Y-J, Weigele P, Moineau S, Hatfull GF, Dedon PC, Hansen LH, de Crécy-Lagard V. 2019. 7-Deazaguanine modifications protect phage DNA from host restriction systems. Nat Commun 10:5442. doi: 10.1038/s41467-019-13384-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11. Manickam N, Joshi K, Bhatt MJ, Farabaugh PJ. 2016. Effects of tRNA modification on translational accuracy depend on intrinsic codon-anticodon strength. Nucleic Acids Res 44:1871–1881. doi: 10.1093/nar/gkv1506 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12. Müller M, Legrand C, Tuorto F, Kelly VP, Atlasi Y, Lyko F, Ehrenhofer-Murray AE. 2019. Queuine links translational control in eukaryotes to a micronutrient from bacteria. Nucleic Acids Res 47:3711–3727. doi: 10.1093/nar/gkz063 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13. Fruchard L, Babosan A, Carvalho A, Lang M, Li B, Duchateau M, Giai-Gianetto Q, Matondo M, Bonhomme F, Fabret C, Namy O, Crécy-Lagard V, Mazel D, Baharoglu Z. 2022. Queuosine modification of tRNA-tyrosine elicits translational reprogramming and enhances growth of Vibrio cholerae with aminoglycosides. bioRxiv. doi: 10.1101/2022.09.26.509455 [DOI]
  • 14. Thiaville JJ, Kellner SM, Yuan Y, Hutinet G, Thiaville PC, Jumpathong W, Mohapatra S, Brochier-Armanet C, Letarov AV, Hillebrand R, Malik CK, Rizzo CJ, Dedon PC, de Crécy-Lagard V. 2016. Novel genomic island modifies DNA with 7-deazaguanine derivatives. Proc Natl Acad Sci U S A 113:E1452–9. doi: 10.1073/pnas.1518570113 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15. Liu Y, Gong R, Liu X, Zhang P, Zhang Q, Cai Y-S, Deng Z, Winkler M, Wu J, Chen W. 2018. Discovery and characterization of the tubercidin biosynthetic pathway from Streptomyces tubercidicus NBRC 13090. Microb Cell Fact 17:131. doi: 10.1186/s12934-018-0978-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16. Lee SA, Jung M. 2007. The nucleoside analog sangivamycin induces apoptotic cell death in breast carcinoma MCF7/adriamycin-resistant cells via protein kinase Cδ and JNK activation. J Biol Chem 282:15271–15283. doi: 10.1074/jbc.M701362200 [DOI] [PubMed] [Google Scholar]
  • 17. McCarty RM, Bandarian V. 2008. Deciphering deazapurine biosynthesis: pathway for pyrrolopyrimidine nucleosides toyocamycin and sangivamycin. Chem Biol 15:790–798. doi: 10.1016/j.chembiol.2008.07.012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18. Nishimura H, Katagiri K, Sato K, Mayama M, Shimaoka N. 1956. Toyocamycin, a new anti-candida antibiotic. J Antibiot (Tokyo) 9:60–62. [PubMed] [Google Scholar]
  • 19. Wu RT, Okabe T, Kim SH, Suzuki H, Tanaka N. 1985. Enhancement of pyrimidine K562 and YAC-1 cells nucleoside uptake into K562 and YAC-1 cells by cadeguomycin. J. Antibiot 38:1588–1595. doi: 10.7164/antibiotics.38.1588 [DOI] [PubMed] [Google Scholar]
  • 20. Naruto S, Uno H, Tanaka A, Kotani H, Takase Y. 1983. Kanagawamicin, a new aminonucleoside analog antibiotic from Actinoplanes kanagawaensis. Heterocycles 20:27. doi: 10.3987/R-1983-01-0027 [DOI] [Google Scholar]
  • 21. Isono K. 1988. Nucleoside antibiotics: structure, biological activity, and biosynthesis. J Antibiot (Tokyo) 41:1711–1739. doi: 10.7164/antibiotics.41.1711 [DOI] [PubMed] [Google Scholar]
  • 22. Nishioka H, Sawa T, Nakamura H, Iinuma H, Ikeda D, Sawa R, Naganawa H, Hayashi C, Hamada M, Takeuchi T, Iitaka Y, Umezawa K. 1991. Isolation and structure determination of novel phosphatidylinositol kinase inhibitors, echiguanines A and B, from Streptomyces sp. J Nat Prod 54:1321–1325. doi: 10.1021/np50077a014 [DOI] [PubMed] [Google Scholar]
  • 23. Isaac BG, Ayer SW, Letendre LJ, Stonard RJ. 1991. Herbicidal nucleosides from microbial sources. J Antibiot (Tokyo) 44:729–732. doi: 10.7164/antibiotics.44.729 [DOI] [PubMed] [Google Scholar]
  • 24. Shomura T, Nishizawa N, Iwata M, Yoshida J, Ito M, Amano S, Koyama M, Kojima M, Inouye S. 1983. Studies on a new nucleoside antibiotic,dapiramicin I. Producing organism,assay method and fermentation. J Antibiot (Tokyo) 36:1300–1304. doi: 10.7164/antibiotics.36.1300 [DOI] [PubMed] [Google Scholar]
  • 25. Seto H, Otake N, Koyama M, Ogino H, Kodama Y, Nishizawa N, Tsuruoka T, Inouye S. 1983. The structure of a novel nucleoside antibiotic, dapiramicin A. Tetrahedron Lett 24:495–498. doi: 10.1016/S0040-4039(00)81446-X [DOI] [Google Scholar]
  • 26. Nishizawa N, Kondo Y, Koyama M, Omoto S, Iwata M, Tsuruoka T, Inouye S. 1984. Studies on a new nucleoside antibiotic, dapiramicin. II Isolation, physico-chemical and biological characterization. J Antibiot (Tokyo) 37:1–5. doi: 10.7164/antibiotics.37.1 [DOI] [PubMed] [Google Scholar]
  • 27. Davies LP, Jamieson DD, Baird-Lambert JA, Kazlauskas R. 1984. Halogenated pyrrolopyrimidine analogues of adenosine from marine organisms: pharmacological activities and potent inhibition of adenosine kinase. Biochem Pharmacol 33:347–355. doi: 10.1016/0006-2952(84)90225-9 [DOI] [PubMed] [Google Scholar]
  • 28. Shuai H, Myronovskyi M, Nadmid S, Luzhetskyy A. 2020. Identification of a biosynthetic gene cluster responsible for the production of a new pyrrolopyrimidine natural product—huimycin. Biomolecules 10:1074. doi: 10.3390/biom10071074 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29. Stewart JB, Bornemann V, Chen JL, Moore RE, Caplan FR, Karuso H, Larsen LK, Patterson GML. 1988. Cytotoxic, fungicidal nucleosides from blue green algae belonging to the Scytonemataceae. J Antibiot (Tokyo) 41:1048–1056. doi: 10.7164/antibiotics.41.1048 [DOI] [PubMed] [Google Scholar]
  • 30. Kato Y, Fusetani N, Matsunaga S, Hashimoto K. 1985. Bioactive marine metabolites IX. mycalisines A and B, novel nucleosides which inhibit cell division of fertilized starfish eggs, from the marine sponge sp. Tetrahedron Lett 26:3483–3486. doi: 10.1016/S0040-4039(00)98670-2 [DOI] [Google Scholar]
  • 31. Ding H, Ruan Z, Kou P, Dong X, Bai J, Xiao Q. 2019. Total synthesis of mycalisine B. Mar Drugs 17:226. doi: 10.3390/md17040226 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32. Kazlauskas R, Murphy P, Wells R, Jamieson D. 1983. Halogenated pyrrolo[2,3-d]pyrimidine nucleosides from marine organisms. Aust J Chem 36:165. doi: 10.1071/CH9830165 [DOI] [Google Scholar]
  • 33. Zabriskie TM, Ireland CM. 1989. The isolation and structure of modified bioactive nucleosides from Jaspis johnstoni. J Nat Prod 52:1353–1356. doi: 10.1021/np50066a032 [DOI] [PubMed] [Google Scholar]
  • 34. Huang R-M, Chen Y-N, Zeng Z, Gao C-H, Su X, Peng Y. 2014. Marine nucleosides: structure, bioactivity, synthesis and biosynthesis. Mar Drugs 12:5817–5838. doi: 10.3390/md12125817 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35. Tsuda M, Nozawa K, Shimbo K, Kobayashi J. 2003. Rigidins B−D, new pyrrolopyrimidine alkaloids from a tunicate Cystodytes species. J Nat Prod 66:292–294. doi: 10.1021/np020393a [DOI] [PubMed] [Google Scholar]
  • 36. Kim J, Pordesimo EO, Toth SI, Schmitz FJ, Van Altena I. 1993. Pantherinine, a cytotoxic aromatic alkaloid, and 7-deazainosine from the ascidian Aplidium pantherinum. J Nat Prod 56:1813–1816. doi: 10.1021/np50100a023 [DOI] [PubMed] [Google Scholar]
  • 37. Mitchell SS, Pomerantz SC, Concepción GP, Ireland CM. 1996. Tubercidin analogs from the ascidian Didemnum voeltzkowi. J Nat Prod 59:1000–1001. doi: 10.1021/np960457f [DOI] [PubMed] [Google Scholar]
  • 38. Tolman RL, Robins RK, Townsend LB. 1968. Pyrrolo[2,3-d]pyrimidine nucleoside antibiotics. Total synthesis and structure of toyocamycin, unamycin B, vengicide, antibiotic E-212, and sangivamycin (BA-90912). J Am Chem Soc 90:524–526. doi: 10.1021/ja01004a076 [DOI] [PubMed] [Google Scholar]
  • 39. Colloc’h N, Poupon A, Mornon JP. 2000. Sequence and structural features of the T-fold, an original tunnelling building unit. Proteins 39:142–154. doi: [DOI] [PubMed] [Google Scholar]
  • 40. Fergus C, Barnes D, Alqasem MA, Kelly VP. 2015. The queuine micronutrient: charting a course from microbe to man. Nutrients 7:2897–2929. doi: 10.3390/nu7042897 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41. Kesh K, Mendez R, Mateo-Victoriano B, Garrido VT, Durden B, Gupta VK, Oliveras Reyes A, Merchant N, Datta J, Banerjee S, Banerjee S. 2022. Obesity enriches for tumor protective microbial metabolites and treatment refractory cells to confer therapy resistance in PDAC. Gut Microbes 14:2096328. doi: 10.1080/19490976.2022.2096328 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42. Yan F, Xiang S, Shi L, Zhu X. 2024. Synthesis of queuine by colonic gut microbiome via cross‐feeding. Food Frontiers 5:174–187. doi: 10.1002/fft2.307 [DOI] [Google Scholar]
  • 43. Reader JS, Metzgar D, Schimmel P, de Crécy-Lagard V. 2004. Identification of four genes necessary for biosynthesis of the modified nucleoside queuosine. J Biol Chem 279:6280–6285. doi: 10.1074/jbc.M310858200 [DOI] [PubMed] [Google Scholar]
  • 44. McCarty RM, Somogyi A, Bandarian V. 2009. Escherichia coli QueD is a 6-carboxy-5,6,7,8-tetrahydropterin synthase. Biochemistry 48:2301–2303. doi: 10.1021/bi9001437 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45. McCarty RM, Somogyi A, Lin G, Jacobsen NE, Bandarian V. 2009. The deazapurine biosynthetic pathway revealed: in vitro enzymatic synthesis of preQ0 from guanosine 5′-triphosphate in four steps. Biochemistry 48:3847–3852. doi: 10.1021/bi900400e [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46. Phillips G, El Yacoubi B, Lyons B, Alvarez S, Iwata-Reuyl D, de Crécy-Lagard V. 2008. Biosynthesis of 7-deazaguanosine-modified tRNA nucleosides: a new role for GTP cyclohydrolase I. J Bacteriol 190:7876–7884. doi: 10.1128/JB.00874-08 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47. Van Lanen SG, Reader JS, Swairjo MA, de Crécy-Lagard V, Lee B, Iwata-Reuyl D. 2005. From cyclohydrolase to oxidoreductase: discovery of nitrile reductase activity in a common fold. Proc Natl Acad Sci U S A 102:4264–4269. doi: 10.1073/pnas.0408056102 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48. Suhadolnik RJ, Uematsu T. 1970. Biosynthesis of the pyrrolopyrimidine nucleoside antibiotic, toyocamycin. J Biol Chem 245:4365–4371. doi: 10.1016/S0021-9258(19)63804-4 [DOI] [PubMed] [Google Scholar]
  • 49. Kuchino Y, Kasai H, Nihei K, Nishimura S. 1976. Biosynthesis of the modified nucleoside Q in transfer RNA. Nucleic Acids Res 3:393–398. doi: 10.1093/nar/3.2.393 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50. Phillips G, Grochowski LL, Bonnett S, Xu H, Bailly M, Blaby-Haas C, El Yacoubi B, Iwata-Reuyl D, White RH, de Crécy-Lagard V. 2012. Functional promiscuity of the COG0720 family. ACS Chem Biol 7:197–209. doi: 10.1021/cb200329f [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51. Sankaran B, Bonnett SA, Shah K, Gabriel S, Reddy R, Schimmel P, Rodionov DA, de Crécy-Lagard V, Helmann JD, Iwata-Reuyl D, Swairjo MA. 2009. Zinc-independent folate biosynthesis: genetic, biochemical, and structural investigations reveal new metal dependence for GTP cyclohydrolase IB. J Bacteriol 191:6936–6949. doi: 10.1128/JB.00287-09 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52. El Yacoubi B, Bonnett S, Anderson JN, Swairjo MA, Iwata-Reuyl D, de Crécy-Lagard V. 2006. Discovery of a new prokaryotic type I GTP cyclohydrolase family. J Biol Chem 281:37586–37593. doi: 10.1074/jbc.M607114200 [DOI] [PubMed] [Google Scholar]
  • 53. Revuelta JL, Serrano-Amatriain C, Ledesma-Amaro R, Jiménez A. 2018. Formation of folates by microorganisms: towards the biotechnological production of this vitamin. Appl Microbiol Biotechnol 102:8613–8620. doi: 10.1007/s00253-018-9266-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54. Zallot R, Harrison KJ, Kolaczkowski B, de Crécy-Lagard V. 2016. Functional annotations of paralogs: a blessing and a curse. Life (Basel) 6:39. doi: 10.3390/life6030039 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55. Miles ZD, Roberts SA, McCarty RM, Bandarian V. 2014. Biochemical and structural studies of 6-Carboxy-5,6,7,8-tetrahydropterin synthase reveal the molecular basis of catalytic promiscuity within the tunnel-fold superfamily. J Biol Chem 289:23641–23652. doi: 10.1074/jbc.M114.555680 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56. Pribat A, Jeanguenin L, Lara-Núñez A, Ziemak MJ, Hyde JE, de Crécy-Lagard V, Hanson AD. 2009. 6-pyruvoyltetrahydropterin synthase paralogs replace the folate synthesis enzyme dihydroneopterin aldolase in diverse bacteria. J Bacteriol 191:4158–4165. doi: 10.1128/JB.00416-09 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57. Jordan MR, Gonzalez-Gutierrez G, Trinidad JC, Giedroc DP. 2022. Metal retention and replacement in QueD2 protect queuosine-tRNA biosynthesis in metal-starved Acinetobacter baumannii Proc Natl Acad Sci U S A 119:e2213630119. doi: 10.1073/pnas.2213630119 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58. Kandlinger F, Plach MG, Merkl R. 2017. AGeNNT: annotation of enzyme families by means of refined neighborhood networks. BMC Bioinformatics 18:274. doi: 10.1186/s12859-017-1689-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59. de Crécy-Lagard V. 2014. Variations in metabolic pathways create challenges for automated metabolic reconstructions: examples from the tetrahydrofolate synthesis pathway. Comput Struct Biotechnol J 10:41–50. doi: 10.1016/j.csbj.2014.05.008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60. Bandarian V, Drennan CL. 2015. Radical-mediated ring contraction in the biosynthesis of 7-deazapurines. Curr Opin Struct Biol 35:116–124. doi: 10.1016/j.sbi.2015.11.005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61. Oberg N, Precord TW, Mitchell DA, Gerlt JA. 2022. RadicalSAM.org: a resource to interpret sequence-function space and discover new radical SAM enzyme chemistry. ACS Bio Med Chem Au 2:22–35. doi: 10.1021/acsbiomedchemau.1c00048 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62. Lachowicz J, Lee J, Sagatova A, Jew K, Grove TL. 2023. The new epoch of structural insights into radical SAM enzymology. Curr Opin Struct Biol 83:102720. doi: 10.1016/j.sbi.2023.102720 [DOI] [PubMed] [Google Scholar]
  • 63. Bruender NA, Young AP, Bandarian V. 2015. Chemical and biological reduction of the radical SAM enzyme CPH4 synthase. Biochemistry 54:2903–2910. doi: 10.1021/acs.biochem.5b00210 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64. Dowling DP, Bruender NA, Young AP, McCarty RM, Bandarian V, Drennan CL. 2014. Radical SAM enzyme QueE defines a new minimal core fold and metal dependent mechanism. Nat Chem Biol 10:106–112. doi: 10.1038/nchembio.1426 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65. Grell TAJ, Bell BN, Nguyen C, Dowling DP, Bruender NA, Bandarian V, Drennan CL. 2019. Crystal structure of AdoMet radical enzyme 7‐carboxy‐7‐deazaguanine synthase from Escherichia coli suggests how modifications near [4Fe–4S] cluster engender flavodoxin specificity. Protein Sci 28:202–215. doi: 10.1002/pro.3529 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66. Nelp MT, Bandarian V. 2015. A single enzyme transforms a carboxylic acid into a nitrile through an amide intermediate. Angew Chem Int Ed Engl 54:10627–10629. doi: 10.1002/anie.201504505 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67. Hennessy AJA, Huang W, Savary C, Campopiano DJ. 2022. Creation of an engineered amide synthetase biocatalyst by the rational separation of a two‐step nitrile synthetase. Chembiochem 23:e202100411. doi: 10.1002/cbic.202100411 [DOI] [PubMed] [Google Scholar]
  • 68. Cicmil N, Huang RH. 2008. Crystal structure of QueC from Bacillus subtilis: an enzyme involved in preQ1 biosynthesis. Proteins 72:1084–1088. doi: 10.1002/prot.22098 [DOI] [PubMed] [Google Scholar]
  • 69. Winkler M, Dokulil K, Weber H, Pavkov-Keller T, Wilding B. 2015. The nitrile-forming enzyme 7-Cyano-7-deazaguanine synthase from Geobacillus kaustophilus: a reverse nitrilase? Chembiochem 16:2373–2378. doi: 10.1002/cbic.201500335 [DOI] [PubMed] [Google Scholar]
  • 70. Gao L, Altae-Tran H, Böhning F, Makarova KS, Segel M, Schmid-Burgk JL, Koob J, Wolf YI, Koonin EV, Zhang F. 2020. Diverse enzymatic activities mediate antiviral immunity in prokaryotes. Science 369:1077–1084. doi: 10.1126/science.aba0372 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71. Millman A, Melamed S, Amitai G, Sorek R. 2020. Diversity and classification of cyclic-oligonucleotide-based anti-phage signalling systems. Nat Microbiol 5:1608–1615. doi: 10.1038/s41564-020-0777-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72. Chikwana VM, Stec B, Lee BWK, de Crécy-Lagard V, Iwata-Reuyl D, Swairjo MA. 2012. Structural basis of biological nitrile reduction. J Biol Chem 287:30560–30570. doi: 10.1074/jbc.M112.388538 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73. Okada N, Noguchi S, Kasai H, Shindo-Okada N, Ohgi T, Goto T, Nishimura S. 1979. Novel mechanism of post-transcriptional modification of tRNA. Insertion of bases of Q precursors into tRNA by a specific tRNA transglycosylase reaction. J Biol Chem 254:3067–3073. [PubMed] [Google Scholar]
  • 74. Lee BWK, Van Lanen SG, Iwata-Reuyl D. 2007. Mechanistic studies of Bacillus subtilis QueF, the nitrile oxidoreductase involved in queuosine biosynthesis. Biochemistry 46:12844–12854. doi: 10.1021/bi701265r [DOI] [PubMed] [Google Scholar]
  • 75. Kim Y, Zhou M, Moy S, Morales J, Cunningham MA, Joachimiak A. 2010. High-resolution structure of the nitrile reductase QueF combined with molecular simulations provide insight into enzyme mechanism. J Mol Biol 404:127–137. doi: 10.1016/j.jmb.2010.09.042 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76. Jung J, Nidetzky B. 2018. Evidence of a sequestered imine intermediate during reduction of nitrile to amine by the nitrile reductase QueF from Escherichia coli. J Biol Chem 293:3720–3733. doi: 10.1074/jbc.M117.804583 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77. Jung J, Czabany T, Wilding B, Klempier N, Nidetzky B. 2016. Kinetic analysis and probing with substrate analogues of the reaction pathway of the nitrile reductase QueF from Escherichia coli. J Biol Chem 291:25411–25426. doi: 10.1074/jbc.M116.747014 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78. Mohammad A, Bon Ramos A, Lee BWK, Cohen SW, Kiani MK, Iwata-Reuyl D, Stec B, Swairjo MA. 2017. Protection of the queuosine biosynthesis enzyme QueF from irreversible oxidation by a conserved intramolecular disulfide. Biomolecules 7:30. doi: 10.3390/biom7010030 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79. Wilding B, Winkler M, Petschacher B, Kratzer R, Egger S, Steinkellner G, Lyskowski A, Nidetzky B, Gruber K, Klempier N. 2013. Targeting the substrate binding site of E. coli nitrile reductase QueF by modeling, substrate and enzyme engineering. Chemistry 19:7007–7012. doi: 10.1002/chem.201300163 [DOI] [PubMed] [Google Scholar]
  • 80. Zallot R, Oberg N, Gerlt JA. 2021. Discovery of new enzymatic functions and metabolic pathways using genomic enzymology web tools. Curr Opin Biotechnol 69:77–90. doi: 10.1016/j.copbio.2020.12.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81. McCown PJ, Liang JJ, Weinberg Z, Breaker RR. 2014. Structural, functional, and taxonomic diversity of three preQ1 riboswitch classes. Chem Biol 21:880–889. doi: 10.1016/j.chembiol.2014.05.015 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82. Roth A, Winkler WC, Regulski EE, Lee BWK, Lim J, Jona I, Barrick JE, Ritwik A, Kim JN, Welz R, Iwata-Reuyl D, Breaker RR. 2007. A riboswitch selective for the queuosine precursor preQ1 contains an unusually small aptamer domain. Nat Struct Mol Biol 14:308–317. doi: 10.1038/nsmb1224 [DOI] [PubMed] [Google Scholar]
  • 83. Meyer MM, Roth A, Chervin SM, Garcia GA, Breaker RR. 2008. Confirmation of a second natural preQ1aptamer class in Streptococcaceae bacteria. RNA 14:685–695. doi: 10.1261/rna.937308 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84. Schroeder GM, Cavender CE, Blau ME, Jenkins JL, Mathews DH, Wedekind JE. 2022. A small RNA that cooperatively senses two stacked metabolites in one pocket for gene control. Nat Commun 13:199. doi: 10.1038/s41467-021-27790-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85. Hu G, Zhou H-X. 2021. Binding free energy decomposition and multiple unbinding paths of buried ligands in a preQ1 riboswitch. PLoS Comput Biol 17:e1009603. doi: 10.1371/journal.pcbi.1009603 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86. Suddala KC, Rinaldi AJ, Feng J, Mustoe AM, Eichhorn CD, Liberman JA, Wedekind JE, Al-Hashimi HM, Brooks CL, Walter NG. 2013. Single transcriptional and translational preQ1 riboswitches adopt similar pre-folded ensembles that follow distinct folding pathways into the same ligand-bound structure. Nucleic Acids Res 41:10462–10475. doi: 10.1093/nar/gkt798 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87. Schroeder GM, Akinyemi O, Malik J, Focht CM, Pritchett EM, Baker CD, McSally JP, Jenkins JL, Mathews DH, Wedekind JE. 2023. A riboswitch separated from its ribosome-binding site still regulates translation. Nucleic Acids Res 51:2464–2484. doi: 10.1093/nar/gkad056 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88. Connelly CM, Numata T, Boer RE, Moon MH, Sinniah RS, Barchi JJ, Ferré-D’Amaré AR, Schneekloth JS. 2019. Synthetic ligands for preQ1 riboswitches provide structural and mechanistic insights into targeting RNA tertiary structure. Nat Commun 10:1501. doi: 10.1038/s41467-019-09493-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89. Wu MC, Lowe PT, Robinson CJ, Vincent HA, Dixon N, Leigh J, Micklefield J. 2015. Rational re-engineering of a transcriptional silencing preQ1 riboswitch. J Am Chem Soc 137:9015–9021. doi: 10.1021/jacs.5b03405 [DOI] [PubMed] [Google Scholar]
  • 90. Harrison KJ, Crécy-Lagard V de, Zallot R. 2018. Gene graphics: a genomic neighborhood data visualization web application. Bioinformatics 34:1406–1408. doi: 10.1093/bioinformatics/btx793 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91. Kawamura T, Hirata A, Ohno S, Nomura Y, Nagano T, Nameki N, Yokogawa T, Hori H. 2016. Multisite-specific archaeosine tRNA-guanine transglycosylase (ArcTGT) from Thermoplasma acidophilum, a thermo-acidophilic archaeon. Nucleic Acids Res 44:1894–1908. doi: 10.1093/nar/gkv1522 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92. Stengl B, Reuter K, Klebe G. 2005. Mechanism and substrate specificity of tRNA-guanine transglycosylases (TGTs): tRNA-modifying enzymes from the three different kingdoms of life share a common catalytic mechanism. Chembiochem 6:1926–1939. doi: 10.1002/cbic.200500063 [DOI] [PubMed] [Google Scholar]
  • 93. Chen Y-C, Brooks AF, Goodenough-Lashua DM, Kittendorf JD, Showalter HD, Garcia GA, Showwalter HD, Garcia GA. 2011. Evolution of eukaryal tRNA-guanine transglycosylase: insight gained from the heterocyclic substrate recognition by the wild-type and mutant human and Escherichia coli tRNA-guanine transglycosylases. Nucleic Acids Res 39:2834–2844. doi: 10.1093/nar/gkq1188 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94. Katoh K, Standley DM. 2013. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol Biol Evol 30:772–780. doi: 10.1093/molbev/mst010 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95. Crooks GE, Hon G, Chandonia J-M, Brenner SE. 2004. WebLogo: a sequence logo generator. Genome Res 14:1188–1190. doi: 10.1101/gr.849004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96. Jumper J, Evans R, Pritzel A, Green T, Figurnov M, Ronneberger O, Tunyasuvunakool K, Bates R, Žídek A, Potapenko A, et al. 2021. Highly accurate protein structure prediction with AlphaFold. Nature 596:583–589. doi: 10.1038/s41586-021-03819-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97. Varadi M, Anyango S, Deshpande M, Nair S, Natassia C, Yordanova G, Yuan D, Stroe O, Wood G, Laydon A, et al. 2022. AlphaFold protein structure database: massively expanding the structural coverage of protein-sequence space with high-accuracy models. Nucleic Acids Res 50:D439–D444. doi: 10.1093/nar/gkab1061 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98. Yuan Y, Hutinet G, Valera JG, Hu J, Hillebrand R, Gustafson A, Iwata-Reuyl D, Dedon PC, de Crécy-Lagard V. 2018. Identification of the minimal bacterial 2′-deoxy-7-amido-7-deazaguanine synthesis machinery. Mol Microbiol 110:469–483. doi: 10.1111/mmi.14113 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99. Gedara SH, Wood E, Gustafson A, Liang C, Hung S-H, Savage J, Phan P, Luthra A, de Crécy-Lagard V, Dedon P, Swairjo MA, Iwata-Reuyl D. 2023. 7-Deazaguanines in DNA: functional and structural elucidation of a DNA modification system. Nucleic Acids Res 51:3836–3854. doi: 10.1093/nar/gkad141 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100. Kot W, Olsen NS, Nielsen TK, Hutinet G, de Crécy-Lagard V, Cui L, Dedon PC, Carstens AB, Moineau S, Swairjo MA, Hansen LH. 2020. Detection of preQ0 deazaguanine modifications in bacteriophage CAjan DNA using Nanopore sequencing reveals same hypermodification at two distinct DNA motifs. Nucleic Acids Res 48:10383–10396. doi: 10.1093/nar/gkaa735 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101. Olsen NS, Nielsen TK, Cui L, Dedon P, Neve H, Hansen LH, Kot W. 2023. A novel queuovirinae lineage of Pseudomonas aeruginosa phages encode preQ0 DNA modifications with a single GA motif that provide restriction and CRISPR Cas9 protection in vitro. Nucleic Acids Res 51:8663–8676. doi: 10.1093/nar/gkad622 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102. Crippen CS, Lee Y-J, Hutinet G, Shajahan A, Sacher JC, Azadi P, de Crécy-Lagard V, Weigele PR, Szymanski CM. 2019. Deoxyinosine and 7-deaza-2-deoxyguanosine as carriers of genetic information in the DNA of campylobacter viruses. J Virol 93:01111–01119. doi: 10.1128/JVI.01111-19 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103. Inglese J, Johnson DL, Shiau A, Smith JM, Benkovic SJ. 1990. Subcloning, characterization, and affinity labeling of Escherichia coli glycinamide ribonucleotide transformylase. Biochemistry 29:1436–1443. doi: 10.1021/bi00458a014 [DOI] [PubMed] [Google Scholar]
  • 104. Slany RK, Bösl M, Crain PF, Kersten H. 1993. A new function of S-adenosylmethionine: the ribosyl moiety of AdoMet is the precursor of the cyclopentenediol moiety of the tRNA wobble base queuine. Biochemistry 32:7811–7817. doi: 10.1021/bi00081a028 [DOI] [PubMed] [Google Scholar]
  • 105. Van Lanen SG, Iwata-Reuyl D. 2003. Kinetic mechanism of the tRNA-modifying enzyme S-adenosylmethionine:tRNA ribosyltransferase-isomerase (QueA). Biochemistry 42:5312–5320. doi: 10.1021/bi034197u [DOI] [PubMed] [Google Scholar]
  • 106. Grimm C, Ficner R, Sgraja T, Haebel P, Klebe G, Reuter K. 2006. Crystal structure of Bacillus subtilis S-adenosylmethionine:tRNA ribosyltransferase-isomerase. Biochem Biophys Res Commun 351:695–701. doi: 10.1016/j.bbrc.2006.10.096 [DOI] [PubMed] [Google Scholar]
  • 107. Mathews I, Schwarzenbacher R, McMullan D, Abdubek P, Ambing E, Axelrod H, Biorac T, Canaves JM, Chiu H-J, Deacon AM, et al. 2005. Crystal structure of S-adenosylmethionine: tRNA ribosyltransferase-isomerase (QueA) from Thermotoga maritima at 2.0 Å resolution reveals a new fold. Proteins 59:869–874. doi: 10.1002/prot.20419 [DOI] [PubMed] [Google Scholar]
  • 108. Kinzie SD, Thern B, Iwata-Reuyl D. 2000. Mechanistic studies of the tRNA-modifying enzyme QueA: a chemical imperative for the use of AdoMet as a “ribosyl” donor. Org Lett 2:1307–1310. doi: 10.1021/ol005756h [DOI] [PubMed] [Google Scholar]
  • 109. Lee YH, Ren D, Jeon B, Liu HW. 2023. S-Adenosylmethionine: more than just a methyl donor. Nat Prod Rep 40:1521–1549. doi: 10.1039/d2np00086e [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110. Chen X, Li S, Zhang B, Sun H, Wang J, Zhang W, Meng W, Chen T, Dyson P, Liu G. 2022. A new bacterial tRNA enhances antibiotic production in Streptomyces by circumventing inefficient wobble base-pairing. Nucleic Acids Res 50:7084–7096. doi: 10.1093/nar/gkac502 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111. Tomasi FG, Kimura S, Rubin EJ, Waldor MK. 2023. A tRNA modification in Mycobacterium tuberculosis facilitates optimal intracellular growth. Elife 12:RP87146. doi: 10.7554/eLife.87146 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112. Koshla O, Vogt LM, Rydkin O, Sehin Y, Ostash I, Helm M, Ostash B. 2023. Landscape of post-transcriptional tRNA modifications in streptomyces albidoflavus j1074 as portrayed by mass spectrometry and genomic data mining. J Bacteriol 205:e0029422. doi: 10.1128/jb.00294-22 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113. Frey B, McCloskey J, Kersten W, Kersten H. 1988. New function of vitamin B12: cobamide-dependent reduction of epoxyqueuosine to queuosine in tRNAs of Escherichia coli and Salmonella typhimurium. J Bacteriol 170:2078–2082. doi: 10.1128/jb.170.5.2078-2082.1988 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114. Miles ZD, McCarty RM, Molnar G, Bandarian V. 2011. Discovery of epoxyqueuosine (oQ) reductase reveals parallels between halorespiration and tRNA modification. Proc Natl Acad Sci U S A 108:7368–7372. doi: 10.1073/pnas.1018636108 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115. Greenhalgh ED, Kincannon W, Bandarian V, Brunold TC. 2022. Spectroscopic and computational investigation of the epoxyqueuosine reductase QueG reveals intriguing similarities with the reductive dehalogenase PceA. Biochemistry 61:195–205. doi: 10.1021/acs.biochem.1c00644 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116. Dowling DP, Miles ZD, Köhrer C, Maiocco SJ, Elliott SJ, Bandarian V, Drennan CL. 2016. Molecular basis of cobalamin-dependent RNA modification. Nucleic Acids Res 44:9965–9976. doi: 10.1093/nar/gkw806 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117. Zallot R, Ross R, Chen W-H, Bruner SD, Limbach PA, de Crécy-Lagard V. 2017. Identification of a novel epoxyqueuosine reductase family by comparative genomics. ACS Chem Biol 12:844–851. doi: 10.1021/acschembio.6b01100 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118. Li Q, Zallot R, MacTavish BS, Montoya A, Payan DJ, Hu Y, Gerlt JA, Angerhofer A, de Crécy-Lagard V, Bruner SD. 2021. Epoxyqueuosine reductase QueH in the biosynthetic pathway to tRNA queuosine is a unique metalloenzyme. Biochemistry 60:3152–3161. doi: 10.1021/acs.biochem.1c00164 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119. Hanson AD, Pribat A, Waller JC, de Crécy-Lagard V. 2009. “Unknown” proteins and “orphan” enzymes: the missing half of the engineering parts list--and how to find it. Biochem J 425:1–11. doi: 10.1042/BJ20091328 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120. Phillipson DW, Edmonds CG, Crain PF, Smith DL, Davis DR, McCloskey JA. 1987. Isolation and structure elucidation of an epoxide derivative of the hypermodified nucleoside queuosine from Escherichia coli transfer RNA. J Biol Chem 262:3462–3471. doi: 10.1016/S0021-9258(18)61373-0 [DOI] [PubMed] [Google Scholar]
  • 121. Zhao X, Ma D, Ishiguro K, Saito H, Akichika S, Matsuzawa I, Mito M, Irie T, Ishibashi K, Wakabayashi K, Sakaguchi Y, Yokoyama T, Mishima Y, Shirouzu M, Iwasaki S, Suzuki T, Suzuki T. 2023. Glycosylated queuosines in tRNAs optimize translational rate and post-embryonic growth. Cell 186:5517–5535. doi: 10.1016/j.cell.2023.10.026 [DOI] [PubMed] [Google Scholar]
  • 122. Campanacci V, Dubois DY, Becker HD, Kern D, Spinelli S, Valencia C, Pagot F, Salomoni A, Grisel S, Vincentelli R, Bignon C, Lapointe J, Giegé R, Cambillau C. 2004. The Escherichia coli yadB gene product reveals a novel aminoacyl-tRNA synthetase like activity. J Mol Biol 337:273–283. doi: 10.1016/j.jmb.2004.01.027 [DOI] [PubMed] [Google Scholar]
  • 123. Blaise M, Becker HD, Keith G, Cambillau C, Lapointe J, Giegé R, Kern D. 2004. A minimalist glutamyl-tRNA synthetase dedicated to aminoacylation of the tRNAAsp QUC anticodon. Nucleic Acids Res 32:2768–2775. doi: 10.1093/nar/gkh608 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124. Dubois DY, Blaise M, Becker HD, Campanacci V, Keith G, Giegé R, Cambillau C, Lapointe J, Kern D. 2004. An aminoacyl-tRNA synthetase-like protein encoded by the Escherichia coli yadB gene glutamylates specifically tRNAAsp. Proc Natl Acad Sci U S A 101:7530–7535. doi: 10.1073/pnas.0401634101 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125. Galperin MY, Koonin EV. 2012. Divergence and convergence in enzyme evolution. J Biol Chem 287:21–28. doi: 10.1074/jbc.R111.241976 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126. Phillips G, Chikwana VM, Maxwell A, El-Yacoubi B, Swairjo MA, Iwata-Reuyl D, de Crécy-Lagard V. 2010. Discovery and characterization of an amidinotransferase involved in the modification of archaeal tRNA. J Biol Chem 285:12706–12713. doi: 10.1074/jbc.M110.102236 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127. Yokogawa T, Nomura Y, Yasuda A, Ogino H, Hiura K, Nakada S, Oka N, Ando K, Kawamura T, Hirata A, Hori H, Ohno S. 2019. Identification of a radical SAM enzyme involved in the synthesis of archaeosine. Nat Chem Biol 15:1148–1155. doi: 10.1038/s41589-019-0390-7 [DOI] [PubMed] [Google Scholar]
  • 128. Boswinkle K, McKinney J, Allen KD. 2022. Highlighting the unique roles of radical S -adenosylmethionine enzymes in methanogenic archaea. J Bacteriol 204:e0019722. doi: 10.1128/jb.00197-22 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129. Phillips G, Swairjo MA, Gaston KW, Bailly M, Limbach PA, Iwata-Reuyl D, de Crécy-Lagard V. 2012. Diversity of archaeosine synthesis in Crenarchaeota. ACS Chem Biol 7:300–305. doi: 10.1021/cb200361w [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130. Bon Ramos A, Bao L, Turner B, de Crécy-Lagard V, Iwata-Reuyl D. 2017. QueF-Like, a non-homologous archaeosine synthase from the Crenarchaeota. Biomolecules 7:36. doi: 10.3390/biom7020036 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131. Mei X, Alvarez J, Bon Ramos A, Samanta U, Iwata-Reuyl D, Swairjo MA. 2017. Crystal structure of the archaeosine synthase QueF-like-Insights into amidino transfer and tRNA recognition by the tunnel fold. Proteins 85:103–116. doi: 10.1002/prot.25202 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132. Iyer LM, Zhang D, Burroughs AM, Aravind L. 2013. Computational identification of novel biochemical systems involved in oxidation, glycosylation and other complex modifications of bases in DNA. Nucleic Acids Res 41:7635–7655. doi: 10.1093/nar/gkt573 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133. He W, Huang T, Tang Y, Liu Y, Wu X, Chen S, Chan W, Wang Y, Liu X, Chen S, Wang L. 2015. Regulation of DNA phosphorothioate modification in Salmonella enterica by DndB. Sci Rep 5:12368. doi: 10.1038/srep12368 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134. Prahlad J, Yuan Y, Lin J, Chang C-W, Iwata-Reuyl D, Liu Y, de Crécy-Lagard V, Wilson MA. 2020. The DUF328 family member YaaA is a DNA-binding protein with a novel fold. J Biol Chem 295:14236–14247. doi: 10.1074/jbc.RA120.015055 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135. Zallot R, Yuan Y, de Crécy-Lagard V. 2017. The Escherichia coli COG1738 member YhhQ is involved in 7-cyanodeazaguanine (preQ) transport. Biomolecules 7:12. doi: 10.3390/biom7010012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136. Rodionov DA, Hebbeln P, Eudes A, ter Beek J, Rodionova IA, Erkens GB, Slotboom DJ, Gelfand MS, Osterman AL, Hanson AD, Eitinger T. 2009. A novel class of modular transporters for vitamins in prokaryotes. J Bacteriol 191:42–51. doi: 10.1128/JB.01208-08 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137. Eitinger T, Rodionov DA, Grote M, Schneider E. 2011. Canonical and ECF-type ATP-binding cassette importers in prokaryotes: diversity in modular organization and cellular functions. FEMS Microbiol Rev 35:3–67. doi: 10.1111/j.1574-6976.2010.00230.x [DOI] [PubMed] [Google Scholar]
  • 138. Quaiyum S, Yuan Y, Sun G, N Ratnayake RMM, Hutinet G, Dedon PC, Minnick MF, de Crécy-Lagard V. 2023. Queuosine salvage in Bartonella henselae Houston 1: a unique evolutionary path. bioRxiv:2023.12.05.570228. doi: 10.1101/2023.12.05.570228 [DOI] [PMC free article] [PubMed]
  • 139. Daley DO, Rapp M, Granseth E, Melén K, Drew D, von Heijne G. 2005. Global topology analysis of the Escherichia coli inner membrane proteome. Science 308:1321–1323. doi: 10.1126/science.1109730 [DOI] [PubMed] [Google Scholar]
  • 140. Granseth E, Daley DO, Rapp M, Melén K, von Heijne G. 2005. Experimentally constrained topology models for 51,208 bacterial inner membrane proteins. J Mol Biol 352:489–494. doi: 10.1016/j.jmb.2005.07.053 [DOI] [PubMed] [Google Scholar]
  • 141. Rempel S, Stanek WK, Slotboom DJ. 2019. ECF-type ATP-binding cassette transporters. Annu Rev Biochem 88:551–576. doi: 10.1146/annurev-biochem-013118-111705 [DOI] [PubMed] [Google Scholar]
  • 142. Patel BI, Heiss M, Samel-Pommerencke A, Carell T, Ehrenhofer-Murray AE. 2022. Queuosine salvage in fission yeast by Qng1-mediated hydrolysis to queuine. Biochem Biophys Res Commun 624:146–150. doi: 10.1016/j.bbrc.2022.07.104 [DOI] [PubMed] [Google Scholar]
  • 143. Hung S-H, Elliott GI, Ramkumar TR, Burtnyak L, McGrenaghan CJ, Alkuzweny S, Quaiyum S, Iwata-Reuyl D, Pan X, Green BD, Kelly VP, de Crécy-Lagard V, Swairjo MA. 2023. Structural basis of Qng1-mediated salvage of the micronutrient queuine from queuosine-5′-monophosphate as the biological substrate. Nucleic Acids Res 51:935–951. doi: 10.1093/nar/gkac1231 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144. Olson RD, Assaf R, Brettin T, Conrad N, Cucinell C, Davis JJ, Dempsey DM, Dickerman A, Dietrich EM, Kenyon RW, et al. 2023. Introducing the bacterial and viral bioinformatics resource center (BV-BRC): a resource combining PATRIC, IRD and ViPR. Nucleic Acids Res 51:D678–D689. doi: 10.1093/nar/gkac1003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145. Letunic I, Bork P. 2021. Interactive tree of life (iTOL) v5: an online tool for phylogenetic tree display and annotation. Nucleic Acids Res 49:W293–W296. doi: 10.1093/nar/gkab301 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146. Puri P, Wetzel C, Saffert P, Gaston KW, Russell SP, Cordero Varela JA, van der Vlies P, Zhang G, Limbach PA, Ignatova Z, Poolman B. 2014. Systematic identification of tRNAome and its dynamics in Lactococcus lactis. Mol Microbiol 93:944–956. doi: 10.1111/mmi.12710 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147. Driscoll TP, Verhoeve VI, Guillotte ML, Lehman SS, Rennoll SA, Beier-Sexton M, Rahman MS, Azad AF, Gillespie JJ. 2017. Wholly Rickettsia! reconstructed metabolic profile of the quintessential bacterial parasite of eukaryotic cells. mBio 8:e00859-17. doi: 10.1128/mBio.00859-17 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148. Jiang Z, Jones DH, Khuri S, Tsinoremas NF, Wyss T, Jander G, Wilson ACC. 2013. Comparative analysis of genome sequences from four strains of the Buchnera aphidicola Mp endosymbiont of the green peach aphid, Myzus persicae. BMC Genomics 14:917. doi: 10.1186/1471-2164-14-917 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149. Castelle CJ, Méheust R, Jaffe AL, Seitz K, Gong X, Baker BJ, Banfield JF. 2021. Protein family content uncovers lineage relationships and bacterial pathway maintenance mechanisms in DPANN archaea. Front Microbiol 12:660052. doi: 10.3389/fmicb.2021.660052 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150. Cragg GM, Newman DJ. 2013. Natural products: a continuing source of novel drug leads. Biochim Biophys Acta 1830:3670–3695. doi: 10.1016/j.bbagen.2013.02.008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151. Chevrette MG, Gutiérrez-García K, Selem-Mojica N, Aguilar-Martínez C, Yañez-Olvera A, Ramos-Aboites HE, Hoskisson PA, Barona-Gómez F. 2020. Evolutionary dynamics of natural product biosynthesis in bacteria. Nat Prod Rep 37:566–599. doi: 10.1039/c9np00048h [DOI] [PubMed] [Google Scholar]
  • 152. Khalifa SAM, Elias N, Farag MA, Chen L, Saeed A, Hegazy M-EF, Moustafa MS, Abd El-Wahed A, Al-Mousawi SM, Musharraf SG, Chang F-R, Iwasaki A, Suenaga K, Alajlani M, Göransson U, El-Seedi HR. 2019. Marine natural products: a source of novel anticancer drugs. Mar Drugs 17:491. doi: 10.3390/md17090491 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153. Cediel Becerra JDD, Suescún Sepúlveda JA, Fuentes JL. 2022. Prodigiosin production and photoprotective/antigenotoxic properties in Serratia marcescens indigenous strains from eastern cordillera of Colombia. Photochem Photobiol 98:254–261. doi: 10.1111/php.13507 [DOI] [PubMed] [Google Scholar]
  • 154. Wright GD. 2017. Opportunities for natural products in 21st century antibiotic discovery. Nat Prod Rep 34:694–701. doi: 10.1039/c7np00019g [DOI] [PubMed] [Google Scholar]
  • 155. Battaglia U, Long JE, Searle MS, Moody CJ. 2011. 7-Deazapurine biosynthesis: NMR study of toyocamycin biosynthesis in Streptomyces rimosus using 2-13C-7-15N-adenine. Org Biomol Chem 9:2227–2232. doi: 10.1039/c0ob01054e [DOI] [PubMed] [Google Scholar]
  • 156. Musumeci F, Sanna M, Grossi G, Brullo C, Fallacara AL, Schenone S. 2017. Pyrrolo [2,3-d]pyrimidines as kinase inhibitors. Curr Med Chem 24:2059–2085. doi: 10.2174/0929867324666170303162100 [DOI] [PubMed] [Google Scholar]
  • 157. Perlíková P, Hocek M. 2017. Pyrrolo[2,3- d ]pyrimidine (7-deazapurine) as a privileged scaffold in design of antitumor and antiviral nucleosides. Med Res Rev 37:1429–1460. doi: 10.1002/med.21465 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158. Liang T, Yang Y, Wang J, Xie Z, Chen X. 2023. The application of pyrrolo[2, 3-d]pyrimidine scaffold in medicinal chemistry from 2017 to 2021. Mini Rev Med Chem 23:1118–1136. doi: 10.2174/1389557523666230111161810 [DOI] [PubMed] [Google Scholar]
  • 159. Xu D, Ma M, Liu Y, Zhou T, Wang K, Deng Z, Hong K. 2015. PreQ0 base, an unusual metabolite with anti-cancer activity from Streptomyces qinglanensis 172205. Anticancer Agents Med Chem 15:285–290. doi: 10.2174/1871520614666141027144653 [DOI] [PubMed] [Google Scholar]
  • 160. Terlouw BR, Blin K, Navarro-Muñoz JC, Avalon NE, Chevrette MG, Egbert S, Lee S, Meijer D, Recchia MJJ, Reitz ZL, et al. 2023. MIBiG 3.0: a community-driven effort to annotate experimentally validated biosynthetic gene clusters. Nucleic Acids Res 51:D603–D610. doi: 10.1093/nar/gkac1049 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161. Blin K, Shaw S, Augustijn HE, Reitz ZL, Biermann F, Alanjary M, Fetter A, Terlouw BR, Metcalf WW, Helfrich EJN, van Wezel GP, Medema MH, Weber T. 2023. AntiSMASH 7.0: new and improved predictions for detection, regulation, chemical structures and visualisation. Nucleic Acids Res 51:W46–W50. doi: 10.1093/nar/gkad344 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162. Urbonavicius J, Stahl G, Durand JMB, Ben Salem SN, Qian Q, Farabaugh PJ, Björk GR. 2003. Transfer RNA modifications that alter +1 frameshifting in general fail to affect -1 frameshifting. RNA 9:760–768. doi: 10.1261/rna.5210803 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163. Urbonavicius J, Qian Q, Durand JMB, Hagervall TG, Björk GR. 2001. Improvement of reading frame maintenance is a common function for several tRNA modifications. EMBO J 20:4863–4873. doi: 10.1093/emboj/20.17.4863 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164. Manickam N, Nag N, Abbasi A, Patel K, Farabaugh PJ. 2014. Studies of translational misreading in vivo show that the ribosome very efficiently discriminates against most potential errors. RNA 20:9–15. doi: 10.1261/rna.039792.113 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165. Beier H, Grimm M. 2001. Misreading of termination codons in eukaryotes by natural nonsense suppressor tRNAs. Nucleic Acids Res 29:4767–4782. doi: 10.1093/nar/29.23.4767 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166. Tittle JM, Schwark DG, Biddle W, Schmitt MA, Fisk JD. 2022. Impact of queuosine modification of endogenous E. coli tRNAs on sense codon reassignment. Front Mol Biosci 9:938114. doi: 10.3389/fmolb.2022.938114 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167. Meier F, Suter B, Grosjean H, Keith G, Kubli E. 1985. Queuosine modification of the wobble base in tRNAHis influences “in vivo” decoding properties. EMBO J 4:823–827. doi: 10.1002/j.1460-2075.1985.tb03704.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168. Díaz-Rullo J, González-Pastor JE. 2023. tRNA queuosine modification is involved in biofilm formation and virulence in bacteria. Nucleic Acids Res 51:9821–9837. doi: 10.1093/nar/gkad667 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169. Dixit S, Kessler AC, Henderson J, Pan X, Zhao R, D’Almeida GS, Kulkarni S, Rubio MAT, Hegedűsová E, Ross RL, Limbach PA, Green BD, Paris Z, Alfonzo JD. 2021. Dynamic queuosine changes in tRNA couple nutrient levels to codon choice in Trypanosoma brucei. Nucleic Acids Res 49:12986–12999. doi: 10.1093/nar/gkab1204 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170. Grosjean H, Westhof E. 2016. An integrated, structure- and energy-based view of the genetic code. Nucleic Acids Res 44:8020–8040. doi: 10.1093/nar/gkw608 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171. Rozov A, Wolff P, Grosjean H, Yusupov M, Yusupova G, Westhof E. 2018. Tautomeric G•U pairs within the molecular ribosomal grip and fidelity of decoding in bacteria. Nucleic Acids Res 46:7425–7435. doi: 10.1093/nar/gky547 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172. Zaborske JM, DuMont VLB, Wallace EWJ, Pan T, Aquadro CF, Drummond DA. 2014. A nutrient-driven tRNA modification alters translational fidelity and genome-wide protein coding across an animal genus. PLoS Biol 12:e1002015. doi: 10.1371/journal.pbio.1002015 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173. Tuorto F, Legrand C, Cirzi C, Federico G, Liebers R, Müller M, Ehrenhofer-Murray AE, Dittmar G, Gröne H-J, Lyko F. 2018. Queuosine‐modified tRNAs confer nutritional control of protein translation. EMBO J 37:e99777. doi: 10.15252/embj.201899777 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174. Michel AM, Baranov PV. 2013. Ribosome profiling: a Hi-Def monitor for protein synthesis at the genome-wide scale. Wiley Interdiscip Rev RNA 4:473–490. doi: 10.1002/wrna.1172 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175. Grosjean HJ, de Henau S, Crothers DM. 1978. On the physical basis for ambiguity in genetic coding interactions. Proc Natl Acad Sci U S A 75:610–614. doi: 10.1073/pnas.75.2.610 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176. Lampi M, Gregorova P, Qasim MS, Ahlblad NCV, Sarin LP. 2023. Bacteriophage infection of the marine bacterium Shewanella glacialimarina induces dynamic changes in tRNA modifications. Microorganisms 11:355. doi: 10.3390/microorganisms11020355 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177. Lei L, Burton ZF. 2022. “Superwobbling” and tRNA-34 wobble and tRNA-37 anticodon loop modifications in evolution and devolution of the genetic code. Life (Basel) 12:252. doi: 10.3390/life12020252 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178. Ehrenhofer-Murray AE. 2017. Cross-talk between Dnmt2-dependent tRNA methylation and queuosine modification. Biomolecules 7:14. doi: 10.3390/biom7010014 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179. Giegé R, Eriani G. 2023. The tRNA identity landscape for aminoacylation and beyond. Nucleic Acids Res 51:1528–1570. doi: 10.1093/nar/gkad007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180. Noguchi S, Nishimura Y, Hirota Y, Nishimura S. 1982. Isolation and characterization of an Escherichia coli mutant lacking tRNA-guanine transglycosylase. Function and biosynthesis of queuosine in tRNA. J Biol Chem 257:6544–6550. [PubMed] [Google Scholar]
  • 181. Pollo-Oliveira L, Davis NK, Hossain I, Ho P, Yuan Y, Salguero García P, Pereira C, Byrne SR, Leng J, Sze M, Blaby-Haas CE, Sekowska A, Montoya A, Begley T, Danchin A, Aalberts DP, Angerhofer A, Hunt J, Conesa A, Dedon PC, de Crécy-Lagard V. 2022. The absence of the queuosine tRNA modification leads to pleiotropic phenotypes revealing perturbations of metal and oxidative stress homeostasis in Escherichia coli K12. Metallomics 14:mfac065. doi: 10.1093/mtomcs/mfac065 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182. Wang X, Matuszek Z, Huang Y, Parisien M, Dai Q, Clark W, Schwartz MH, Pan T. 2018. Queuosine modification protects cognate tRNAs against ribonuclease cleavage. RNA 24:1305–1313. doi: 10.1261/rna.067033.118 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183. Ogawa T. 2016. tRNA-targeting ribonucleases: molecular mechanisms and insights into their physiological roles. Biosci Biotechnol Biochem 80:1037–1045. doi: 10.1080/09168451.2016.1148579 [DOI] [PubMed] [Google Scholar]
  • 184. Boccaletto P, Machnicka MA, Purta E, Piatkowski P, Baginski B, Wirecki TK, de Crécy-Lagard V, Ross R, Limbach PA, Kotter A, Helm M, Bujnicki JM. 2018. MODOMICS: a database of RNA modification pathways. 2017 update. Nucleic Acids Res 46:D303–D307. doi: 10.1093/nar/gkx1030 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185. Durand JM, Dagberg B, Uhlin BE, Björk GR. 2000. Transfer RNA modification, temperature and DNA superhelicity have a common target in the regulatory network of the virulence of Shigella flexneri: the expression of the virF gene. Mol Microbiol 35:924–935. doi: 10.1046/j.1365-2958.2000.01767.x [DOI] [PubMed] [Google Scholar]
  • 186. Thibessard A, Borges F, Fernandez A, Gintz B, Decaris B, Leblond-Bourget N. 2004. Identification of Streptococcus thermophilus CNRZ368 genes involved in defense against superoxide stress. Appl Environ Microbiol 70:2220–2229. doi: 10.1128/AEM.70.4.2220-2229.2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187. Nairn BL, Lonergan ZR, Wang J, Braymer JJ, Zhang Y, Calcutt MW, Lisher JP, Gilston BA, Chazin WJ, de Crécy-Lagard V, Giedroc DP, Skaar EP. 2016. The response of Acinetobacter baumannii to zinc starvation. Cell Host Microbe 19:826–836. doi: 10.1016/j.chom.2016.05.007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188. Jo J, Jang YS, Kim KY, Kim MH, Kim IJ, Chung WI. 1997. Isolation of ALU1-P gene encoding a protein with aluminum tolerance activity from Arthrobacter viscosus. Biochem Biophys Res Commun 239:835–839. doi: 10.1006/bbrc.1997.7567 [DOI] [PubMed] [Google Scholar]
  • 189. Águila-Clares B, Castiblanco LF, Quesada JM, Penyalver R, Carbonell J, López MM, Marco-Noales E, Sundin GW. 2018. Transcriptional response of Erwinia amylovora to copper shock: in vivo role of the copA gene. Mol Plant Pathol 19:169–179. doi: 10.1111/mpp.12510 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190. Pawlik M-C, Hubert K, Joseph B, Claus H, Schoen C, Vogel U. 2012. The zinc-responsive regulon of Neisseria meningitidis comprises 17 genes under control of a zur element. J Bacteriol 194:6594–6603. doi: 10.1128/JB.01091-12 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191. Quaranta D, McCarty R, Bandarian V, Rensing C. 2007. The copper-Inducible cin operon encodes an unusual methionine-rich azurin-Like protein and a Pre-Q0 reductase in Pseudomonas putida KT2440. J Bacteriol 189:5361–5371. doi: 10.1128/JB.00377-07 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192. Heindl JE, Hibbing ME, Xu J, Natarajan R, Buechlein AM, Fuqua C. 2016. Discrete responses to limitation for iron and manganese in Agrobacterium tumefaciens: influence on attachment and biofilm formation. J Bacteriol 198:816–829. doi: 10.1128/JB.00668-15 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193. Durand JM, Okada N, Tobe T, Watarai M, Fukuda I, Suzuki T, Nakata N, Komatsu K, Yoshikawa M, Sasakawa C. 1994. VACC, a virulence-associated chromosomal locus of Shigella flexneri, is homologous to tgt, a gene encoding tRNA-guanine transglycosylase (Tgt) of Escherichia coli K-12. J Bacteriol 176:4627–4634. doi: 10.1128/jb.176.15.4627-4634.1994 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194. Marchetti M, Capela D, Poincloux R, Benmeradi N, Auriac MC, Le Ru A, Maridonneau-Parini I, Batut J, Masson-Boivin C. 2013. Queuosine biosynthesis is required for Sinorhizobium meliloti-induced cytoskeletal modifications on HeLa cells and symbiosis with Medicago truncatula. PLoS One 8:e56043. doi: 10.1371/journal.pone.0056043 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195. Liu Y, Dong Y, Chen Y-Y, Burne RA. 2008. Environmental and growth phase regulation of the Streptococcus gordonii arginine deiminase genes. Appl Environ Microbiol 74:5023–5030. doi: 10.1128/AEM.00556-08 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196. Babosan A, Fruchard L, Krin E, Carvalho A, Mazel D, Baharoglu Z. 2022. Nonessential tRNA and rRNA modifications impact the bacterial response to sub-MIC antibiotic stress. Microlife 3:uqac019. doi: 10.1093/femsml/uqac019 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197. Morawska LP, Kuipers OP. 2022. Transcriptome analysis and prediction of the metabolic state of stress-induced viable but non-culturable Bacillus subtilis cells. Sci Rep 12:18015. doi: 10.1038/s41598-022-21102-w [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198. Price MN, Ray J, Iavarone AT, Carlson HK, Ryan EM, Malmstrom RR, Arkin AP, Deutschbauer AM. 2019. Oxidative pathways of deoxyribose and deoxyribonate catabolism. mSystems 4:e00297-18. doi: 10.1128/mSystems.00297-18 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199. Gonçalves LG, Santos S, Gomes LP, Armengaud J, Miragaia M, Coelho AV. 2022. Skin-to-blood pH shift triggers metabolome and proteome global remodelling in Staphylococcus epidermidis. Front Microbiol 13:1000737. doi: 10.3389/fmicb.2022.1000737 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200. Huber SM, Begley U, Sarkar A, Gasperi W, Davis ET, Surampudi V, Lee M, Melendez JA, Dedon PC, Begley TJ. 2022. Arsenite toxicity is regulated by queuine availability and oxidation-induced reprogramming of the human tRNA epitranscriptome. Proc Natl Acad Sci U S A 119:e2123529119. doi: 10.1073/pnas.2123529119 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201. Nagaraja S, Cai MW, Sun J, Varet H, Sarid L, Trebicz-Geffen M, Shaulov Y, Mazumdar M, Legendre R, Coppée J-Y, Begley TJ, Dedon PC, Gourinath S, Guillen N, Saito-Nakano Y, Shimokawa C, Hisaeda H, Ankri S. 2021. Queuine is a nutritional regulator of Entamoeba histolytica response to oxidative stress and a virulence attenuator. mBio 12:e03549-20. doi: 10.1128/mBio.03549-20 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202. Gu C, Begley TJ, Dedon PC. 2014. tRNA modifications regulate translation during cellular stress. FEBS Lett 588:4287–4296. doi: 10.1016/j.febslet.2014.09.038 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203. Pollo-Oliveira L, de Crécy-Lagard V. 2019. Can protein expression be regulated by modulation of tRNA Modification profiles? Biochemistry 58:355–362. doi: 10.1021/acs.biochem.8b01035 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204. Rajakovich LJ, Balskus EP. 2019. Metabolic functions of the human gut microbiota: the role of metalloenzymes. Nat Prod Rep 36:593–625. doi: 10.1039/c8np00074c [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 205. Haas CE, Rodionov DA, Kropat J, Malasarn D, Merchant SS, de Crécy-Lagard V. 2009. A subset of the diverse COG0523 family of putative metal chaperones is linked to zinc homeostasis in all kingdoms of life. BMC Genomics 10:470. doi: 10.1186/1471-2164-10-470 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206. Helm M, Alfonzo JD. 2014. Post-transcriptional RNA modifications: playing metabolic games in a cell’s chemical legoland. Chem Biol 21:174–185. doi: 10.1016/j.chembiol.2013.10.015 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207. Xie D, Wen Y, Chen J, Lu H, He H, Liu Z. 2022. Probing queuosine modifications of transfer RNA in single living cells via plasmonic affinity sandwich assay. Anal Chem 94:12828–12835. doi: 10.1021/acs.analchem.2c02784 [DOI] [PubMed] [Google Scholar]
  • 208. Huber LB, Kaur N, Henkel M, Marchand V, Motorin Y, Ehrenhofer-Murray AE, Marx A. 2023. A dual-purpose polymerase engineered for direct sequencing of pseudouridine and queuosine. Nucleic Acids Res 51:3971–3987. doi: 10.1093/nar/gkad177 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209. Cirzi C, Tuorto F. 2021. Analysis of queuosine tRNA modification using APB northern blot assay. Methods Mol Biol 2298:217–230. doi: 10.1007/978-1-0716-1374-0_14 [DOI] [PubMed] [Google Scholar]
  • 210. Zhang W, Xu R, Matuszek Ż, Cai Z, Pan T. 2020. Detection and quantification of glycosylated queuosine modified tRNAs by acid denaturing and APB gels. RNA 26:1291–1298. doi: 10.1261/rna.075556.120 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211. Katanski CD, Watkins CP, Zhang W, Reyer M, Miller S, Pan T. 2022. Analysis of queuosine and 2-thio tRNA modifications by high throughput sequencing. Nucleic Acids Res 50:e99–e99. doi: 10.1093/nar/gkac517 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212. Sun Y, Piechotta M, Naarmann-de Vries I, Dieterich C, Ehrenhofer-Murray AE. 2023. Detection of queuosine and queuosine precursors in tRNAs by direct RNA sequencing. Nucleic Acids Res 51:11197–11212. doi: 10.1093/nar/gkad826 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213. Nedialkova DD, Leidel SA. 2015. Optimization of codon translation rates via tRNA modifications maintains proteome integrity. Cell 161:1606–1618. doi: 10.1016/j.cell.2015.05.022 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 214. Caballero VC, Toledo VP, Maturana C, Fisher CR, Payne SM, Salazar JC. 2012. Expression of Shigella flexneri gluQ-RS gene is linked to dksA and controlled by a transcriptional terminator. BMC Microbiol 12:226. doi: 10.1186/1471-2180-12-226 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215. Ames BN. 2018. Prolonging healthy aging: longevity vitamins and proteins. Proc Natl Acad Sci U S A 115:10836–10844. doi: 10.1073/pnas.1809045115 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216. Varghese S, Cotter M, Chevot F, Fergus C, Cunningham C, Mills KH, Connon SJ, Southern JM, Kelly VP. 2017. In vivo modification of tRNA with an artificial nucleobase leads to full disease remission in an animal model of multiple sclerosis. Nucleic Acids Res 45:2029–2039. doi: 10.1093/nar/gkw847 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217. Cirzi C, Dyckow J, Legrand C, Schott J, Guo W, Perez Hernandez D, Hisaoka M, Parlato R, Pitzer C, van der Hoeven F, Dittmar G, Helm M, Stoecklin G, Schirmer L, Lyko F, Tuorto F. 2023. Queuosine‐tRNA promotes sex‐dependent learning and memory formation by maintaining codon‐biased translation elongation speed. EMBO J 42:e112507. doi: 10.15252/embj.2022112507 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218. Skolnick SD, Greig NH. 2019. Microbes and monoamines: potential neuropsychiatric consequences of dysbiosis. Trends Neurosci 42:151–163. doi: 10.1016/j.tins.2018.12.005 [DOI] [PubMed] [Google Scholar]
  • 219. Magnúsdóttir S, Ravcheev D, de Crécy-Lagard V, Thiele I. 2015. Systematic genome assessment of B-vitamin biosynthesis suggests co-operation among gut microbes. Front Genet 6:148. doi: 10.3389/fgene.2015.00148 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 220. Rodionov DA, Arzamasov AA, Khoroshkin MS, Iablokov SN, Leyn SA, Peterson SN, Novichkov PS, Osterman AL. 2019. Micronutrient requirements and sharing capabilities of the human gut microbiome. Front Microbiol 10:1316. doi: 10.3389/fmicb.2019.01316 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 221. Varriale L, Coretti L, Dipineto L, Green BD, Pace A, Lembo F, Menna LF, Fioretti A, Borrelli L. 2022. An outdoor access period improves chicken cecal microbiota and potentially increases micronutrient biosynthesis. Front Vet Sci 9:904522. doi: 10.3389/fvets.2022.904522 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222. Blaby IK, Phillips G, Blaby-Haas CE, Gulig KS, El Yacoubi B, de Crécy-Lagard V. 2010. Towards a systems approach in the genetic analysis of archaea: accelerating mutant construction and phenotypic analysis in Haloferax volcanii. Archaea 2010:426239. doi: 10.1155/2010/426239 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223. Orita I, Futatsuishi R, Adachi K, Ohira T, Kaneko A, Minowa K, Suzuki M, Tamura T, Nakamura S, Imanaka T, Suzuki T, Fukui T. 2019. Random mutagenesis of a hyperthermophilic archaeon identified tRNA modifications associated with cellular hyperthermotolerance. Nucleic Acids Res 47:1964–1976. doi: 10.1093/nar/gky1313 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 224. Kulikov EE, Golomidova AK, Letarova MA, Kostryukova ES, Zelenin AS, Prokhorov NS, Letarov AV. 2014. Genomic sequencing and biological characteristics of a novel Escherichia coli bacteriophage 9g, a putative representative of a new Siphoviridae genus. Viruses 6:5077–5092. doi: 10.3390/v6125077 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225. Tsai R, Corrêa IR, Xu MY, Xu SY. 2017. Restriction and modification of deoxyarchaeosine (dG+)-containing phage 9 g DNA. Sci Rep 7:8348. doi: 10.1038/s41598-017-08864-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 226. Waters TR, Connolly BA. 1994. Interaction of the restriction endonuclease EcoRV with the deoxyguanosine and deoxycytidine bases in Its recognition sequence. Biochemistry 33:1812–1819. doi: 10.1021/bi00173a026 [DOI] [PubMed] [Google Scholar]
  • 227. ACS G, REICH E, MORI M. 1964. Biological and biochemical properties of the analogue antibiotic tubercidin. Proc Natl Acad Sci U S A 52:493–501. doi: 10.1073/pnas.52.2.493 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 228. Igloi GL, Kössel H. 1985. Affinity electrophoresis for monitoring terminal phosphorylation and the presence of queuosine in RNA. Application of polyacrylamide containing a covalently bound boronic acid. Nucleic Acids Res 13:6881–6898. doi: 10.1093/nar/13.19.6881 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229. Chan CTY, Dyavaiah M, DeMott MS, Taghizadeh K, Dedon PC, Begley TJ. 2010. A quantitative systems approach reveals dynamic control of tRNA modifications during cellular stress. PLoS Genet 6:e1001247. doi: 10.1371/journal.pgen.1001247 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 230. Gregorova P, Sipari NH, Sarin LP. 2021. Broad-range RNA modification analysis of complex biological samples using rapid C18-UPLC-MS. RNA Biol 18:1382–1389. doi: 10.1080/15476286.2020.1853385 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 231. Kimura S, Dedon PC, Waldor MK. 2020. Comparative tRNA sequencing and RNA mass spectrometry for surveying tRNA modifications. Nat Chem Biol 16:964–972. doi: 10.1038/s41589-020-0558-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 232. Bessler L, Kaur N, Vogt LM, Flemmich L, Siebenaller C, Winz ML, Tuorto F, Micura R, Ehrenhofer-Murray AE, Helm M. 2022. Functional integration of a semi-synthetic azido-queuosine derivative into translation and a tRNA modification circuit. Nucleic Acids Res 50:10785–10800. doi: 10.1093/nar/gkac822 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 233. Sheldon RA, Woodley JM. 2018. Role of biocatalysis in sustainable chemistry. Chem Rev 118:801–838. doi: 10.1021/acs.chemrev.7b00203 [DOI] [PubMed] [Google Scholar]
  • 234. Hoye RC. 1999. Reductions by the alumino- and borohydrides in organic synthesis, 2nd edition (Seyden-Penne, Jacqueline). J Chem Educ 76:33. doi: 10.1021/ed076p33.2 [DOI] [Google Scholar]
  • 235. Laval S, Dayoub W, Favre-Reguillon A, Berthod M, Demonchaux P, Mignani G, Lemaire M. 2009. A mild and efficient method for the reduction of nitriles. Tetrahedron Lett 50:7005–7007. doi: 10.1016/j.tetlet.2009.09.164 [DOI] [Google Scholar]
  • 236. Bornschein C, Werkmeister S, Junge K, Beller M. 2013. TBAF-catalyzed hydrosilylation for the reduction of aromatic nitriles. New J Chem 37:2061. doi: 10.1039/c3nj00171g [DOI] [Google Scholar]
  • 237. Mukherjee A, Srimani D, Chakraborty S, Ben-David Y, Milstein D. 2015. Selective hydrogenation of nitriles to primary amines catalyzed by a cobalt pincer complex. J Am Chem Soc 137:8888–8891. doi: 10.1021/jacs.5b04879 [DOI] [PubMed] [Google Scholar]
  • 238. Arnold FH. 2018. Directed evolution: bringing new chemistry to life. Angew Chem Int Ed Engl 57:4143–4148. doi: 10.1002/anie.201708408 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 239. Isaev AB, Musharova OS, Severinov KV. 2021. Microbial arsenal of antiviral defenses. Part II. Biochemistry (Mosc) 86:449–470. doi: 10.1134/S0006297921040064 [DOI] [PubMed] [Google Scholar]
  • 240. Schmitt SM, Stefan K, Wiese M. 2016. Pyrrolopyrimidine derivatives as novel inhibitors of multidrug resistance-associated protein 1 (MRP1, ABCC1). J Med Chem 59:3018–3033. doi: 10.1021/acs.jmedchem.5b01644 [DOI] [PubMed] [Google Scholar]
  • 241. Stipković Babić M, Makuc D, Plavec J, Martinović T, Kraljević Pavelić S, Pavelić K, Snoeck R, Andrei G, Schols D, Wittine K, Mintas M. 2015. Novel halogenated 3-deazapurine, 7-deazapurine and alkylated 9-deazapurine derivatives of l-ascorbic or imino-l-ascorbic acid: synthesis, antitumour and antiviral activity evaluations. Eur J Med Chem 102:288–302. doi: 10.1016/j.ejmech.2015.08.008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 242. Ribociclib (Kisqali). Available from: https://www.fda.gov/drugs/resources-information-approved-drugs/ribociclib-kisqali
  • 243. Utility of deazapurines in drug discovery. Available from: https://www.pharmablock.com/web/upload/2020/10/30/16040241028608o6qc1.pdf
  • 244. Lawhorn BG, Philp J, Zhao Y, Louer C, Hammond M, Cheung M, Fries H, Graves AP, Shewchuk L, Wang L, Cottom JE, Qi H, Zhao H, Totoritis R, Zhang G, Schwartz B, Li H, Sweitzer S, Holt DA, Gatto GJ, Kallander LS. 2015. Identification of purines and 7-deazapurines as potent and selective type I inhibitors of troponin I-interacting kinase (TNNI3K). J Med Chem 58:7431–7448. doi: 10.1021/acs.jmedchem.5b00931 [DOI] [PubMed] [Google Scholar]
  • 245. Aleksander SA, Balhoff J, Carbon S, Cherry JM, Drabkin HJ, Ebert D, Feuermann M, Gaudet P, Harris NL, Hill DP, et al. 2023. The gene ontology knowledgebase in 2023. Genetics 224:iyad031. doi: 10.1093/genetics/iyad031 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 246. Kanehisa M, Furumichi M, Sato Y, Ishiguro-Watanabe M, Tanabe M. 2021. KEGG: integrating viruses and cellular organisms. Nucleic Acids Res 49:D545–D551. doi: 10.1093/nar/gkaa970 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 247. Caspi R, Billington R, Fulcher CA, Keseler IM, Kothari A, Krummenacker M, Latendresse M, Midford PE, Ong Q, Ong WK, Paley S, Subhraveti P, Karp PD. 2018. The MetaCyc database of metabolic pathways and enzymes. Nucleic Acids Res 46:D633–D639. doi: 10.1093/nar/gkx935 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 248. Zhu J, Yang F, Du K, Wei ZL, Wu QF, Chen Y, Li WF, Li Q, Zhou CZ. 2023. Phylogenomics of five Pseudanabaena cyanophages and evolutionary traces of horizontal gene transfer. Environ Microbiome 18:3. doi: 10.1186/s40793-023-00461-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 249. Zhang J, Lu R, Zhang Y, Matuszek Ż, Zhang W, Xia Y, Pan T, Sun J. 2020. tRNA Queuosine modification enzyme modulates the growth and microbiome recruitment to breast tumors. Cancers (Basel) 12:628. doi: 10.3390/cancers12030628 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 250. Zhang J, Zhang Y, McGrenaghan CJ, Kelly VP, Xia Y, Sun J. 2023. Disruption to tRNA modification by queuine contributes to inflammatory bowel disease. Cell Mol Gastroenterol Hepatol 15:1371–1389. doi: 10.1016/j.jcmgh.2023.02.006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 251. Husso A, Pessa-Morikawa T, Koistinen VM, Kärkkäinen O, Kwon HN, Lahti L, Iivanainen A, Hanhineva K, Niku M. 2023. Impacts of maternal microbiota and microbial metabolites on fetal intestine, brain, and placenta. BMC Biol 21:207. doi: 10.1186/s12915-023-01709-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 252. Aučynaitė A, Rutkienė R, Tauraitė D, Meškys R, Urbonavičius J. 2018. Identification of a 2′-O-methyluridine nucleoside hydrolase using the metagenomic libraries. Molecules 23:E2904. doi: 10.3390/molecules23112904 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 253. Preumont A, Snoussi K, Stroobant V, Collet J-F, Van Schaftingen E. 2008. Molecular identification of pseudouridine-metabolizing enzymes. J Biol Chem 283:25238–25246. doi: 10.1074/jbc.M804122200 [DOI] [PubMed] [Google Scholar]
  • 254. Chevrette MG, Currie CR. 2019. Emerging evolutionary paradigms in antibiotic discovery. J Ind Microbiol Biotechnol 46:257–271. doi: 10.1007/s10295-018-2085-6 [DOI] [PubMed] [Google Scholar]

Articles from Microbiology and Molecular Biology Reviews : MMBR are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES