Abstract 6-thioguanine (6-TG) is a therapeutic medication for childhood acute lymphoblastic leukemia (ALL) and a potent antimicrobial agent. Its biosynthesis relies on the YcfA-YcfC system, yet the formation of its critical thioamide moiety remains incompletely understood. Here, we provide a detailed biochemical and structural characterization of YcfA, including apo and substrate-bound crystal structures, which reveal that substrate adenylation and L-cysteine addition are key initial steps in the reaction cascade. Cryo-electron microscopy (cryo-EM) and functional analyses highlight YcfA’s assembly into a two-layered heptameric structure, essential for the enzymatic function. GTP serves a dual role as a substrate and oligomerization enhancer. Additionally, pyridoxal 5’-phosphate (PLP), a cofactor for YcfC, the partner enzyme in this system, promotes YcfA oligomerization but inhibits its activity by obstructing GTP binding. Biochemical and structural evidence confirms that YcfC acts as a C‒S lyase, which is essential for thioamide formation in the presence of PLP. Exploiting substrate flexibility, we synthesized a seleno analog with antimicrobial properties. Multi-omics analyses of the biosynthetic precursor underscore its potential as an antibiotic. Collectively, our findings unravel the distinct architecture and functionality of the YcfA-YcfC system, offering an evolutionary perspective on noncanonical thioamide biosynthesis and a foundation for synthetic biology applications in drug development. Subject terms: Biocatalysis, X-ray crystallography, Enzyme mechanisms __________________________________________________________________ Here, the authors structurally and biochemically characterise the YcfA-YcfC enzyme system, crucial for biosynthesis of the anticancer and antimicrobial agent 6-thioguanine. Introduction Thioamides, characterized by the substitution of sulfur for oxygen in the amide group (C(=S)NH), are critical components of numerous metabolites, exhibiting antimicrobial and anticancer properties^[44]1–[45]3. Their diverse bioactivities make them compelling targets for drug discovery. Prominent examples include methanobactin, a metal-chelating compound^[46]4–[47]6, and closthioamide, an antibiotic agent^[48]7,[49]8. Among these, 6-thioguanine (6-TG) stands out as a clinically utilized antimetabolite for treating acute lymphoblastic leukemia (ALL)^[50]9. Beyond oncology, 6-TG demonstrates antimicrobial activity against Staphylococcus aureus^[51]10 and antiviral effects against coronaviruses^[52]11,[53]12. Intriguingly, it also functions as a virulence factor in Erwinia amylovora, the causative agent of fire blight in rosaceous plants^[54]13,[55]14. The biosynthesis of the thioamide group in 6-TG, catalyzed by the YcfA-YcfC system^[56]13,[57]15, was initially thought to follow the canonical pathways for thioamide formation in nucleobases^[58]16. These involve alpha adenine nucleotide hydrolase (AANH), such as ThiI, MnmA, TtuA, TtcA, or Ncs6, and a cysteine desulfurase. Together, they facilitate the transfer of sulfur from L-cysteine or cystine via a persulfide intermediate, often mediated by persulfide carrier proteins like TusA, TusBCD, or TusE. The reactive persulfide intermediate then reacts with an ATP-adenylated nucleobase, to form the thioamide group, requiring multiple steps and complex machinery. Emerging evidence suggests that the YcfA‒YcfC system employs a more streamlined mechanism^[59]17. Unlike typical AANH enzymes, YcfA bypasses the persulfide intermediate, directly utilizing L-cysteine to produce a cysteine–guanine nucleotide adduct. YcfC, functioning as a C‒S lyase, then cleaves this adduct to generate the thioamide group in 6-TG. This process eliminates the need for mobile persulfide intermediates, presenting a biochemical strategy for thioamidation. Despite these advances, key questions remain: (1) How does YcfA specifically recognize and utilize L-cysteine for nucleotide modification? (2) What are the molecular mechanisms underlying YcfC’s interaction with the cysteine-conjugate intermediate? (3) How do YcfA and YcfC coordinate their activities to ensure efficient biosynthesis? Addressing these questions is crucial for advancing our understanding of thioamide biosynthesis and its applications. This research opens avenues for the rational design of bioactive compounds, enabling the engineering of 6-TG analogs for diverse therapeutic applications, such as cancer treatment, antiviral therapies, and antimicrobial development. In this work, we determine a series of crystal structures of YcfA and YcfC, complemented by molecular docking and biochemical assays. Together with the cryo-EM analysis, we reveal YcfA’s two-layered heptameric architecture, which is essential for its catalytic activity. We also examine the regulatory interplay within the YcfA-YcfC system, discovering that YcfA oligomerization is promoted by its nucleotide substrates and the YcfC cofactor pyridoxal 5’-phosphate (PLP). Interestingly, PLP exhibits dual effects, enhancing oligomerization while inhibiting enzymatic activity, suggesting a sophisticated feedback mechanism. Furthermore, we synthesize a seleno analog, 6-seleno-GTP, which exhibits potent antimicrobial activity against multidrug-resistant S. aureus. Multi-omics analyses further elucidate its antibacterial mechanism. These findings offer insights into this biosynthetic pathway and its potential applications in drug discovery. Results In vitro reconstitution of the YcfA-YcfC system To elucidate the catalytic mechanism of the YcfA-YcfC system in synthesizing 6-thioguanine (6-TG) nucleotides, we reconstituted the system in vitro using various substrates, including sulfur and selenium sources—L-cysteine (L-Cys) and L-selenocysteine (L-SeCys)—and guanine nucleotides (GTP, GDP, and GMP). Consistent with previous studies^[60]15,[61]17, combining YcfA, YcfC, ATP, Mg²⁺, and L-Cys with each guanine nucleotide yielded products with absorption peaks at 340 nm (Fig. [62]1a; Supplementary Fig. [63]1a). High-performance liquid chromatography (HPLC) and high-resolution mass spectrometry (HRMS) identified these products as 6-thio-GTP, 6-thio-GDP, and 6-thio-GMP (Fig. [64]1a; Supplementary Fig. [65]1b, c). With more 6-thio-GTP produced, GTP was selected as the optimal substrate for further studies. Substituting L-Cys with L-SeCys resulted in corresponding 6-selenoguanine nucleotides with a redshifted absorption peak at 360 nm (Fig. [66]1b; Supplementary Fig. [67]1d). Fig. 1. In vitro reconstitution of the YcfA-YcfC system. [68]Fig. 1 [69]Open in a new tab a UV–Vis spectra detection of the thioamide formation catalyzed by the YcfA-YcfC system using L-Cys and GTP as substrates. A characteristic absorption peak at 340 nm was observed. The molecular weight (as determined by MS) and structure of the product are shown. b UV–Vis spectra detection of the selenoamide formation catalyzed by the YcfA-YcfC system using L-SeCys and GTP as substrates. The resulting product exhibits an absorption peak at 360 nm. The molecular weight (as determined by MS) and structure of the product are shown. c, d Continuous monitoring of the reaction process of the YcfA-YcfC system. c Incubation of YcfA with ATP, GTP, and L-Cys results in increased absorbance at 315 nm over 200 s. Subsequent addition of YcfC leads to a decrease in absorbance at 315 nm, concomitant with the emergence of a new peak at 340 nm. d YcfC alone showed no activity with ATP, GTP, and L-Cys. Addition of YcfA produced simultaneous absorption peaks at 315 nm and 340 nm. e The formation of the 315 nm intermediate requires the adenylation activity of YcfA. f Adenylation activity assay of YcfA with diverse amino acids and 3MP. Only L-Cys and 3MP yielded the 315 nm peak. g Proposed catalytic steps of the YcfA-YcfC system in 6-Thioguanine nucleotides formation, however, the detailed regulatory and catalytic mechanisms of the two enzymes remain elusive. Source data are provided as a Source Data file. To delineate the individual roles of YcfA and YcfC, we monitored the reaction kinetics using high-frequency UV‒visible absorption scans. Sequential enzyme addition experiments revealed that YcfA catalyzes the initial formation of a 315 nm intermediate, which YcfC subsequently converts into the final product, 6-thio-GTP (340 nm) (Fig. [70]1c). Starting with YcfC produced no detectable product; however, adding YcfA afterward led to the sequential formation of both the 315 nm intermediate and the 340 nm final product (Fig. [71]1d). These results underscore YcfA’s role in initiating the reaction and YcfC’s role in processing the intermediate. Further analysis revealed that ATP, GTP, and L-Cys are essential for forming the 315 nm intermediate (Fig. [72]1e). As an AANH-like enzyme^[73]15, YcfA likely catalyzes the adenylation of GTP, activating its amide carbonyl group. The resulting adenylated GTP, though unstable and undetectable via UV‒visible absorption^[74]18, reacts with the thiol group of L-Cys to form the 315 nm intermediate. Testing alternative amino acids confirmed the critical role of the thiol group, as only L-Cys produced the 315 nm peak (Fig. [75]1f). Similarly, replacing L-Cys with mercaptopyruvate (3MP), which lacks only an amine group, also yielded a product absorbing at 315 nm (Fig. [76]1f). These findings suggest that the thiol group of L-Cys interacts with the activated amide carbonyl of guanine in adenylated GTP, forming a 6-Cys-GTP S-adduct. Tandem mass spectrometry confirmed this adduct (Supplementary Fig. [77]1e, f). YcfA’s dual catalytic roles are thus evident: it catalyzes the adenylation of guanine nucleotides and facilitates the formation of 6-Cys-guanine or 6-SeCys-guanine nucleotide S-adducts. YcfC, acting as a C–S or C–Se lyase, cleaves these intermediates to produce 6-thio-guanine or 6-seleno-guanine nucleotides, as depicted in Fig. [78]1g. However, the inability to isolate and independently measure the adenylation step precluded precise differentiation of YcfA’s contributions to each step. These results align with recent studies highlighting YcfA’s dual catalytic functions and YcfC’s specific role as a C–S lyase^[79]17. The dual roles of YcfA in adenylation and L-Cys adduct formation To elucidate the dual catalytic roles of YcfA, we determined the crystal structures of apo-YcfA, YcfA in complexes with ATP (YcfA–ATP) or GTP (YcfA–GTP), and an inactive variant, YcfA^D19A complexed with ATP and GTP (YcfA^D19A–ATP–GTP) (Supplementary Table [80]1). Unexpectedly, YcfA assembles into a two-layered heptameric ring, a distinct arrangement in this enzyme family (Fig. [81]2a). Structural analyses revealed that catalytic cavities are located at the interface of adjacent monomers across the two heptameric layers, forming a catalytic dimer unit (Fig. [82]2b). Notably, a single monomer lacks a complete cavity for guanine substrates, highlighting the interdependence of monomers for substrate binding. Fig. 2. The dual roles of YcfA in adenylation and L-Cys adduct formation. [83]Fig. 2 [84]Open in a new tab a A two-layered heptameric architecture of YcfA, with each monomer distinctly colored. b The catalytic cavities of adjacent monomers across the two heptameric layers. c ATP binding mode in the YcfA-ATP complex structure. Interaction networks formed between ATP and surrounding residues of YcfA, with hydrogen bonds represented as blue sticks. d Adenylation activity assay of YcfA mutants targeting ATP binding. The WT YcfA is regarded as possessing 100% enzyme activity. The mean is displayed ±SD for n = 3 biological replicates. e GTP binding mode in the YcfA-GTP complex structure. The polar interactions formed between GTP and surrounding residues are represented as blue sticks, and the π-π stacking interaction is shown as yellow stick. f Adenylation activity assay of YcfA mutants targeting GTP binding. The WT YcfA is regarded as possessing 100% enzyme activity. The mean is displayed ±SD for n = 3 biological replicates. g Surface representation of YcfA^D19A-ATP-GTP complex. ATP and GTP are shown as sticks, Mg^2+ is shown as green sphere. h The cooperative binding of ATP and GTP in the complex structure of YcfA^D19A-ATP-GTP. Interaction networks in the active site involving ATP, GTP, Mg^2+, and surrounding residues are presented, with polar contacts represented as blue sticks. i Docking simulations of YcfA with adenylated-GTP and L-Cys. The surrounding residues of L-Cys binding are presented, with potential polar contacts represented as blue sticks. j Mutagenesis analysis of the L-Cys binding site shows YcfA’s relative activity in forming the 6-Cys-GTP S-adduct. The WT YcfA is regarded as possessing 100% enzyme activity. The mean is displayed ±SD for n = 3 biological replicates. k Proposed catalytic mechanism of YcfA. Key residues involved in adenylation and L-Cys nucleophilic addition are indicated. Source data are provided as a Source Data file. In the YcfA–ATP complex, ATP is deeply embedded in the catalytic cavity and forms critical hydrogen bonds with Asp19, Gln129, and Glu183, stabilizing the transition state during ATP cleavage (Fig. [85]2c; Supplementary Fig. [86]2a). Mutations at any of these residues abolished 6-Cys-GTP S-adduct production, underscoring their essential roles in adenylation. Pyrophosphate is the leaving group in this reaction, facilitated via these residues by promoting the phosphodiester bond cleavage (Fig. [87]2d). In the YcfA–GTP structure, GTP binds near the catalytic cavity entrance, with its guanine ring stabilized by forming interactions with Phe132 and Glu134 (Fig. [88]2e; Supplementary Fig. [89]2b). Glu134 serves as a general base, deprotonating the guanine ring to activate the amide carbonyl for adenylation (Fig. [90]2f). The triphosphate tail of GTP is anchored by Arg18, Lys169, and Lys180 from the adjacent monomer. Mutation of Arg18 significantly reduced S-adduct formation, suggesting its dual roles in GTP binding and facilitating subsequent steps in catalysis (Fig. [91]2f). The YcfA^D19A–ATP–GTP complex revealed the interactions between ATP and GTP, with Gln135 aligning the substrates by interacting with the α-phosphate of ATP and the guanine base of GTP (Fig. [92]2g, h). Mutation of Gln135 reduced S-adduct production (Fig. [93]2f). In this inactive mutant, Glu183 coordinates with all three phosphate groups of ATP via Mg²⁺, but the absence of interaction with Asp19 and Gln129 prevents the stabilization of ATP transition state, underscoring the cooperative roles of these residues in adenylation (Fig. [94]2c, h). Following adenylation, YcfA catalyzes the incorporation of L-Cys into adenylated GTP. Substituting L-Cys with inorganic sulfides failed to produce detectable thiolated GTP (Supplementary Fig. [95]3), highlighting YcfA’s specificity for L-Cys. Structural data from the YcfA^D19A–ATP–GTP complex suggest that L-Cys binding is sterically hindered until pyrophosphate release creates sufficient space. Molecular docking simulations revealed that adenylated GTP undergoes a conformational change upon pyrophosphate release, enabling L-Cys binding (Fig. [96]2h, i). The carboxylate group of L-Cys is anchored by Arg18, while Tyr128 and Gln129 stabilize its orientation (Fig. [97]2i). Substitution of Tyr128 with alanine abolished S-adduct formation, whereas phenylalanine substitution had minimal effect, underscoring the aromatic ring’s role in L-Cys positioning. Gln129 likely facilitates thiol nucleophilic attack by deprotonating L-Cys (Fig. [98]2j) and is unable to catalyze this step with L-serine (Fig. [99]1f), further confirming substrate specificity. YcfA’s dual catalytic steps, adenylation and L-Cys adduct formation, are mediated by distinct interdependent residues. Asp19, Gln129, and Glu183 stabilize the ATP transition state, while Glu134 deprotonates the guanine ring to drive adenylation (Fig. [100]2k). Pyrophosphate release creates space for L-Cys accommodation, where Gln129 facilitate nucleophilic addition to adenylated GTP. Disruption of any key residue compromises one or both catalytic steps, highlighting the intricate coordination required for YcfA’s enzymatic function. These findings provide mechanistic insights into YcfA’s dual roles and broader implications for this enzyme family. Tetradecamer assembly is essential for YcfA activity Beyond its dual catalytic functions, YcfA adopts a rare tetradecamer arrangement, comprising a two-layered heptameric architecture as revealed by crystal structures and cryo-EM (Figs. [101]2a, [102]3a; Supplementary Fig. [103]4; Supplementary Table [104]2). This physiologically relevant structure could not be predicted by computational methods like AlphaFold3^[105]19 without experimental confirmation. Retrospective size-exclusion chromatography (SEC) data revealed that increasing protein concentration decreased the apparent elution volume (Fig. [106]3b), suggesting concentration-dependent oligomerization. Fig. 3. Nucleotide substrates drive functional YcfA assembly. [107]Fig. 3 [108]Open in a new tab a Cryo-EM density map of YcfA with the fitted atomic model. Side view (top) shows the two layers, and the top view (bottom) reveals the heptameric arrangement in each layer, forming a two-layered heptameric (tetradecameric) assembly. b SEC analysis of YcfA at varying concentrations (0.5 mg/mL, 1 mg/mL, 2 mg/mL and 4 mg/mL). Inset presents a calibration curve for SEC using proteins of known sizes. c SEC analysis of YcfA^L173D and YcfA^D166R/Y170N. d The adenylation activity assays of YcfA^L173D and YcfA^D166R/Y170N using ATP, GTP, and L-Cys as substrates. The WT YcfA is regarded as possessing 100% enzyme activity. The mean is displayed ± SD for n = 3 biological replicates. e The effects of different substrates (Mg^2+, L-Cys, ATP, and GTP) in promoting YcfA oligomerization were analyzed by SEC. Both ATP and GTP can promote the assembly of YcfA. f SEC analysis of YcfA with GMP, GDP, and GTP showed that GDP had a weaker effect than GTP on oligomerization, whereas GMP had a negligible effect. Source data are provided as a Source Data file. To probe the functional significance of the tetradecamer, we introduced mutations at the oligomeric interfaces based on structural analysis (Supplementary Fig. [109]5a–f). The K180A mutation promoted polymerization but severely reduced enzymatic activity (Supplementary Fig. [110]5e, f). As Lys180 is critical for guanine substrate binding at the interface, this mutation likely disrupts optimal substrate alignment (Fig. [111]2e, f). To mitigate substrate-binding confounders, we further introduced the L173D mutation, disrupting a hydrophobic interface region. This resulted in a stable dimer (confirmed by analytical ultracentrifugation, AUC) but without enzymatic activity (Fig. [112]3c, d; Supplementary Fig. [113]5g–i). Conversely, the D166R/Y170N double mutant enhanced interfacial stability through synergistic modifications: The Y170N substitution was designed to introduce a more favorable hydrogen-bonding network with Asp190 and potentially introduce a new polar interaction with Arg78. This change was expected to increase the number of contacts between the two layers of YcfA. As for Asp166, while it is located near Interface III, it is positioned at the interlayer boundary. The D166R mutation changes the charge from negative to positive, which could enhance interfacial stability by forming a polar interaction with Asp190. (Fig. [114]3c, d; Supplementary Fig. [115]5d, j). These results indicate that YcfA’s tetradecameric assembly is crucial for activity, as it forms a continuous ring of catalytic sites, improving substrate access and residue positioning. Notably, the D166R/Y170N mutant exhibited stable oligomerization compared to the variable forms of the wild type, presenting the potential for industrial applications, particularly in enhancing 6-TG production. Nucleotide substrates drive functional YcfA assembly The active tetradecamer form of YcfA contrasts with its presence in multiple oligomeric states in solution, suggesting a substrate-induced mechanism for functional assembly. To test this, we analyzed the effects of ATP, GTP, Mg²⁺, and L-Cys on oligomerization. SEC showed that ATP and GTP facilitated assembly, with GTP exhibiting a stronger effect (Fig. [116]3e). AUC further confirmed that both nucleotides predominantly promoted tetradecamer formation, with increased resistance to trypsin digestion supporting the enhanced oligomerization (Supplementary Figs. [117]6a–c). The stronger effect of GTP arises from its binding characteristics. While ATP binds deeply within a monomer cavity, GTP binds at the catalytic interface, with its guanine base interacting with one monomer and its phosphate tail engaging Lys169, Lys180, and Arg18 of an adjacent monomer (Fig. [118]2e). This inter-monomer linkage is critical, as mutations at Lys169 or Arg18 disrupted GTP-induced assembly and binding (Supplementary Fig. [119]6d). Surprisingly, the K180A mutant promoted stable oligomerization independent of GTP, though the mechanism remains unclear, highlighting the regulatory role of subtle interface modifications (Supplementary Fig. [120]6d). Our biochemical assays demonstrated that the YcfA-YcfC system exhibits a strong preference for GTP over GDP and GMP, with GTP being the optimal substrate. This preference is linked to their phosphate group compositions (Supplementary Fig. [121]6e). SEC analysis confirmed that GTP strongly facilitated oligomerization, GDP had a weaker effect, and GMP had negligible influence (Fig. [122]3f). These findings suggest that GTP’s interactions at the catalytic interface drive the transition from diverse oligomeric states to a stable tetradecamer, modulating catalytic activity. PLP promotes YcfA assembly by competing with GTP Building on our investigation of YcfA substrates, we examined whether molecules associated with YcfC could modulate YcfA assembly. Our results identified three basic residues at the catalytic interface, which are essential for GTP binding and pivotal to YcfA’s functional assembly. Given that PLP, the cofactor of YcfC, is known to covalently bind to lysine residues^[123]20–[124]24, we explored its potential influence on YcfA oligomerization. Remarkably, PLP facilitated YcfA assembly in a manner comparable to GTP. SEC analysis revealed that PLP coeluted with YcfA, demonstrated by overlapping absorbance peaks at 420 nm (characteristic of Lys-PLP) and 280 nm (protein), suggesting a covalent attachment (Fig. [125]4a; Supplementary Fig. [126]6f). This attachment, likely via a protonated Schiff base^[127]20,[128]24, was localized to Lys169 of YcfA by mass spectrometry (Supplementary Fig. [129]6g, h). A Lys169-to-alanine mutation abolished PLP binding and its effect on oligomerization (Fig. [130]4b). As Lys169 mediates GTP interactions, PLP binding at this residue significantly disrupts GTP recruitment (Fig. [131]4c, d). Fig. 4. PLP promotes YcfA assembly by competing with GTP. [132]Fig. 4 [133]Open in a new tab a SEC analysis of YcfA (1 mg/mL) with or without the presence of PLP (200 μM). b SEC analysis of the YcfA^K169A (1 mg/mL) mutation with or without the presence of PLP (200 μM). c, d ITC measurement determined the binding affinity of GTP to YcfA alone and in the presence of PLP, respectively. The outer panel shows the raw titration traces; the inner panel shows the fitted titration curve. e Adenylation activity assay of YcfA with different concentrations of PLP over 150 s. The absorption peak at 315 nm gradually increased with decreasing concentration, showing that PLP inhibits the enzymatic activity of YcfA in a dose-dependent manner. Source data are provided as a Source Data file. Although PLP induces oligomeric organization similar to GTP, it blocks substrate binding and thereby inhibits YcfA’s enzymatic activity (Fig. [134]4e). This inhibition, while seemingly detrimental, likely functions as a feedback mechanism within the YcfA-YcfC system. Specifically, YcfC converts the unstable 6-Cys-GTP S-adduct produced by YcfA into 6-thio-GTP. In the absence of YcfC, PLP stabilizes YcfA’s oligomeric structure but suppresses its catalytic function, thereby preventing the wastage of ATP, guanine substrates, sulfur sources, and unstable intermediates. When conditions improve, the reversible PLP-lysine linkage can be cleaved, restoring YcfA’s active state and enabling its catalytic activity alongside YcfC. Structural basis for YcfC C–S lyase To elucidate our understanding of the catalytic mechanism of YcfC, we resolved its crystal structure in complex with PLP at a resolution of 2.09 Å (Supplementary Table [135]3). The structure revealed that YcfC forms a homodimer, with each monomer comprising small and large domains that create two identical active sites. These sites accommodate both the PLP cofactor and the substrates (Fig. [136]5a; Supplementary Fig. [137]7a). PLP forms a covalent Schiff base bond with Lys148, stabilized by π stacking interactions with Tyr76 and Ser124, and hydrogen bonding with His125, and Asp122 (Fig. [138]5b; Supplementary Fig. [139]7b). Mutations at Lys148, which covalently binds PLP, and His125, which interacts with PLP’s hydroxyl group, drastically reduced YcfC activity. In contrast, mutations affecting other stabilizing interactions had minimal impact, suggesting the covalent PLP-Lys148 bond is sufficient to orient PLP for catalysis (Fig. [140]5c). Fig. 5. Structural basis for YcfC C-S lyase. [141]Fig. 5 [142]Open in a new tab a Surface representation of the YcfC homodimer, creating an active site for accommodating PLP and substrates. b Interaction network between PLP and YcfC, depicting π-π/anion-π stacking (yellow dashed lines) and hydrogen bonds (blue dashed lines). c Impact of mutations in putative PLP-binding residues on thioamide formation. The WT YcfC is regarded as possessing 100% enzyme activity. The mean is displayed ±SD for n = 3 biological replicates. d Docking model of the YcfC with 6-Cys-GTP S-adduct, showing putative binding residues. e Activity assay of YcfC^Y19A and YcfC^E179A for thioamide formation. The WT YcfC is regarded as possessing 100% enzyme activity. The mean is displayed ±SD for n = 3 biological replicates. f Electrostatic potential map of YcfC, with 6-Cys-GTP S-adduct represented as sticks bound to the active site. g Proposed catalytic mechanism for YcfC-catalyzed C–S bond cleavage. The reaction begins with His147 deprotonating the substrate’s α-amine group, facilitating external aldimine formation by releasing PLP from Lys148. Freed Lys148 then mediates essential proton transfer reactions critical for C–S bond cleavage. Following reversible proton exchange with the solvent, His147 returns to its deprotonated state and subsequently acts again as a general base, deprotonating the substrate’s thiol (–SH) group to yield a thiolate anion. The protonated His147 can then transfer a proton to the N1 position of the resulting 6-thioguanine product. Finally, Lys148 regenerates the internal Schiff base by nucleophilically attacking the C4’ atom of PLP, releasing the aminoacrylate product and resetting the enzyme for subsequent catalytic cycles. Source data are provided as a Source Data file. Although YcfC shares similar structural features with cysteine desulfurases^[143]25,[144]26, a DALI search identified C-DES^[145]27,[146]28, a C–S lyase, as its closest homolog (Supplementary Fig. [147]8a). Unlike cysteine desulfurases, YcfC lacks the conserved cysteine residue required for persulfide transfer (Supplementary Fig. [148]8b). Serine substitution of all cysteine residues in YcfC had no effect on product yield (Supplementary Fig. [149]9a), ruling out a desulfurase-like pathway. Furthermore, no direct interactions between YcfA and YcfC were detected (Supplementary Fig. [150]9b), unlike typical persulfide transfer systems. These findings confirm that YcfC functions as a C–S lyase, catalyzing the conversion of the 6-Cys-GTP S-adduct produced by YcfA into 6-thio-GTP. Molecular docking simulations provided insights into YcfC’s substrate recognition. Tyr19 and Glu179 were identified as key residues, interacting with the guanosine and ribose moieties, respectively. Tyr19 forms a π–π stacking interaction with the guanine base, while Glu179 forms hydrogen bonds with the ribose hydroxyl group (Fig. [151]5d). Substituting Tyr19 with alanine significantly impaired 6-thio-GTP formation, whereas mutating Glu179 had a minor effect (Fig. [152]5e), underscoring the critical role of Tyr19 in substrate positioning. A positively charged region comprising Arg17, Lys20, Lys77, Arg175, Lys247, and His256 at the entrance of the substrate-binding pocket, likely facilitates electrostatic interactions with phosphate groups of the substrate (Fig. [153]5f). However, mutations in these residues had minimal impact, indicating sufficient flexibility within this charged patch for substrate recruitment (Supplementary Fig. [154]9c). His147, positioned adjacent to the PLP–Schiff base, is essential for catalytic activity, as demonstrated by the complete loss of enzymatic function upon alanine substitution (Supplementary Fig. [155]9d, e). Aligning with the established mechanisms of PLP-dependent C–S bond cleavage enzymes^[156]29–[157]31, we propose a detailed catalytic mechanism for YcfC (Fig. [158]5g). The reaction cycle initiates as His147 deprotonates the α-amine group of the incoming 6-Cys-guanine nucleotide S-adduct, promoting substrate activation. This facilitates transaldimination with PLP, resulting in the displacement of the ε-amino group of Lys148 and formation of an external aldimine intermediate. Subsequently, the liberated Lys148 acts as a general base to abstract the α-proton from the substrate, triggering electron delocalization into the PLP conjugated system. This electron redistribution destabilizes the adjacent C–S bond, preparing it for cleavage–a critical step toward forming the PLP-aminoacrylate intermediate. Following reversible proton exchange with the solvent, His147 returns to its deprotonated state and subsequently acts again as a general base, deprotonating the substrate’s thiol (–SH) group to yield a thiolate anion. The protonated His147 can then transfer a proton to the N1 position of the resulting 6-thioguanine product. Finally, the ε-amino group of Lys148 executes a nucleophilic attack at the C4’ atom of the PLP–aminoacrylate intermediate, restoring the internal Schiff base, releasing the aminoacrylate product, and regenerating the active enzyme state. While we cannot definitively rule out a primary role of His147 in substrate positioning or stabilizing the catalytic site, we highlight its mechanistic resemblance to His114 in the previously reported PLP-dependent L-cystine C–S lyase from Synechocystis^[159]29, which acts as a general base and stabilizes a labile persulfide intermediate. However, definitive experimental evidence confirming an analogous catalytic role for His147 in YcfC remains to be established. Antibacterial activity and potential mechanism of 6-seleno-GTP Building on the antibacterial effect against S. aureus of 6-thioguanine^[160]10, we investigated the antimicrobial potential of enzymatically synthesized 6-thio-GTP and 6-seleno-GTP against ESKAPE pathogens^[161]32. While 6-thio-GTP exhibited negligible activity, 6-seleno-GTP demonstrated potent antibacterial effects specifically against S. aureus (Fig. [162]6a), indicating a selenoamide-dependent mechanism. To elucidate this mechanism, we performed metabolomic, transcriptomic, and proteomic analyses on the S. aureus Newman strain treated with 6-seleno-GTP. Metabolomics identified differentially expressed metabolites, primarily affecting amino acids, nucleotide, and purine metabolism (Supplementary Fig. [163]10a, d; Supplementary Data [164]1). Transcriptomic analysis revealed differentially expressed genes (Supplementary Fig. [165]10b, e; Supplementary Data [166]1), with upregulation in pathways related to purine, amino acids, and vitamin B₆ metabolism, and downregulated in glycolysis/gluconeogenesis, two-component signaling, and quorum sensing pathways (Fig. [167]6b). Proteomics further identified differentially expressed proteins, including upregulated transporters and transcriptional regulators, alongside downregulated permeases and antiporters (Supplementary Fig. [168]10c, f; Supplementary Data [169]1). Fig. 6. Antibacterial activity and potential mechanism of 6-seleno-GTP. [170]Fig. 6 [171]Open in a new tab a MICs of 6-Seleno-GTP and 6-Thio-GTP against ESKAPE pathogens are shown. Blue arrows indicate MICs exceeding the highest tested concentration. b KEGG pathway enrichment analysis of differentially expressed genes is presented. Red represents upregulation; blue represents downregulation. c, d Venn diagrams show the overlap between differentially expressed genes (transcriptomics) and proteins (proteomics). c Nine upregulated genes were identified in both datasets. d One downregulated gene was identified in both datasets. Source data are provided as a Source Data file. Integrative analysis highlighted significant upregulation of purine biosynthesis genes (purQ, purF, purN, purH, purD) and arginine metabolism genes (arcA, arcB) (Fig. [172]6c), while the cold shock protein cspC was markedly downregulated (Fig. [173]6d). These findings suggest that 6-seleno-GTP disrupts purine and arginine metabolism while impairing stress response pathways in S. aureus. Given the efficacy of this disruption in suppressing the S. aureus Newman strain, these pathways represent promising targets for developing antimicrobial therapies against drug-resistant S. aureus. Discussion Our findings, alongside recent studies, redefine the paradigm of thioamide biosynthesis by revealing a mechanism that diverges from the classical persulfide transfer model typically associated with cysteine desulfurases^[174]33. In E. amylovora, the YcfA–YcfC system employs a distinctive catalytic strategy: YcfA first adenylates guanine nucleotides and subsequently facilitates the direct addition of L-Cys to produce a 6-Cys-guanine nucleotide S-adduct (Fig. [175]2k). YcfC then acts as a C–S lyase, cleaving this intermediate to generate the thioamide product (Fig. [176]5g). Intriguingly, this system also synthesizes selenoamides, offering an alternative route for producing rare selenium-containing compounds^[177]34 with potential application in antimicrobial therapy (Fig. [178]1b; Supplementary Fig. [179]1d). Unlike canonical AANH-like enzymes that utilize reactive persulfides derived from L-Cys via desulfurases as sulfur donors^[180]16, YcfA directly employs L-Cys as the nucleophile to attack the adenylated guanine nucleotide substrates (Fig. [181]1c–e). Since L-Cys is inherently more stable but less reactive than persulfides, YcfA must overcome a significantly higher activation energy barrier. This challenge is mitigated by Gln129 in YcfA, which facilitates a proton bridge between L-Cys and the adenylated guanine, enabling L-Cys addition and subsequent AMP release (Fig. [182]2i, k). This dual function distinguishes YcfA from other AANH-like enzymes. Phylogenetic analyses (Supplementary Fig. [183]11) suggest that YcfC and canonical cysteine desulfurases likely diverged from ancestral protein but evolved under distinct selective pressures. Structurally, YcfC shares similarities with cysteine desulfurases^[184]25,[185]26 and C–S/C–Se lyases, such as CsdA, CsdB, and IscS (Supplementary Fig. [186]8a). However, unlike desulfurases, YcfC lacks the conserved cysteine residue critical for enzyme–persulfide intermediate formation (Supplementary Fig. [187]8b). Instead, YcfC operates as a C–S lyase, directly cleaving the C–S bond in the 6-Cys-guanine S-adduct (Fig. [188]5g), and a reaction that occurs with a lower activation energy compared to the desulfurization of L-Cys by cysteine desulfurases. This suggests that the YcfA–YcfC system has evolved to bypass the need for reactive sulfur carriers, reflecting an adaptation to improve catalytic efficiency. Our study also elucidates the dynamic oligomerization of YcfA, which assembles into a two-layered heptameric structure with full enzymatic activity (Fig. [189]3; Supplementary Fig. [190]5). Modifications to surface residues near the interface significantly influence this assembly. For instance, the L173D mutation stabilizes a dimeric form, while the D166R/Y170N double mutation enhances interfacial interactions, leading to a tetradecameric assembly with improved catalytic efficiency (Fig. [191]3c, d). Two key residues, Lys169 and Lys180, play pivotal roles in regulating oligomerization (Supplementary Fig. [192]6d). We identified four conditions that stabilize the tetradecameric structure: nucleotide substrates binding, PLP attachment to Lys169, mutation at Lys180, and the D166R/Y170N double mutation (Figs. [193]3c, e, f and [194]4a, b; Supplementary Fig. [195]5e). All these factors converge at the catalytic interface, facilitating the formation of a complete catalytic cavity. The highly ordered tetradecameric architecture of YcfA is critical for constructing functional catalytic cavities. This organization not only increases the local concentration of catalytic units and substrates but also safeguards reactive intermediates, particularly adenylated guanine nucleotides. The dynamic nature of this assembly acts as a regulatory mechanism, fine-tuning 6-TG production. Nucleotide substrates and the cofactor PLP exert opposing influences on the active assembly of YcfA: while guanine substrates promote active oligomerization, PLP inhibits the enzymatic activity despite promoting assembly (Figs. [196]3e, f and [197]4c, d). This competitive interplay ensures that the catalytic process remains efficient and synchronized with substrate availability, preventing the unnecessary expenditure of ATP and protecting transient intermediates. We hypothesize that this regulatory mechanism between YcfA and YcfC may allow E. amylovora to dynamically balance 6-TG production under varying metabolic conditions, potentially linking thioamide biosynthesis to bacterial virulence modulation during infection. In the early stages, when bacterial populations are low, the inhibitory effect of PLP could limit 6-TG production, delaying the expression of virulence factors and preventing premature activation of host defenses. As the infection progresses and guanine substrates accumulate, 6-TG production could be amplified, enabling the bacterial population to overcome host defenses. While these findings establish a foundation for understanding PLP’s role in this pathway, additional in vivo studies are required to validate its physiological relevance and broader implications for metabolic adaptation in pathogenic bacteria. Our discovery of 6-Se-GTP’s potent anti-methicillin-resistant S. aureus (MRSA) activity highlights its potential as a precision antimicrobial agent. Multi-omics analyses uncovered significant disruptions in key metabolic pathways of S. aureus, including purine and arginine metabolism, alongside impaired stress response pathways. While the exact selectivity of 6-Se-GTP remains to be fully characterized, its ability to target bacterial purine metabolism suggests that it could serve as a promising therapeutic candidate for combating antibiotic-resistant pathogens. These findings align with ongoing efforts to develop precision-based antimicrobial agents that target specific metabolic vulnerabilities to pathogens, offering a potential pathway for the treatment of resistant bacterial infections. Beyond structural biology, our work bridges fundamental enzymology with translational applications. The ability of the YcfA-YcfC system to bypass persulfide intermediates offers a biotechnological advantage for scalable thioamide production. Variants such as the D166R/Y170N mutant, which demonstrate enhanced catalytic efficiency, hold potential for the industrial-scale synthesis of 6-TG, reducing reliance on conventional chemical approaches that generate toxic byproducts. Furthermore, the modularity of this system, which is capable of incorporating both sulfur and selenium, opens opportunities for generating diverse thioamide and selenoamide libraries for drug discovery. In conclusion, our study provides a comprehensive view of thioamide biosynthesis via the YcfA–YcfC system, emphasizing its catalytic efficiency and regulatory mechanisms. We unveil a strategy involving YcfA oligomerization, wherein nucleotide substrates and PLP exert antagonistic effects to control catalytic activity (Fig. [198]7). These mechanisms may be exploited by E. amylovora to regulate 6-TG production and modulate virulence during host infection. Exploring this efficient thioamide biosynthesis pathway and its regulation could inform the development of antimicrobial agents, like rare selenium-containing compounds, and strategies for enhancing biosynthetic efficiency. Fig. 7. A proposed working model of the YcfA-YcfC system. [199]Fig. 7 [200]Open in a new tab This illustration depicts the dynamic interplay within the YcfA-YcfC system, emphasizing its dual catalytic roles and regulation by GTP and PLP. YcfA alternates between two conformational states: an active GTP-bound form and an inactive PLP-bound form. Binding of PLP induces oligomerization while simultaneously blocking the GTP binding site, acting as a regulatory switch. In its active state, YcfA facilitates a two-step catalytic process involving guanine nucleotides. First, ATP is used to adenylate the guanine nucleotide, releasing pyrophosphate (PPi). This modification enables the subsequent recruitment of L-Cys or L-SeCys, forming a 6-Cys/SeCys-guanine nucleotide S-adduct. YcfC, functioning as a C–S/C-Se lyase, cleaves the S-adduct to yield the final thioamide product. While in vitro data establishes GTP/PLP-mediated regulation of this system, the physiological relevance of PLP-induced oligomerization and metabolite homeostasis maintenance remains to be verified in vivo. Dashed arrows indicate hypothesized connections requiring experimental validation in cellular contexts. Methods Gene cloning and protein expression The ycfA and ycfC genes from E. amylovora ATCC 49946 were codon-optimized and synthesized by GENEWIZ (Hangzhou, China). YcfA, YcfC, and their mutants were overexpressed using the pET-21b vector. The interaction between YcfA and YcfC was investigated by co-expressing His-tagged YcfA and untagged YcfC in the pET-Duet-1 vector, which enabled subsequent affinity purification. Site-directed mutagenesis was introduced via quick-change PCR, and all the constructs were verified by Sanger sequencing (the plasmids and primers used are listed in the Supplementary Tables [201]4 and [202]5, respectively). The recombinant plasmids were transformed into E. coli BL21 (DE3) cells, which were subsequently cultured at 37 °C and 220 rpm to an OD[600] of 0.6. Protein expression was induced with 200 μM isopropyl-β-D-thiogalactopyranoside (IPTG) at 16 °C for 16 h. Purification Harvested E. coli BL21(DE3) cell pellets were resuspended in buffer A (200 mM NaCl, 25 mM HEPES pH 8.0, and 10% glycerol) and lysed using a French press (AH-1500, ATS, China) at 4 °C. Following centrifugation (285,000 × g, 30 min), the supernatant was purified using nickel-affinity column chromatography (Ni-NTA; GE Healthcare, Little Chalfont, UK) with an imidazole gradient (10 and 20 mM), followed by elution with buffer B (200 mM NaCl, 25 mM HEPES pH 8.0, 300 mM imidazole, 5% glycerol). Further purification involved anion exchange chromatography (Source Q; GE Healthcare, Sweden; Q buffer A: 25 mM HEPES pH 8.0, 5% glycerol; Q buffer B: 1 M NaCl, 25 mM HEPES pH 8.0, 5% glycerol) and size-exclusion chromatography (Superdex 200 Increase 10/30 GL; GE Healthcare, Sweden; gel filtration buffer: 15 mM HEPES pH 8.0, 100 mM NaCl, 5% glycerol). Protein purity was assessed via sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE). Crystallization Crystallization trials were conducted using purified YcfA (10 mg/mL) and YcfC (10 mg/mL). To obtain substrate-bound structures, YcfA and YcfA^D19A were pre-incubated with 5 mM substrates for 2 h at 4 °C. Crystals were grown by the hanging-drop vapor diffusion method at 16 °C. Apo-YcfA crystallized in 3.5% Tacsimate and 25% PEG3350. The YcfA complexes (YcfA-ATP, YcfA-GTP, and YcfA^D19A-ATP-GTP) crystallized in 0.2 M sodium malonate pH 7.0 and 20% PEG3350. YcfC crystallized in 50 mM sodium formate. Crystals were cryoprotected with 20% glycerol and flash-frozen in liquid nitrogen for data collection. X-ray diffraction and structure determination X-ray diffraction data were collected at beamlines 18U1 and 19U1 of the Shanghai Synchrotron Radiation Facility at a wavelength of 0.9791 Å and processed using DIALS^[203]35 or XDS^[204]36. Structures were determined by molecular replacement using models generated by AlphaFold2^[205]37 and refined using COOT^[206]38 and Phenix^[207]39. The representative electron density maps for each crystal structure are illustrated in Supplementary Fig. [208]12 and omit maps for all ligand-bound crystal structures are illustrated in Supplementary Fig. [209]13. Cryo-EM sample preparation For cryo-EM analysis, the YcfA protein sample (0.1 mg/mL) was applied to glow-discharged Quantifoil Au grids (40 s, 200-mesh, R2/1). The grids were blotted for 2 s at 100% humidity and plunge-frozen in liquid ethane using a Vitrobot Mark IV (Thermo Fisher). Cryo-EM single particle data acquisition and data processing Data were acquired (Titan Krios microscope, 300 kV, Thermo Fisher) using EPU software (K2 Bioquantum director, 20 eV energy filter, 5 frames/s, total exposure for 6 s, 49 e−/Å^2) to yield movie stacks (−1.2 to −1.7 μm defocus). Motion correction (Motioncor2^[210]40), CTF correction (CTFFIND4^[211]41), and particle extraction (EMAN2.31^[212]42) resulted in 85,057 particles. After 2D classification (Relion3.1^[213]43), the best classes were selected by visual examination. Then, a total of 67,556 particles were subjected to cryoSPARC4.4.1 to build the initial model and then subjected to heterogeneous refinement to yield a major class including 16,708 particles. The major class including 16,708 particles were subjected to cryoSPARC4.4.1 homogeneous refinement. The particles were subjected to symmetry alignment and homogeneous refinement with C7 symmetry string to yield the final map. A sharpening B-factor of −102.6 Å^2 was applied to the resulting cryo-EM map to yield the final sharpened map at 3.44 Å resolution estimated by the 0.143 criterion of FSC curve. Data collection and processing statistics were included in Supplementary Table [214]2. Representative model building with cryo-EM densities of the YcfA are illustrated in Supplementary Fig. [215]14. UV-Vis spectrophotometric activity assays YcfA and YcfC activity assays were performed using UV-Vis spectrophotometry. For endpoint assays, purified YcfA, YcfC, or their mutants (10 μM or 5 μM, respectively) were incubated in a reaction buffer (10 mM HEPES pH 8.0, 100 mM NaCl, and 5 mM MgCl[2]) with 100 μM substrates (ATP, GTP/GDP/GMP, and L-Cys/L-SeCys/other amino acids) at 30 °C for 30 min. The reactions were quenched with methanol (80 μL for each 20 μL of reaction mixture) and centrifuged, and the supernatant was analyzed by a Nanodrop One spectrophotometer (Thermo Scientific, MA). For kinetic assays, reactions (1 mL) containing 20 μM YcfA, 500 μM ATP, 500 μM GTP, and 100 μM L-Cys in reaction buffer (100 mM NaCl, 15 mM HEPES pH 8.0, 5% glycerol, and 5 mM MgCl[2]) were monitored at 5-second intervals in a quartz cuvette using an Ultrospec 2100 pro UV-Vis spectrometer (Biochrom, UK). After the first 40 readings, 30 nM YcfC was added, and the reaction was monitored for an additional 60 readings. In separate experiments, 42 nM YcfC, 500 μM ATP, 500 μM GTP, and 100 μM L-Cys were mixed, recorded at 5-s intervals for 40 reads, followed by the addition of 20 μM YcfA for another 60 readings. To analyze the effect of PLP on the enzyme activity of YcfA, we incubated YcfA with various concentrations of PLP (0 μM, 40 μM, 80 μM and 160 μM) for 10 min, followed by kinetic analysis. The PLP concentration is consistent with reported in vivo PLP levels^[216]44. The data were analyzed using GraphPad Prism 8.3. High-performance liquid chromatography (HPLC) analysis of the reaction products The reaction products of the YcfA-YcfC system were analyzed by HPLC following methanol treatment and centrifugation to remove precipitates. A Shimadzu LC-2030 system equipped with a Prodigy ODS-3 analytical column (4.6 × 250 mm, 5 μm) was used for analysis. The mobile phase consisted of solvent A (30 mM KH[2]PO[4], 10 mM tetrabutylammonium hydrogen sulfate in water, pH 7.5) and solvent B (methanol). Gradient elution was performed as follows: 0.5% solvent B (0-10 min), a linear gradient to 70% solvent B (10–23 min), and a final increase to 99% (23–28 min). Detection was performed at 340/360/315 nm, with a flow rate of 1 mL/min at room temperature. Product identification by mass spectrometry The reaction products, prepared as described above, were dissolved in a 1:1 (v/v) methanol: water mixture. An Ultimate 3000 system coupled with a Q Exactive Plus Q-Orbitrap high-resolution mass spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) was used for analysis. After injection of 10 μL of sample, isocratic elution was performed using solvent A (water containing 10 mM NH[4]Ac) and solvent B (acetonitrile: water = 90:10 (v/v) with 10 mM NH[4]Ac). A heated electrospray ionization (HESI) source was used with the following parameters: sheath gas, 35 arb; auxiliary gas, 10 arb; spray voltage, 3.2 kV (positive/negative mode); capillary temperature 320 °C; and auxiliary gas heater temperature, 350 °C. Data were collected within a mass range of 50–750 m/z with a mass resolution of 70,000. Identification of the 6-Cys-GTP S-adduct intermediate The reaction mixture (50 μL), containing reaction buffer (100 mM NaCl, 15 mM HEPES, pH 8.0, and 5 mM MgCl[2]), 100 μM YcfA, 1 mM ATP, 1 mM GTP, and 1 mM L-Cys, was incubated at room temperature for 30 min. Then, the mixture was quenched with methanol (100 μL) and centrifuged (160,000 × g, 20 min). For MS analysis, 5 µL samples were separated on a BEH amide column (2.1 × 100 mm, 1.7 µm, Waters, USA) at 40 °C at a flow rate of 0.3 mL/min). The mobile phase comprised solvent A (H[2]O) and solvent B (90:10 acetonitrile: H[2]O), both of which contained 10 mM ammonium formate and 0.15% formic acid. The gradient elution profile was as follows: 100% B (0–2 min), 100% B to 85% B (2–9 min), 85% B to 50% B (9–14 min), isocratic elution with 50% B (14–19 min), and 100% B (19–27 min). A heated electrospray ionization (HESI) source was used with the following parameters: sheath gas, 35 arb; auxiliary gas, 10 arb; spray voltage, 3.2 kV (positive/negative mode); capillary temperature 320 °C; and auxiliary gas heater temperature, 350 °C. Data were collected within a mass range of 50–750 m/z with a mass resolution of 70,000. The dehydride ion at m/z 624.9925 was fragmented using DDA mode with specific MS/MS parameters. Limited proteolysis assay of YcfA The YcfA used in this assay was adjusted to a molar concentration of about 20 µM. YcfA with or without ATP and GTP was subjected to trypsin digestion. After 30 min of incubation on ice, the resulting proteolyzed products were analyzed via SDS-PAGE. MS analysis of YcfA-PLP covalent binding Covalent binding of PLP to YcfA was analyzed using a previously described method^[217]24. Briefly, 100 µg of YcfA was incubated with 1 mM PLP for 30 min, followed by precipitation with 1 mL of acetone, and storage overnight at −80 °C. After centrifugation (13,000 × g, 15 min, 7 °C), the protein pellet was solubilized in 100 mM HEPES (pH 7.8), reduced with 10 μL of 1 M NaBH[4] (37 °C, 1 h), and then precipitated again. The resulting sample was dissolved in S-Trap lysis buffer, reduced with 4.5 mM DTT, alkylated with 10 mM iodoacetamide (quenched with DTT), and purified using an S-Trap column. Trypsin digestion was performed overnight at 37 °C. The resulting peptides were separated using a Thermo Scientific EASY nLC-1200 system and analyzed by a Q Exactive Hybrid mass spectrometer. The data was processed using Thermo Xcalibur and search parameters included carbamidomethyl (+57.021 Da) as a static modification and PLP-binding (+231.0297 Da) as a dynamic modification; all other parameters were set to the default values. Molecular docking simulations Molecular dynamics simulations were performed to obtain the YcfA-adenylated-GTP complex. YcfA with a 2.34 Å crystal structure resolution was used as a template. The protonation states of the residues were determined using H^++ server, in accordance with the hydrogen bonding network, and the force fields for amino acid residues and solvent water molecules were assigned using GROMACS 2022.5^[218]45. The system was placed in a cubic box with a minimum distance of 10 Å between the protein surface and the box edge, then solvated with TIP3P water molecules and neutralized with sodium and chloride ions. Energy minimization was conducted using a 5000-step steepest descent method to remove unreasonable contacts, followed by 5000 steps of gradient conjugation. Subsequently, the system was heated from 0 K to 310 K over 100 ps using the NVT ensemble. This was followed by 100 ps of NPT ensemble equilibration at 1 bar pressure, 310 K. Hydrogens bonds were constrained using the LINCS algorithm and the time step for all simulations was 2 fs. A 100 ns production MD simulation was then performed under NPT ensemble conditions without any restraints. Trajectory analysis was performed with the VMD 1.9.3 software. The average RMSD of YcfA, adenylated-GTP, and YcfA-adenylated-GTP complex is provided in Supplementary Fig. [219]15. Molecular docking simulations were performed using AutoDock Vina 4.2.0^[220]46. Ligands were prepared using IQmol Molecular Viewer. For the YcfA-adenylated-GTP docking with L-Cys, the binding site was defined with the following coordinates: center_x: 46.396, center_y: 38.633, and center_z: 6.731. For YcfC docking with the 6-Cys-GTP S-adduct, the binding site coordinates were as follows: center_x: 12.399, center_y: 47.636, and center_z: -5.996. All other parameters were set the default values. PyMOL 2.5^[221]47 was used for visualization and image processing. All the coordinate files of the final docked models and MD data are provided in Supplementary Data [222]2. Size-exclusion chromatography (SEC) analysis Size-exclusion chromatography was performed using a Superdex 200 Increase 10/30 GL column (GE Healthcare) equilibrated with gel filtration buffer containing 200 μM of the appropriate substrates. Purified YcfA and its variants were preincubated with the corresponding substrates (200 μM) before injection. As for the interaction between PLP and YcfA, YcfA (1 mg/mL) and PLP (200 μM) were used for SEC analysis. Isothermal titration calorimetry (ITC) ITC experiments were conducted using a Nano ITC (TA instruments, USA) at 16 °C. To analyze the binding of GTP to YcfA (or PLP-bound YcfA), the concentration of YcfA was used at 260 μM, and the GTP solution was diluted to 4 mM. GTP (2.5 µL) was injected into 300 μL of YcfA (or PLP-bound YcfA) twenty times every 120 seconds with a stirring speed of 200 rpm. The data were analyzed using the one-site binding mode in NanoAnalyze Data Analysis software (version 3.8.0). Analytical ultracentrifugation (AUC) The SV-AUC experiments were conducted using a 12 mm charcoal-filled Epon centerpiece and an eight-hole An50 Ti rotor at 56,500 × g (Optima AUC analytical ultracentrifuge, Beckman Coulter). The absorbance was monitored at 280 nm. The data were analyzed using SEDFlIT software to determine the sedimentation coefficient distribution C(S). Phylogenetic analysis of YcfC All protein sequences used for analysis were obtained from the NCBI server ([223]https://www.ncbi.nlm.nih.gov/). The protein data used for phylogenetic analysis is listed in Supplementary Table [224]6. The amino acid sequences of these proteins were aligned using the MAFFT ver. 7 server ([225]https://mafft.cbrc.jp/alignment/server/) with the default settings. The unrooted fast maximum-likelihood-based tree was generated using IQ-TREE^[226]47 with automatic model selection mode (ModelFinder)^[227]48, where LG + R4 was selected. Phylogenetic bootstrap analysis was performed by ultrafast approximate bootstrap with 1000 bootstrap replicates. The tree was displayed using MEGA7^[228]49 software and then beautified with iTOL. Minimum inhibitory concentration (MIC) determination MIC values were determined via a standard broth microdilution method. The bacterial cultures were adjusted to ~5 × 10^5 CFU/mL, and 2-fold serial dilutions of each antibiotic were dispensed into 96-well plates. After a 24 h of incubation at 37 °C, the MIC was defined as the lowest concentration of antibiotic with no visible bacterial growth. The experiments were performed in triplicate. Omics sample preparation S. aureus Newman was grown overnight then diluted to an OD[600] of 0.1 in TSB medium. For the drug treatment groups, 6-seleno-GTP was added at a final concentration of 4 μg/mL, and the cultures grew to the mid-exponential phase. Ten duplicate samples were prepared for metabolomics analysis, whereas three duplicate samples were prepared for transcriptomics and proteomics analyses. Metabolite extraction Treated bacterial suspensions were centrifuged and washed twice with PBS. Then, 100 μL of prechilled methanol: water (8:2) mixture and 10 μL of internal standard were added. The mixture was subjected to three freeze-thaw cycles (liquid nitrogen and ice), followed by 15 min of sonication in an ice bath. An additional 400 μL of prechilled methanol: water (8:2) was added, and the mixture was vortexed for 3 min at 4 °C and 2500 g. After quenching at −80 °C for 20 min, the sample was sonicated for 5 min at 4 °C and centrifuged at 4 °C and 189,000 × g for 20 min. The resulting supernatant (200 μL) was collected, vacuum-dried, and prepared for GC-MS analysis. Metabolomics analysis The samples were analyzed via GC–TOF–MS using an Agilent 8890/LECO Pegasus BT instrument. One microliter of each sample was separated on an Rxi-5ms column (30 m × 0.25 mm × 0.25 μm) with helium as the carrier gas (1 mL/min). The inlet was operated in splitless mode at 270 °C. The GC oven temperature program was as follows: 50 °C for 0.5 min, followed by an increase of 5 °C/min to 300 °C, where it was held for 6 min. The transfer line and ion source temperatures were 280 °C and 250 °C, respectively. MS analysis was conducted in the m/z range of 60–700 with an ionization energy of 70 eV. The data were compared to an in-house data library and analyzed using tidyverse^[229]50. Metabolites with |log[2]FoldChange|> 0.5 and P-value < 0.05 were considered significantly differentially expressed. Transcriptomics analysis RNA extraction and transcriptome sequencing were performed by Shanghai Personal Biotechnology Co., Ltd. (Shanghai, China). Total RNA was isolated using TRIzol Reagent, and rRNA was removed using the Zymo-Seq RiboFree Kit. cDNA was synthesized using SuperScript III and DNA Polymerase I. cDNA fragments were purified using the AMPure XP system (Beckman Coulter, Beverly, CA, USA), and sequencing libraries were prepared using the Illumina PCR Primer Cocktail and sequenced on the NovaSeq 6000 platform (Illumina). The transcriptomic datas were aligned to the S. aureus Newman genome using Bowtie2. Differentially expressed mRNAs were identified using DESeq2^[230]51, with significance defined as |log[2]FoldChange|>0.5 and P-value < 0.05. GO enrichment analysis was performed using topGO^[231]52, and KEGG pathway enrichment analysis was performed using clusterProfiler 4.0^[232]53, with significance defined as a P-value < 0.05. Proteomics analysis Protein extraction and proteomics analysis were performed by Shanghai Personal Biotech Co., Ltd. (Shanghai, China). Samples were extracted via homogenization and SDT lysis, and the protein was quantified via the BCA method, and separated by SDS-PAGE. Use FASP to digest the protein. After drying the digested peptides, dissolve them with 0.1% FA, and determine the peptide concentration for LC-MS analysis. The Vanquish Neo UHPLC (Thermo Scientific) system and an Orbitrap Astral high-resolution mass spectrometer were utilized for the purpose of proteomic analysis. The sample is injected into the Trap Column (PepMap Neo 5 µm C18 300 µm × 5 mm, Thermo Scientific) and then subjected to gradient separation using the analytical column (μPAC Neo High Throughput column, Thermo Scientific). The mass spectrometry was performed in Data Independent Acquisition (DIA) mode with parameters including a scan range of m/z 380–980, a resolution of 240,000 for the first stage, a resolution of 80,000 for the second stage, an isolation window of 2 Th, and an HCD collision energy of 25%. DIA raw files were analyzed via DIA-NN^[233]54. NCBI S. aureus Newman (access number 426430) as reference database. The analysis yielded 2246 protein groups and 32,906 PSMs. Proteins with |log[2]FoldChange|>0.5 and P-value < 0.05 were considered significantly differentially expressed. All the gene models were annotated and classified using our draft genome database draft for COG and KEGG annotation. Reporting summary Further information on research design is available in the [234]Nature Portfolio Reporting Summary linked to this article. Supplementary information [235]Supplementary Information^ (21.4MB, pdf) [236]41467_2025_63937_MOESM2_ESM.pdf^ (88.8KB, pdf) Description of Additional Supplementary Files [237]Supplementary Data 1^ (47.7KB, xlsx) [238]Supplementary Data 2^ (1.5MB, zip) [239]Reporting Summary^ (75.1KB, pdf) [240]Transparent Peer Review file^ (677.3KB, pdf) Source data [241]Source data^ (29.8MB, xlsx) Acknowledgements