Abstract
Helicobacter pylori is a Gram-negative bacterium that is responsible for gastric and duodenal ulcers. H. pylori uses the unusual mqn pathway with aminofutalosine (AFL) as an intermediate for menaquinone biosynthesis. Previous reports indicate that hydrolysis of AFL by 5′-methylthioadenosine nucleosidase (HpMTAN) is the direct path for producing downstream metabolites in the mqn pathway. However, genomic analysis indicates jhp0252 is a candidate for encoding AFL deaminase (AFLDA), an activity for deaminating aminofutolasine. The product, futalosine, is not a known substrate for bacterial MTANs. Recombinant jhp0252 was expressed and characterized as an AFL deaminase (HpAFLDA). Its catalytic specificity includes AFL, 5′-methylthioadenosine, 5′-deoxyadenosine, adenosine, and S-adenosylhomocysteine. The kcat/Km value for AFL is 6.8 × 104 M−1 s−1, 26-fold greater than that for adenosine. 5′-Methylthiocoformycin (MTCF) is a slow-onset inhibitor for HpAFLDA and demonstrated inhibitory effects on H. pylori growth. Supplementation with futalosine partially restored H. pylori growth under MTCF treatment, suggesting AFL deamination is significant for cell growth. The crystal structures of apo-HpAFLDA and with MTCF at the catalytic sites show a catalytic site Zn2+ or Fe2+ as the water-activating group. With bound MTCF, the metal ion is 2.0 Å from the sp3 hydroxyl group of the transition state analogue. Metabolomics analysis revealed that HpAFLDA has intracellular activity and is inhibited by MTCF. The mqn pathway in H. pylori bifurcates at aminofutalosine with HpMTAN producing adenine and depurinated futalosine and HpAFLDA producing futalosine. Inhibition of cellular HpMTAN or HpAFLDA decreased the cellular content of menaquinone-6, supporting roles for both enzymes in the pathway.
Graphical Abstract

Individuals infected with Helicobacter pylori have an increased risk of developing gastric malignancies, which are the third most deadly cancers according to GLOBOCAN 2018 data.1 Menaquinone (MK or vitamin K2) is a membranous quinone and a bacterial redox cofactor that transfers electrons in the electron transport chain of prokaryotes.2 It is an essential component of anaerobic respiration in Gram-negative bacteria.2–5 The classical MK biosynthetic pathway in Escherichia coli consists of nine men genes and is well conserved in prokaryotes.6 More recently, an alternative pathway, the futalosine pathway, was discovered and is found in Streptomyces coelicolor, Chlamydia trachomatis, Thermus thermophilus, Acidothermus cellulolyticus, Wolinella succinogenes, Helicobacter pylori, Campylobacter jejuni, and others [mqn genes (Figure 1)].5–9 Analyses of the taxonomic distribution of mqn genes suggest that the futalosine pathway may have appeared earlier than the classical men pathway in evolution.6
Figure 1.

Futalosine pathway route that diverges at aminofutalosine to form futalosine (red, deamination) and nucleoside hydrolysis by MTAN (colored blue to form dehypoxanthinefutalosine). Other organisms are reported to express an MqnB, but this activity was not observed in H. pylori.
H. pylori is a Gram-negative pathogenic bacterium that colonizes the human gastric epithelium.10 More than half of the global population is infected by H. pylori, which can lead to duodenal and gastric ulcers.11,12 H. pylori-induced gastritis is responsible for >60% of the world’s gastric carcinogenesis patients, making it the only bacterium identified as a class I carcinogen.13–15 Frequent antibiotic use is causing drug resistance in H. pylori to compromise the effectiveness of conventional antibiotic-based therapy.16 Clarithromycin-resistant H. pylori strains are listed in the high-priority group requiring new antibiotics by the World Health Organization (WHO).17 H. pylori solely relies on the futalosine pathway for MK production, as most of the men genes are absent.5,7,18 The futalosine pathway is rarely utilized in human gut microbiota. Thus, the futalosine pathway provides potential for developing H. pylori-targeted therapeutic agents that can treat ulcers and reduce gastric cancer incidence with minimal impact on healthy gut flora.
One of the unique steps in the mqn pathway is the formation of dehypoxanthine futalosine (DHFL). Previous studies have reported that aminofutalosine (AFL) is converted into futalosine (FL) by AFL deaminase followed by MqnB-catalyzed hydrolysis to form DHFL in T. thermophilus and S. coelicolor (Figure 1).5,7 In contrast, the presence of an active 5′-methylthioadenosine nucleosidase (MTAN) in H. pylori directly hydrolyzes AFL to form DHFL, while HpMTAN cannot utilize FL as a substrate.7,19 Moreover, it was originally reported that H. pylori does not possess a typical ortholog of Acel_2064, the gene encoding AFL deaminase in A. cellulolyticus.7
MTAN is a multifunctional enzyme that can hydrolyze the N-ribosidic bond of 5′-methylthioadenosine (MTA), S-adenosylhomocysteine (SAH), 5′-deoxyadenosine (5′dAdo), and AFL.7,19,20 The broad substrate specificity of MTAN allows it to participate in several biological processes, including polyamine biosynthesis (MTA degradation), S-adenosyl-L-methionine (SAM) metabolism, the methionine salvage pathway, bacterial quorum sensing, and MK biosynthesis. Organisms lacking MTANs utilize a 5′-methylthioadenosine → 5′-methylthioinosine → hypoxanthine + ribose pathway for MTA degradation.21 Previously, we reported an MTA deaminase in Pseudomonas aeruginosa that catalyzes the deamination of both adenosine and MTA.21 Interestingly, a previous study reported an adenosine deaminase capable of catalyzing SAH deamination in H. pylori strain 26695,22 leading to the question of whether deamination of MTA/AFL also takes place in H. pylori.
Here, we report the protein encoded by jhp0252 to be an AFL/MTA deaminase of H. pylori strain J99. In vitro kinetic analysis characterized substrate specificity, and inhibitor screening indicates 5′-methylthiocoformycin (MTCF) as a slow-onset inhibitor for AFL deaminase. MTCF was shown to have inhibitory effects on pylori growth. The crystal structure of MTCF-bound jhp0252 revealed structural interactions at the active site. Additionally, metabolite analyses in intact H. pylori cells established the intracellular activity and inhibition of jhp0252 activity by MTCF.
MATERIALS AND METHODS
Chemicals.
Coformycin (CF), 5′-methylthiocoformycin (MTCF), 5′-methylthio-2′-deoxycoformycin (MTDCF), 5′-propylthio-2′-deoxycoformycin (PrTDCF), and 5′-phenylthio-2′-deoxycoformycin (PhTDCF) were synthesized by methods reported previously.23 Inhibitors were dissolved in distilled water, and the concentrations determined from the extinction coefficient (8.25 mM−1 cm−1 at 282 nm). 6-Amino-6-deoxyfutalosine (AFL) and futalosine (FL) were synthesized by methods reported previously.6,19,23,24 All other chemicals and reagents were purchased from Sigma or Fisher Scientific and were of reagent grade.
Plasmid Construction and Enzyme Purification.
The jhp0252 gene from the H. pylori strain J99 genome database25 was predicted to encode an AFL deaminase (HpAFLDA) and was used for plasmid design. The E. coli codon-optimized jhp0252 gene was cloned into the N-terminal TEV cleavable N-terminal His6 tag containing vector pJ express414 from ATUM DNA2.0 Gene Design. E. coli BL21 (DE3) was transformed with plasmid pJ411-HpAFLDA for recombinant protein expression. LB medium (50 mL) with 100 μg/mL ampicillin was inoculated with one single colony, and the culture was incubated at 37 °C overnight before being transferred into 1 L of LB medium. When the OD600 reached 0.6–0.7, 1 mM IPTG was added to induce protein expression at 23 °C for 10 h. The cells were harvested by centrifugation at 4500g for 30 min. The cell pellet was resuspended in 20 mL of lysis buffer containing 1× BugBuster (Millipore Sigma), 100 μg/mL lysozyme (Sigma), 25 μg/mL DNaseI (Sigma), and 1 tablet of EDTA-free protease inhibitor (Roche). After being stirred for 20 min, the lysed cells were centrifuged at 16000g for 30 min at 4 °C to remove cell debris. The collected supernatant was incubated with Ni-NTA agarose [1.7 mL of slurry/g of cell pellet (Qiagen)] for 1 h with rocking. The mixture was then transferred into a column, and the resin was washed with 200 mL of wash buffer [50 mM HEPES (pH 7.4), 50 mM KCl, and 1 mM DTT]. The target protein was eluted with 50 mL of elution buffer [50 mM HEPES (pH 7.4), 50 mM KCl, 1 mM DTT, and 50 mM imidazole]. Sodium dodecyl sulfate–polyacrylamide gel electrophoresis analysis was used to determine the presence and purity of the target protein. Size-exclusion chromatography (Hi-Load Superdex 200 column) was used as the final purification step. The eluted protein was stored in 50 mM HEPES (pH 8.0), 50 mM KCl, and 0.5 mM DTT.
Enzymatic Assays for HpAFLDA.
The deaminase activity on AFL, MTA, 5′dAdo, adenosine, and SAH was measured by the absorbance change (Δε = 8012 M−1 cm−1) at 263 nm.26 Reactions were performed in 1 cm cuvettes at 25 °C. Assay mixtures (1 mL) contained 50 mM HEPES (pH 7.0), variable concentrations of the substrate, and appropriate amounts of purified HpAFLDA. Reactions were started by addition of HpAFLDA, and the initial rates were monitored on a CARY 300 ultraviolet–visible spectrophotometer for steady state kinetic calculation. The kinetic constants were obtained by fitting initial rates to the Michaelis–Menten equation in Graphpad Prism 8.
Inhibition Assays.
Inhibition assay reactions were initiated with 77 nM HpAFLDA in an assay mixture containing 400 μM MTA and variable concentrations of the inhibitor in 50 mM HEPES (pH 7.0) at 25 °C. Controls without enzymes were included in all of the inhibition assays. Inhibition constants were obtained by fitting initial rates with variable inhibitor concentrations to eq 1 using Graphpad Prism 8.
| (1) |
where vi and v0 are the initial rates with and without an inhibitor, respectively, [S] and [I] are the substrate and inhibitor concentrations, respectively, Km is the Michaelis constant, and Ki is the inhibition constant. The inhibitor concentration was corrected by eq 2 when it was <10 times the enzyme concentration.27
| (2) |
where [I]′ is the effective inhibitor concentration, [I] is the inhibitor concentration in the assay mixture, and Et is the total enzyme concentration.
Inhibition of H. pylori Growth.
H. pylori J99 was cultured as described previously.28 Briefly, it was grown under microaerophilic conditions (5% O2, 10% CO2, and 85% N2) at 37 °C in brain heart infusion medium (Oxoid) with 10% fetal bovine serum (Sigma). The half-maximal inhibitory concentration (IC50) for MTCF was determined in an H. pylori culture in the exponential growth phase diluted to an OD600 of 0.010. Samples of 100 μL were transferred to 96-well plates. Increasing final concentrations of MTCF (1–250 μM) were added to wells followed by incubation for 72 h under microaerophilic conditions at 37 °C.
A SpectraMax M5 plate reader (Molecular Devices) was used to determine the OD600 value of each well, and the IC50 value was obtained using a nonlinear regression curve fit in GraphPad Prism 8. Rescue of H. pylori growth from MTCF treatment experiments used 6, 13, 25, 50, and 100 μM MTCF supplemented with 50 μM FL. Actively growing H. pylori was inoculated in each treatment with the initial OD600 of 0.005. Cultures were maintained under microaerophilic conditions at 37 °C for 96 h followed by OD600 analysis.
Intracellular Metabolite Analysis.
The metabolite extraction method was described previously.29,30 Briefly, ice-cold extraction buffer (500 μL; 40:40:20 acetonitrile/methanol/water mixture and 0.1 M formic acid) with an internal standard at known concentrations [(RS)-S-adenosyl-L-methionine-d3 (CDN Isotopes)] was added to pelleted H. pylori cells. Samples were vortexed and incubated on ice for 10 min followed by centrifugation at 20200g for 10 min at 4 °C. The supernatant was transferred to a fresh tube and dried under vacuum overnight. Prior to injection, samples were dissolved in 50 μL of distilled H2O. The lipid extraction for menaquinone-6 analysis was carried out using a chloroform/methanol-based method described by Bligh and Dyer.31 H. pylori cells (~2.5 million per sample) were washed in PBS buffer after being harvested; 750 μL of a 1:2 chloroform/methanol mixture with added menaquinone-4 at known concentrations (internal standard) was added to cell pellets followed by a 2 min vortex. Then, 250 μL of chloroform was added to the mixture followed by a 30 s vortex. After that, 250 μL of 1.5 M NaCl was added to the extraction mixture followed by a 30 s vortex. After centrifugation at 20200g for 20 min, the bottom layer (chloroform) was collected and lyophilized in separate tubes, and 100% methanol was used to dissolve extracted lipids before LC-MS analysis. Biological triplicates were prepared for each treatment group in this study.
High-performance liquid chromatography (HPLC) with detection by mass spectrometry (LC-MS) was conducted on an Agilent Technologies (Santa Clara, CA) 1200 system coupled with an Agilent Technologies 6410 triple quadrupole mass spectrometer. The system was operated with the associated MassHunter software package, which was also used for data collection and analysis. Assay mixtures were separated on a Zorbax Rapid Resolution SB-C18 column (2.4 mm × 35 mm, 3.5 μm particle size) equilibrated in 99% buffer A [5 mM perfluoroheptanoic acid and 6 mM ammonium formate (pH 3)] and 1% buffer B (100% acetonitrile). A gradient of 1% to 12% buffer B was applied from 0 to 2 min, followed by a gradient of 12% to 30% buffer B from 2 to 3 min. Then, a gradient of 30% to 50% buffer B was applied from 3 to 5 min before returning to 1% buffer A from 5 to 6 min. The column was allowed to re-equilibrate for 1.5 min under initial conditions before subsequent sample injections. Detection of products was performed by two separate injections using electrospray ionization in negative mode (ESI−) with the following MRM methods (Table S1). For the analysis of menaquinone-6, the extracted lipids were separated on the SB-C18 column equilibrated in 12% buffer A (0.1% formic acid) and 88% buffer B (100% acetonitrile). A gradient of 88% to 99% buffer B was applied from 0 to 0.5 min followed by 22 min with 1% buffer A and 99% buffer B. The column was equilibrated back to initial conditions within 1.5 min before the next injection.
AFL and FL Hydrolysis by HpMTAN and H. pylori Cell Mixture.
AFL and FL hydrolysis by HpMTAN was measured in a discontinuous assay in 1 mL mixtures of 50 mM HEPES (pH 7.0), variable concentrations of the substrate, and purified HpMTAN. Reactions were started by addition of HpMTAN and quenched by addition of sulfuric acid to a final concentration of 0.01 M. Reactions were analyzed by LC-MS using the same method as for metabolite analysis. Initial rates were used for steady state kinetic calculations in GraphPad Prism 8. Live H. pylori cell mixtures were prepared from overnight cultures and washed twice with PBS prior to incubations. AFL or FL (25 μM) was added to ~108 cells in PBS (pH 7.0, 700 μL) to initiate incubations. Supernatants were collected at various time points and analyzed by LC-MS to determine metabolites.
Crystallization.
HpAFLDA (5 mg/mL) was screened for crystallization with commercially available Microlytic (MCSG1–4) and Hampton (crystal screenHT) crystallization conditions. The crystallization of the apoenzyme was performed by sitting drop vapor diffusion at 22 °C. Crystallization used the CRYSTAL-GRYPHON crystallization robot (Art Robbins) in 96-well INTELLI plates (Art Robbins). Crystallization drops contained 0.5 μL of apo HpAFLDA and 0.5 μL of a well solution. The well solution volume was 70 μL. Diffraction-quality crystals were obtained in 2 weeks from 100 mM sodium malonate (pH 4.0) and 12% (w/v) polyethylene glycol 3350.
Co-crystallization of HpAFLDA with MTCF used 5 mg/mL enzyme mixed with MTCF in a 1:2 molar ratio, incubated on ice for 2 h. Crystallization was performed as described above for apo HpAFLDA. Crystals appeared in 2 weeks from a crystallization condition of 100 mM sodium acetate (pH 4.6) and 8% (w/v) polyethylene glycol 4000. Crystals were cryoprotected with 20% ethylene glycol and frozen in liquid nitrogen before X-ray diffraction data collection.
Data Collection and Data Processing.
Diffraction data of apo and MTCF-bound HpAFLDA crystals were collected at the LRL-CAT beamline (Argonne National Laboratory, Lemont, IL) irradiated at a wavelength of 0.97931 Å to resolutions of 2.79 and 1.89 Å, respectively (Table 4). The data were processed using iMOSFLM and scaled by the AIMLESS program of the CCP4 suite in the P1211 space group (Table 4).32,33 The quality of the crystal diffraction data was analyzed by SFCHECK and XTRIAGE.33,34 The Matthews coefficient (Vm) calculations were performed to calculate monomers present in the asymmetric unit. The data collection and processing statistics are summarized in Table 4.
Table 4.
Data Collection and Refinement Statistics of HpAFLDA Structuresa
| apo-HpAFLDA | complex of HpAFLDA with MTCF | |
|---|---|---|
| Unit Cell Data | ||
| space group | P1211 | P1211 |
| cell parameters | a = 73.29 Å, b = 73.29 Å, c = 157.49 Å | a = 72.90 Å, b = 73.07 Å, c = 157.44 Å |
| α = 90.0°, β = 98.6°, γ = 90.0° | α = 90.0°, β = 98.3°, γ = 90.0° | |
| Vm (Å3/Da) | 2.2 | 2.2 |
| no. of subunits in the asymmetric unit | 4.0 | 4.0 |
| Data Collection | ||
| beamline | LRL-CAT | LRL-CAT |
| wavelength (Å) | 0.97931 | 0.97931 |
| temperature (K) | 100 | 100 |
| resolution range (Å) | 155.73–2.79 (2.90–2.79) | 155.79–1.89 (1.92–1.89) |
| total no. of observed reflections | 140666 (14721) | 494028 (24293) |
| no. of unique reflections | 39040 (4432) | 129546 (6388) |
| Rmerge (%)b | 22.7 (105.7) | 12.1 (68.3) |
| CC1/2 (%) | 97.7 (54.7) | 99.2 (66.4) |
| 〈I/σ(I)〉c | 5.0 (1.3) | 6.5 (1.9) |
| completeness (%) | 94.2 (95.7) | 98.8 (98.1) |
| multiplicity | 3.6 (3.3) | 3.8 (3.8) |
| Wilson B-factor (Å2) | 34.7 | 21.0 |
| Refinement | ||
| Rwork (%)d | 25.0 | 23.1 |
| Rfree (%)e | 28.7 | 26.9 |
| no. of atoms | 12744 | 14029 |
| protein | 12570 | 12904 |
| ligand | – | 84 |
| solvent | 174 | 1041 |
| model quality | ||
| root-mean-square deviation from ideal values | ||
| bond lengths (Å) | 0.006 | 0.003 |
| bond angles (deg) | 0.98 | 0.73 |
| average B-factor | ||
| protein atoms (Å2) | 41.3 | 26.2 |
| ligand atoms (Å2) | – | 28.8 |
| waters (Å2) | 30.5 | 31.4 |
| Ramachandran plotf | ||
| most favored regions (%) | 94.1 | 96.6 |
| allowed regions (%) | 5.4 | 3.2 |
| outlier regions (%) | 0.5 | 0.2 |
| PDB entry | 7LKJ | 7LKK |
Values in parentheses refer to the highest-resolution shell.
Rmerge = [ΣhklΣi|Ii(hkl) − 〈I(hkl)〉|]/ΣhklΣi〈Ii(hkl)〉, where Ii(hkl) is the intensity of the ith measurement of reflection (hkl) and 〈I(hkl)〉 is its mean intensity.
I is the integrated intensity, and σ(I) is its estimated standard deviation.
Rwork = (Σhkl|Fo − Fc|)/ΣhklFo, where Fo and Fc are the observed and calculated structure factors, respectively.
Rfree is calculated as described for Rwork but from a randomly selected subset of the data (5%), which were excluded from the refinement calculation.
Calculated by MOLPROBITY.
Structure Determination and Refinement.
Crystal structures of apo and MTCF-bound HpAFLDA were determined by molecular replacement using PHASER.35 Chain A of wild-type Nitratiruptor amidohydrolase [aminodeoxyfutalosine deaminase, Protein Data Bank (PDB) entry 3V7P, UniProt entry A6Q234] was used as the initial phasing model. The model obtained from PHASER was autobuilt using BUCCANEER and then manually adjusted and completed using COOT.36,37 Refinement of the structures was performed by PHENIX-REFINE.34 The final refinement statistics of the structures are summarized in Table 4.
Structure Analysis.
The crystal structure of apo-amidohydrolase (aminodeoxyfutalosine deaminase, PDB entry 3V7P, chain A, UniProt entry A6Q234) from Nitratiruptor was used for structural comparisons. The MTCF complex with P. aeruginosa adenosine deaminase (PDB entry 4GBD, chain A) was also used in the structural comparisons. Structural superimpositions used the SSM protocol of COOT. The geometry analyses of the final model used MolProbity.38 The B-factors of the structures were calculated with the BAVERAGE program of the CCP4 suite. Structural figures were produced by PyMOL. For HpAFLDA structures, subunit A was used for all structural analyses and comparisons.
RESULTS AND DISCUSSION
jhp0252 Encodes an Aminofutalosine Deaminase.
It was previously reported that H. pylori lacks AFL deaminase orthologs, and its DHFL production depends solely on AFL hydrolysis by HpMTAN.7 However, the KEGG database for the H. pylori J99 genome contained a gene entry, jhp0252, annotated as a potential aminofutalosine deaminase (AFL deaminase). We expressed and purified the recombinant jhp0252 protein and investigated its substrate specificity (Table 1). The recombinant enzyme displayed catalytic activity for deamination of AFL, MTA, 5′dAdo, adenosine, and SAH, suggesting a high degree of substrate promiscuity. Among the tested substrates, AFL was the most favorable substrate with a kcat of 0.62 s−1 and a Km of 9 μM (kcat/Km of 6.8 × 104 M−1 s−1). The jhp0252 deaminase showed comparable activity for deamination of MTA and 5′dAdo with kcat/Km values of 1.3 × 104 and 1.2 × 104 M−1 s−1, respectively. The catalytic efficiency for adenosine was 2.6 × 103 M−1 s−1, which was 26-fold lower than that for AFL. Recombinant jhp0252 showed lower activity in catalyzing SAH deamination (kcat/Km of 5.5 × 102 M−1 s−1). On the basis of the substrate profile, we concluded that jhp0252 encodes an active AFL deaminase (HpAFLDA) in H. pylori J99.
Table 1.
Substrate Specificities of HpAFLDA and Steady State Kinetic Constants
| substrate | Km (μM) | kcat (s−1) | kcat/Km (M−1 s−1) |
|---|---|---|---|
| AFL | 9 ± 1 | 0.62 ± 0.03 | 6.8 × 104 |
| MTA | 100 ± 10 | 1.24 ± 0.07 | 1.3 × 104 |
| 5′dAdo | 130 ± 50 | 1.6 ± 0.3 | 1.2 × 104 |
| adenosine | 170 ± 40 | 0.44 ± 0.06 | 2.6 × 103 |
| SAH | 220 ± 50 | 0.12 ± 0.02 | 5.5 × 102 |
Deamination activity on AFL has been reported in Sav2595 from Steptomyces avermitilis as well as its distantly related orthologs (Acel0264, Dr0824, and Nis0429) from three thermophilic bacteria.39 The characterized catalytic efficiencies for AFL deamination range from 4.8 × 106 to 7.9 × 105 M−1 s−1, which are approximately 1–2 orders of magnitude higher than that for jhp0252. However, jhp0252 has greater kcat/Km values for MTA and 5′Ado than other AFL deaminases except Dr0824 from Deinococcus radiodurans. It is worth noting that the kinetic characterization of these four enzymes was carried out at 30 °C instead of 25 °C, which was used to assay jhp0252. The catalytic activity for jhp0252 on AFL (6.8 × 104 M−1 s−1) is similar to that for the activity of Plasmodium falciparum adenosine/MTA deaminase on MTA (9 × 104 M−1 s−1).23 Despite the well-defined catalytic potential for the substrates listed in Table 1, other substrates remain an alternative possibility for this enzyme.
Inhibitors of HpAFLDA.
Coformycin (CF) and 2′-deoxycoformycin (2′d-CF), originally isolated from Streptomyces, are natural product transition state analogues of human adenosine deaminase (HsADA) with submicromolar Ki values.23,40 Coformycin analogues, methylthio-coformycin (MTCF) and methylthio-2′-deoxy-coformycin (MT-2′d-CF), were designed to be specific inhibitors of MTA deaminase from P. falciparum, as the 5′-methylthio group provided additional target selectivity. We tested (8R)-CF, (8R)-MTCF, (8R)-MT-2′d-CF, (8R)-PrT-2′d-CF, and (8R)-PhT-2′d-CF as inhibitors of HpAFLDA (Figure 2A). Initial rates were determined and used to calculate initial inhibition constants (Ki). The MTCF-inhibited reactions showed a second phase of inhibition after reaction for 30 min (Figure 2B). A similar inhibition pattern for MTCF has been reported for PaMTADA,21 establishing MTCF as a slow-onset inhibitor of HpAFLDA. Reaction rates following slow-onset inhibition were used to obtain the second-phase inhibition constant (Ki*). MTCF was the most potent HpAFLDA inhibitor with a Ki* of 63 nM (Table 2). Surprisingly, MT-2′d-CF did not inhibit HpAFLDA activity, even at 10 μM, establishing that the 2′-hydroxyl group is essential for binding of the CF analogue to HpAFLDA. Additional 2′d-CF analogues with modifications at the 5′-subsitutent, including PrT-2′d-CF and PhT-2′d-CF, showed some inhibition, indicating that 5′-substituents also play a critical role in inhibitor affinity.
Figure 2.

(A) Coformycin analogues tested for inhibition of HpAFLDA. (B) Inhibition of HpAFLDA by MTCF. The initial inhibition constant (Ki) was calculated from data collected for the initial rates of the reactions. The slow-onset inhibition constant (Ki*) was calculated from data collected following completion of the slow-onset inhibition.
Table 2.
Summary of Ki Values for HpAFLDAa
| inhibitor | Ki (nM) | Ki* (nM) |
|---|---|---|
| (8R)-CF | N/O | N/O |
| (8R)-MTCF | 560 ± 170 | 63 ± 4 |
| (8R)-MT-2′d-CF | N/O | N/O |
| (8R)-PrT-2′d-CF | 420 ± 140 | N/O |
| (8R)-PhT-2′d-CF | 850 ± 200 | N/O |
N/O means inhibition not observed.
Inhibition of H. pylori Growth.
Is FL essential for bacterial growth? Addition of MTCF to H. pylori cultures demonstrated an inhibitory effect on H. pylori growth in liquid medium with an IC50 of 14 ± 1 μM (Figure 3A). Moreover, an addition of 50 μM FL partially reversed the growth inhibition (Figure 3B). Thus, FL is an important metabolite in H. pylori growth, and impairment of HpAFLDA intracellular activity inhibits H. pylori growth.
Figure 3.

(A) Inhibition of H. pylori growth by MTCF. (B) The inhibitory effect of MTCF on H. pylori growth is reduced by supplementation with 50 μM futalosine (FL).
Inhibition of Intracellular HpAFLDA Activity by MTCF.
Inhibition of cell growth, assumed to result from inhibition of HpAFLDA, requires experimental validation. Cells were cultured with and without MTCF at 14 μM (IC50 value), and cell extracts were analyzed for MTA, MTI, AFL, FL, and DHFL. In the MTCF-treated groups, MTI and FL levels were significantly reduced [p < 0.0001 and p < 0.005, respectively (Figure 4)]. Depletion of the HpAFLDA products in MTCF treatment cells establishes that HpAFLDA activity was reduced.
Figure 4.

Metabolic analysis of cellular inhibition of HpAFLDA. (A) Inhibitory effect on the conversion of MTA to MTI. (B) Inhibitory effect on the conversion of AFL to FL. (C) Reduced HpAFLDA or HpMTAN activity led to decreased DHFL production. (D) Menaquinone-6 biosynthesis was negatively modulated by HpAFLDA or HpMTAN inhibition.
MTA and AFL both serve as substrates for HpAFLDA, and both are expected to accumulate upon inhibition of HpAFLDA if HpAFLDA catalyzes their major metabolic path in vivo. MTCF treatment caused an increase in AFL abundance [p < 0.0001 (Figure 4B)] caused by decreased HpAFLDA activity. However, the level of cellular MTA did not increase following MTCF treatment. MTA and AFL are also both substrates of HpMTAN, a nucleoside hydrolase with broad substrate specificity.20 The kcat/Km values for MTA and AFL hydrolysis by HpMTAN were 8.3 × 106 and 6.9 × 104 M−1 s−1, respectively (Supporting Information). As the catalytic efficiency of MTA hydrolysis is >100-fold higher than that of AFL hydrolysis, this metabolic analysis indicates that HpMTAN is efficient at the hydrolysis of MTA but not AFL when intracellular HpAFLDA activity is inhibited.
DHFL is the reaction product of HpMTAN-catalyzed AFL hydrolysis7,19 and is an essential menaquinone precursor in the futalosine pathway (Figure 1).5–9 The role of HpAFLDA in menaquinone biosynthesis was examined by treatment with MTCF at the IC50 concentration. The cellular DHFL levels decreased significantly [p < 0.005 (Figure 4C)] in the MTCF-treated groups relative to the untreated controls. Significant decreases in cellular DHFL concentrations were also observed in cells cultured with BuT-DADMe-ImmA (BTDIA), an HpMTAN transition state analogue,19 at its half-maximal inhibitory concentration (14 nM). Additionally, we measured the levels of menaquinone-6, the dominant isoprenoid quinone in Helicobacter,41 in lipid extracts. The menaquinone-6 concentrations decreased significantly in both MTCF- and BTDIA-treated groups [p < 0.005 and p < 0.005, respectively (Figure 4D)]. It is known that HpMTAN activity is critical to MK production in H. pylori.7,19 The inhibition of HpAFLDA by MTCF and the inhibition of HpMTAN by BTDIA both led to decreased levels of DHFL, a downstream menaquinone precursor, and menaquinone-6, the final product of the pathway. Overall, the metabolic analysis suggests that HpAFLDA plays a significant, but as yet undefined, role in modulating menaquinone biosynthesis in H. pylori.
Exploring Futalosine N-Ribosyl Hydrolase in H. pylori.
In microorganisms utilizing the AFL → FL → DHFL pathway for menaquinone biosynthesis, the MqnB hydrolase converts FL to DHFL and hypoxanthine (Figure 1).5,7 However, no MqnB activity or genome coding region has been reported in H. pylori. In the past, HpMTAN has been misannotated as HpMqnB and has been reported to be active for hydrolysis.7 We tested HpMTAN for FL hydrolase activity and confirmed that HpMTAN was capable of catalyzing only AFL hydrolysis but not FL hydrolysis, in agreement with published studies (Table 3).7 Metabolism of AFL and FL in H. pylori was investigated by adding AFL or FL to live H. pylori cells, followed by metabolite analysis. H. pylori cells readily converted AFL to FL and adenine (Figure 5A). Thus, AFL hydrolysis and deamination occur simultaneously with accumulation of products. Cells incubated with FL caused no significant changes in the FL concentration, and no hypoxanthine was formed (Figure 5B). Under these conditions, H. pylori cells contain no functionally significant FL hydrolase.
Table 3.
Kinetic Analysis of AFL and FL Hydrolysis by HpMTAN
| substrate | Km (μM) | kcat (s−1) | kcat/Km (M−1 s−1) |
|---|---|---|---|
| AFL | 1.6 ± 0.3 | 0.022 ± 0.02 | 1.4 × 104 |
| FL | n/a | <1 × 10−5 | n/a |
Figure 5.

Metabolism of AFL and FL by H. pylori cell culture. (A) Conversion of AFL to adenine by hydrolysis and to FL by deamination in H. pylori cells. (B) Incubation of FL with H. pylori cells.
Investigating Potential Feedback Mechanisms in the Futalosine Pathway.
The involvement of HpAFLDA in MK biosynthesis is supported by the inhibition of HpAFLDA causing metabolic perturbations in the futalosine pathway (Figure 4). We examined the possibility of potential feedback mechanisms involving HpAFLDA activity, namely, the production of FL as a feedback control when the pathway was overproducing AFL. Supplementation of culture growth medium with 200 μM AFL did not cause a significant impact on cell growth when BTDIA was used to stress MK biosynthesis (Figure 6A). Supplementation with FL at concentrations ranging from 1 to 200 μM did not affect cell growth (Figure 6B). A simple feedback regulation involving HpAFLDA activity within the futalosine pathway was not observed.
Figure 6.

Investigation of the potential effect of metabolite supplementation on cell growth. (A) Supplementation with AFL in the presence of BTDIA treatment. (B) Supplementation of FL as a potential feedback regulator of the pathway.
HpAFLDA Structure and Comparisons with Homology.
The crystal structures of apo-HpAFLDA and MTCF-bound structures were determined in space group P1211 with acceptable R and Rfree values (Table 4). The PISA analysis supports a dimer with two catalytic sites as the crystallographic state. Except for a few residues at the N-terminus and surface side chain residues, the entire polypeptide chain of apo and HpAFLDA structures was well-defined in the electron density.
The crystal structure of apo-HpAFLDA (PDB entry 7LKJ) was determined to 2.79 Å resolution with four monomers (chains A–D) in the asymmetric unit. Polypeptide chains A and C were well-defined in the electron density map, whereas chains B and D were disordered in Phe70–Lys74 and Gly94–Asn99. The solvent accessible surface area of the dimer interface is 1196 Å2. HpAFLDA contains two domains, including a small N-terminal β-barrel domain positioned between two α-helices (Met1–Leu58 and Glu354–Ile409) and a C-terminally distorted TIM barrel domain (Pro59–Leu360) (Figure 7). The density for metal binding was observed in the active site of HpAFLDA. Both Fe2+ and Zn2+ were modeled and refined on the basis of the coordination geometry and refinement parameters. The precedent of the Fe2+ and Zn2+ atoms was established in the original description of the closely related enzyme (Hp0267, 96% identical) from H. pylori 26695 and reported to contain both Fe2+ and Zn2+ by direct metal analysis and by metal substitution experiments.22,39,42
Figure 7.

Crystal structure of HpAFLDA. (A) Quaternary homodimeric structure of HpAFLDA. The active site of the enzyme is deeply buried in the structure and highlighted with black arrows. The two domains including the TIM barrel domain (blue) and β-barrel domain (magenta) are highlighted. (B) The active site residues involved in inhibitor binding are shown with bound MTCF. The metal ion (brown) and water (red) molecules are also indicated. The hydrogen bond and metal coordination bonds are shown with the black dotted lines. (C) Stereoview structural comparison of apo-HpAFLDA (PDB entry 7LKJ) with the MTCF-bound structure (PDB entry 7LKK). The gray structure is the unliganded monomer A of HpAFLDA, and the blue structure is the MTCF-bound monomer. There are no major structural changes observed between apo and MTCF-bound structures.
Most adenosine deaminases (both eukaryotes and prokaryotes) use Zn2+ as the metal cofactor.43–45 The metal content of HpAFLDA was quantified for both Fe2+ and Zn2+ content (Supporting Information). The results showed a molar ratio of 0.38 for Fe2+ and enzyme active sites and a molar ratio of 0.60 for Zn2+ and enzyme active sites. This heterogeneous Fe2+ and Zn2+ content for HpAFLDA was similar to that found for the adenosine deaminase from Hp0267.22 Fe2+- or Zn2+-bound only enzymes were prepared for kinetic comparison in metal removal and replacement experiments (Supporting Information). The kcat for AFL with fully Fe2+-substituted HpAFLDA is roughly 40% faster than that for fully Zn2+-substituted HpAFLDA. However, the overall catalytic efficiency for AFL for fully Fe2+-bound HpAFLDA is lower than that for the Zn2+-bound enzyme (kcat/Km values of 2.0 and 5.0 × 104 M−1 s−1, respectively) (Table S2).
The structure of HpAFLDA bound to MTCF (PDB entry 7LKK) was also determined with four monomers (chains A–D) in the asymmetric unit in space group P1211 at 1.89 Å resolution with the entire polypeptide chain well resolved in the electron density. MTCF is clearly resolved in the electron density map (Figure 7 and SI-1). Inhibitor binding does not cause major conformational changes or loop movements, as superposition of apo-HpAFLDA and MTCF-bound (chain-A) structures gives a root-mean-square deviation (RMSD) of 0.256 Å. The solvent accessible surface area of the dimer interface with MTCF bound was 1239 Å2, greater than that of apo-HpAFLDA.
The structure of HpAFLDA was compared to those of AFLDA from Nitratiruptor and MTADA from P. aeruginosa (PaMTADA).21,39 Human adenosine deaminase (PDB entry 3IAR) shares only 11.9% sequence identity with HpAFLDA and was not included in the comparison. HpAFLDA shares 37.2% and 25.0% sequence identity with the AFLDA from Nitratiruptor and MTADA from P. aeruginosa, respectively. Upon comparison of HpAFLDA and Nitratiruptor AFLDA (PDB entry 3V7P), the RMSD for Cα was 0.854 Å, suggesting nearly identical secondary structures. The Leu12–Glu18 and Lys222–Thr239 loops of HpAFLDA differ slightly from those of Nitratiruptor AFLDA (Figure 8.). Comparison of HpAFLDA with PaMTADA (PDB entry 4GBD) gives a RMSD value of 1.924 Å, and significant differences between HpAFLDA and PaMTADA structural folds. The substrate binding loop (Leu146–Ala160 of HpAFLDA) is open but a closed conformation in PaMTADA consistent with structural specificity for the smaller substrate (MTA). The helix covering the catalytic site of HpAFLDA (Lys222–Thr239) is missing in PaMTADA (Figure 8). The amino acid residues involved in metal coordination are conserved in all three structures.
Figure 8.

Structural comparisons of HpAFLDA with orthologues. (A) Stereoview of the structural comparison of HpAFLDA (blue, PDB entry 7LKK) with Nitratiruptor AFLDA (cyan, PDB entry 3V7P). The minor structural differences between Leu12–Glu18 and Lys222–Thr239 loops are denoted with the black and red stars, respectively. (B) Structural comparison (stereoview) of HpAFLDA (blue, PDB entry 7LKK) with PaMTADA (gray, PDB entry 4GBD). The open conformation of the HpAFLDA substrate binding loop from residue Leu146 to Ala160 is highlighted with a black star. The catalytic sitecovering helix of HpAFLDA from Lys222 to Thr239 is highlighted with a red star.
Catalytic Site and MTCF Binding to HpAFLDA.
The catalytic site of HpAFLDA is located deep in the monomer (Figure 7). In apo-HpAFLDA, only a few ordered water molecules are observed (Figure SI-2). The Me2+ binding site is coordinated with NE2(His65), NE2(His67), NE2(His207), OD2(Asp314), and O8 of the MTCF inhibitor (Figures 9 and 10). The residues interacting with the MTCF inhibitor with 4 Å are His65, His67, Phe70, Trp85, Asn124, Glu144, His180, His207, Glu210, His261, Leu292, Asp314, and Ser317 (Figure 9). With respect to the 5′-alkyl group, a hydrophobic packet is formed by Leu86, Val89, Leu90, Leu146, Ser148, Tyr183, Ser184, Phe228, Tyr229, and Leu233 (Figure 9). This hydrophobic pocket provides adequate space for binding the 5′-group of the AFL substrate (Figure 1). The adenine mimic of the inhibitor is sandwiched between His207 and Trp85 with Phe70 providing a hydrophobic interaction with the β-ribosyl face of the inhibitor. Phe70 is disordered in the apo-HpAFLDA structure. N6 of MTCF is hydrogen bonded with OD1 and OD2 of Glu201. The OG(Ser317) and a water molecule have hydrogen bond interactions with N1 of MTCF. The ribosyl O2 and O3 atoms of MTCF are hydrogen bonded with OE1(Glu144) and OD1(Asn124). A solvent ethylene glycol is found in the hydrophobic packet beyond the 5′-alkyl thio group of the inhibitor (Figures 9 and 10).
Figure 9.

Binding of MTCF to HpAFLDA (stereoviews). (A) Active site of HpAFLDA in complex with MTCF (yellow). The amino acid residues interacting with MTCF are indicated. The binding of metal (brown), water (red), and ethylene glycol (EG, light blue) is also highlighted. Selected hydrogen bond interactions and metal coordinations are shown as black dotted lines (see also Figure 8). (B) Amino acids forming the hydrophobic pocket extending beyond the 5′-methylthio group of MTCF. This pocket provides adequate space for the substrate (AFL) to bind to the enzyme. (C) Surface representation of the HpAFLDA binding site bound with the MTCF. The hydrophobic pocket extends from the 5′-methylthio group of MTCF toward the viewer and the surface of the protein.
Figure 10.

Two-dimensional active site distance map comparing (A) PaMTADA and (B) HpAFLDA. The hydrogen bonds (dotted black lines) and metal coordination (dotted red lines) are indicated. The distances are in angstroms.
A comparison of the catalytic site contacts to MTCF in PaMTADA and HpAFLDA shows conservation for metal binding. Adenine interactions differ with residues Phe70, Phe82, Val89, Leu90, and Ser317 in HpAFLDA being replaced by Met77, Leu93, Leu89, Trp98, and Ala312, respectively, in PaMTADA. The Ser317 hydrogen bond interaction with N1 of MTCF in HpAFLDA is replaced with a water molecule in PaMTADA.21 The substrate binding pocket of PaMTADA is smaller than that in HpAFLDA to allow 5′-substituent binding with hydrophobic interactions. The binding conformation of ribose and the 5′-substituent for MTCF also differ in PaMTADA such that MTCF is a more potent inhibitor for PaMTADA.
Structure–Inhibition Relationship of MTCF.
MTCF is a slow-onset tight-binding inhibitor of HpAFLDA with a Ki* of 63 ± 4 nM. However, it binds 13000-fold tighter to PaMTADA with a Ki* of 4.8 ± 0.5 pM.21 The active site of PaMTADA is designed to accommodate a 5′-methylthio group from its 5′-methylthioadenosine substrate (Figure 8). The catalytic pocket of HpAFLDA is much larger to accommodate the 5′-substituent of AFL, leaving an unfavorable water–hydrophobic site interaction with this inhibitor (Figure 9). Larger hydrophobic groups attached to the 5′-alkyl of MTCF have the potential to increase the binding affinity and provide an ongoing synthetic chemistry challenge.
CONCLUSIONS
Chronic H. pylori infections cause gastric and duodenal ulcers and are considered a strong risk factor for developing gastric adenocarcinoma.46 The current standard-of-care treatment for H. pylori infection fails at a rate of 30–40% due to resistance against commonly used antibiotics.16 Thus, there is growing interest in developing narrow-spectrum therapeutics for H. pylori infection. Menaquinone is an essential component in the electron transport chain in bacteria that adopt anaerobic respiration.47 MTAN-catalyzed AFL hydrolysis has been reported to be a drug target for anti-H. pylori therapeutic development.19,48,49 Here, we have characterized jhp0252 to be an AFL deaminase in H. pylori J99. HpAFLDA catalyzes AFL deamination to form FL, and the AFL → FL process is a biologically significant, possibly regulatory, step for growth of H. pylori. MTCF is a nanomolar inhibitor of HpAFLDA and is functional in growing cells. MTCF demonstrated an anti-H. pylori growth activity in bacterial culture, partially reversed by FL, highlighting the importance of FL production to bacterial survival. The crystal structure of apo-HpAFLDA and in complex with MTCF revealed favorable inhibitor binding interactions at the catalytic site. A large hydrophobic pocket in the crystal structure is present extending from the 5′-methylthio group of the inhibitor. The binding mode of MTCF in the HpAFLDA binding site provides the basis for designing the next-generation inhibitors.
The catalytic specificity for AFL implies that the primary biological function of HpAFLDA resides in the futalosine pathway. However, an active AFL → FL → DHFL pathway has not been established in H. pylori. No whole cell activity for the FL → DHFL step could be detected. The current literature favors HpMTAN in the primary metabolic path for DHFL production in H. pylori. Transition state analogues of HpMTAN inhibit H. pylori growth with IC90 values as low as 6 ng/mL, supporting the essential role of HpMTAN in AFL hydrolysis.7,19,48 Kinetic and metabolic analysis of HpAFLDA demonstrates its capability of deaminating AFL in vitro and in vivo. The formation of intracellular FL from AFL is readily detected in viable cell assays, establishing the entry of AFL into cells and release of FL following AFL deamination. Intracellular AFL and FL levels changed substantially when HpAFLDA was inhibited by MTCF, further establishing AFL deamination as a physiological reaction in H. pylori. H. pylori cell growth inhibition by MTCF and cell growth rescue by supplementation with FL support a role for HpAFLDA in menaquinone synthesis. A significant decrease in DHFL and menaquinone-6 concentrations was observed when cells were cultured under HpAFLDA inhibition. The missing link in establishing a role for HpAFLDA in the pathway is the lack of a measurable catalytic activity for FL conversion to DHFL in H. pylori, possibly suggesting an activity needed for other environmental conditions. A potential feedback regulation involving HpAFLDA formation of FL did not identify growth modulation via MK biosynthesis. The possibility that formation of FL by HpAFLDA plays a regulatory role in other pathways essential for growth of H. pylori remains interesting.
HpAFLDA catalyzes the deamination of MTA and 5′dAdo in addition to AFL. These metabolites are products of the SAM-dependent polyamine pathway and of radical SAM reactions, respectively,50 allowing the reaction products to enter salvage pathways.51 MTI, converted from MTA, can undergo N-ribosyl phosphorolysis to yield hypoxanthine and 5-methylthioribose 1-phosphate, a methionine salvage precursor in P. aeruginosa.21 Deamination of 5′dAdo produces 5′-deoxyinosine, a precursor for purine recycling in P. falciparum.23,52 The MTA metabolite is reported to cause potent product inhibition on polyamine biosynthesis,53 and accumulation of 5′dAdo inhibits radical SAM reactions, such as biotin synthesis.54 Being capable of metabolizing MTA and 5′dAdo efficiently suggests that HpAFLDA has a function in recycling end products in several pathways in addition to a potential regulatory role in menaquinone biosynthesis.
Supplementary Material
ACKNOWLEDGMENTS
The Albert Einstein Crystallographic Core X-ray diffraction facility is supported by National Institutes of Health Shared Instrumentation Grant S10 OD020068. Data collection also involved resources of the Advanced Photon Source, a U.S. Department of Energy (DOE) Office of Science User Facility operated for the DOE Office of Science by Argonne National Laboratory under Contract DE-AC02-06CH11357. Use of the Lilly Research Laboratories Collaborative Access Team (LRL-CAT) beamline at Sector 31 of the Advanced Photon Source was provided by Eli Lilly Co., which operates the facility. The authors thank Dr. Tyler L. Grove for helpful scientific advice about LC-MS-related experiments.
Funding
This work was supported by National Institutes of Health Grant GM041916.
Footnotes
The authors declare no competing financial interest.
ASSOCIATED CONTENT
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.biochem.1c00215.
Continuous assay for HpMTAN, ferrozine-based spectrophotometric assay for Fe2+ and Fe3+ detection, zincon-based spectrophotometric assay for Zn2+, preparation of Fe2+- or Zn2+-bound only HpAFLDA for kinetic comparison, additional crystallographic figures, and mass spectrometry parameters for detection of cellular metabolites (PDF)
Accession Codes
UniProt entry A6Q234 and PDB entries 7LKJ and 7LKK.
Complete contact information is available at: https://pubs.acs.org/10.1021/acs.biochem.1c00215
Contributor Information
Mu Feng, Department of Biochemistry, Albert Einstein College of Medicine, Bronx, New York 10461, United States.
Rajesh K. Harijan, Department of Biochemistry, Albert Einstein College of Medicine, Bronx, New York 10461, United States
Lawrence D. Harris, The Ferrier Research Institute, Victoria University of Wellington, Wellington 6140, New Zealand.
Peter C. Tyler, The Ferrier Research Institute, Victoria University of Wellington, Wellington 6140, New Zealand.
Richard F. G. Fröhlich, The Ferrier Research Institute, Victoria University of Wellington, Wellington 6140, New Zealand
Morais Brown, Department of Biochemistry, Albert Einstein College of Medicine, Bronx, New York 10461, United States.
Vern L. Schramm, Department of Biochemistry, Albert Einstein College of Medicine, Bronx, New York 10461, United States.
REFERENCES
- (1).Bray F, Ferlay J, Soerjomataram I, Siegel RL, Torre LA, and Jemal A (2018) Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. Ca-Cancer J. Clin. 68, 394–424. [DOI] [PubMed] [Google Scholar]
- (2).Bishop DHL, Pandya KP, and King HK (1962) Ubiquinone and vitamin K in bacteria. Biochem. J. 83, 606–614. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (3).Bentley R, and Meganathan R (1982) Biosynthesis of vitamin K (menaquinone) in bacteria. Microbiol. Rev. 46, 241–280. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (4).Meganathan R, and Kwon O (2009) Biosynthesis of menaquinone (vitamin K2) and ubiquinone (Coenzyme Q). EcoSal Plus 3, 2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (5).Hiratsuka T, Furihata K, Ishikawa J, Yamashita H, Itoh N, Seto H, and Dairi T (2008) An alternative menaquinone biosynthetic pathway operating in microorganisms. Science 321, 1670–1673. [DOI] [PubMed] [Google Scholar]
- (6).Zhi X-Y, Yao J-C, Tang S-K, Huang Y, Li H-W, and Li W-J (2014) The futalosine pathway played an important role in menaquinone biosynthesis during early prokaryote evolution. Genome Biol. Evol. 6, 149–160. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (7).Arakawa C, Kuratsu M, Furihata K, Hiratsuka T, Itoh N, Seto H, and Dairi T (2011) Diversity of the early step of the futalosine pathway. Antimicrob. Agents Chemother. 55, 913–916. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (8).Li X, Apel D, Gaynor EC, and Tanner ME (2011) 5′-methylthioadenosine nucleosidase is implicated in playing a key role in a modified futalosine pathway for menaquinone biosynthesis in Campylobacter jejuni. J. Biol. Chem 286, 19392–19398. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (9).Barta ML, Thomas K, Yuan H, Lovell S, Battaile KP, Schramm VL, and Hefty PS (2014) Structural and biochemical characterization of Chlamydia trachomatis hypothetical protein CT263 supports that menaquinone synthesis occurs through the futalosine pathway. J. Biol. Chem. 289, 32214–32229. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (10).Wroblewski LE, Peek RM, and Wilson KT (2010) Helicobacter pylori and gastric cancer: factors that modulate disease risk. Clin. Microbiol. Rev. 23, 713–739. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (11).Peek RM, and Blaser MJ (2002) Helicobacter pylori and gastrointestinal tract adenocarcinomas. Nat. Rev. Cancer 2, 28–37. [DOI] [PubMed] [Google Scholar]
- (12).Chey WD, Leontiadis GI, Howden CW, and Moss SF (2017) ACG clinical guideline: treatment of Helicobacter pylori Infection. Am. J. Gastroenterol. 112, 212–239. [DOI] [PubMed] [Google Scholar]
- (13).Forman D, Newell DG, Fullerton F, Yarnell JW, Stacey AR, Wald N, and Sitas F (1991) Association between infection with Helicobacter pylori and risk of gastric cancer: evidence from a prospective investigation. BMJ. 302, 1302–1305. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (14).Chang AH, and Parsonnet J (2010) Role of Bacteria in Oncogenesis. Clin. Microbiol. Rev. 23, 837–857. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (15).Nomura A, Stemmermann GN, Chyou PH, Kato I, Perez-Perez GI, and Blaser MJ (1991) Helicobacter pylori infection and gastric carcinoma among Japanese Americans in Hawaii. N. Engl. J. Med. 325, 1132–1136. [DOI] [PubMed] [Google Scholar]
- (16).Rimbara E, Fischbach LA, and Graham DY (2011) Optimal therapy for Helicobacter pylori infections. Nat. Rev. Gastroenterol. Hepatol. 8, 79–88. [DOI] [PubMed] [Google Scholar]
- (17).Tacconelli E, Carrara E, Savoldi A, Harbarth S, Mendelson M, Monnet DL, Pulcini C, Kahlmeter G, Kluytmans J, Carmeli Y, Ouellette M, Outterson K, Patel J, Cavaleri M, Cox EM, Houchens CR, Grayson ML, Hansen P, Singh N, Theuretzbacher U, Magrini N, Aboderin AO, Al-Abri SS, Awang Jalil N, Benzonana N, Bhattacharya S, Brink AJ, Burkert FR, Cars O, Cornaglia G, Dyar OJ, Friedrich AW, Gales AC, Gandra S, Giske CG, Goff DA, Goossens H, Gottlieb T, Guzman Blanco M, Hryniewicz W, Kattula D, Jinks T, Kanj SS, Kerr L, Kieny M-P, Kim YS, Kozlov RS, Labarca J, Laxminarayan R, Leder K, Leibovici L, Levy-Hara G, Littman J, Malhotra-Kumar S, Manchanda V, Moja L, Ndoye B, Pan A, Paterson DL, Paul M, Qiu H, Ramon-Pardo P, Rodríguez-Baño J, Sanguinetti M, Sengupta S, Sharland M, Si-Mehand M, Silver LL, Song W, Steinbakk M, Thomsen J, Thwaites GE, van der Meer JW, Van Kinh N, Vega S, Villegas MV, Wechsler-Fördös A, Wertheim HFL, Wesangula E, Woodford N, Yilmaz FO, and Zorzet A (2018) Discovery, research, and development of new antibiotics: the WHO priority list of antibiotic-resistant bacteria and tuberculosis. Lancet Infect. Dis. 18, 318–327. [DOI] [PubMed] [Google Scholar]
- (18).Dairi T (2012) Menaquinone biosyntheses in microorganisms. Methods Enzymol. 515, 107–122. [DOI] [PubMed] [Google Scholar]
- (19).Wang S, Haapalainen AM, Yan F, Du Q, Tyler PC, Evans GB, Rinaldo-Matthis A, Brown RL, Norris GE, Almo SC, and Schramm VL (2012) A picomolar transition state analogue inhibitor of MTAN as a specific antibiotic for H. pylori. Biochemistry 51, 6892–6894. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (20).Mishra V, and Ronning DR (2012) Crystal structures of the Helicobacter pylori MTAN enzyme reveal specific interactions between S-adenosylhomocysteine and the 5′-alkylthio binding subsite. Biochemistry 51, 9763–9772. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (21).Guan R, Ho M-C, Fröhlich RFG, Tyler PC, Almo SC, and Schramm VL (2012) Methylthioadenosine deaminase in an alternative quorum sensing pathway in Pseudomonas aeruginosa. Biochemistry 51, 9094–9103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (22).Miller EF, and Maier RJ (2013) Identification of a new class of adenosine deaminase from Helicobacter pylori with homologs among diverse taxa. J. Bacteriol. 195, 4154–4160. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (23).Tyler PC, Taylor EA, Fröhlich RFG, and Schramm VL (2007) Synthesis of 5′-methylthio coformycins: specific inhibitors for malarial adenosine deaminase. J. Am. Chem. Soc. 129, 6872–6879. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (24).Li X, and Tanner ME (2010) An efficient synthesis of futalosine. Tetrahedron Lett. 51, 6463–6465. [Google Scholar]
- (25).Alm RA, Ling LS, Moir DT, King BL, Brown ED, Doig PC, Smith DR, Noonan B, Guild BC, de Jonge BL, Carmel G, Tummino PJ, Caruso A, Uria-Nickelsen M, Mills DM, Ives C, Gibson R, Merberg D, Mills SD, Jiang Q, Taylor DE, Vovis GF, and Trust TJ (1999) Genomic-sequence comparison of two unrelated isolates of the human gastric pathogen Helicobacter pylori. Nature 397, 176–180. [DOI] [PubMed] [Google Scholar]
- (26).Burgos ES, Walters RO, Huffman DM, and Shechter D (2017) A simplified characterization of S-adenosyl-L-methionine-consuming enzymes with 1-Step EZ-MTase: a universal and straightforward coupled-assay for in vitro and in vivo setting. Chem. Sci. 8, 6601–6612. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (27).Morrison JF, and Walsh CT (2006) The behavior and significance of slow-binding enzyme inhibitors. Adv. Enzymol. Relat. Areas Mol. Biol, 201–301. [DOI] [PubMed] [Google Scholar]
- (28).Namanja-Magliano HA, Evans GB, Harijan RK, Tyler PC, and Schramm VL (2017) Transition state analogue inhibitors of 5′-deoxyadenosine/5′-methylthioadenosine nucleosidase from Mycobacterium tuberculosis. Biochemistry 56, 5090–5098. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (29).Bennett BD, Yuan J, Kimball EH, and Rabinowitz JD (2008) Absolute quantitation of intracellular metabolite concentrations by an isotope ratio-based approach. Nat. Protoc. 3, 1299–1311. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (30).Gizzi AS, Grove TL, Arnold JJ, Jose J, Jangra RK, Garforth SJ, Du Q, Cahill SM, Dulyaninova NG, Love JD, Chandran K, Bresnick AR, Cameron CE, and Almo SC (2018) A naturally occurring antiviral ribonucleotide encoded by the human genome. Nature 558, 610–614. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (31).Bligh EG, and Dyer WJ (1959) A rapid method of total lipid extraction and purification. Can. J. Biochem. Physiol. 37, 911–917. [DOI] [PubMed] [Google Scholar]
- (32).Battye TGG, Kontogiannis L, Johnson O, Powell HR, and Leslie AGW (2011) iMOSFLM: a new graphical interface for diffraction-image processing with MOSFLM. Acta Crystallogr., Sect. D: Biol. Crystallogr. 67, 271–281. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (33).Collaborative Computational Project, Number 4 (1994) The CCP4 suite: programs for protein crystallography. Acta Crystallogr., Sect. D: Biol. Crystallogr. 50, 760–763. [DOI] [PubMed] [Google Scholar]
- (34).Adams PD, Afonine PV, Bunkóczi G, Chen VB, Davis IW, Echols N, Headd JJ, Hung L-W, Kapral GJ, Grosse-Kunstleve RW, McCoy AJ, Moriarty NW, Oeffner R, Read RJ, Richardson DC, Richardson JS, Terwilliger TC, and Zwart PH (2010) PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr., Sect. D: Biol. Crystallogr. 66, 213–221. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (35).McCoy AJ, Grosse-Kunstleve RW, Adams PD, Winn MD, Storoni LC, and Read RJ (2007) Phaser crystallographic software. J. Appl. Crystallogr. 40, 658–674. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (36).Cowtan K (2006) The Buccaneer software for automated model building. 1. Tracing protein chains. Acta Crystallogr., Sect. D: Biol. Crystallogr. 62, 1002–1011. [DOI] [PubMed] [Google Scholar]
- (37).Emsley P, Lohkamp B, Scott WG, and Cowtan K (2010) Features and development of Coot. Acta Crystallogr., Sect. D: Biol. Crystallogr. 66, 486–501. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (38).Chen VB, Arendall WB, Headd JJ, Keedy DA, Immormino RM, Kapral GJ, Murray LW, Richardson JS, and Richardson DC (2010) MolProbity: all-atom structure validation for macromolecular crystallography. Acta Crystallogr., Sect. D: Biol. Crystallogr. 66, 12–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (39).Goble AM, Toro R, Li X, Ornelas A, Fan H, Eswaramoorthy S, Patskovsky Y, Hillerich B, Seidel R, Sali A, Shoichet BK, Almo SC, Swaminathan S, Tanner ME, and Raushel FM (2013) Deamination of 6-aminodeoxyfutalosine in menaquinone biosynthesis by distantly related enzymes. Biochemistry 52, 6525–6536. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (40).Hovi T, Smyth JF, Allison AC, and Williams SC (1976) Role of adenosine deaminase in lymphocyte proliferation. Clin. Exp. Immunol. 23, 395–403. [PMC free article] [PubMed] [Google Scholar]
- (41).Marcelli SW, Chang HT, Chapman T, Chalk PA, Miles RJ, and Poole RK (1996) The respiratory chain of Helicobacter pylori: identification of cytochromes and the effects of oxygen on cytochrome and menaquinone levels. FEMS Microbiol. Lett. 138, 59–64. [DOI] [PubMed] [Google Scholar]
- (42).Barber-Zucker S, Shaanan B, and Zarivach R (2017) Transition metal binding selectivity in proteins and its correlation with the phylogenomic classification of the cation diffusion facilitator protein family. Sci. Rep. 7, 16381. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (43).Wilson DK, Rudolph FB, and Quiocho FA (1991) Atomic structure of adenosine deaminase complexed with a transition-state analog: understanding catalysis and immunodeficiency mutations. Science 252, 1278–1284. [DOI] [PubMed] [Google Scholar]
- (44).Luo M, and Schramm VL (2008) Transition state structure of E. coli tRNA-specific adenosine deaminase. J. Am. Chem. Soc. 130, 2649–2655. [DOI] [PubMed] [Google Scholar]
- (45).Niu W, Shu Q, Chen Z, Mathews S, Di Cera E, and Frieden C (2010) The role of Zn2+ on the structure and stability of murine adenosine deaminase. J. Phys. Chem. B 114, 16156–16165. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (46).Correa P, and Piazuelo MB (2011) Helicobacter pylori infection and gastric adenocarcinoma. U.S. Gastroenterology and Hepatology Review 7, 59–64. [PMC free article] [PubMed] [Google Scholar]
- (47).Paudel A, Hamamoto H, Panthee S, and Sekimizu K (2016) Menaquinone as a potential target of antibacterial agents. Drug Discoveries Ther. 10, 123–128. [DOI] [PubMed] [Google Scholar]
- (48).Wang S, Cameron SA, Clinch K, Evans GB, Wu Z, Schramm VL, and Tyler PC (2015) New antibiotic candidates against Helicobacter pylori. J. Am. Chem. Soc. 137, 14275–14280. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (49).Carl AG, Harris LD, Feng M, Nordstrøm LU, Gerfen GJ, Evans GB, Silakov A, Almo SC, and Grove TL (2020) Narrow-spectrum antibiotic targeting of the radical SAM enzyme MqnE in menaquinone biosynthesis. Biochemistry 59, 2562–2575. [DOI] [PubMed] [Google Scholar]
- (50).Parveen N, and Cornell KA (2011) Methylthioadenosine/S-adenosylhomocysteine nucleosidase, a critical enzyme for bacterial metabolism. Mol. Microbiol. 79, 7–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (51).Madrid DC, Ting LM, Waller KL, Schramm VL, and Kim K (2008) Plasmodium falciparum purine nucleoside phosphorylase is critical for viability of malaria parasites. J. Biol. Chem. 283, 35899–35907. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (52).Miller D, O’Brien K, Xu H, and White RH (2014) Identification of a 5′-deoxyadenosine deaminase in Methanocaldococcus jannaschii and its possible role in recycling the radical S-adenosylmethionine enzyme reaction product 5′-deoxyadenosine. J. Bacteriol. 196, 1064–1072. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (53).Casero RA, Stewart TM, and Pegg AE (2018) Polyamine metabolism and cancer: treatments, challenges and opportunities. Nat. Rev. Cancer 18, 681–695. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (54).Booker SJ, and Grove TL (2010) Mechanistic and functional versatility of radical SAM enzymes. F1000 Biol. Rep. 2, 52. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
