research communications\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoSTRUCTURAL BIOLOGY
COMMUNICATIONS
ISSN: 2053-230X

Biochemical and X-ray analyses of the players involved in the faRel2/aTfaRel2 toxin–antitoxin operon

crossmark logo

aCellular and Molecular Microbiology, Faculté des Sciences, Université Libre de Bruxelles (ULB), Boulevard du Triomphe, Building BC (1C4 203), 1050 Brussels, Belgium, bDepartment of Experimental Medical Science, Lund University, Lund, Sweden, cUniversity of Tartu Institute of Technology, Tartu, Estonia, and dScience for Life Laboratory, Lund, Sweden
*Correspondence e-mail: vasili.hauryliuk@med.lu.se, abel.garcia.pino@ulb.be

Edited by K. K. Kim, Sungkyunkwan University School of Medicine, Republic of Korea (Received 11 July 2023; accepted 19 August 2023; online 20 September 2023)

The aTfaRel2/faRel2 operon from Coprobacillus sp. D7 encodes a bicistronic type II toxin–antitoxin (TA) module. The FaRel2 toxin is a toxic small alarmone synthetase (toxSAS) that inhibits translation through the pyrophosphorylation of uncharged tRNAs at the 3′-CCA end. The toxin is neutralized by the antitoxin ATfaRel2 through the formation of an inactive TA complex. Here, the production, biophysical analysis and crystallization of ATfaRel2 and FaRel2 as well as of the ATfaRel2–FaRel2 complex are reported. ATfaRel2 is monomeric in solution. The antitoxin crystallized in space group P21212 with unit-cell parameters a = 53.3, b = 34.2, c = 37.6 Å, and the best crystal diffracted to a resolution of 1.24 Å. Crystals of FaRel2 in complex with APCPP, a non­hydrolysable ATP analogue, belonged to space group P21, with unit-cell parameters a = 31.5, b = 60.6, c = 177.2 Å, β = 90.6°, and diffracted to 2.6 Å resolution. The ATfaRel2–FaRel2Y128F complex forms a heterotetramer in solution composed of two toxins and two antitoxins. This complex crystallized in two space groups: F4132, with unit-cell parameters a = b = c = 227.1 Å, and P212121, with unit-cell parameters a = 51.7, b = 106.2, c = 135.1 Å. The crystals diffracted to 1.98 and 2.1 Å resolution, respectively.

1. Introduction

Bacterial toxin–antitoxin (TA) systems are encoded by (typically) bicistronic bacterial operons and contribute to phage defence and stress resistance as well as to the stabilization of plasmids and genomic islands (LeRoux & Laub, 2022[LeRoux, M. & Laub, M. T. (2022). Annu. Rev. Microbiol. 76, 21-43.]; Zhang et al., 2022[Zhang, T., Tamman, H., Coppieters 't Wallant, K., Kurata, T., LeRoux, M., Srikant, S., Brodiazhenko, T., Cepauskas, A., Talavera, A., Martens, C., Atkinson, G. C., Hauryliuk, V., Garcia-Pino, A. & Laub, M. T. (2022). Nature, 612, 132-140.]). Based on their nature (protein or RNA) and their mode of inhibition, TA systems have been classified into eight types (Jurėnas et al., 2022[Jurėnas, D., Fraikin, N., Goormaghtigh, F. & Van Melderen, L. (2022). Nat. Rev. Microbiol. 20, 335-350.]; Page & Peti, 2016[Page, R. & Peti, W. (2016). Nat. Chem. Biol. 12, 208-214.]). The best characterized TA modules belong to type II, where both components are proteins and the antitoxin neutralizes the toxin by direct interaction. TA toxins target essential cellular processes, replication, transcription and translation, commonly causing irreversible modification of the target and rapid growth arrest (Harms et al., 2018[Harms, A., Brodersen, D. E., Mitarai, N. & Gerdes, K. (2018). Mol. Cell, 70, 768-784.]; Jurėnas et al., 2017[Jurėnas, D., Garcia-Pino, A. & Van Melderen, L. (2017). Plasmid, 93, 30-35.], 2022[Jurėnas, D., Fraikin, N., Goormaghtigh, F. & Van Melderen, L. (2022). Nat. Rev. Microbiol. 20, 335-350.]). While toxin neutralization can be achieved via different mechanisms, in type II TA systems inhibition commonly occurs by the formation of stable toxin–antitoxin complexes, which are often engaged co-translationally via an intrinsically disordered region (IDR) of the antitoxin (Loris & Garcia-Pino, 2014[Loris, R. & Garcia-Pino, A. (2014). Chem. Rev. 114, 6933-6947.]; Page & Peti, 2016[Page, R. & Peti, W. (2016). Nat. Chem. Biol. 12, 208-214.]).

The expression of TA systems is often tightly autoregulated, with the antitoxin binding to inverted repeats in the promoter region to inhibit transcription of the operon (Garcia-Pino et al., 2016[Garcia-Pino, A., De Gieter, S., Talavera, A., De Greve, H., Efremov, R. G. & Loris, R. (2016). Nat. Chem. Biol. 12, 490-496.]; Talavera et al., 2019[Talavera, A., Tamman, H., Ainelo, A., Konijnenberg, A., Hadži, S., Sobott, F., Garcia-Pino, A., Hõrak, R. & Loris, R. (2019). Nat. Commun. 10, 972.]). This interaction is mediated by the dedicated DNA-binding domain of the antitoxin (Loris & Garcia-Pino, 2014[Loris, R. & Garcia-Pino, A. (2014). Chem. Rev. 114, 6933-6947.]). In this dual functionality of the antitoxins underlies the extreme promiscuity of TA modules observed across genomes, with both functional modules (the DNA-binding domain and the toxin-neutralization element) being exchangeable across different TA families (Jurėnas et al., 2022[Jurėnas, D., Fraikin, N., Goormaghtigh, F. & Van Melderen, L. (2022). Nat. Rev. Microbiol. 20, 335-350.]; Loris & Garcia-Pino, 2014[Loris, R. & Garcia-Pino, A. (2014). Chem. Rev. 114, 6933-6947.]).

The high degree of horizontal mobility of TA operons combined with the sheer diversity of neutralization mechanisms, often involving unstructured domains, fuels the high rates of evolution leading to the diversification of TA operons (Aakre et al., 2015[Aakre, C. D., Herrou, J., Phung, T. N., Perchuk, B. S., Crosson, S. & Laub, M. T. (2015). Cell, 163, 594-606.]). Many TA toxins are evolutionarily related to bacterial housekeeping enzymes. Notable examples include toxins from the Fic/Doc family (Garcia-Pino et al., 2008[Garcia-Pino, A., Christensen-Dalsgaard, M., Wyns, L., Yarmolinsky, M., Magnuson, R. D., Gerdes, K. & Loris, R. (2008). J. Biol. Chem. 283, 30821-30827.], 2014[Garcia-Pino, A., Zenkin, N. & Loris, R. (2014). Trends Biochem. Sci. 39, 121-129.]), toxic small alarmone synthetases (toxSAS), which are members of the RelA–SpoT homolog (RSH) protein family (Brown et al., 2016[Brown, A., Fernández, I. S., Gordiyenko, Y. & Ramakrishnan, V. (2016). Nature, 534, 277-280.]; Steinchen et al., 2015[Steinchen, W., Schuhmacher, J. S., Altegoer, F., Fage, C. D., Srinivasan, V., Linne, U., Marahiel, M. A. & Bange, G. (2015). Proc. Natl Acad. Sci. USA, 112, 13348-13353.]; Tamman et al., 2023[Tamman, H., Ernits, K., Roghanian, M., Ainelo, A., Julius, C., Perrier, A., Talavera, A., Ainelo, H., Dugauquier, R., Zedek, S., Thureau, A., Pérez, J., Lima-Mendez, G., Hallez, R., Atkinson, G. C., Hauryliuk, V. & Garcia-Pino, A. (2023). Nat. Chem. Biol. 19, 334-345.]; Zhang et al., 2022[Zhang, T., Tamman, H., Coppieters 't Wallant, K., Kurata, T., LeRoux, M., Srikant, S., Brodiazhenko, T., Cepauskas, A., Talavera, A., Martens, C., Atkinson, G. C., Hauryliuk, V., Garcia-Pino, A. & Laub, M. T. (2022). Nature, 612, 132-140.]), as well as bacterial acetyltransferases (Jurėnas et al., 2017[Jurėnas, D., Garcia-Pino, A. & Van Melderen, L. (2017). Plasmid, 93, 30-35.], 2019[Jurėnas, D., Van Melderen, L. & Garcia-Pino, A. (2019). Nat. Chem. Biol. 15, 285-294.]). toxSAS serve as bacterial stress response factors that produce a (pp)pGpp alarmone (Nanamiya et al., 2008[Nanamiya, H., Kasai, K., Nozawa, A., Yun, C. S., Narisawa, T., Murakami, K., Natori, Y., Kawamura, F. & Tozawa, Y. (2008). Mol. Microbiol. 67, 291-304.]; Geiger et al., 2014[Geiger, T., Kästle, B., Gratani, F. L., Goerke, C. & Wolz, C. (2014). J. Bacteriol. 196, 894-902.]). CapRelSJ46, a fused toxSAS TA, has recently been demonstrated to mediate bacterial defence against phages via abortive infection (Zhang et al., 2022[Zhang, T., Tamman, H., Coppieters 't Wallant, K., Kurata, T., LeRoux, M., Srikant, S., Brodiazhenko, T., Cepauskas, A., Talavera, A., Martens, C., Atkinson, G. C., Hauryliuk, V., Garcia-Pino, A. & Laub, M. T. (2022). Nature, 612, 132-140.]). Once activated, CapRelSJ46 (and the majority of known toxSAS) corrupts protein synthesis by pyrophosphorylating the 3′-OH of the adenine residue of 3′-CCA uncharged tRNA (Kurata et al., 2021[Kurata, T., Brodiazhenko, T., Alves Oliveira, S. R., Roghanian, M., Sakaguchi, Y., Turnbull, K. J., Bulvas, O., Takada, H., Tamman, H., Ainelo, A., Pohl, R., Rejman, D., Tenson, T., Suzuki, T., Garcia-Pino, A., Atkinson, G. C. & Hauryliuk, V. (2021). Mol. Cell, 81, 3160-3170.]). Other toxSAS synthesize the (pp)pApp alarmone, which results in depletion of the cellular ATP pool (Jimmy et al., 2020[Jimmy, S., Saha, C. K., Kurata, T., Stavropoulos, C., Oliveira, S. R. A., Koh, A., Cepauskas, A., Takada, H., Rejman, D., Tenson, T., Strahl, H., Garcia-Pino, A., Hauryliuk, V. & Atkinson, G. C. (2020). Proc. Natl Acad. Sci. USA, 117, 10500-10510.]). Both groups of TA toxins can be neutralized directly by type II antitoxins or by small alarmone hydrolase (SAH) antitoxins that remove the pyrophosphate group (Jimmy et al., 2020[Jimmy, S., Saha, C. K., Kurata, T., Stavropoulos, C., Oliveira, S. R. A., Koh, A., Cepauskas, A., Takada, H., Rejman, D., Tenson, T., Strahl, H., Garcia-Pino, A., Hauryliuk, V. & Atkinson, G. C. (2020). Proc. Natl Acad. Sci. USA, 117, 10500-10510.]).

The core of these enzymes is conserved and consists of an arrangement of β-strands and α-helices that fold together to form the active pocket. This results in a twisted β-sheet surrounded by α-helices, where the catalytic residues are situated to ensure the transfer of the pyrophosphate group of ATP to the corresponding substrate ATP, GDP, GTP or tRNA (Ahmad et al., 2019[Ahmad, S., Wang, B., Walker, M. D., Tran, H. R., Stogios, P. J., Savchenko, A., Grant, R. A., McArthur, A. G., Laub, M. T. & Whitney, J. C. (2019). Nature, 575, 674-678.]; Steinchen et al., 2015[Steinchen, W., Schuhmacher, J. S., Altegoer, F., Fage, C. D., Srinivasan, V., Linne, U., Marahiel, M. A. & Bange, G. (2015). Proc. Natl Acad. Sci. USA, 112, 13348-13353.]; Tamman et al., 2020[Tamman, H., Van Nerom, K., Takada, H., Vandenberk, N., Scholl, D., Polikanov, Y., Hofkens, J., Talavera, A., Hauryliuk, V., Hendrix, J. & Garcia-Pino, A. (2020). Nat. Chem. Biol. 16, 834-840.]; Zhang et al., 2022[Zhang, T., Tamman, H., Coppieters 't Wallant, K., Kurata, T., LeRoux, M., Srikant, S., Brodiazhenko, T., Cepauskas, A., Talavera, A., Martens, C., Atkinson, G. C., Hauryliuk, V., Garcia-Pino, A. & Laub, M. T. (2022). Nature, 612, 132-140.]).

The regulation of these toxSAS has only been structurally characterized for Tas1 (Ahmad et al., 2019[Ahmad, S., Wang, B., Walker, M. D., Tran, H. R., Stogios, P. J., Savchenko, A., Grant, R. A., McArthur, A. G., Laub, M. T. & Whitney, J. C. (2019). Nature, 575, 674-678.]) and CapRel SJ46. Tas1 [type VI secretion effector (p)ppApp synthetase 1] catalyses the synthesis of (pp)pApp, which leads to ATP depletion in the targeted cell. The immunity factor Tis1 binds to Tas1 and hinders its activity by obstructing access of the acceptor ATP nucleotide (Ahmad et al., 2019[Ahmad, S., Wang, B., Walker, M. D., Tran, H. R., Stogios, P. J., Savchenko, A., Grant, R. A., McArthur, A. G., Laub, M. T. & Whitney, J. C. (2019). Nature, 575, 674-678.]). In contrast, CapRelSJ46 features an autoinhibitory domain that, unlike Tis1, hinders the binding of the pyrophosphate donor ATP (Zhang et al., 2022[Zhang, T., Tamman, H., Coppieters 't Wallant, K., Kurata, T., LeRoux, M., Srikant, S., Brodiazhenko, T., Cepauskas, A., Talavera, A., Martens, C., Atkinson, G. C., Hauryliuk, V., Garcia-Pino, A. & Laub, M. T. (2022). Nature, 612, 132-140.]).

FaRel2 from Coprobacillus sp. D7 uses ATP to pyrophosphorylate the 3′-CCA end of uncharged tRNA (Kurata et al., 2021[Kurata, T., Brodiazhenko, T., Alves Oliveira, S. R., Roghanian, M., Sakaguchi, Y., Turnbull, K. J., Bulvas, O., Takada, H., Tamman, H., Ainelo, A., Pohl, R., Rejman, D., Tenson, T., Suzuki, T., Garcia-Pino, A., Atkinson, G. C. & Hauryliuk, V. (2021). Mol. Cell, 81, 3160-3170.]). Its toxicity is counteracted by its type II cognate antitoxin ATfaRel2 via the formation of a tight complex (Kurata et al., 2021[Kurata, T., Brodiazhenko, T., Alves Oliveira, S. R., Roghanian, M., Sakaguchi, Y., Turnbull, K. J., Bulvas, O., Takada, H., Tamman, H., Ainelo, A., Pohl, R., Rejman, D., Tenson, T., Suzuki, T., Garcia-Pino, A., Atkinson, G. C. & Hauryliuk, V. (2021). Mol. Cell, 81, 3160-3170.]). In this work, we describe the purification, crystallization and X-ray diffraction experiments of the antitoxin ATfaRel2, FaRel2–ATfaRel2 and the complex of FaRel2 with the ATP analogue APCPP. We expect that the resulting structures will shed light on the molecular basis of the toxicity and regulation mechanisms of the FaRel2–ATfaRel2 system and reveal which residues of FaRel2 are involved in the binding and hydrolysis of ATP. Furthermore, the structure of the FaRel2–ATfaRel2 complex will contribute to elucidating the mechanism of inhibition of the toxic effect of FaRel2.

2. Materials and methods

2.1. Protein expression and purification

2.1.1. Antitoxin ATfaRel2

ATfaRel2 was expressed from pET-24d_His1a-HisTEV-Coprobacillus sp. (VHp364), where the ATfaRel2 antitoxin has a 6×His tag followed by a TEV protease recognition site. Cultures were grown in LB medium supplemented with kanamycin (50 µg ml−1) at 37°C with aeration. Expression was induced with 0.5 mM isopropyl β-D-1-thiogalactopyranoside (IPTG) when the cells carrying the plasmid reached an OD600 nm of ∼0.5–0.8. The cells were harvested 16 h after induction by centrifugation and resuspended in buffer (25 mM HEPES pH 7.6, 1 M NaCl, 5 mM MgCl2, 1 mM TCEP, 0.0002% mellitic acid) supplied with cOmplete Protease-Inhibitor Cocktail (Roche). The resuspended cells were flash-frozen in liquid nitrogen and stored at −80°C prior to further applications.

The cell extracts were lysed using an Emulsiflex cell disruptor and the lysate was centrifuged to remove cell debris for 45 min at 25 000g. The supernatant was loaded onto a 1 ml HiTrap Ni–NTA column (Cytiva) coupled to an FPLC (ÄKTAexplorer) equilibrated with buffer A (25 mM HEPES pH 7.6, 1 M NaCl, 5 mM MgCl2, 1 mM TCEP, 0.0002% mellitic acid, 20 mM imidazole). The column was washed with a linear gradient of buffer B (25 mM HEPES pH 7.6, 1 M NaCl, 5 mM MgCl2, 1 mM TCEP, 0.0002% mellitic acid, 500 mM imidazole). The fractions containing His6-TEV-ATfaRel2 were concentrated using 3 kDa spin filters (Amicon) and loaded onto a Superdex 75 Increase 10/30 size-exclusion chromatography (SEC) column (Cytiva) equilibrated with 25 mM HEPES pH 7.6, 200 mM NaCl, 2 mM MgCl2, 0. 0002% mellitic acid, 1 mM TCEP. The purity of the protein was analysed by SDS–PAGE.

2.1.2. Generation of catalytically impaired FaRel2Y128F for structural biology

To counter the intrinsic toxicity of FaRel2, we generated a catalytically impaired version of the toxin with a Y128F substitution in the G-loop (FaRel2Y128F). The mutation in faRel2 was initially introduced into the pBAD33-FaRel2 toxSAS plasmid (Jimmy et al., 2020[Jimmy, S., Saha, C. K., Kurata, T., Stavropoulos, C., Oliveira, S. R. A., Koh, A., Cepauskas, A., Takada, H., Rejman, D., Tenson, T., Strahl, H., Garcia-Pino, A., Hauryliuk, V. & Atkinson, G. C. (2020). Proc. Natl Acad. Sci. USA, 117, 10500-10510.]) using the primers F-BmtI SAS Copro (GATCgctagcATGTACATCCTGGATAAGAT) and R-SAS Copro (gccgaagctTAAATTTTCTTGCAGTG). The PCR product was treated with DpnI to remove the template plasmid, purified on a PCR purification column (Sigma), phosphorylated and ligated by T4 ligase. The ligation mixture was transformed into Escherichia coli strain MC1061 by electroporation. The sequence of the resulting pBAD33-FaRel2Y128F_toxSAS plasmid was confirmed by sequencing (Eurofins Genomics).

The faRel2Y128F gene was amplified from pBAD33-FaRel2Y128F using Q5 High-Fidelity Polymerase and treated with DpnI. The faRel2Y128F gene was inserted into a modified pET-28b vector by restriction cloning with the enzymes BmtI and HindIII. The modified pET-28b already contained a six-His tag (6×His) as well as a SUMO cleavage site. The sequence of pET-28b-6×His-SUMO-faRel2Y128F (Table 1[link]) was confirmed by sequencing (Genewiz).

Table 1
Macromolecule-production information

Construct 6×His-TEV-ATfaRel2 His-SUMO-FaRel2Y128F
Source organism Coprobacillus sp. D7 Coprobacillus sp. D7
DNA source Synthetic Synthetic
Cloning vector pET-24d pET-28b
Expression vector pET-24d pET-28b
Expression host E. coli BL21 (DE3) E. coli BL21 (DE3)
Tag 6×His + TEV cleavage site: MKHHHHHHPMSDYDIPTTENLYFQG 6×His + SUMO cleavage site: MGSSHHHHHHGSDSEVNQEAKPEVKPEVKPETHINLKVSDGSSEIFFKIKKTTPLRRLMEAFAKRQGKEMDSLRFLYDGIRIQADQTPEDLDMEDNDIIEAHREQIGG
Complete amino-acid sequence of the construct produced MKHHHHHHPMSDYDIPTTENLYFQGAMCYIIAKRFKKSGCVALKAKRGKELADFATDLQKKLGYDIQIVAITRPTAYGEYEPYKFVNSFEEFSIEASRL MGSSHHHHHHGSDSEVNQEAKPEVKPEVKPETHINLKVSDGSSEIFFKIKKTTPLRRLMEAFAKRQGKEMDSLRFLYDGIRIQADQTPEDLDMEDNDIIEAHREQIGGASMYILDKIGLNIEILESLSYESKLGMSFKRTLSHFNKEEVLKEIELINNWYFSLEIIDDLPLDSRIKSVSSAKMKFERYYPNATYNRVFNDILGFRVICKSYDEVLELEKEDKIRVVDMSRGKSNDDGFRGIHVYYQRDNHHYPIEIQFNTYYDRQLNDWLHDKFYKRGYDSSCGQLLRKYYENGKIKSAEELEEVLEDVLYHCKKI
2.1.3. Preparation of FaRel2Y128F

Cultures of E. coli strain BL21(DE3) transformed with pET-28b-6×His-SUMO-faRel2Y128F were grown in LB medium supplemented with kanamycin (50 µg ml−1) at 37°C with aeration. When the cell culture reached an OD600 nm of ∼0.5–0.8, expression was induced with 0.5 mM IPTG. After 3 h of induction, the cells were harvested by centrifugation and resuspended in buffer (100 mM K2HPO4/KH2PO4 pH 8.0, 500 mM NaCl, 500 mM KCl, 1 mM TCEP, 1% glycerol, 20 mM imidazole, 0.0002% mellitic acid) supplied with cOmplete Protease-Inhibitor Cocktail (Roche). The cells were flash-frozen in liquid nitrogen and stored at –80°C prior to further applications.

The purification was carried out similarly to that of ATfaRel2, except for the buffers: buffer A consisted of 100 mM K2HPO4/KH2PO4 pH 8.0, 500 mM NaCl, 500 mM KCl, 1 mM TCEP, 20 mM imidazole and buffer B consisted of 100 mM K2HPO4/KH2PO4 pH 8.0, 500 mM NaCl, 500 mM KCl, 1 mM TCEP, 500 mM imidazole. The fractions containing His6-SUMO-FaRel2 were pooled together, concentrated and loaded onto a Superdex 75 Increase 10/30 SEC column (Cytiva) equilibrated in buffer GF (25 mM HEPES pH 7.6, 1 M NaCl, 2 mM MgCl2, 1 mM TCEP).

The His6-SUMO tag was cleaved from His6-SUMO-FaRel2Y128F by adding UlpI protease in a 1:100 molar ratio and incubating overnight at 4°C. To separate FaRel2Y128F from UlpI protease and free His6-SUMO, the mixture was applied onto a gravity-flow TALON column previously equilibrated with buffer GF. FaRel2Y128F was recovered after washing the column with buffer C (25 mM HEPES pH 7.6, 1 M NaCl, 2 mM MgCl2, 1 mM TCEP, 20 mM imidazole) for elution. The quality of cleavage was then analysed by SDS–PAGE and complete removal of the tag was confirmed by a Western blot using an anti-6×His-Tag Monoclonal Antibody (Thermo Fisher, catalogue No. MA1-21315-D680). The elution fractions containing FaRel2Y128F were then concentrated using 10 kDa cutoff spin filters and loaded onto a Superdex 75 10/30 SEC column (Cytiva) equilibrated with buffer GF.

2.1.4. Preparation of the ATfaRel2–FaRel2Y128F complex

The ATfaRel2–FaRel2Y128F complex was prepared by adding ATfaRel2 to FaRel2Y128F in a 1.2:1 molar ratio. The mixture was injected onto a Superdex 75 16/60 column (Cytiva) and the peak corresponding to the ATfaRel2Y128F–FaRel2Y128F complex was collected and analysed by SDS–PAGE.

2.2. Crystallization

Crystallization conditions were screened at 20 and 4°C by the sitting-drop vapour-diffusion method. The drops were set up in Swissci (MRC) 96-well 2-drop UVP sitting-drop plates. The drops consisted of 0.1 µl protein solution plus 0.1 µl reservoir solution and were equilibrated against 80 µl reservoir solution. Crystallization conditions were tested with several commercially available screens: Crystal Screen, Crystal Screen 2 (Hampton Research), HELIX, PACT premier (Molecular Dimensions), LMB Crystallization Screen (Molecular Dimensions), SG1 (Molecular Dimensions) and NeXtal (Qiagen). The protein concentration was determined from the absorbance at 280 nm and was corrected using the theoretical extinction coefficients estimated by ProtParam (Gasteiger et al., 2003[Gasteiger, E., Gattiker, A., Hoogland, C., Ivanyi, I., Appel, R. D. & Bairoch, A. (2003). Nucleic Acids Res. 31, 3784-3788.]).

For the co-crystallization of FaRel2Y128F with the ATP analogue APCPP (Jena Bioscience catalogue No. NU-421), fresh FaRel2Y128F concentrated to 19 mg ml−1 was mixed with APCPP at 100 mM and incubated for 10 min at room temperature. ATfaRel2 was used at 12 mg ml−1, while the ATfaRel2–FaRel2Y128 F complex was concentrated to 10 mg ml−1 for crystallization.

2.3. X-ray data collection and analysis

Prior to data collection, the crystals were transferred to a suitable cryoprotectant solution (Table 2[link]) and flash-cooled in liquid nitrogen. Data were collected on the PROXIMA-1 and PROXIMA-2 beamlines at the SOLEIL synchrotron facility, Gif-sur-Yvette, France (ATfaRel2 and the FaRel2Y128F–APCPP and ATfaRel2–FaRel2Y128F complexes) and at the Diamond Light Source (DLS) synchrotron, Oxfordshire, UK (ATfaRel2–FaRel2Y128F complex) and were recorded on EIGER detectors. All data sets were indexed, integrated and scaled using autoPROC (Vonrhein et al., 2011[Vonrhein, C., Flensburg, C., Keller, P., Sharff, A., Smart, O., Paciorek, W., Womack, T. & Bricogne, G. (2011). Acta Cryst. D67, 293-302.]). AutoPROC also makes use of the programs XDS (Kabsch, 2010[Kabsch, W. (2010). Acta Cryst. D66, 125-132.]), POINTLESS (Evans, 2007[Evans, P. R. (2007). Acta Cryst. D63, 58-61.]), AIMLESS (Evans & Murshudov, 2013[Evans, P. R. & Murshudov, G. N. (2013). Acta Cryst. D69, 1204-1214.]) and others from the CCP4 suite (Agirre et al., 2023[Agirre, J., Atanasova, M., Bagdonas, H., Ballard, C. B., Baslé, A., Beilsten-Edmands, J., Borges, R. J., Brown, D. G., Burgos-Mármol, J. J., Berrisford, J. M., Bond, P. S., Caballero, I., Catapano, L., Chojnowski, G., Cook, A. G., Cowtan, K. D., Croll, T. I., Debreczeni, J. É., Devenish, N. E., Dodson, E. J., Drevon, T. R., Emsley, P., Evans, G., Evans, P. R., Fando, M., Foadi, J., Fuentes-Montero, L., Garman, E. F., Gerstel, M., Gildea, R. J., Hatti, K., Hekkelman, M. L., Heuser, P., Hoh, S. W., Hough, M. A., Jenkins, H. T., Jiménez, E., Joosten, R. P., Keegan, R. M., Keep, N., Krissinel, E. B., Kolenko, P., Kovalevskiy, O., Lamzin, V. S., Lawson, D. M., Lebedev, A. A., Leslie, A. G. W., Lohkamp, B., Long, F., Malý, M., McCoy, A. J., McNicholas, S. J., Medina, A., Millán, C., Murray, J. W., Murshudov, G. N., Nicholls, R. A., Noble, M. E. M., Oeffner, R., Pannu, N. S., Parkhurst, J. M., Pearce, N., Pereira, J., Perrakis, A., Powell, H. R., Read, R. J., Rigden, D. J., Rochira, W., Sammito, M., Sánchez Rodríguez, F., Sheldrick, G. M., Shelley, K. L., Simkovic, F., Simpkin, A. J., Skubak, P., Sobolev, E., Steiner, R. A., Stevenson, K., Tews, I., Thomas, J. M. H., Thorn, A., Valls, J. T., Uski, V., Usón, I., Vagin, A., Velankar, S., Vollmar, M., Walden, H., Waterman, D., Wilson, K. S., Winn, M. D., Winter, G., Wojdyr, M. & Yamashita, K. (2023). Acta Cryst. D79, 449-461.]).

Table 2
Crystallization conditions

Enzyme Protein solution Reservoir solution Temperature (K) Cryoprotectant (collected data sets)
(HisTEV)ATfaRel2 11.65 mg ml−1 in 25 mM HEPES pH 7.6, 300 mM NaCl, 1 mM TCEP, 2 mM MgCl2 LMB C10: 15%(w/v) PEG 2000 MME, 0.1 M bis-Tris propane pH 6.9 277 20% glycerol
(HisTEV)ATfaRel2 11.65 mg ml−1 in 25 mM HEPES pH 7.6, 300 mM NaCl, 1 mM TCEP, 2 mM MgCl2 LMB F5: 29%(w/v) PEG 4000, 0.1 M sodium citrate pH 6.5, 0.1 M magnesium acetate tetrahydrate, 0.1 M ammonium sulfate 277 20% glycerol
(HisTEV)ATfaRel2 11.65 mg ml−1 in 25 mM HEPES pH 7.6, 300 mM NaCl, 1 mM TCEP, 2 mM MgCl2 LMB D6: 21%(w/v) PEG 3350, 0.1 M MES 6.0, 0.15 M sodium chloride 277 20% glycerol
FaRel2Y128F + APCPP 19 mg ml−1 + 100 mM APCPP in 25 mM HEPES pH 7.6, 1 M NaCl, 1 mM TCEP, 2 mM MgCl2 Helix G9: 0.05 M MES pH 5.5, 12%(w/v) PEG 2000 277 20% glycerol
FaRel2Y128F + APCPP 19 mg ml−1 + 100 mM APCPP in 25 mM HEPES pH 7.6, 1 M NaCl, 1 mM TCEP, 2 mM MgCl2 Crystal Screen 1&2 E4: 35%(v/v) 1,4-dioxane 277 20% glycerol
FaRel2Y128F + APCPP 19 mg ml−1 + 100 mM APCPP in 25 mM HEPES pH 7.6, 1 M NaCl, 1 mM TCEP, 2 mM MgCl2 Nextal H10: 0.05 M magnesium chloride, 0.1 M MES pH 6.5, 10%(v/v) 2-propanol, 5%(w/v) PEG 4000 293 20% glycerol
FaRel2Y128F+ (HisTEV)ATfaRel2 10 mg ml−1 in 25 mM HEPES pH 7.6, 1 M NaCl, 1 mM TCEP, 2 mM MgCl2 Crystal Screen 1&2 E4: 35%(v/v) 1,4-dioxane 277 20% glycerol
FaRel2Y128F+ (HisTEV)ATfaRel2 10 mg ml−1 in 25 mM HEPES pH 7.6, 1 M NaCl, 1 mM TCEP, 2 mM MgCl2 LMB G2: 3.5%(w/v) PEG 6000, 0.1 M bis-Tris propane pH 7.1, 0.1 M potassium chloride 277 40% MPD
FaRel2Y128F+ (HisTEV)ATfaRel2 10 mg ml−1 in 25 mM HEPES pH 7.6, 1 M NaCl, 1 mM TCEP, 2 mM MgCl2 SG1 A5: 0.2 M sodium citrate tribasic dihydrate, 20%(w/v) PEG 3350 277  

2.4. Isothermal titration calorimetry

The 150 bp DNA fragment used for ITC is listed in Table 3[link]. The duplex was reconstituted from synthetic single-stranded oligonucleotides (Sigma) that were annealed by heating to 85 C° and slowly cooling to room temperature.

Table 3
farel2 promoter sequences used for ITC

Promoter_forward (5′–3′) ATATTATTTTGGAATAGAAGAATAAAAAGCAAGAAAAGAGGTATTAAATATATGATAAAAAAATATCTGTAATAAGTTTGATTATCAGCAGTTAAGAGATAAACTATGTGAGGTGAATTATTTATGTATATACTTGATAAAATAGGACTT
Promoter_reverse (5′–3′) AAGTCCTATTTTATCAAGTATATACATAAATAATTCACCTCACATAGTTTATCTCTTAACTGCTGATAATCAAACTTATTACAGATATTTTTTTATCATATATTTAATACCTCTTTTCTTGCTTTTTATTCTTCTATTCCAAAATAATAT

The ITC measurements were carried out at 20°C in an Affinity ITC calorimeter (TA Instruments). Prior to the experiments, ATfaRel2 and the 150 bp operator fragment were dialyzed, in the same reservoir, against 25 mM HEPES pH 7.6, 300 mM NaCl, 2 mM MgCl2, 1 mM TCEP. The concentration of ATfaRel2 in the 177 µl cell was 10 µM and the 2 µl injections contained 115 µM of the faRel2/aTfaRel2 DNA operator. The complete data processing and analysis was performed with NanoAnalyze (TA Instruments).

2.5. In vivo toxicity neutralization assays

Three vectors were generated for this study. Two pBAD33 vectors with arabinose-inducible promoters were generated: one encoding FaRel2 (designated pBAD33_FaRel2) and the other encoding the FaRel2Y128F mutant (designated pBAD33_FaRel2Y128F). Additionally, a pKKD vector with an IPTG-inducible promoter was created to express ATfaRel2 (designated pKKD_ATfaRel2).

E. coli DJ624Δara cells were subjected to individual transformations with each of the vectors. Moreover, co-transformations were performed by introducing pKKD_ATfaRel2 along with either pBAD33_FaRel2 or pBAD33_FaRel2Y128F into the cells. pBAD33 vectors provide resistance to chloramphenicol, while pKKD vectors provide resistance to ampicillin.

The cell cultures were grown overnight at 37°C in liquid LB medium supplemented with the appropriate antibiotics. Cultures were diluted serially (tenfold) and 10 µl of each dilution were spotted into LB agar plates supplied with the required antibiotics and either 1% glucose (repressing conditions) or 0.2% arabinose plus 1 mM IPTG (inducing conditions). Plates were scored after overnight incubation at 37°C.

3. Results

3.1. Production of ATfaRel2, FaRel2Y128F and the ATfaRel2–FaRel2Y128F complex

Due to their extreme toxicity, TA toxins are notoriously difficult to produce recombinantly (Sterckx et al., 2015[Sterckx, Y. G., De Gieter, S., Zorzini, V., Hadži, S., Haesaerts, S., Loris, R. & Garcia-Pino, A. (2015). Protein Expr. Purif. 108, 30-40.]). This is particularly true of toxSAS such as FaRel2 (Figs. 1[link]a and 1[link]b), FaRel and Tas1 (Jimmy et al., 2020[Jimmy, S., Saha, C. K., Kurata, T., Stavropoulos, C., Oliveira, S. R. A., Koh, A., Cepauskas, A., Takada, H., Rejman, D., Tenson, T., Strahl, H., Garcia-Pino, A., Hauryliuk, V. & Atkinson, G. C. (2020). Proc. Natl Acad. Sci. USA, 117, 10500-10510.]; Ahmad et al., 2019[Ahmad, S., Wang, B., Walker, M. D., Tran, H. R., Stogios, P. J., Savchenko, A., Grant, R. A., McArthur, A. G., Laub, M. T. & Whitney, J. C. (2019). Nature, 575, 674-678.]). A common strategy to overcome this challenge is the use of catalytically impaired versions of the toxins that have limited enzymatic activity but still retain structural integrity (Garcia-Pino et al., 2016[Garcia-Pino, A., De Gieter, S., Talavera, A., De Greve, H., Efremov, R. G. & Loris, R. (2016). Nat. Chem. Biol. 12, 490-496.]; Garcia-Rodriguez et al., 2021[Garcia-Rodriguez, G., Charlier, D., Wilmaerts, D., Michiels, J. & Loris, R. (2021). Nucleic Acids Res. 49, 7164-7178.]; Jurėnas et al., 2019[Jurėnas, D., Van Melderen, L. & Garcia-Pino, A. (2019). Nat. Chem. Biol. 15, 285-294.]; Talavera et al., 2018[Talavera, A., Hendrix, J., Versees, W., Jurenas, D., Van Nerom, K., Vandenberk, N., Singh, R. K., Konijnenberg, A., De Gieter, S., Castro-Roa, D., Barth, A., De Greve, H., Sobott, F., Hofkens, J., Zenkin, N., Loris, R. & Garcia-Pino, A. (2018). Sci. Adv. 4, eaap9714.]). The substitution of the conserved tyrosine of the G-loop of SYNTH and toxSYNTH domains (i.e. the loop involved in the coordination of pyrophosphate acceptors; Steinchen et al., 2018[Steinchen, W., Vogt, M. S., Altegoer, F., Giammarinaro, P. I., Horvatek, P., Wolz, C. & Bange, G. (2018). Sci. Rep. 8, 2195.]) by alanine residues leads to complete inactivation of the toxSAS CapRel and FaRel2 as well as long RSHs (Kurata et al., 2021[Kurata, T., Brodiazhenko, T., Alves Oliveira, S. R., Roghanian, M., Sakaguchi, Y., Turnbull, K. J., Bulvas, O., Takada, H., Tamman, H., Ainelo, A., Pohl, R., Rejman, D., Tenson, T., Suzuki, T., Garcia-Pino, A., Atkinson, G. C. & Hauryliuk, V. (2021). Mol. Cell, 81, 3160-3170.]; Tamman et al., 2020[Tamman, H., Van Nerom, K., Takada, H., Vandenberk, N., Scholl, D., Polikanov, Y., Hofkens, J., Talavera, A., Hauryliuk, V., Hendrix, J. & Garcia-Pino, A. (2020). Nat. Chem. Biol. 16, 834-840.]; Zhang et al., 2022[Zhang, T., Tamman, H., Coppieters 't Wallant, K., Kurata, T., LeRoux, M., Srikant, S., Brodiazhenko, T., Cepauskas, A., Talavera, A., Martens, C., Atkinson, G. C., Hauryliuk, V., Garcia-Pino, A. & Laub, M. T. (2022). Nature, 612, 132-140.]). This is likely due to the loss of coordination of the purine base of the acceptor nucleotide. Therefore, to study the toxin FaRel2 we decided to use a more conservative substitution and mutated the aforementioned tyrosine (residue 128) to phenylalanine (FaRel2Y128F). We hypothesized that with this substitution, FaRel2Y128F would become catalytically impaired but would still be able to bind uncharged tRNA. Indeed, in vivo toxicity neutralization assays show that while FaRel2Y128F remains toxic to E. coli, it is dramatically less potent than wild-type FaRel2, and its toxicity is counteracted by the co-expression of the ATfaRel2 antitoxin (Figs. 1[link]c and 1[link]d). Thus, this strategy provided us with a system to produce biologically active FaRel2Y128F in sufficient amounts for structural biology studies that could be used to investigate both toxin neutralization and substrate binding.

[Figure 1]
Figure 1
Microbiological and biochemical characterization of the proteins from the faRel2/aTfaRel2 TA operon. (ad) In vivo toxicity neutralization assays were used to assess the effects of substitutions on FaRel2 toxicity. Serial dilutions of E. coli strains were plated on solid LB medium and scored after 16 h at 37°C uninduced (grey) and under induction conditions (blue). (e) Coomassie-stained SDS–PAGE of purified ATfaRel2 (lane 2; molecular-weight markers are shown in lane 1). (f) Coomassie-stained SDS–PAGE of the ATfaRel2–FaRel2Y128F complex (lane 2) and FaRel2Y128F (lane 3); molecular-weight markers are shown in lane 1. The migration of the band corresponding to ATfaRel2 is highlighted by a solid black triangle and that for FaRel2Y128F by an open triangle. (g) Analytical SEC of ATfaRel2, FaRel2Y128F and the ATfaRel2–FaRel2Y128F complex on a Superdex 75 Increase column in 25 mM HEPES, 300 mM NaCl, 1 mM TCEP, 2 mM MgCl2. (h) ITC titration to monitor the interaction of ATfaRel2 with its 150 bp operator region.

ATfaRel2 and FaRel2Y128F were both expressed in E. coli BL21(DE3) cells using pET-24d and pET-28b expression vectors. Both proteins have a His6 tag at the N-terminus followed by a protease cleavage site (Figs. 1[link]e and 1[link]f). For ATfaRel2, the cleavage site is specific for TEV protease, while FaRel2Y128F has a SUMO tag specific to UlpI (His6-SUMO-FaRel2Y128F). In both cases, a two-step purification process was employed. It includes an initial step of affinity chromatography using a Ni–NTA column followed by a second step of size-exclusion chromatography (SEC). Collectively, these steps rendered homogeneous samples. His6-SUMO-Farel2Y128F was further subjected to removal of the 6×His-SUMO tag by incubation with UlpI, providing tagless FaRel2Y128F. The ATfaRel2–FaRel2Y128F complex was obtained by mixing both proteins followed by SEC to separate the complex from the excess of antitoxin and was confirmed by SDS–PAGE. All preparations were highly pure as judged by SDS–PAGE (Figs. 1[link]e and 1[link]f).

3.2. Characterization and crystallization of ATfaRel2

We used analytical SEC to determine the oligomeric state of ATfaRel2 (Fig. 1[link]g). ATfaRel2 eluted at 14.2 ml, which corresponds to an estimated molecular weight of 11.5 kDa consistent with a monomeric species in solution.

In most type II TA operons, transcription is autoregulated via the direct binding of the antitoxin to an operator region in the operon promoter (Hayes & Kędzierska, 2014[Hayes, F. & Kędzierska, B. (2014). Toxins, 6, 337-358.]). To assess whether ATfaRel2 could interact with its own promoter, we monitored the potential interaction of the antitoxin with a 150 bp DNA fragment derived from the upstream region of the operon by isothermal titration calorimetry (ITC). Analysis of the binding isotherm confirmed the interaction of ATfaRel2 with DNA with an affinity of 484 nM (Fig. 1[link]h).

To screen for crystallization conditions for ATfaRel2, the protein was concentrated to 12 mg ml−1 in 25 mM HEPES pH 7.6, 200 mM NaCl, 2 mM MgCl2, 1 mM TCEP. Crystallization hits were observed in several conditions from the LMB screen (Molecular Dimensions; Table 2[link]). These crystals were directly harvested, soaked in a suitable cryoprotectant solution and vitrified in liquid nitrogen prior to data collection (Table 2[link]). The crystals grown in 15% PEG 2000 MME and 0.1 M bis-Tris propane pH 6.9 (Figs. 2[link]a–2[link]c) diffracted to 1.2 Å resolution (Tables 2[link] and 4[link]). The data collected on the PROXIMA-1 beamline of the SOLEIL synchrotron were indexed as orthorhombic. The systematic absences indicated space group P21212, with unit-cell parameters a = 53.2, b = 34.2, c = 37.6 Å. Matthews calculation (Kantardjieff & Rupp, 2003[Kantardjieff, K. A. & Rupp, B. (2003). Protein Sci. 12, 1865-1871.]; Matthews, 1968[Matthews, B. W. (1968). J. Mol. Biol. 33, 491-497.]) suggested that there is only one molecule in the asymmetric unit, with a low solvent content of 17.4%, VM = 1.49 Å3 Da−1 and probability 1, in agreement with the oligomeric state determined by SEC.

Table 4
Data collection and processing

  ATfaRel2 ATfaRel2–FaRel2Y128F ATfaRel2–FaRel2Y128F FaRel2Y128F–APCPP
Diffraction source SOLEIL SOLEIL DLS SOLEIL
Beamline PROXIMA-1 PROXIMA-2 I24 PROXIMA-2
Wavelength (Å) 0.9801 1.0000 0.9795 1.0000
Temperature (K) 90 90 90 90
Detector EIGER EIGER EIGER EIGER
Crystal-to-detector distance (mm) 255.730 171.55 289.30 322.94
Rotation range per image (°) 0.1 0.1 0.1 0.1
Total rotation range (°) 720 360 360 360
Space group P21212 F4132 P212121 P21
a, b, c (Å) 53.3, 34.3, 37.6 227.1, 227.1, 227.1 51.7, 106.2, 135.1 31.5, 60.6, 177.2
α, β, γ (°) 90.0, 90.0, 90.0 90.0, 90.0, 90.0 90.0 90.0, 90.0 90.0, 90.6, 90.0
Mosaicity (°) 0.2 0.2 0.2 0.2
Resolution range (Å) 30.52–1.24 (1.37–1.24) 80.3–1.98 (2.05–1.98) 80.4–2.14 (3.36–2.14) 59.10–2.62 (2.83–2.62)
Total No. of reflections 191338 (5867) 2656572 (135051) 1074368 (52414) 220686 (8909)
No. of unique reflections 15171 (759) 33494 (1670) 26913 (1346) 16332 (818)
Completeness (%) 92.6 (46.5) 95.3 (52.1) 94.1 (64.3) 90.9 (40.6)
Mutliplicity 12.6 (7.7) 79.3 (80.9) 39.9 (38.9) 13.5 (10.9)
I/σ(I)〉 13.0 (1.3) 14.6 (1.3) 11.3 (1.5) 7.3 (1.4)
Rmerge 0.087 (1.205) 0.591 (6.727) 0.33 (4.77) 0.446 (2.730)
Rmeas 0.091 (1.289) 0.594 (6.769) 0.33 (4.8) 0.464 (2.864)
CC1/2 0.998 (0.711) 0.997 (0.477) 0.99 (0.66) 0.985 (0.332)
Overall B factor from Wilson plot (Å2) 17.18 27.51 40.97 67.37
[Figure 2]
Figure 2
Representative crystals of ATfaRel2 (a, b, c), ATfaRel2 in complex with FaRel2Y128F, space groups P212121 (d) and F4132 (e, f), and the FaRel2Y128F–APCPP complex (g, h, i).

A BLAST search against the Protein Data Bank (PDB) did not return relevant hits that could be used as search models for molecular replacement. Given the high resolution of the diffraction data and the relatively small size of the protein, we use ARCIMBOLDO_LITE (Sammito et al., 2015[Sammito, M., Millán, C., Frieske, D., Rodríguez-Freire, E., Borges, R. J. & Usón, I. (2015). Acta Cryst. D71, 1921-1930.]) to directly obtain the initial phases. ARCIMBOLDO_LITE performs ab initio phasing by placing and evaluating single polyalanine α-helices with Phaser (McCoy et al., 2007[McCoy, A. J., Grosse-Kunstleve, R. W., Adams, P. D., Winn, M. D., Storoni, L. C. & Read, R. J. (2007). J. Appl. Cryst. 40, 658-674.]) and performing density modification and model extension with SHELXE (Thorn, 2017[Thorn, A. (2017). Methods Mol. Biol. 1607, 357-376.]; Usón & Sheldrick, 2018[Usón, I. & Sheldrick, G. M. (2018). Acta Cryst. D74, 106-116.]). The initial search was performed using the default parameters of the program, searching for three generic α-helices each of 16 residues. After density modification, the final solution contained 75 of 99 residues traced into different stretches of alanine peptides with a final Phaser Z-score of 9.7, which indicates a definite solution (Fig. 3[link]a).

[Figure 3]
Figure 3
Initial electron-density maps obtained after phasing of ATfaRel2 (a), the ATfaRel2–FaRel2Y128F complex in space groups P212121 (b) and F4132 (c), and the FaRel2Y128F–APCPP complex (d).

3.3. Characterization and crystallization of the ATfaRel2–FaRel2Y128F complex

To reconstitute the toxin–antitoxin complex in vitro, FaRel2Y128F and ATfaRel2 were mixed in a 1:1.2 molar ratio and injected onto an analytical SEC column to separate the complex from excess ATfaRel2. The ATfaRel2–FaRel2Y128F complex eluted at 10.5 ml, which corresponds to an estimated molecular weight of 60 kDa (Fig. 1[link]g). Therefore, given the molecular weights of the individual proteins, the complex is most likely to be an (ATfaRel2)2–(FaRel2Y128F)2 heterotetramer in solution.

Crystals of the complex grew at 4°C in four different conditions (Figs. 2[link]g–2[link]i and Table 2[link]). Crystals grown in a condition consisting of 35%(v/v) 1,4-dioxane as the precipiting agent diffracted to 2.1 Å resolution (Table 4[link]), while crystals that grew in 1 M potassium phosphate monobasic, 3%(v/v) 2-propanol, 0.1 M sodium cacodylate pH 6.5 diffracted to 1.9 Å resolution (Table 4[link]). For the former, diffraction data were collected at DLS. The data were indexed in space group P212121, with unit-cell parameters a = 51.7, b = 106.6, c = 135.1 Å. Matthews calculation strongly supported the presence of an (ATfaRel2)2–(FaRel2Y128F)2 heterotetramer in the asymmetric unit (VM = 2.8 Å3 Da−1, with 52% solvent content and probability 0.72). The crystals from the second condition belonged to the cubic space group F4132, with unit-cell parameters a = b = c = 227.1 Å. In this case Matthews analysis could not clearly discern whether the asymmetric unit was composed of one or two ATfaRel2–FaRel2Y128F complexes. For one complex in the asymmetric unit VM = 3.67 Å3 Da−1 with 66% solvent content and probability 0.48, while for two complexes VM = 1.84 Å3 Da−1 with 33% solvent content and probability 0.52.

To further investigate the unit-cell content of the cubic crystals, we used the partial solution of ATfaRel2 as a search model for molecular replacement with Phaser. The MR search could detect only one ATfaRel2 molecule in the asymmetric unit (TFZ = 28.1, PAK = 0, LLG = 457). On the other hand, for the orthorhombic crystals of the complex preliminary MR analysis confirmed the presence of two molecules in the asymmetric unit (TFZ = 15.5 and LLG = 198). These results suggest that ATfaRel2 and FaRel2 are likely to form a symmetrical heterotetramer that crystallizes in space groups P212121 and F4132 (in the latter the full complex is likely to be generated by crystallographic symmetry). Thereafter, we used the partial solution found in space group F4132 as input for MR-Rosetta (Terwilliger et al., 2012[Terwilliger, T. C., DiMaio, F., Read, R. J., Baker, D., Bunkóczi, G., Adams, P. D., Grosse-Kunstleve, R. W., Afonine, P. V. & Echols, N. (2012). J. Struct. Funct. Genomics, 13, 81-90.]). The program successfully built 216 full residues plus 18 without side chains of a total of 289 (FaRel2Y128F + ATfaRel2; Figs. 3[link]b and 3[link]c).

3.4. Characterization and crystallization of FaRel2Y128F–APCPP complex

FaRel2 hydrolyses the α- and β-phosphate bond of ATP and transfers the βγ pyrophosphate to the 3′-CCA end of uncharged tRNAs (Kurata et al., 2021[Kurata, T., Brodiazhenko, T., Alves Oliveira, S. R., Roghanian, M., Sakaguchi, Y., Turnbull, K. J., Bulvas, O., Takada, H., Tamman, H., Ainelo, A., Pohl, R., Rejman, D., Tenson, T., Suzuki, T., Garcia-Pino, A., Atkinson, G. C. & Hauryliuk, V. (2021). Mol. Cell, 81, 3160-3170.]). To study the inter­action between the toxin and the nucleotide substrate, and to avoid its hydrolysis during the structural experiments, we used adenosine-5′-[(α,β)-methyleno]triphosphate (APCPP), a nonhydrolysable analogue of ATP. The FaRel2Y128F–APCPP complex was formed by mixing FaRel2Y128F at 19 mg ml−1 (562 µM) with 100 mM APCPP and was screened for crystallization conditions. Crystals grew in a condition consisting of 50 mM MES pH 5.5 plus 12%(w/v) PEG 2000 and were tested for diffraction on the PROXIMA-2 beamline at the SOLEIL synchrotron (Figs. 2[link]d–2[link]f). The best crystal diffracted to 2.6 Å resolution and belonged to space group P21, with unit-cell parameters a = 31.5, b = 60.5, c = 177.2 Å, β = 90.6° (Table 4[link]). Analysis of the unit-cell content (Matthews coefficient VM = 2.28 Å3 Da−1, with a solvent content of 46% and probability 0.86) suggests the presence of three molecules in the asymmetric unit. Using the partial structure of FaRel2Y128F obtained from the ATfaRel2–FaRelY128F complex as the search model, Phaser found a solution with final TFZ = 34.6, LLG = 1509 and PAK = 0 (Fig. 3[link]d).

4. Conclusions

The structural basis of the neutralization of toxSAS by their cognate antitoxins is poorly understood. In particular, it is unclear how homologous antitoxin domains can neutralize similar toxins in cis (in the case of fused toxSAS TAs such as CapRelSJ46; Zhang et al., 2022[Zhang, T., Tamman, H., Coppieters 't Wallant, K., Kurata, T., LeRoux, M., Srikant, S., Brodiazhenko, T., Cepauskas, A., Talavera, A., Martens, C., Atkinson, G. C., Hauryliuk, V., Garcia-Pino, A. & Laub, M. T. (2022). Nature, 612, 132-140.]) and in trans (in the case of the majority of toxSAS TAs). To overcome the high toxicity of FaRel2 from Coprobacillus sp. D7, we used a catalytically impaired version of the toxin, FaRel2Y128F. The production of FaRel2Y128F enabled studies of the neutralization of FaRel2 by the ATfaRel2 antitoxin and of how FaRel2 interacts with ATP. Our crystal structures advance the understanding of the function and regulation of the catalytic activity of toxSAS enzymes.

Footnotes

These authors contributed equally to this work.

Acknowledgements

The authors acknowledge the use of beamlines PROXIMA-1 and PROXIMA-2A and SWING at the SOLEIL synchrotron, Gif-sur-Yvette, France and I24 at Diamond Light Source, UK.

Funding information

This work was supported by the Fonds National de Recherche Scientifique (FNRS CDR J.0068.19 and J.0065.23F; FNRS-EQP UN.025.19; FNRS PDR T.0066.18 and PDR T.0090.22 to AGP); ERC (CoG DiStRes, No. 864311 to AGP); Fonds Jean Brachet and the Fondation Van Buuren (AGP). This work was also supported by a Swedish Research Council (Vetenskapsrådet) grant (2021-01146 to VH), the Crafoord Foundation (project grant No. 20220562 to VH), the Estonian Research Council (PRG335 to VH), Cancerfonden (20 0872 Pj to VH) and a project grant from the Knut and Alice Wallenberg Foundation (2020-0037 to VH). AC was supported by the Fund for Research in Industry and Agronomy (FRIA FC31211 to AC).

References

First citationAakre, C. D., Herrou, J., Phung, T. N., Perchuk, B. S., Crosson, S. & Laub, M. T. (2015). Cell, 163, 594–606.  Web of Science CrossRef CAS PubMed Google Scholar
First citationAgirre, J., Atanasova, M., Bagdonas, H., Ballard, C. B., Baslé, A., Beilsten-Edmands, J., Borges, R. J., Brown, D. G., Burgos-Mármol, J. J., Berrisford, J. M., Bond, P. S., Caballero, I., Catapano, L., Chojnowski, G., Cook, A. G., Cowtan, K. D., Croll, T. I., Debreczeni, J. É., Devenish, N. E., Dodson, E. J., Drevon, T. R., Emsley, P., Evans, G., Evans, P. R., Fando, M., Foadi, J., Fuentes-Montero, L., Garman, E. F., Gerstel, M., Gildea, R. J., Hatti, K., Hekkelman, M. L., Heuser, P., Hoh, S. W., Hough, M. A., Jenkins, H. T., Jiménez, E., Joosten, R. P., Keegan, R. M., Keep, N., Krissinel, E. B., Kolenko, P., Kovalevskiy, O., Lamzin, V. S., Lawson, D. M., Lebedev, A. A., Leslie, A. G. W., Lohkamp, B., Long, F., Malý, M., McCoy, A. J., McNicholas, S. J., Medina, A., Millán, C., Murray, J. W., Murshudov, G. N., Nicholls, R. A., Noble, M. E. M., Oeffner, R., Pannu, N. S., Parkhurst, J. M., Pearce, N., Pereira, J., Perrakis, A., Powell, H. R., Read, R. J., Rigden, D. J., Rochira, W., Sammito, M., Sánchez Rodríguez, F., Sheldrick, G. M., Shelley, K. L., Simkovic, F., Simpkin, A. J., Skubak, P., Sobolev, E., Steiner, R. A., Stevenson, K., Tews, I., Thomas, J. M. H., Thorn, A., Valls, J. T., Uski, V., Usón, I., Vagin, A., Velankar, S., Vollmar, M., Walden, H., Waterman, D., Wilson, K. S., Winn, M. D., Winter, G., Wojdyr, M. & Yamashita, K. (2023). Acta Cryst. D79, 449–461.  Web of Science CrossRef IUCr Journals Google Scholar
First citationAhmad, S., Wang, B., Walker, M. D., Tran, H. R., Stogios, P. J., Savchenko, A., Grant, R. A., McArthur, A. G., Laub, M. T. & Whitney, J. C. (2019). Nature, 575, 674–678.  CrossRef CAS PubMed Google Scholar
First citationBrown, A., Fernández, I. S., Gordiyenko, Y. & Ramakrishnan, V. (2016). Nature, 534, 277–280.  CrossRef CAS PubMed Google Scholar
First citationEvans, P. R. (2007). Acta Cryst. D63, 58–61.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationEvans, P. R. & Murshudov, G. N. (2013). Acta Cryst. D69, 1204–1214.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationGarcia-Pino, A., Christensen-Dalsgaard, M., Wyns, L., Yarmolinsky, M., Magnuson, R. D., Gerdes, K. & Loris, R. (2008). J. Biol. Chem. 283, 30821–30827.  Web of Science CrossRef PubMed CAS Google Scholar
First citationGarcia-Pino, A., De Gieter, S., Talavera, A., De Greve, H., Efremov, R. G. & Loris, R. (2016). Nat. Chem. Biol. 12, 490–496.  Web of Science CAS PubMed Google Scholar
First citationGarcia-Pino, A., Zenkin, N. & Loris, R. (2014). Trends Biochem. Sci. 39, 121–129.  CAS PubMed Google Scholar
First citationGarcia-Rodriguez, G., Charlier, D., Wilmaerts, D., Michiels, J. & Loris, R. (2021). Nucleic Acids Res. 49, 7164–7178.  CAS PubMed Google Scholar
First citationGasteiger, E., Gattiker, A., Hoogland, C., Ivanyi, I., Appel, R. D. & Bairoch, A. (2003). Nucleic Acids Res. 31, 3784–3788.  Web of Science CrossRef PubMed CAS Google Scholar
First citationGeiger, T., Kästle, B., Gratani, F. L., Goerke, C. & Wolz, C. (2014). J. Bacteriol. 196, 894–902.  Web of Science CrossRef PubMed Google Scholar
First citationHarms, A., Brodersen, D. E., Mitarai, N. & Gerdes, K. (2018). Mol. Cell, 70, 768–784.  Web of Science CrossRef CAS PubMed Google Scholar
First citationHayes, F. & Kędzierska, B. (2014). Toxins, 6, 337–358.  Web of Science CrossRef CAS PubMed Google Scholar
First citationJimmy, S., Saha, C. K., Kurata, T., Stavropoulos, C., Oliveira, S. R. A., Koh, A., Cepauskas, A., Takada, H., Rejman, D., Tenson, T., Strahl, H., Garcia-Pino, A., Hauryliuk, V. & Atkinson, G. C. (2020). Proc. Natl Acad. Sci. USA, 117, 10500–10510.  CrossRef CAS PubMed Google Scholar
First citationJurėnas, D., Fraikin, N., Goormaghtigh, F. & Van Melderen, L. (2022). Nat. Rev. Microbiol. 20, 335–350.  PubMed Google Scholar
First citationJurėnas, D., Garcia-Pino, A. & Van Melderen, L. (2017). Plasmid, 93, 30–35.  PubMed Google Scholar
First citationJurėnas, D., Van Melderen, L. & Garcia-Pino, A. (2019). Nat. Chem. Biol. 15, 285–294.  Web of Science PubMed Google Scholar
First citationKabsch, W. (2010). Acta Cryst. D66, 125–132.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationKantardjieff, K. A. & Rupp, B. (2003). Protein Sci. 12, 1865–1871.  Web of Science CrossRef PubMed CAS Google Scholar
First citationKurata, T., Brodiazhenko, T., Alves Oliveira, S. R., Roghanian, M., Sakaguchi, Y., Turnbull, K. J., Bulvas, O., Takada, H., Tamman, H., Ainelo, A., Pohl, R., Rejman, D., Tenson, T., Suzuki, T., Garcia-Pino, A., Atkinson, G. C. & Hauryliuk, V. (2021). Mol. Cell, 81, 3160–3170.  CrossRef CAS PubMed Google Scholar
First citationLeRoux, M. & Laub, M. T. (2022). Annu. Rev. Microbiol. 76, 21–43.  CrossRef PubMed Google Scholar
First citationLoris, R. & Garcia-Pino, A. (2014). Chem. Rev. 114, 6933–6947.  Web of Science CrossRef CAS PubMed Google Scholar
First citationMatthews, B. W. (1968). J. Mol. Biol. 33, 491–497.  CrossRef CAS PubMed Web of Science Google Scholar
First citationMcCoy, A. J., Grosse-Kunstleve, R. W., Adams, P. D., Winn, M. D., Storoni, L. C. & Read, R. J. (2007). J. Appl. Cryst. 40, 658–674.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationNanamiya, H., Kasai, K., Nozawa, A., Yun, C. S., Narisawa, T., Murakami, K., Natori, Y., Kawamura, F. & Tozawa, Y. (2008). Mol. Microbiol. 67, 291–304.  CrossRef PubMed CAS Google Scholar
First citationPage, R. & Peti, W. (2016). Nat. Chem. Biol. 12, 208–214.  Web of Science CrossRef CAS PubMed Google Scholar
First citationSammito, M., Millán, C., Frieske, D., Rodríguez-Freire, E., Borges, R. J. & Usón, I. (2015). Acta Cryst. D71, 1921–1930.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSteinchen, W., Schuhmacher, J. S., Altegoer, F., Fage, C. D., Srinivasan, V., Linne, U., Marahiel, M. A. & Bange, G. (2015). Proc. Natl Acad. Sci. USA, 112, 13348–13353.  Web of Science CrossRef CAS PubMed Google Scholar
First citationSteinchen, W., Vogt, M. S., Altegoer, F., Giammarinaro, P. I., Horvatek, P., Wolz, C. & Bange, G. (2018). Sci. Rep. 8, 2195.  CrossRef PubMed Google Scholar
First citationSterckx, Y. G., De Gieter, S., Zorzini, V., Hadži, S., Haesaerts, S., Loris, R. & Garcia-Pino, A. (2015). Protein Expr. Purif. 108, 30–40.  CrossRef CAS PubMed Google Scholar
First citationTalavera, A., Hendrix, J., Versees, W., Jurenas, D., Van Nerom, K., Vandenberk, N., Singh, R. K., Konijnenberg, A., De Gieter, S., Castro-Roa, D., Barth, A., De Greve, H., Sobott, F., Hofkens, J., Zenkin, N., Loris, R. & Garcia-Pino, A. (2018). Sci. Adv. 4, eaap9714.  CrossRef PubMed Google Scholar
First citationTalavera, A., Tamman, H., Ainelo, A., Konijnenberg, A., Hadži, S., Sobott, F., Garcia-Pino, A., Hõrak, R. & Loris, R. (2019). Nat. Commun. 10, 972.  Web of Science CrossRef PubMed Google Scholar
First citationTamman, H., Ernits, K., Roghanian, M., Ainelo, A., Julius, C., Perrier, A., Talavera, A., Ainelo, H., Dugauquier, R., Zedek, S., Thureau, A., Pérez, J., Lima-Mendez, G., Hallez, R., Atkinson, G. C., Hauryliuk, V. & Garcia-Pino, A. (2023). Nat. Chem. Biol. 19, 334–345.  CrossRef CAS PubMed Google Scholar
First citationTamman, H., Van Nerom, K., Takada, H., Vandenberk, N., Scholl, D., Polikanov, Y., Hofkens, J., Talavera, A., Hauryliuk, V., Hendrix, J. & Garcia-Pino, A. (2020). Nat. Chem. Biol. 16, 834–840.  CrossRef CAS PubMed Google Scholar
First citationTerwilliger, T. C., DiMaio, F., Read, R. J., Baker, D., Bunkóczi, G., Adams, P. D., Grosse-Kunstleve, R. W., Afonine, P. V. & Echols, N. (2012). J. Struct. Funct. Genomics, 13, 81–90.  CrossRef CAS PubMed Google Scholar
First citationThorn, A. (2017). Methods Mol. Biol. 1607, 357–376.  CrossRef CAS PubMed Google Scholar
First citationUsón, I. & Sheldrick, G. M. (2018). Acta Cryst. D74, 106–116.  Web of Science CrossRef IUCr Journals Google Scholar
First citationVonrhein, C., Flensburg, C., Keller, P., Sharff, A., Smart, O., Paciorek, W., Womack, T. & Bricogne, G. (2011). Acta Cryst. D67, 293–302.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationZhang, T., Tamman, H., Coppieters 't Wallant, K., Kurata, T., LeRoux, M., Srikant, S., Brodiazhenko, T., Cepauskas, A., Talavera, A., Martens, C., Atkinson, G. C., Hauryliuk, V., Garcia-Pino, A. & Laub, M. T. (2022). Nature, 612, 132–140.  CrossRef CAS PubMed Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

Journal logoSTRUCTURAL BIOLOGY
COMMUNICATIONS
ISSN: 2053-230X
Follow Acta Cryst. F
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds