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

IUCrJ
Volume 9| Part 5| September 2022| Pages 625-631
ISSN: 2052-2525

The Haemophilus influenzae HipBA toxin–antitoxin system adopts an unusual three-com­ponent regulatory mechanism

crossmark logo

aResearch Institute of Pharmaceutical Sciences, College of Pharmacy, Seoul National University, Seoul 08826, Republic of Korea, bCollege of Pharmacy, Duksung Women's University, Seoul 01369, Republic of Korea, cJeju Research Institute of Pharmaceutical Sciences, College of Pharmacy, Jeju National University, Jeju 63243, Republic of Korea, and dInterdisciplinary Graduate Program in Advanced Convergence Technology and Science, Jeju National University, Jeju 63243, Republic of Korea
*Correspondence e-mail: lbj@nmr.snu.ac.kr

Edited by Z.-J. Liu, Chinese Academy of Sciences, China (Received 10 May 2022; accepted 6 July 2022; online 29 July 2022)

Type II toxin–antitoxin (TA) systems encode two proteins: a toxin that inhibits cell growth and an antitoxin that neutralizes the toxin by direct inter­molecular protein–protein inter­actions. The bacterial HipBA TA system is implicated in persister formation. The Haemophilus influenzae HipBA TA system consists of a HipB antitoxin and a HipA toxin, the latter of which is split into two fragments, and here we investigate this novel three-com­ponent regulatory HipBA system. Structural and functional analysis revealed that HipAN corresponds to the N-ter­minal part of HipA from other bacteria and toxic HipAC is inactivated by HipAN, not HipB. This study will be helpful in understanding the detailed regulatory mechanism of the HipBAN+C system, as well as why it is constructed as a three-com­ponent system.

1. Introduction

Prokaryotic toxin–antitoxin (TA) systems are encoded by small operons that are com­posed of two elements, a toxic subunit that causes cell-growth arrest and an antitoxic subunit that neutralizes the harmful effect of the toxin (Cheverton et al., 2016[Cheverton, A. M., Gollan, B., Przydacz, M., Wong, C. T., Mylona, A., Hare, S. A. & Helaine, S. (2016). Mol. Cell, 63, 86-96.]). Depending on the mol­ecular mechanism by which the antitoxins neutralize their cognate toxins, TA systems have been classified into diverse types (Hayes, 2003[Hayes, F. (2003). Science, 301, 1496-1499.]). In the most abundant type II TA systems, the antitoxin is a protein that directly binds and neutralizes its cognate toxin through strong mol­ecular inter­actions (Hayes & Van Melderen, 2011[Hayes, F. & Van Melderen, L. (2011). Crit. Rev. Biochem. Mol. Biol. 46, 386-408.]). When freed from inhibition by antitoxins, the toxin acts on various targets to suppress cell growth or even induce cell death (Hall et al., 2017[Hall, A. M., Gollan, B. & Helaine, S. (2017). Curr. Opin. Microbiol. 36, 102-110.]). The toxin is relatively long-lived, whereas the antitoxin is degraded by cellular proteases due to its labile nature. Accordingly, the antitoxin should be constantly produced to regulate the toxin (Hõrak & Tamman, 2017[Hõrak, R. & Tamman, H. (2017). Curr. Genet. 63, 69-74.]). Type II antitoxins commonly contain a DNA-binding domain to bind at operators in the promoter region and repress the TA operon (Chan et al., 2016[Chan, W. T., Espinosa, M. & Yeo, C. C. (2016). Front. Mol. Biosci. 3, 9.]). Therefore, as the amount of antitoxin is reduced, repression stops and the operon is transcribed to supply the antitoxin (Loris & Garcia-Pino, 2014[Loris, R. & Garcia-Pino, A. (2014). Chem. Rev. 114, 6933-6947.]). The cellular activity of a toxin is thus driven by the qu­antity of the cognate antitoxin (Hõrak & Tamman, 2017[Hõrak, R. & Tamman, H. (2017). Curr. Genet. 63, 69-74.]).

Type II TA systems are greatly abundant in nearly all free-living bacteria, raising debates about their biological functions (Pandey & Gerdes, 2005[Pandey, D. P. & Gerdes, K. (2005). Nucleic Acids Res. 33, 966-976.]). Many studies point to additional functions of TA systems, including phage resistance, stress responses, biofilm formation and antibiotic persistence (Kędzierska & Hayes, 2016[Kędzierska, B. & Hayes, F. (2016). Molecules, 21, 790.]; Lobato-Márquez et al., 2016[Lobato-Márquez, D., Díaz-Orejas, R. & García-del Portillo, F. (2016). FEMS Microbiol. Rev. 40, 592-609.]; Díaz-Orejas et al., 2017[Díaz-Orejas, R., Espinosa, M. & Yeo, C. C. (2017). Front. Microbiol. 8, 1479.]). Stress-induced activation of TA systems protects bacteria from adverse environmental conditions, such as antibiotic treatment, by causing persister formation (Radzikowski et al., 2016[Radzikowski, J. L., Vedelaar, S., Siegel, D., Ortega, A. D., Schmidt, A. & Heinemann, M. (2016). Mol. Syst. Biol. 12, 882.]). Anti­biotic persistence results in a transient slow-growing state of bacteria, which allows antibiotics to be tolerated (Harms et al., 2016[Harms, A., Maisonneuve, E. & Gerdes, K. (2016). Science, 354, aaf4268.]). This persister formation is a bet-hedging strategy in which the bacteria temporarily enters a nonreplicating state to gain stress tolerance against unfavourable environmental threats (Balaban et al., 2004[Balaban, N. Q., Merrin, J., Chait, R., Kowalik, L. & Leibler, S. (2004). Science, 305, 1622-1625.]; Veening et al., 2008[Veening, J. W., Smits, W. K. & Kuipers, O. P. (2008). Annu. Rev. Microbiol. 62, 193-210.]; Lewis, 2010[Lewis, K. (2010). Annu. Rev. Microbiol. 64, 357-372.]). Moreover, antibiotic persistence plays a crucial role in chronic infections and stimulates the evolution to antibiotic resistance (Levin-Reisman et al., 2017[Levin-Reisman, I., Ronin, I., Gefen, O., Braniss, I., Shoresh, N. & Balaban, N. Q. (2017). Science, 355, 826-830.]).

The HipBA TA system was first identified in Escherichia coli, in which an increase in persister formation is associated with the hipBA operon (Moyed & Bertrand, 1983[Moyed, H. S. & Bertrand, K. P. (1983). J. Bacteriol. 155, 768-775.]). As an abbreviation of `high incidence of persistence', hipB (antitoxin gene) and hipA (toxin gene) constitute a type II TA module that encodes two proteins: HipB (antitoxin) and HipA (toxin) (Korch & Hill, 2006[Korch, S. B. & Hill, T. M. (2006). J. Bacteriol. 188, 3826-3836.]). HipA acts as a serine–threonine kinase, and the expression of HipA triggers considerable cell-growth arrest (Hanks et al., 1988[Hanks, S. K., Quinn, A. M. & Hunter, T. (1988). Science, 241, 42-52.]; Black et al., 1991[Black, D. S., Kelly, A. J., Mardis, M. J. & Moyed, H. S. (1991). J. Bacteriol. 173, 5732-5739.], 1994[Black, D. S., Irwin, B. & Moyed, H. S. (1994). J. Bacteriol. 176, 4081-4091.]; Stancik et al., 2018[Stancik, I. A., Šestak, M. S., Ji, B., Axelson-Fisk, M., Franjevic, D., Jers, C., Domazet-Lošo, T. & Mijakovic, I. (2018). J. Mol. Biol. 430, 27-32.]). On the other hand, HipB counteracts HipA by directly binding it and forming a stable protein com­plex (Korch & Hill, 2006[Korch, S. B. & Hill, T. M. (2006). J. Bacteriol. 188, 3826-3836.]). This HipBA protein com­plex binds to operators in the promoter region via the DNA-binding domain of HipB antitoxin and represses the transcription of hipBA (Schumacher et al., 2015[Schumacher, M. A., Balani, P., Min, J., Chinnam, N. B., Hansen, S., Vulić, M., Lewis, K. & Brennan, R. G. (2015). Nature, 524, 59-64.]). Upon HipB degradation, free HipA induces multidrug tolerance, leading to the dormant state of cells (Veening et al., 2008[Veening, J. W., Smits, W. K. & Kuipers, O. P. (2008). Annu. Rev. Microbiol. 62, 193-210.]; Wood et al., 2013[Wood, T. K., Knabel, S. J. & Kwan, B. W. (2013). Appl. Environ. Microbiol. 79, 7116-7121.]).

This study elucidates the crystal structure of HI0666, which exhibits sequence similarity with the N-terminal part of HipA (HipAN). To date, the HipBA system has been structurally studied in only E. coli (Schumacher et al., 2015[Schumacher, M. A., Balani, P., Min, J., Chinnam, N. B., Hansen, S., Vulić, M., Lewis, K. & Brennan, R. G. (2015). Nature, 524, 59-64.]) and Shewanella oneidensis (Wen et al., 2014[Wen, Y., Behiels, E., Felix, J., Elegheert, J., Vergauwen, B., Devreese, B. & Savvides, S. N. (2014). Nucleic Acids Res. 42, 10134-10147.]). Structural information obtained in this study shows that HI0666 is part of an unusual three-com­ponent TA system resembling the HipBA system. Compared with sequences of HipAs, whose structures were known previously, the HI0665 sequence exhibits similarity with the C-terminal sequence of HipA (HipAC) from other species. In the genome map, hipAC (encoding the C-terminus of HipA) is preceded by hipAN (encoding the N-terminus of HipA), and hipAN is preceded by hipB (encoding HipB). The results confirmed that HipAN and HipAC form a binary com­plex, and that HipB, HipAN and HipAC form a tertiary com­plex. HipAC-mediated growth inhibition was rescued by the expression of HipAN but not by HipB. HipAN functions as an untraditional antitoxin to HipAC, and HipB increases the ability of HipAN to counteract HipAC. Taken together, these results outline a nontraditional three-com­ponent TA system, and a putative regulatory mechanism of the H. influenzae HipBA system is proposed.

2. Materials and methods

2.1. Gene cloning and transformation

hi0666 (hipAN), a gene encoding HipAN from H. influenzae, was amplified by polymerase chain reaction (PCR) using H. influenzae (KW20 strain, ATCC 51907) genomic DNA as a template. The oligonucleotide primers used for PCR are given in Table 1[link]. After PCR amplification, hipAN was cloned into the pET28a vector (Novagen) using the restriction enzymes NdeI and XhoI. hi0666.1 (hipB), a gene encoding HipB, was amplified in the same way and cloned into the pET21a vector (Novagen) using the same enzymes. To express HipAN and HipAC together, hipAN was first cloned into the pETDuet vector (Addgene), and hi0665 (hipAC), a gene encoding HipAC, was subsequently introduced to follow hipAN in the pETDuet vector. For the cloning of hipAC into the pETDuet vector, the restriction enzymes BamHI and HindIII were used to digest PCR products and vector. A recombinant pET28a vector containing hipAC was also produced using the same approach. All recombinant plasmids were transformed into E. coli BL21(DE3) com­petent cells (Novagen).

Table 1
Primers designed for cloning the genes encoding HipAN, HipAC and HipB

Proteins (vector)   Primer sequence (5′→3′)
HipAN (pET28a) Forward GGAATTC CATATG ATG CGC GAT TTA GTC CG
  Reverse CCG CTCGAG TTA TTG TTT TTC TTC CTG AAT TTG C
HipB (pET21a) Forward GGAATTC CATATG ATG GAC AAT CTT AGT GCA C
  Reverse CCG CTCGAG TTA AAT CGC GCA TAG TGA AAC
HipAN (pETDuet) Forward GGAATTC CATATG ATG CGC GAT TTA GTC
  Reverse CCG CTCGAG TTA TTG TTT TTC TTC CTG AAT TTG
HipAC (pETDuet) Forward CGC GGATCC G ATG AAT TTT TGT CGT ATT TTA
  Reverse CCC AAGCTT TTA TAG TTC AGG TTC ATT TAA TAG
HipAC (pET28a) Forward CGC GGATCC G ATG AAT TTT TGT CGT ATT TTA TTA AAG CCA
  Reverse CCC AAGCTT TTA TAG TTC AGG TTC ATT TAA TAG GTT AAG CA

2.2. Protein expression and purification

Transformed cells harbouring hipAN were grown to an optical density at 600 nm (OD600) of approximately 0.5 in LB medium containing 50 mg l−1 kanamycin at 37°C. At this point, expression was induced by adding 0.5 mM isopropyl β-D-1-thiogalactopyran­oside (IPTG) and cells were grown for an additional 4 h at the same temperature. Grown cells were harvested by centrifugation at 5600g. The harvested cells were resuspended in buffer A [50 mM Tris–HCl, pH 7.9, and 500 mM NaCl containing 10%(v/v) glycerol] and lysed by sonication. The cell lysate was centrifuged at 18 000g for 1 h, and the supernatant was applied to an affinity chromatography column of nickel–nitrilo­tri­acetic acid (Roche) equilibrated with buffer A. The trapped protein in the column was washed with buffer A containing 20 mM imidazole and eluted with buffer A containing 100–500 mM imidazole. Final purification of HipAN was achieved by size-exclusion chromatography using a Superdex 200 (16/600PG) column (GE Healthcare) equilibrated with 20 mM HEPES, pH 7.5, and 500 mM NaCl. The purity of HipAN was verified by SDS–PAGE.

2.3. Crystallization and diffraction data collection

The purified HipAN was concentrated to 10 mg ml−1 using an Amicon Ultra Centrifugal Filter Unit (Millipore) for crystallization. Initial crystallization was performed with publicly available screening kits using a 384-well plate by the sitting-drop vapour-diffusion method. Each well of plates contained 0.5 µl of protein solution and 0.5 µl of reservoir solution. The plates were cultured at 4°C. The crystals used for data collection were grown in reservoirs with 2 M ammonium sulfate and 0.1 M Tris, pH 8.5. The X-ray diffraction data used for structure calculation were collected at 100 K using an ADSC Quantum 315r CCD detector at the 5C beamline of the Pohang Light Source, Republic of Korea. Collected raw data were processed and scaled with the HKL2000 program suite (Otwinowski & Minor, 1997[Otwinowski, Z. & Minor, W. (1997). Methods Enzymol. 276, 307-326.]). Statistics for the data collection are described in Table 2[link].

Table 2
Statistics for data collection and model refinement

Values in parentheses refer to the highest-resolution shell.

  HipAN (PDB code: 7czo)
Data collection  
X-ray wavelength (Å) 0.9795
Space group P41212
Unit-cell length (a, b, c, Å) 66.25, 66.25, 103.62
Unit-cell angle (α, β, γ, °) 90.00, 90.00, 90.00
Resolution range (Å) 50.00–2.70 (2.75–2.70)
Total/unique reflections 94,246/6,880 (345)
Completeness (%) 99.9 (100.0)
CC1/2 0.998 (0.973)
I/σI 27.9 (3.2)
Rmerge 0.096 (0.766)
   
Model refinement  
Rwork/Rfree§ 0.242/0.274
No./average B factor (Å2)  
Protein atoms 882/41.43
Water oxygen atoms 30/45.88
R.m.s. deviation from ideal geometry  
Bond lengths (Å)/bond angles (°) 0.011/1.473
Ramachandran plot (%)  
Most favorable 99.42
Allowed 0.58
Disallowed 0.00
†CC1/2 is described in Karplus & Diederichs (2012[Karplus, P. A. & Diederichs, K. (2012). Science, 336, 1030-1033.]).
Rmerge = [\textstyle\sum_h\textstyle\sum_i|I(h)i - \langle I(h)\rangle | / \textstyle\sum_h\textstyle\sum_i I(h)i], where I(h) is the intensity of reflection h, [\textstyle\sum_h] is the sum over all reflections and [\textstyle\sum_i] is the sum over i measurements of reflection h.
§R = [\textstyle\sum || F_{\rm obs}|-| F_{\rm calc}||/\textstyle\sum|F_{\rm obs}|], where Rfree is calculated for a randomly chosen 5% of reflections that were not used for structure refinement and Rwork is calculated for the remaining reflections.

2.4. Structure determination and refinement

To determine the structure of HipAN, PhaserMR in Phenix (Adams et al., 2010[Adams, P. D., Afonine, P. V., Bunkóczi, G., Chen, V. B., Davis, I. W., Echols, N., Headd, J. J., Hung, L.-W., Kapral, G. J., Grosse-Kunstleve, R. W., McCoy, A. J., Moriarty, N. W., Oeffner, R., Read, R. J., Richardson, D. C., Richardson, J. S., Terwilliger, T. C. & Zwart, P. H. (2010). Acta Cryst. D66, 213-221.]) was employed using the Protein Data Bank (PDB) entry 2wiu (Evdokimov et al., 2009[Evdokimov, A., Voznesensky, I., Fennell, K., Anderson, M., Smith, J. F. & Fisher, D. A. (2009). Acta Cryst. D65, 875-879.]) structure as a model. Iterative cycles of initial model building were performed by Coot (Emsley et al., 2010[Emsley, P., Lohkamp, B., Scott, W. G. & Cowtan, K. (2010). Acta Cryst. D66, 486-501.]), and further refinement was conducted using REFMAC5 (Murshudov et al., 1997[Murshudov, G. N., Vagin, A. A. & Dodson, E. J. (1997). Acta Cryst. D53, 240-255.]). The HipAN crystal structure belongs to the P41212 space group at 2.70 Å. The coordinates and structure factors are deposited in the PDB under the accession code 7czo for HipAN, the N-terminal com­ponent of HipA toxin from H. influenzae. Detailed refinement statistics are available in Table 2[link].

2.5. Bioinformatics

Structure figures were generated using PyMol (The PyMOL Mol­ecular Graphics System, Version 1.8, Schrödinger, LLC). Superimpositions of structures were performed with the SSM (Krissinel & Henrick, 2004[Krissinel, E. & Henrick, K. (2004). Acta Cryst. D60, 2256-2268.]) option within Coot. The solvent-accessible areas of the protein surface were calculated by PISA (Krissinel & Henrick, 2007[Krissinel, E. & Henrick, K. (2007). J. Mol. Biol. 372, 774-797.]). Genome maps were utilized from the KEGG database (Kanehisa et al., 2017[Kanehisa, M., Furumichi, M., Tanabe, M., Sato, Y. & Morishima, K. (2017). Nucleic Acids Res. 45, D353-D361.]). Sequence and structural similarity were searched using DALI (Holm & Rosenstrom, 2010[Holm, L. & Rosenström, P. (2010). Nucleic Acids Res. 38, W545-W549.]). Structure-assisted alignment was carried out using ESPript (Robert & Gouet, 2014[Robert, X. & Gouet, P. (2014). Nucleic Acids Res. 42, W320-W324.]), with the help of ClustalW (Chenna et al., 2003[Chenna, R., Sugawara, H., Koike, T., Lopez, R., Gibson, T. J., Higgins, D. G. & Thompson, J. D. (2003). Nucleic Acids Res. 31, 3497-3500.]) for sequence alignments. The quality of the final structure was assessed at the wwPDB X-ray structure validation server (Berman et al., 2003[Berman, H., Henrick, K. & Nakamura, H. (2003). Nat. Struct. Mol. Biol. 10, 980.]).

2.6. Copurification of HipAN, HipAC and HipB

To determine whether HipB, HipAN and HipAC might form a protein com­plex, the pETDuet vector containing both hipAN and hipAC, and the pET21a vector containing hipB were used. For this study, a hexa­histidine tag was attached to the N-ter­minus of HipAC. These plasmids were cotransformed and bacterial cells were cultured in the same way as described above, with 50 mg l−1 streptomycin. Protein expression and affinity chromatography were also carried out in a similar manner to the above methods, but the three proteins were eluted carefully from being bound to the column with a gradient of 50–700 mM imidazole.

2.7. Analytical size-exclusion chromatography

The purified protein mixture of HipB, HipAN and HipAC was subjected to size-exclusion chromatography under the same conditions as the above mixture. A standard curve was obtained using standards from gel-filtration calibration kits (GE Healthcare), including aldolase (158 kDa), conalbumin (75 kDa) and ovalbumin (43 kDa), and the curve was com­pared with the peak position obtained from the protein mixture.

2.8. Cell-growth assay

To validate the three-com­ponent regulatory mechanism of the H. influenzae HipBA system, a cell-growth assay was performed using the following plasmids resistant to different antibiotics: pET21a containing hipB, pETDuet containing only hipAN and pET28a containing hipAC. For this assay, all plasmids were transformed into E. coli strain BL21(DE3). Single colonies from transformed cells grown in M9 medium plates containing 0.1% glucose were further grown overnight. These overnight cultures were then diluted to an OD600 of 0.1. The diluted cells were freshly grown until the OD600 reached 0.5; 0.5 mM IPTG was then added to the culture medium to induce protein expression. The cells were incubated at 37°C and monitored at 1 h inter­vals.

3. Results and discussion

3.1. Overall structure of H. influenzae HipAN and the hipBA genome map

There are four α-helices and five β-strands arranged as a β-barrel in the structure of H. influenzae HipAN [Fig. 1[link](a)]. Among those β-strands, five β-strands (β1–β5) form a β-sheet antiparallel to each other and these antiparallel β-sheets are flanked by four α-helices. The total solvent-accessible surface area of the monomeric structure is 7119 Å2.

[Figure 1]
Figure 1
The overall structure of the H. influenzae HipAN and hipBA genome maps. (a) 180° rotated views of H. influenzae HipAN. α-Helices, β-strands and loops are coloured red, yellow and green, respectively. (b) Schematic diagram showing a com­parison of the hipBA operons of H. influenzae, E. coli and S. oneidensis.

In the reported structures of HipBA systems, we discovered an additional locus in the genome map of H. influenzae immediately upstream from hipAC. This locus is called hipAN in this article and encodes a 106 aa protein corresponding to the N-terminal part of HipA from E. coli and S. oneidensis [Fig. 1[link](b)]. In all these cases, hipB was located upstream of hipA and these putative HipB homologs might thus autoregulate the hipBAN+C operons. In summary, the hipBA operon contains two genes, hipB and hipA, while the hipBA operon from H. influenzae contains three genes, hipB, hipAN and hipAC.

3.2. Comparative structural analysis of the HipBA system

To date, structures of HipA have been reported from two species: E. coli (PDB code 5k98; Z score of 9.2, r.m.s. deviation of 2.4 Å, sequence identity of 24%) (Schumacher et al., 2015[Schumacher, M. A., Balani, P., Min, J., Chinnam, N. B., Hansen, S., Vulić, M., Lewis, K. & Brennan, R. G. (2015). Nature, 524, 59-64.]) and S. oneidensis (PDB code 4pu3; Z score of 9.9, r.m.s. deviation of 2.7 Å, sequence identity of 23%) (Wen et al., 2014[Wen, Y., Behiels, E., Felix, J., Elegheert, J., Vergauwen, B., Devreese, B. & Savvides, S. N. (2014). Nucleic Acids Res. 42, 10134-10147.]) [Fig. 2[link](a)]. To com­pare the structural characteristics of H. influenzae HipAN and its homologs, a com­parative analysis was performed with these two reported structures. H. influenzae HipAN superimposed well with the N-terminal parts of the HipA proteins from E. coli and S. oneidensis [Fig. 2[link](b)]. In addition, H. influenzae HipB showed a 48% sequence similarity to HipB of the E. coli O127:H6 tripartite HipBA system (PDB code: 7ab3), which in turn showed high structural homology with E. coli HipB (PDB code: 5k98, Z score of 12.0, r.m.s. deviation of 1.2 Å) (Schumacher et al., 2015[Schumacher, M. A., Balani, P., Min, J., Chinnam, N. B., Hansen, S., Vulić, M., Lewis, K. & Brennan, R. G. (2015). Nature, 524, 59-64.]) and S.oneidensis HipB (PDB code: 4pu3, Z score of 9.7, r.m.s. deviation of 1.8 Å) (Wen et al., 2014[Wen, Y., Behiels, E., Felix, J., Elegheert, J., Vergauwen, B., Devreese, B. & Savvides, S. N. (2014). Nucleic Acids Res. 42, 10134-10147.]) [Fig. 2[link](c)]. Regardless of the binary or tertiary nature of the HipBA systems, dimeric HipB antitoxins were highly conserved in their DNA-binding HTH motif [Fig. 2[link](c)], which also indicated their conserved role as transcriptional regulators. Furthermore, although the HipA toxins had similar folds, the N-terminal of HipA bound HipB antitoxins differently with respect to E. coli (loop between β3 and β4, containing α1) and S. oneidensis (α3–α4) proteins. This may suggest that the toxin neutralization mechanism of the HipBA system might be different among its homologues.

[Figure 2]
Figure 2
Comparative structural analysis of HipBA systems. (a) Structure-guided sequence alignment of H. influenzae HipAN with the N-terminal parts of other HipAs. The secondary structural organization of H. influenzae HipAN is displayed on the alignment. Conserved residues are highlighted in red and yellow. (b) Structural com­parison of H. influenzae HipAN and HipBA systems from other organisms. Residues showing high conservation are marked in the overlay. HipBA binding inter­faces of E. coli and S. oneidensis are identified in circles and enlarged in squares. (c) Structural com­parison of HipB antitoxins from E. coli O127:H6 (PDB code: 7ab3) and S. oneidensis (PDB code: 4pu3) using E. coli HipB antitoxin bound to the DNA (PDB code: 5k98) structure as model.

3.3. Complex formation of HipB, HipAN and HipAC

To show that HipB, HipAN and HipAC from H. influenzae form a com­plex similar to those of other type II TA systems, analytical gel filtration was conducted [Fig. 3[link](a)], and the eluted fractions were analyzed by SDS–PAGE [Fig. 3[link](b)]. His-tagged HipAC was pulled down with both HipAN and HipB. Size-exclusion chromatography further confirmed that two forms of com­plex [(HipB + HipAN + HipAC) and (HipAN + HipAC)] are monodispersed in solution. The second peak eluted between the 75 and 43 kDa mol­ecular weight standards, which corresponded to the heterodimeric HipAN+C com­plex (∼52.3 kDa). This was consistent with the relative masses of HipA toxin structures in the binary E. coli HipBA systems (Schumacher et al., 2015[Schumacher, M. A., Balani, P., Min, J., Chinnam, N. B., Hansen, S., Vulić, M., Lewis, K. & Brennan, R. G. (2015). Nature, 524, 59-64.]). In addition, chemical crosslinking of H. influenzae HipB showed the predominant multimeric form of HipB as the dimer [Fig. 3[link](c)], as the other HipB antitoxins and thus the results of the first peak which eluted between the 158 and 75 kDa mol­ecular weight standards were presumed to be the hexa­meric com­plex consisting of one HipB dimer bound to two HipAN+C heterodimers. As predicted, the recently deposited structure of the tertiary HipBA com­plex (PDB code: 7ab3) in E. coli O127:H6 revealed a hexa­meric assembly in which the HipB dimer was bound to two HipAN+C hetero­dimers (Baerentsen et al., 2022[Baerentsen, R. L., Nielsen, S. V., Lyngsø, J., Bisiak, F., Pedersen, J. S., Gerdes, K., Sørensen, M. A. & Brodersen, D. E. (2022). bioRxiv, https://doi.org/10.1101/2022.01.28.478185.]). Indeed, the expected mol­ecular weights of the copurified three proteins suggested that H. influenzae (HipB + HipAN + HipAC) forms a hetero­hexamer and (HipAN + HipAC) forms a heterodimer in solution.

[Figure 3]
Figure 3
Complex formation of HipB, HipAN and HipAC. (a) Purification of tertiary mixtures using size exclusion chromatography. Elution locations of standard proteins with known mass are shown with vertical arrows. The two major peaks of the protein mixture are emphasized with star symbols. (b) SDS–PAGE analysis of size exclusion chromatography fractions. Star symbols indicate the lanes to which the starred peaks belong. (c) SDS–PAGE analysis of chemical crosslinking of H. influenzae HipB using DMA and DMS crosslinkers. Experimental conditions are indicated above and the corresponding multimer form is labelled on the right side of the gel.

3.4. Three-com­ponent regulatory mechanism of the H. influenzae HipBA system

We validated the com­ponents constituting the H. influenzae HipBA system and their regulatory mechanism through a cell-growth assay [Fig. 4[link](a)]. Expression of HipAC resulted in strong inhibition of cell growth, indicating that HipAC functions as a toxin. Surprisingly, the growth was not rescued by the traditional HipB antitoxin but was rescued upon the expression of HipAN, suggesting that HipAN functions as the antitoxin. In addition, co-induction of HipB and HipAN provided a slightly advantageous growth rescue com­pared to HipAN alone, giving rise to the possibility that HipB may augment the antitoxin activity of HipAN to quickly restore cell growth [Fig. 4[link](b)]. The co-induction of HipB and HipAN in the E. coli O127:H6 HipBA tripartite TA system also showed improved growth over induction of HipAN alone (Vang Nielsen et al., 2019[Vang Nielsen, S., Turnbull, K. J., Roghanian, M., Baerentsen, R., Semanjski, M., Brodersen, D. E., Macek, B. & Gerdes, K. (2019). mBio, 10, e01138-19.]), further supporting the augmentative role of HipB along with the HipAN com­ponent to further inhibit the toxic activity of HipAC toxins.

[Figure 4]
Figure 4
Three-com­ponent regulatory mechanism of the H. influenzae HipBA system. (a) Cell-growth analysis showing the effects of expression of different proteins. Each curve represents the growth of cells with a different combination of proteins, as presented within the graph. The data show the average values obtained by triplicate assays; the standard deviations are indicated by error bars. (b) Schematic overview of the presumptive regulatory mechanism according to the inter­actions between the com­ponents of the H. influenzae HipBA system.

4. Conclusions

Of the three proteins constituting the H. influenzae HipBAN+C system, the function of HipAN as the third com­ponent is quite striking. We found that HipAN counteracts the toxin HipAC, while the conserved HipB does not have an antitoxic effect. However, HipB is proposed to strengthen the activity of the HipAN-mediated neutralization of HipAC, supporting its augmentative role. For example, HipB might intensify the efficacy of HipAN by stabilizing its inter­action with HipAC. This hypothesis is consistent with the com­plex formation of HipB, HipAN and HipAC. In addition, hipA might have been split into two fragments during evolution, but the reason for this is unknown.

In conclusion, our work here reveals a novel type of three-com­ponent TA module with unknown regulator properties that is important and exciting to study.

Supporting information


Footnotes

These authors contributed equally and share first authorship.

Acknowledgements

We thank the beamline (BL) staff members at the Pohang Light Source, Republic of Korea (BL-5 C and BL-11 C), and SPring-8, Japan [beamline BL44XU under the approval of the Japan Synchrotron Radiation Research Institute (proposal No. 2019B6972)], for assistance with the X-ray diffraction experiments. This work was funded by the Korea Ministry of Science, Information, Communication, Technology, and Future Planning and the National Research Foundation (NRF) of Korea. This work was also supported by the Inter­national Program (ICR-21-05) of the Institute for Protein Research (IPR), Osaka University.

Funding information

Funding for this research was provided by: National Research Foundation of Korea (grant Nos. NRF-2018R1A5A2024425 and NRF-2021R1F1A1050961 to BJ.L; grant No. NRF-2019R1C1C1002128 to SMK; grant No. NRF-2022R1C1C1006354 to DHK).

References

First citationAdams, P. D., Afonine, P. V., Bunkóczi, G., Chen, V. B., Davis, I. W., Echols, N., Headd, J. J., Hung, L.-W., Kapral, G. J., Grosse-Kunstleve, R. W., McCoy, A. J., Moriarty, N. W., Oeffner, R., Read, R. J., Richardson, D. C., Richardson, J. S., Terwilliger, T. C. & Zwart, P. H. (2010). Acta Cryst. D66, 213–221.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationBaerentsen, R. L., Nielsen, S. V., Lyngsø, J., Bisiak, F., Pedersen, J. S., Gerdes, K., Sørensen, M. A. & Brodersen, D. E. (2022). bioRxiv, https://doi.org/10.1101/2022.01.28.478185.  Google Scholar
First citationBalaban, N. Q., Merrin, J., Chait, R., Kowalik, L. & Leibler, S. (2004). Science, 305, 1622–1625.  Web of Science CrossRef PubMed CAS Google Scholar
First citationBerman, H., Henrick, K. & Nakamura, H. (2003). Nat. Struct. Mol. Biol. 10, 980.  Web of Science CrossRef Google Scholar
First citationBlack, D. S., Irwin, B. & Moyed, H. S. (1994). J. Bacteriol. 176, 4081–4091.  CrossRef CAS PubMed Web of Science Google Scholar
First citationBlack, D. S., Kelly, A. J., Mardis, M. J. & Moyed, H. S. (1991). J. Bacteriol. 173, 5732–5739.  CrossRef PubMed CAS Web of Science Google Scholar
First citationChan, W. T., Espinosa, M. & Yeo, C. C. (2016). Front. Mol. Biosci. 3, 9.  Web of Science CrossRef PubMed Google Scholar
First citationChenna, R., Sugawara, H., Koike, T., Lopez, R., Gibson, T. J., Higgins, D. G. & Thompson, J. D. (2003). Nucleic Acids Res. 31, 3497–3500.  Web of Science CrossRef PubMed CAS Google Scholar
First citationCheverton, A. M., Gollan, B., Przydacz, M., Wong, C. T., Mylona, A., Hare, S. A. & Helaine, S. (2016). Mol. Cell, 63, 86–96.  Web of Science CrossRef CAS PubMed Google Scholar
First citationDíaz-Orejas, R., Espinosa, M. & Yeo, C. C. (2017). Front. Microbiol. 8, 1479.  Web of Science PubMed Google Scholar
First citationEmsley, P., Lohkamp, B., Scott, W. G. & Cowtan, K. (2010). Acta Cryst. D66, 486–501.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationEvdokimov, A., Voznesensky, I., Fennell, K., Anderson, M., Smith, J. F. & Fisher, D. A. (2009). Acta Cryst. D65, 875–879.  Web of Science CrossRef IUCr Journals Google Scholar
First citationHall, A. M., Gollan, B. & Helaine, S. (2017). Curr. Opin. Microbiol. 36, 102–110.  Web of Science CrossRef CAS PubMed Google Scholar
First citationHanks, S. K., Quinn, A. M. & Hunter, T. (1988). Science, 241, 42–52.  CrossRef CAS PubMed Web of Science Google Scholar
First citationHarms, A., Maisonneuve, E. & Gerdes, K. (2016). Science, 354, aaf4268.  Web of Science CrossRef PubMed Google Scholar
First citationHayes, F. (2003). Science, 301, 1496–1499.  Web of Science CrossRef PubMed CAS Google Scholar
First citationHayes, F. & Van Melderen, L. (2011). Crit. Rev. Biochem. Mol. Biol. 46, 386–408.  Web of Science CrossRef CAS PubMed Google Scholar
First citationHolm, L. & Rosenström, P. (2010). Nucleic Acids Res. 38, W545–W549.  Web of Science CrossRef CAS PubMed Google Scholar
First citationHõrak, R. & Tamman, H. (2017). Curr. Genet. 63, 69–74.  Web of Science PubMed Google Scholar
First citationKanehisa, M., Furumichi, M., Tanabe, M., Sato, Y. & Morishima, K. (2017). Nucleic Acids Res. 45, D353–D361.  Web of Science CrossRef CAS PubMed Google Scholar
First citationKarplus, P. A. & Diederichs, K. (2012). Science, 336, 1030–1033.  Web of Science CrossRef CAS PubMed Google Scholar
First citationKędzierska, B. & Hayes, F. (2016). Molecules, 21, 790.  Google Scholar
First citationKorch, S. B. & Hill, T. M. (2006). J. Bacteriol. 188, 3826–3836.  Web of Science CrossRef PubMed CAS Google Scholar
First citationKrissinel, E. & Henrick, K. (2004). Acta Cryst. D60, 2256–2268.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationKrissinel, E. & Henrick, K. (2007). J. Mol. Biol. 372, 774–797.  Web of Science CrossRef PubMed CAS Google Scholar
First citationLevin-Reisman, I., Ronin, I., Gefen, O., Braniss, I., Shoresh, N. & Balaban, N. Q. (2017). Science, 355, 826–830.  Web of Science CAS PubMed Google Scholar
First citationLewis, K. (2010). Annu. Rev. Microbiol. 64, 357–372.  Web of Science CrossRef CAS PubMed Google Scholar
First citationLobato-Márquez, D., Díaz-Orejas, R. & García-del Portillo, F. (2016). FEMS Microbiol. Rev. 40, 592–609.  Web of Science PubMed Google Scholar
First citationLoris, R. & Garcia-Pino, A. (2014). Chem. Rev. 114, 6933–6947.  Web of Science CrossRef CAS PubMed Google Scholar
First citationMoyed, H. S. & Bertrand, K. P. (1983). J. Bacteriol. 155, 768–775.  CrossRef CAS PubMed Web of Science Google Scholar
First citationMurshudov, G. N., Vagin, A. A. & Dodson, E. J. (1997). Acta Cryst. D53, 240–255.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationOtwinowski, Z. & Minor, W. (1997). Methods Enzymol. 276, 307–326.  CrossRef CAS PubMed Web of Science Google Scholar
First citationPandey, D. P. & Gerdes, K. (2005). Nucleic Acids Res. 33, 966–976.  Web of Science CrossRef PubMed CAS Google Scholar
First citationRadzikowski, J. L., Vedelaar, S., Siegel, D., Ortega, A. D., Schmidt, A. & Heinemann, M. (2016). Mol. Syst. Biol. 12, 882.  Web of Science CrossRef PubMed Google Scholar
First citationRobert, X. & Gouet, P. (2014). Nucleic Acids Res. 42, W320–W324.  Web of Science CrossRef CAS PubMed Google Scholar
First citationSchumacher, M. A., Balani, P., Min, J., Chinnam, N. B., Hansen, S., Vulić, M., Lewis, K. & Brennan, R. G. (2015). Nature, 524, 59–64.  Web of Science CrossRef CAS PubMed Google Scholar
First citationStancik, I. A., Šestak, M. S., Ji, B., Axelson-Fisk, M., Franjevic, D., Jers, C., Domazet-Lošo, T. & Mijakovic, I. (2018). J. Mol. Biol. 430, 27–32.  Web of Science CrossRef CAS PubMed Google Scholar
First citationVang Nielsen, S., Turnbull, K. J., Roghanian, M., Baerentsen, R., Semanjski, M., Brodersen, D. E., Macek, B. & Gerdes, K. (2019). mBio, 10, e01138-19.  Web of Science CrossRef PubMed Google Scholar
First citationVeening, J. W., Smits, W. K. & Kuipers, O. P. (2008). Annu. Rev. Microbiol. 62, 193–210.  Web of Science CrossRef PubMed CAS Google Scholar
First citationWen, Y., Behiels, E., Felix, J., Elegheert, J., Vergauwen, B., Devreese, B. & Savvides, S. N. (2014). Nucleic Acids Res. 42, 10134–10147.  Web of Science CrossRef CAS PubMed Google Scholar
First citationWood, T. K., Knabel, S. J. & Kwan, B. W. (2013). Appl. Environ. Microbiol. 79, 7116–7121.  Web of Science 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.

IUCrJ
Volume 9| Part 5| September 2022| Pages 625-631
ISSN: 2052-2525