Next Article in Journal
Antibacterial Secondary Metabolites from Marine-Derived Fungus Aspergillus sp. IMCASMF180035
Next Article in Special Issue
Optical DNA Mapping of Plasmids Reveals Clonal Spread of Carbapenem-Resistant Klebsiella pneumoniae in a Large Thai Hospital
Previous Article in Journal
Targeting Biofilm of MDR Providencia stuartii by Phages Using a Catheter Model
Previous Article in Special Issue
Sorting out the Superbugs: Potential of Sortase A Inhibitors among Other Antimicrobial Strategies to Tackle the Problem of Antibiotic Resistance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Targeting the ATP Synthase in Staphylococcus aureus Small Colony Variants, Streptococcus pyogenes and Pathogenic Fungi

1
Department of Veterinary and Animal Sciences, Faculty of Health and Medical Sciences, University of Copenhagen, Stigbøjlen 4, DK-1870 Frederiksberg C, Denmark
2
Department of Food Science and Technology, Faculty of Agriculture, Ferdowsi University of Mashhad, Mashhad 9177948974, Iran
*
Author to whom correspondence should be addressed.
Submission received: 26 February 2021 / Revised: 25 March 2021 / Accepted: 31 March 2021 / Published: 2 April 2021
(This article belongs to the Special Issue Solutions to Antimicrobial Resistance)

Abstract

:
The ATP synthase has been validated as a druggable target with the approval of the ATP synthase inhibitor, bedaquiline, for treatment of drug-resistant Mycobacterium tuberculosis, a bacterial species in which the ATP synthase is essential for viability. Gene inactivation studies have also shown that the ATP synthase is essential among Streptococci, and some studies even suggest that inhibition of the ATP synthase is a strategy for the elimination of Staphylococcus aureus small colony variants with deficiencies in the electron transport chain, as well as pathogenic fungi, such as Candida albicans. Here we investigated five structurally diverse ATP synthase inhibitors, namely N,N′-dicyclohexylcarbodiimide (DCCD), oligomycin A, tomatidine, resveratrol and piceatannol, for their growth inhibitory activity against the bacterial strains Streptococcus pyogenes, S. aureus and two isogenic small colony variants, as well as the pathogenic fungal species, C. albicans and Aspergillus niger. DCCD showed broad-spectrum inhibitory activity against all the strains (minimum inhibitory concentration (MIC) 2–16 µg/mL), except for S. aureus, where the ATP synthase is dispensable for growth. Contrarily, oligomycin A selectively inhibited the fungal strains (MIC 1–8 µg/mL), while tomatidine showed very potent, but selective, activity against small colony variants of S. aureus with compromised electron transport chain activity (MIC 0.0625 µg/mL). Small colony variants of S. aureus were also more sensitive to resveratrol and piceatannol than the wild-type strain, and piceatannol inhibited S. pyogenes at 16–32 µg/mL. We previously showed that transposon inactivation of the ATP synthase sensitizes S. aureus towards polymyxin B and colistin, and here we demonstrate that treatment with structurally diverse ATP synthase inhibitors sensitized S. aureus towards polymyxin B. Collectively, our data show that ATP synthase inhibitors can have selective inhibitory activity against pathogenic microorganisms in which the ATP synthase is essential. The data also show that the inhibition of the ATP synthase in Streptococcus pyogenes may be a new strategy for development of a narrow-spectrum antibiotic class. In other major bacterial pathogens, such as S. aureus and potentially Escherichia coli, where the ATP synthase is dispensable, the ATP synthase inhibitors may be applied in combination with antimicrobial peptides to provide new therapeutic options.

1. Introduction

The majority of the clinically available antibiotics interfere with a limited set of targets, including inhibition of cell wall biosynthesis, protein synthesis, DNA synthesis, RNA synthesis, folic acid metabolism and disruption of cell membrane integrity [1]. In the past decade, targeting bacterial energetics has been added to the list with the approval of the anti-mycobacterial antibiotic, bedaquiline, which targets the membrane-bound F1Fo-ATP synthase [2]. ATP synthesis by the membrane-bound ATP synthase is essential for both growing and nongrowing mycobacteria, including Mycobacterium tuberculosis [3] and Mycobacterium smegmatis [4]. Despite resemblance to the human mitochondrial ATP synthase, bedaquiline selectively inhibits the ATP synthase of Mycobacterial spp. [5].
The ATP synthase is also essential in Streptococci, where all eight genes encoding for subunits of the ATP synthase in Streptococcus sanguinis were identified as being essential in a genome-wide screen [6]. Three of the ATP synthase genes were subsequently confirmed to be essential in Streptococcus pneumoniae and Streptococcus mutants [6]. Furthermore, essentiality of the ATP synthase has been confirmed by others in S. pneumoniae [7,8] and Streptococcus pyogenes [9]. Streptococci lack a respiratory chain to establish the proton motive force and the ATP synthase is indispensable for establishing the proton motive force needed for various transport processes and to maintain pH homeostasis [6].
In contrast, the ATP synthase is not essential in the facultative anaerobic pathogen Staphylococcus aureus [10]. However, the ATP synthase was recently established as the target of tomatidine, which selectively inhibits growth of small colony variants with a compromised electron transport chain (ETC-SCVs), while having no inhibitory activity against wild-type cells [11]. ETC-SCVs often carry mutations in genes that contribute to the biosynthesis of the electron carrier menaquinone or the heme group in cytochromes [12]. Even though the ATP synthase is not essential for viability in S. aureus, the genetic inactivation of subunits of the ATP synthase leads to increased susceptibility towards different cationic antimicrobials, including aminoglycosides [13,14], polymyxins [15] and certain human antimicrobial peptides [16].
Besides tomatidine, multiple ATP synthase inhibitors have been identified, such as N,N′-dicyclohexylcarbodiimide (DCCD) and oligomycin A, which non-selectively inhibits ATP synthases in S. aureus, Escherichia coli and human mitochondria [17]. Furthermore, the polyphenolic compounds, resveratrol and piceatannol, inhibit the ATP synthase in E. coli [18]. Consistent with the genetic studies, the ATP synthase inhibitors tomatidine [19], DCCD [20] and resveratrol [21] sensitize S. aureus towards aminoglycosides, while resveratrol has been shown to sensitize S. aureus to polymyxin B [16].
The ATP synthase consists of multiple subunits, with an overall assembly of a membrane-embedded F0-domain and a cytoplasmic F1-domain. The composition of bacterial ATP synthases generally consists of the subunits α3β3γδε, constituting the F1-domain and the F0-domain that comprises the ab2c9–15 subunits [22]. Proton translocation across the lipid bilayer occurs at the F0 c-ring, an oligomer of the subunit c, which is encoded by atpE in bacteria [23].
Oligomycin A was originally discovered as an antifungal in 1954 [24] and its binding to the yeast mitochondrial ATP synthase subunit c-ring has been elucidated [25]. Oligomycin A effectively inhibits a range of pathogenic fungal species, including Candida albicans and Aspergillus niger [26], suggesting that the ATP synthase could also be an essential target in some pathogenic fungi.
In the present study, we investigated the impact of structurally diverse ATP synthase inhibitors on a wide range of microorganisms, namely Streptococcus pyogenes, Candida albicans and Aspergillus niger, as well as towards the S. aureus ETC-SCVs (menD::Tn and hemB::Tn mutants [10]). Our results point to interesting, novel therapeutic possibilities and suggest the ATP synthase may be a druggable target in many organisms.

2. Results and Discussion

2.1. DCCD Displays Broad-Spectrum Inhibitory Activity

The ATP synthase inhibitor DCCD showed broad-spectrum activity and inhibited growth of both the S. aureus ETC-SCVs, S. pyogenes and all fungal isolates (minimum inhibitory concentration (MIC) 2–16 µg/mL) (Table 1). DCCD did not inhibit the growth of the wild-type S. aureus, in which the ATP synthase is dispensable for growth [10]. Even though DCCD did not inhibit the growth of wild-type S. aureus (MIC > 128 µg/mL), at a subinhibitory concentration (16 µg/mL), this compound sensitized S. aureus towards polymyxin B (Table 2).
DCCD reacts covalently with carboxylic groups in both the F1 β-subunit and F0 subunit c [27]. In E. coli, DCCD reacts with the carboxylic group in aspartic acid at position 61 in the F0 subunit c [27], while this residue in the organisms that we tested was substituted with the carboxyl-containing amino acid, glutamic acid (Figure 1A, the position is marked with an asterisk and bold letters). This carboxylic group is directly involved in proton translocation and its modification with DCCD abolishes the function of F0 [28]. The carboxyl-containing amino acids in the subunit c are well conserved in bacterial, mitochondrial and chloroplast enzymes [11,29], indicating that DCCD could be a broad-spectrum ATP synthase inhibitor.
Even though DCCD non-selectively inhibits ATP synthases in human mitochondria, S. aureus and E. coli [11,17], which precludes it from clinical use, it remains a valuable tool for initial investigations on the implications of ATP synthase inhibition in various organisms.

2.2. Oligomycin A Selectively Inhibits Fungal Pathogens

Oligomycin A selectively inhibited the growth of the fungal pathogens at concentrations ranging from 1 to 8 µg/mL, while showing no inhibitory activity against the S. aureus ETC-SCVs nor S. pyogenes.
The oligomycin binding site was determined in the yeast Saccharomyces cerevisiae and involves interactions with nine amino acids [25]. The binding site spanned the carboxyl-containing amino acid of the subunit c of F0, namely Glu59, corresponding to Asp61 in E. coli. The amino acids of the S. cerevisiae subunit c involved in binding of oligomycin are highlighted in red in Figure 1A. The amino acids were completely conserved between S. cerevisiae and C. albicans, while a single amino acid was changed in the A. niger protein sequence. Only few of the residues were conserved in the bacterial isolates, which may explain the lack of inhibitory activity against S. pyogenes and S. aureus ETC-SCVs.
Even though oligomycin A had no inhibitory activity against S. aureus ETC-SCVs, oligomycin A still sensitized wild-type S. aureus to polymyxin B.

2.3. Tomatidine Is Highly Selective for S. aureus Small Colony Variants

Tomatidine was the most potent inhibitor against S. aureus ETC-SCVs, with the MIC being 0.0625 µg/mL, without having inhibitory activity against the wild-type S. aureus (Table 1), corroborating previous results of Mitchell and colleagues [30]. Contrarily, tomatidine did not inhibit growth of S. pyogenes nor the fungal strains (Table 1), which supports the observation that tomatidine primarily inhibits members of the order Bacillales, which includes pathogens such as S. aureus, Listeria monocytogenes and Bacillus anthrasis [31].
Tomatidine was also shown to target the F0 subunit c [11]. The exact binding site was not determined; however, resistance-conferring mutations were identified (Figure 1A, green markings highlight amino acids that upon mutation confer resistance to tomatidine in S. aureus ETC-SCVs [11]). Members of the order Bacillales show highly similar F0 subunit c sequences, and it was hypothesized by Boulet and colleagues that, for example, Streptococcus pneumoniae is not targeted by tomatidine because of low sequence identity of the target F0 subunit c between S. aureus and S. pneumoniae [11].
In Figure 1A, we show that the sequence similarity of F0 subunit c between S. aureus and S. pyogenes, C. albicans or A. niger is also low (31%–37% amino acid identity), which may also explain the lack of inhibitory activity of tomatidine against these strains.
Tomatidine also sensitized wild-type S. aureus to polymyxin B, similar to the other ATP synthase inhibitors, DCCD and oligomycin A, showing that structurally diverse ATP synthase inhibitors in general sensitize S. aureus to polymxin B.

2.4. Resveratrol and Piceatannol Inhibits S. aureus Small Colony Variants and S. pyogenes

Resveratrol and its hydroxylated analogue piceatannol, were 2–8 times more active against the S. aureus ETC-SCVs and S. pyogenes than the wild-type S. aureus. Piceatannol inhibited S. pyogenes at 16–32 µg/mL, while resveratrol and piceatannol did not prevent the growth of the fungal strains at the tested concentrations (MIC > 32 µg/mL). Resveratrol has been widely studied for its antibacterial and antifungal properties [32] and others have shown similar levels of inhibition of S. pyogenes by resveratrol [33]. Piceatannol is less studied for its antimicrobial properties than resveratrol; however, our data suggest that piceatannol is more potent than resveratrol, at least against S. pyogenes and S. aureus ETC-SCVs.
Resveratrol and piceatannol were determined to bind to the F1-domain in the bovine ATP synthase in a pocket between a β-, α- and the γ-subunit [34]. The residues involved in the interaction with resveratrol and piceatannol are highlighted in Figure 1B [34]. Resveratrol was also shown to inhibit the ATP synthase in E. coli and six residues of the eight residues involved in the interaction are conserved between the bovine and the E. coli sequence, suggesting a similar binding pocket in the E. coli F1-domain [18]. The residues that are important for the resveratrol/piceatannol interaction with the bovine F1-domain, and hypothesized to constitute a similar binding site in E. coli [18], are identical between E. coli, S. aureus and S. pyogenes (Figure 1), suggesting a similar binding pocket in the Gram-positive pathogens as in E. coli.

3. Conclusions

In this study, we showed that structurally diverse ATP synthase inhibitors display variability in the inhibitory activity against bacterial and fungal species, in which the ATP synthase is indispensable for growth. The ATP synthase inhibitor, DCCD, had the broadest spectrum of activity compared with the other ATP synthase inhibitors investigated. Tomatidine was the most active against S. aureus ETC-SCVs, while oligomycin A was the most potent antifungal inhibitor. The ATP synthase is essential in Streptococci [6,7,8,9], and we showed that resveratrol and piceatannol display intermediary inhibitory activity against two S. pyogenes isolates (Table 1). Others have also identified inhibitors of the ATP synthase to inhibit the growth of S. pneumoniae [17,35] and other Streptococcal species [35]. Collectively, these data suggest that the ATP synthase in Streptococcal species is a druggable target for novel antibiotics with the potential for narrow-spectrum activity.
Our study also supports the accumulating evidence that the ATP synthase can be targeted to sensitize S. aureus to polymyxins, and here we showed that structurally diverse ATP synthase inhibitors all worked as polymyxin B sensitizers (Table 2), corroborating previous studies with the ATP synthase inhibitors venturicidin A [36] and resveratrol [16]. Recent reports show that resveratrol can also sensitize several Gram-negative species, including E. coli and K. pneumoniae, to colistin; however, the mechanism of potentiation was not investigated in these studies [37,38]. It remains to be explored if, in general, other ATP synthase inhibitors sensitize Gram-negative bacteria to polymyxins, where this antibiotic class is considered as last-resort antibiotics [39].

4. Materials and Methods

4.1. Bacterial Strains, Growth Conditions and Chemicals

The bacterial strains used in this study are highlighted in Table 1. Chemicals used in this study include tomatidine hydrochloride (Merck KGaA, Darmstadt, Germany), oligomycin A (Santa Cruz Biotechnology, Dallas, TX, USA), N,N′-Dicyclohexylcarbodiimide (Merck KGaA, Darmstadt, Germany), piceatannol (Santa Cruz Biotechnology, Dallas, TX, USA) and resveratrol (Santa Cruz Biotechnology, Dallas, TX, USA).
All bacterial strains and the two Candida albicans isolates were routinely cultured at 37 °C in tryptic soy broth (TSB) or on tryptic soy agar (TSA), whereas Aspergillus niger was cultured in nutrient broth (NB).

4.2. Minimum Inhibitory Concentration

The minimum inhibitory concentrations (MICs) for the ATP synthase inhibitors were determined using the twofold broth microdilution assay.

4.3. Bacteria

Overnight cultures of the bacterial isolates were diluted in physiological saline (0.9% NaCl) to reach a turbidity of 0.5 McFarland (Sensititre® nephelometer and the Sensititre® McFarland Standard). Bacterial suspensions were adjusted to 5 × 105 CFU/mL in TSB broth containing twofold dilutions of the inhibitors in a final volume of 100 µL. The plates were incubated for 24 h at 37 °C without shaking. MIC was defined as the concentration of the agents that completely prevented visible growth. All experiments were performed with biological triplicates.

4.4. Fungi

From the agar plates, A. niger and C. albicans were diluted in physiological saline (0.9% NaCl) to reach an initial cell density of 106 CFU/mL. Fungal suspensions were adjusted to 1 × 105 CFU/mL in NB (A. niger) or TSB (C. albicans) solutions containing twofold dilutions of the inhibitors in a final volume of 100 µl. The plates were incubated for 48 h at 30 °C without shaking. MIC was defined as the concentration of the agents that completely prevented visible growth. All experiments were performed with biological triplicates.
Growth controls with the solvents (DMSO and ethanol) of the ATP synthase inhibitors were performed at the highest concentration (3.2%) that the microorganisms were exposed to.

4.5. Combination of ATP Synthase Inhibitors and Polymyxin B

The MIC for polymyxin B was determined using Etest (bioMérieux, France) on TSA plates in the absence and in the presence of ATP synthase inhibitors (DCCD, oligomycin A and tomatidine) supplemented at a single subinhibitory concentration (16 µg/mL). From overnight cultures, strains were diluted to approximately 108 CFU/mL and then distributed on TSA plates using a sterile cotton swab. MIC was determined upon incubation at 37 °C for 24 h. All experiments were performed with biological triplicates.

4.6. Protein Sequence Alignment of ATP Synthase Subunits

Protein sequences of the c-, γ-, α- and β-subunits from S. aureus JE2, S. pyogenes MGAS315, S. pyogenes SF370, E. coli K12, C. albicans ATCCMYA-2876 (sequences of tested strains could not be retrieved), A. niger CBS 513.88 (sequences of tested strain could not be retrieved), Saccharamyces cerevisiae ATCC204508 and Bos Taurus were retrieved from the KEGG GENOME Database. Multiple sequence alignment was conducted with Clustal Omega [40].

Author Contributions

M.V. and H.I. conceived and designed the study. Experiments were performed by M.V. and S.R.; M.V. and H.I. contributed to the analysis of data and drafting of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

Novo Nordisk Fonden NNF20OC0062531 and the Olav Thon Foundation.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

This work was supported by Olav Thon Foundation and Novo Nordisk Foundation NNF20OC0062531 to HI. The funding source had no influence on study design, data analysis or decision to submit the article for publication. We appreciate Professor emeritus Gunnar Lindahl for kindly providing the Streptococcus pyogenes strains.

Conflicts of Interest

The authors declare that there is no conflict of interest.

References

  1. Lewis, K. Platforms for antibiotic discovery. Nat. Rev. Drug Discov. 2013, 12, 371–387. [Google Scholar] [CrossRef] [PubMed]
  2. Chan, B.; Khadem, T.M.; Brown, J. A review of tuberculosis: Focus on bedaquiline. Am. J. Health Syst. Pharm. 2013, 70, 1984–1994. [Google Scholar] [CrossRef] [PubMed]
  3. Sassetti, C.M.; Boyd, D.H.; Rubin, E.J. Genes required for mycobacterial growth defined by high density mutagenesis. Mol. Microbiol. 2003, 48, 77–84. [Google Scholar] [CrossRef] [PubMed]
  4. Tran, S.L.; Cook, G.M. The F1Fo-ATP Synthase of Mycobacterium smegmatis Is Essential for Growth. J. Bacteriol. 2005, 187, 5023–5028. [Google Scholar] [CrossRef] [Green Version]
  5. Haagsma, A.C.; Abdillahi-Ibrahim, R.; Wagner, M.J.; Krab, K.; Vergauwen, K.; Guillemont, J.; Andries, K.; Lill, H.; Koul, A.; Bald, D. Selectivity of TMC207 towards Mycobacterial ATP Synthase Compared with That towards the Eukaryotic Homologue. Antimicrob. Agents Chemother. 2009, 53, 1290–1292. [Google Scholar] [CrossRef] [Green Version]
  6. Xu, P.; Ge, X.; Chen, L.; Wang, X.; Dou, Y.; Xu, J.Z.; Patel, J.R.; Stone, V.; Trinh, M.; Evans, K. Genome-wide essential gene identification in Streptococcus sanguinis. Sci. Rep. 2011, 1, 125. [Google Scholar] [CrossRef] [Green Version]
  7. Song, J.-H.; Ko, K.S.; Lee, J.-Y.; Baek, J.Y.; Oh, W.S.; Yoon, H.S.; Jeong, J.-Y.; Chun, J. Identification of Essential Genes in Streptococcus pneumoniae by Allelic Replacement Mutagenesis. Mol. Cells 2005, 19, 365–374. [Google Scholar]
  8. Ferrándiz, M.J.; de la Campa, A.G. The membrane-associated F0F1 ATPase is essential for the viability of Streptococcus pneumoniae. FEMS Microbiol. Lett. 2002, 212, 133–138. [Google Scholar] [CrossRef] [Green Version]
  9. Le Breton, Y.; Belew, A.T.; Valdes, K.M.; Islam, E.; Curry, P.; Tettelin, H.; Shirtliff, M.E.; El-Sayed, N.M.; McIver, K.S. Essential Genes in the Core Genome of the Human Pathogen Streptococcus pyogenes. Sci. Rep. 2015, 5, 9838. [Google Scholar] [CrossRef] [Green Version]
  10. Fey, P.D.; Endres, J.L.; Yajjala, V.K.; Widhelm, T.J.; Boissy, R.J.; Bose, J.L.; Bayles, K.W. A Genetic Resource for Rapid and Comprehensive Phenotype Screening of Nonessential Staphylococcus aureus Genes. mBio 2013, 4, e00537-12. [Google Scholar] [CrossRef] [Green Version]
  11. Boulet, M.L.; Isabelle, C.; Guay, I.; Brouillette, E.; Langlois, J.-P.; Jacques, P.-É.; Rodrigue, S.; Brzezinski, R.; Beauregard, P.B.; Bouarab, K. Tomatidine, a lead antibiotic molecule that targets Staphylococcus aureus ATP synthase subunit C. Antimicrob. Agents Chemother. 2018, 62, e02197-17. [Google Scholar] [CrossRef] [Green Version]
  12. Kahl, B.C.; Becker, K.; Löffler, B. Clinical Significance and Pathogenesis of Staphylococcal Small Colony Variants in Persistent Infections. Clin. Microbiol. Rev. 2016, 29, 401–427. [Google Scholar] [CrossRef] [Green Version]
  13. Vestergaard, M.; Leng, B.; Haaber, J.; Bojer, M.S.; Vegge, C.S.; Ingmer, H. Genome-Wide Identification of Antimicrobial Intrinsic Resistance Determinants in Staphylococcus aureus. Front. Microbiol. 2016, 7. [Google Scholar] [CrossRef] [Green Version]
  14. Grosser, M.R.; Paluscio, E.; Thurlow, L.R.; Dillon, M.M.; Cooper, V.S.; Kawula, T.H.; Richardson, A.R. Genetic requirements for Staphylococcus aureus nitric oxide resistance and virulence. PLoS Pathog. 2018, 14, e1006907. [Google Scholar] [CrossRef]
  15. Vestergaard, M.; Nøhr-Meldgaard, K.; Bojer, M.S.; Nielsen, C.K.; Meyer, R.L.; Slavetinsky, C.; Peschel, A.; Ingmer, H. Inhibition of the ATP synthase Eliminates the Intrinsic Resistance of Staphylococcus aureus towards Polymyxins. mBio 2017, 8, e01114-17. [Google Scholar] [CrossRef] [Green Version]
  16. Liu, L.; Beck, C.; Nøhr-Meldgaard, K.; Peschel, A.; Kretschmer, D.; Ingmer, H.; Vestergaard, M. Inhibition of the ATP synthase sensitizes Staphylococcus aureus towards human antimicrobial peptides. Sci. Rep. 2020, 10. [Google Scholar] [CrossRef]
  17. Balemans, W.; Vranckx, L.; Lounis, N.; Pop, O.; Guillemont, J.; Vergauwen, K.; Mol, S.; Gilissen, R.; Motte, M.; Lançois, D. Novel Antibiotics Targeting Respiratory ATP Synthesis in Gram-Positive Pathogenic Bacteria. Antimicrob. Agents Chemother. 2012, 56, 4131–4139. [Google Scholar] [CrossRef] [Green Version]
  18. Dadi, P.K.; Ahmad, M.; Ahmad, Z. Inhibition of ATPase activity of Escherichia coli ATP synthase by polyphenols. Int. J. Biol. Macromol. 2009, 45, 72–79. [Google Scholar] [CrossRef]
  19. Mitchell, G.; Lafrance, M.; Boulanger, S.; Séguin, D.L.; Guay, I.; Gattuso, M.; Marsault, É.; Bouarab, K.; Malouin, F. Tomatidine acts in synergy with aminoglycoside antibiotics against multiresistant Staphylococcus aureus and prevents virulence gene expression. J. Antimicrob. Chemother. 2012, 67, 559–568. [Google Scholar] [CrossRef] [Green Version]
  20. Miller, M.H.; Edberg, S.C.; Mandel, L.J.; Behar, C.F.; Steigbigel, N.H. Gentamicin Uptake in Wild-Type and Aminoglycoside-Resistant Small-Colony Mutants of Staphylococcus aureus. Antimicrob. Agents Chemother. 1980, 18, 722–729. [Google Scholar] [CrossRef] [Green Version]
  21. Nøhr-Meldgaard, K.; Ovsepian, A.; Ingmer, H.; Vestergaard, M. Resveratrol enhances the efficacy of aminoglycosides against Staphylococcus aureus. Int. J. Antimicrob. Agents 2018, 52, 390–396. [Google Scholar] [CrossRef]
  22. Guo, H.; Rubinstein, J.L. Cryo-EM of ATP synthases. Curr. Opin. Struct. Biol. 2018, 52, 71–79. [Google Scholar] [CrossRef]
  23. Rühle, T.; Leister, D. Assembly of F1F0-ATP synthases. Biochim. Biophys. Acta 2015, 1847, 849–860. [Google Scholar] [CrossRef] [Green Version]
  24. Smith, R.; Peterson, W.; McCoy, E. Oligomycin, a new antifungal antibiotic. Antibiot. Chemother. 1954, 4, 962–970. [Google Scholar]
  25. Symersky, J.; Osowski, D.; Walters, D.E.; Mueller, D.M. Oligomycin frames a common drug-binding site in the ATP synthase. Proc. Natl. Acad. Sci. USA 2012, 109, 13961–13965. [Google Scholar] [CrossRef] [Green Version]
  26. Lysenkova, L.N.; Saveljev, O.Y.; Grammatikova, N.E.; Tsvetkov, V.B.; Bekker, O.B.; Danilenko, V.N.; Dezhenkova, L.G.; Bykov, E.E.; Omelchuk, O.A.; Korolev, A.M. Verification of oligomycin A structure: Synthesis and biological evaluation of 33-dehydrooligomycin A. J. Antibiot. 2017, 70, 871–877. [Google Scholar] [CrossRef]
  27. Hong, S.; Pedersen, P.L. ATP Synthase and the Actions of Inhibitors Utilized To Study Its Roles in Human Health, Disease, and Other Scientific Areas. Microbiol. Mol. Biol. Rev. 2008, 72, 590–641. [Google Scholar] [CrossRef] [Green Version]
  28. Fillingame, R.H. H+ transport and coupling by the F 0 sector of the ATP synthase: Insights into the molecular mechanism of function. J. Bioenerg. Biomembr. 1992, 24, 485–491. [Google Scholar] [CrossRef]
  29. Arechaga, I.; Jones, P.C. The rotor in the membrane of the ATP synthase and relatives. FEBS Lett. 2001, 494, 1–5. [Google Scholar] [CrossRef] [Green Version]
  30. Mitchell, G.; Gattuso, M.; Grondin, G.; Marsault, É.; Bouarab, K.; Malouin, F. Tomatidine inhibits replication of Staphylococcus aureus small-colony variants in cystic fibrosis airway epithelial cells. Antimicrob. Agents Chemother. 2011, 55, 1937–1945. [Google Scholar] [CrossRef] [Green Version]
  31. Guay, I.; Boulanger, S.; Isabelle, C.; Brouillette, E.; Chagnon, F.; Bouarab, K.; Marsault, E.; Malouin, F. Tomatidine and analog FC04–100 possess bactericidal activities against Listeria, Bacillus and Staphylococcus spp. BMC Pharmacol. Toxicol. 2018, 19, 7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Vestergaard, M.; Ingmer, H. Antibacterial and antifungal properties of resveratrol. Int. J. Antimicrob. Agents 2019, 53, 716–723. [Google Scholar] [CrossRef] [PubMed]
  33. Macé, S.; Truelstrup Hansen, L.; Rupasinghe, H. Anti-Bacterial Activity of Phenolic Compounds against Streptococcus pyogenes. Medicines 2017, 4, 25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Gledhill, J.R.; Montgomery, M.G.; Leslie, A.G.; Walker, J.E. Mechanism of inhibition of bovine F1-ATPase by resveratrol and related polyphenols. Proc. Natl. Acad. Sci. USA 2007, 104, 13632–13637. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Wang, X.; Zeng, Y.; Sheng, L.; Larson, P.; Liu, X.; Zou, X.; Wang, S.; Guo, K.; Ma, C.; Zhang, G.; et al. A Cinchona Alkaloid Antibiotic That Appears To Target ATP Synthase in Streptococcus pneumoniae. J. Med. Chem. 2019, 62, 2305–2332. [Google Scholar] [CrossRef] [PubMed]
  36. Yarlagadda, V.; Medina, R.; Wright, G.D. Venturicidin A, A Membrane-active Natural Product Inhibitor of ATP synthase Potentiates Aminoglycoside Antibiotics. Sci. Rep. 2020, 10, 8134. [Google Scholar] [CrossRef]
  37. Liu, L.; Yu, J.; Shen, X.; Cao, X.; Zhan, Q.; Guo, Y.; Yu, F. Resveratrol enhances the antimicrobial effect of polymyxin B on Klebsiella pneumoniae and Escherichia coli isolates with polymyxin B resistance. BMC Microbiol. 2020, 20, 306. [Google Scholar] [CrossRef]
  38. Cannatelli, A.; Principato, S.; Colavecchio, O.L.; Pallecchi, L.; Rossolini, G.M. Synergistic Activity of Colistin in Combination With Resveratrol Against Colistin-Resistant Gram-Negative Pathogens. Front. Microbiol. 2018, 9, 1808. [Google Scholar] [CrossRef]
  39. Al-Tawfiq, J.A.; Laxminarayan, R.; Mendelson, M. How should we respond to the emergence of plasmid-mediated colistin resistance in humans and animals? Int. J. Infect. Dis. 2017, 54, 77–84. [Google Scholar] [CrossRef] [Green Version]
  40. Sievers, F.; Wilm, A.; Dineen, D.; Gibson, T.J.; Karplus, K.; Li, W.; Lopez, R.; McWilliam, H.; Remmert, M.; Söding, J. Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega. Mol. Syst. Biol. 2011, 7, 539. [Google Scholar] [CrossRef]
Figure 1. Amino acid sequence alignments of four ATP synthase subunits involved in the binding of tested ATP synthase inhibitors. (A) Amino acid sequence alignments of the F0 subunit c for bacterial and fungal isolates. The essential carboxyl-containing amino acid (glutamic acid (E) or aspartic acid (D)) for proton translocation is indicated with the asterisk and bold letters. Amino acids that upon mutation in S. aureus conferred resistance to tomatidine are marked in green [11], along with shared amino acids in the other microorganisms. The amino acid residues constituting the binding site of oligomycin A in S. cerevisiae are marked in red [25], along with shared amino acids in the other microorganisms. (B) Amino acid sequence alignments of short regions of the ATP synthase subunits α, β and γ involved in the binding of resveratrol to the F1-domain of the bovine ATP synthase [34], with corresponding regions of E. coli as previously aligned [18]. Here, alignments of corresponding sequences for S. aureus and S. pyogenes are added. Residues marked in green are involved in hydrophobic interactions with resveratrol/piceatannol, while residues marked in red provide additional nonpolar interactions [34]. The residues involved in resveratrol binding are identical between the E. coli, S. aureus and S. pyogenes sequences.
Figure 1. Amino acid sequence alignments of four ATP synthase subunits involved in the binding of tested ATP synthase inhibitors. (A) Amino acid sequence alignments of the F0 subunit c for bacterial and fungal isolates. The essential carboxyl-containing amino acid (glutamic acid (E) or aspartic acid (D)) for proton translocation is indicated with the asterisk and bold letters. Amino acids that upon mutation in S. aureus conferred resistance to tomatidine are marked in green [11], along with shared amino acids in the other microorganisms. The amino acid residues constituting the binding site of oligomycin A in S. cerevisiae are marked in red [25], along with shared amino acids in the other microorganisms. (B) Amino acid sequence alignments of short regions of the ATP synthase subunits α, β and γ involved in the binding of resveratrol to the F1-domain of the bovine ATP synthase [34], with corresponding regions of E. coli as previously aligned [18]. Here, alignments of corresponding sequences for S. aureus and S. pyogenes are added. Residues marked in green are involved in hydrophobic interactions with resveratrol/piceatannol, while residues marked in red provide additional nonpolar interactions [34]. The residues involved in resveratrol binding are identical between the E. coli, S. aureus and S. pyogenes sequences.
Antibiotics 10 00376 g001
Table 1. Minimum inhibitory concentrations of ATP synthase inhibitors against S. aureus small colony variants, S. pyogenes and the fungal pathogens C. albicans and A. niger.
Table 1. Minimum inhibitory concentrations of ATP synthase inhibitors against S. aureus small colony variants, S. pyogenes and the fungal pathogens C. albicans and A. niger.
MIC (µg/mL)
TomatidineOligomycin ADCCDPiceatannolResveratrol
S. aureus JE2>128>128>128128256
S. aureus JE2 menD::Tn [10]0.0625>1282–432128
S. aureus JE2 hemB::Tn [10]0.0625>1284–864128
S. pyogenes SF370>128>1288–1616–3264
S. pyogenes MGAS315>128>12881664
C. albicans ATCC 64548>324–84–8>32>32
C. albicans ATCC 64550>324–84–8>32>32
A. niger IBT 28144>321–216>32>32
Table 2. ATP synthase inhibitors sensitizes S. aureus JE2 to polymyxin B (PMB). Minimum inhibitory concentration (MIC) in µg/mL. TSA: tryptic soy agar.
Table 2. ATP synthase inhibitors sensitizes S. aureus JE2 to polymyxin B (PMB). Minimum inhibitory concentration (MIC) in µg/mL. TSA: tryptic soy agar.
Supplementation in PlatePMB MIC (Fold Reduction)
TSA128
TSA + 16 µg/mL Oligomycin A32 (4)
TSA + 16 µg/mL Tomatidine16–24 (8)
TSA + 16 µg/mL DCCD48 (2.6)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Vestergaard, M.; Roshanak, S.; Ingmer, H. Targeting the ATP Synthase in Staphylococcus aureus Small Colony Variants, Streptococcus pyogenes and Pathogenic Fungi. Antibiotics 2021, 10, 376. https://0-doi-org.brum.beds.ac.uk/10.3390/antibiotics10040376

AMA Style

Vestergaard M, Roshanak S, Ingmer H. Targeting the ATP Synthase in Staphylococcus aureus Small Colony Variants, Streptococcus pyogenes and Pathogenic Fungi. Antibiotics. 2021; 10(4):376. https://0-doi-org.brum.beds.ac.uk/10.3390/antibiotics10040376

Chicago/Turabian Style

Vestergaard, Martin, Sahar Roshanak, and Hanne Ingmer. 2021. "Targeting the ATP Synthase in Staphylococcus aureus Small Colony Variants, Streptococcus pyogenes and Pathogenic Fungi" Antibiotics 10, no. 4: 376. https://0-doi-org.brum.beds.ac.uk/10.3390/antibiotics10040376

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop