Next Article in Journal
Role of Urological Botulinum Toxin-A Injection for Overactive Bladder and Voiding Dysfunction in Patients with Parkinson’s Disease or Post-Stroke
Previous Article in Journal
A Review on Genotoxic and Genoprotective Effects of Biologically Active Compounds of Animal Origin
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Type II Toxin–Antitoxin Systems in Pseudomonas aeruginosa

Provincial Key Laboratory of Biotechnology, Key Laboratory of Resource Biology and Biotechnology in Western China, Ministry of Education, College of Life Sciences, Northwest University, Xi’an 710069, China
*
Authors to whom correspondence should be addressed.
Toxins 2023, 15(2), 164; https://doi.org/10.3390/toxins15020164
Submission received: 4 January 2023 / Revised: 3 February 2023 / Accepted: 14 February 2023 / Published: 17 February 2023
(This article belongs to the Section Bacterial Toxins)

Abstract

:
Toxin–antitoxin (TA) systems are typically composed of a stable toxin and a labile antitoxin; the latter counteracts the toxicity of the former under suitable conditions. TA systems are classified into eight types based on the nature and molecular modes of action of the antitoxin component so far. The 10 pairs of TA systems discovered and experimentally characterised in Pseudomonas aeruginosa are type II TA systems. Type II TA systems have various physiological functions, such as virulence and biofilm formation, protection host against antibiotics, persistence, plasmid maintenance, and prophage production. Here, we review the type II TA systems of P. aeruginosa, focusing on their biological functions and regulatory mechanisms, providing potential applications for the novel drug design.
Key Contribution: In this review; we mainly summarise the current knowledge of the functions and regulatory mechanisms of type II TA systems in P. aeruginosa.

1. Introduction

Toxin–antitoxin (TA) systems were first discovered on a conjugative plasmid in 1983, functioning as plasmid maintenance systems by post-segregational killing (PSK) [1]. They are small genetic modules composed of a stable toxin and a labile cognate antitoxin. Currently, eight types of TA systems (types I–VIII) are classified based on the nature of the antitoxin and the molecular mode of action. The toxins of all known TA systems are proteins, with the exception of type VIII toxins (RNAs); the antitoxins are RNA or protein [2]. In type I, III, and VIII TA systems, antitoxins are RNAs, and type I and III antitoxins bind to toxin mRNAs and proteins, thereby inhibiting toxin expression and toxicity, respectively [2]. Type VIII antitoxins and toxins are RNAs that bind to each other to repress RNA toxin expression [3]. The type II, IV, V, VI, and VII TA system toxins and antitoxins are proteins. Antitoxin proteins block the toxicity of the toxins via the following mechanisms: direct protein–protein interactions (type II) [4,5,6], compete with and remove the toxin from its target (type IV) [7,8], cleave the toxin mRNA (type V) [9], promote toxin degradation (type VI) [10], or regulate post-translational chemical modifications (type VII) [11,12].
The toxin inhibits an essential cellular process (kills cells or inhibits their growth), while the antitoxin counteracts the toxicity of its cognate toxin [2,13]. TA systems are widespread in bacterial and archaeal chromosomes and mobile genetic elements (plasmids, prophages, transposons, and integrate and conjugate elements) and have diverse roles in bacterial physiology and pathogenicity [2]. The plasmid-encoded PrpT/PrpA TA system controls the plasmid copy number; these new insights improve the understanding of the TA systems in the role of bacterial physiology [14]. Among all types of TA systems, type II TA systems are highly prevalent in bacterial genomes and have been evaluated extensively. They are implicated in the maintenance of genetic material, virulence and pathogenesis, biofilm formation, phage inhibition, and stress response [15,16].
Pseudomonas aeruginosa is one of the ESKAPE pathogens (Enterococcus faecium, Staphylococcus aureus, Klebsiella pneumoniae, Acinetobacter baumannii, Pseudomonas aeruginosa, and Enterobacter sp.) and is a critical priority pathogen [17]. Dai et al. found 16,963 TA gene hits in 5432 genomic sequences of P. aeruginosa from the NCBI Genome database, most of which were in chromosome sequences (94%) and belonged to type II TA systems (16,808 hits) [18]. To date, at least 10 pairs of type II TA systems have been experimentally characterised in P. aeruginosa, 8 of which are located on the chromosome [19,20,21,22,23,24,25,26], whereas PumA/PumB and PfiT/PfiA were identified in plasmid pUM505 and prophage Pf4, respectively [27,28]. These TA systems have multiple biological functions and regulatory mechanisms, which makes TA research more interesting and attractive. Here, we summarise recent findings on the functions and regulatory mechanism of type II TA systems in P. aeruginosa.

2. Transcriptional Regulation of Type II TA Systems in P. aeruginosa

Ten pairs of type II TA systems have been experimentally verified in P. aeruginosa (Table 1), as follows: HigB/HigA [19], ParD/ParE [20], HicA/HicB [21], RelB/RelE [22], ResA/XreA [23], PrrT/PrrA [24], CrlT/CrlA [25], PacT/PacA [26], PumA/PumB [27], and PfiT/PfiA [28]. In type II TA systems, genes encoding the toxin and antitoxin are within a single operon (Figure 1). In the pfiTA, parDE, resA/xreA, prrTA, crlTA, and relBE TA systems, the antitoxin gene precedes the toxin gene, and the whole operon is transcribed by a single promoter. By contrast, in the higBA, hicAB, pumAB, and pacTA TA systems, the toxin gene is upstream of the antitoxin gene; of the two promoters, the first promoter is responsible for the expression of the whole operon, and the second promoter controls the expression of the antitoxin. TA system expression is tightly autoregulated at the transcriptional level [6]. Antitoxins are the critical regulator and typically comprise a DNA-binding domain that recognises and binds to palindromic sequences in the promoter regions of the cognate TA operon [6]. The antitoxin alone or the TA complex as a transcriptional repressor autoregulates TA operon expression [29]. The higBA, crlTA, prrTA, and resA/xreA TA system toxins are negatively regulated by the corresponding antitoxins. In the pfiTA TA system, pfiT and pfiA are co-transcribed, and the PfiTA complex binds to the palindrome sequence (5′-AATTCN5GTTAA-3′) overlapping the −35 region of the pfiTA promoter, thereby repressing the expression of the pfiTA operon [28]. Interestingly, the relBE operon is regulated by both the antitoxin RelB and the TA complex RelBE; antitoxin RelB is a weak repressor, whereas RelE is an efficient co-repressor to further increase transcriptional regulation [30,31]. However, the transcriptional regulatory mechanisms of the pumAB, pacTA, parDE, and hicAB TA systems in P. aeruginosa are unclear. The stoichiometries of TA complexes depend on the toxin/antitoxin ratio and affect the affinity for the operon. In the presence of excess antitoxin relative to a toxin, the operon is weakly repressed. However, excess toxin leads to the formation of saturated TA complexes, which cannot bind operons [6,32,33].

3. Biological Functions of Type II TA Systems in P. aeruginosa

In P. aeruginosa, most type II TA systems are encoded on the chromosome [19,20,21,22,23,24,25,26], the exceptions being PfiT/PfiA and PumA/PumB (phage and plasmid, respectively) [27,28]. These TA systems have diverse biological functions, such as virulence and biofilm formation [34,35,36], protection host against antibiotics [20], persistence [23], plasmid maintenance [27], and prophage production [25,37] (Figure 2).

3.1. Virulence and Biofilm Formation

Type II TA systems are involved in the virulence and biofilm formation of bacteria. Many chronic infections with pathogenic bacteria are associated with biofilm formation [38]. Biofilms cause chronic infection because it protects bacteria against a wide range of antimicrobial agents and enhances bacteria’s adaptable ability to survive in different environments [39,40]. The cumulative evidence suggests that TA systems are involved in biofilm formation in P. aeruginosa [41,42].
HigBA is prevalent in P. aeruginosa clinical isolates and was the first TA system characterised in this species [19,22]. The toxin HigB regulates multiple virulence factors, such as pyocyanin, pyochelin, swarming motility, and biofilm formation [19] (Figure 3). The higA operon is induced by treatment with ciprofloxacin. HigB is involved in ciprofloxacin-induced the formation of persister cell and activates the expression of type III secretion system (T3SS) genes [43]. HigB modulates the expression of the T3SS genes and biofilm formation by regulating the expression of cyclic di-GMP (c-di-GMP) hydrolysis genes [35]. The antitoxin HigA functions as a transcriptional repressor of mvfR, exsA, and amrZ by directly binding to their promoter regions, and it controls pyocyanin synthesis, T3SS, and type VI secretion system (T6SS) expression [34,44]. mvfR is a key virulence-related regulator of P. aeruginosa that activates the expression of the phnAB and pqsA-E operons [45]. exsA is the master regulator that activates the expression of all the T3SS genes [46]. amrZ is a global T6SS transcriptional regulator that activates H1-T6SS and H3-T6SS but represses H2-T6SS expression by binding to the promoter regions of T6SS genes [47].
The RelBE TA system was discovered in the chromosome of Escherichia coli K-12 [48,49], and relBE genes were detected in 100% of P. aeruginosa [50]. RelBE TA systems are associated with biofilm formation, antibiotic resistance, oxidative stress, and persistence [51]. relBE expression is increased in isolates that are sensitive, compared with isolates that are resistant to aztreonam [50]. Zadeh et al. showed that the expression of relBE promotes persister cell formation in biofilms in the presence of ciprofloxacin and colistin in P. aeruginosa [52]. RelBE TA systems control biofilm formation by indirectly regulating the expression of biofilm-associated genes [36].
The PrrTA TA system is encoded by the chromosome of the clinical P. aeruginosa isolate 39016, 3504 bp upstream of prophage att sites [24]. Deletion of prrA significantly increased biofilm formation and reduced motility [24]. The amrZ transcript level was significantly decreased in a ΔprrA mutant, a global regulator of genes associated with a virulence that indirectly modulates the c-di-GMP level by repressing diguanylate cyclase genes or activating phosphodiesterase genes [53,54]. Moreover, AmrZ regulates bacterial motility and alginate synthesis [55]. In addition, the prrTA system is involved in prophage regulation and production. Most prophage genes are significantly downregulated in the ΔprrA mutant, and phage production and infectivity are significantly lower than those in the wild type [24].
The novel type II TA system PacTA has been characterised by P. aeruginosa PA14. The acetyltransferase toxin PacT inhibits the DNA-binding activity of Fur (ferric uptake regulator) to maintain iron homeostasis by directly binding to the HTH domain of Fur [26]. Fur is a central transcriptional repressor that modulates iron uptake processes by repressing the expression of genes responsible for iron acquisition and storage [56]. Moreover, PacTA contributes to pyocyanin biosynthesis and biofilm formation in P. aeruginosa. Compared with the wild-type, pyocyanin production and biofilm formation were markedly decreased in the ΔpacT and ΔpacTA mutants [26]. Fur represses biofilm formation in pathogens such as Yersinia pestis [57] and Stenotrophomonas maltophilia [58]; PacTA may contribute to biofilm formation by the path. PacTA also contributes to the virulence of P. aeruginosa; compared with the wild type, the ΔpacT and ΔpacTA mutants caused lower mortality.

3.2. Protection Host against Antibiotics

The ParDE TA system was characterised in plasmid RK2 as a plasmid stabilisation element [59]. This TA system is distributed in the IncI and IncF-type antibiotic resistance and virulence plasmids in E. coli and Salmonella species [60]. ParDE confers a survival advantage to the host under antibiotic and other stress conditions [60]. In P. aeruginosa, the toxin ParE promoted survival in the presence of quinolone antibiotics, e.g., ciprofloxacin, levofloxacin, and novobiocin [20]. However, higher concentrations of ParE decrease cell viability and alter cell morphology by inhibiting DNA gyrase [20]. ParDE is the only TA system capable of switching from a protective to a toxic effect depending on the toxin concentration [20]. This is the first example that the TA system can exert either protective or toxic effects, depending on the amount of toxin present. Nine chromosomal ParE toxins from various bacterial species alter cellular morphology from rods to filaments, consistent with disruption of DNA topology [61]. This phenotype is a marker of ParE toxin activity.

3.3. Persistence

Persister cells are slow-growing or growth-arrested cells that can resume growth after the stress disappears [62]. Persister formation is related to antibiotic tolerance and to refractory and chronic infections. Persistence is associated with TA systems, and TA systems play a major role in persistence formation [63]. The ResA/XreA (PA14_51010/PA14_51020) TA system was discovered in P. aeruginosa strain PA14. The toxin ResA belongs to the RES family, and the antitoxin XreA is a xenobiotic response element (Xre) [23]. ResA/XreA affects bacterial survival in the presence of tobramycin and ciprofloxacin [23]. Overexpression of ResA promotes persistence and reduces the intracellular NAD+ level, a key component of the respiratory machinery. Therefore, a reduced intracellular NAD+ level could increase antibiotic resistance. Overproduction of NAD+ counteracts the effect of overexpression of ResA on persister formation, suggesting that ResA/XreA contributes to persister formation by reducing the intracellular NAD+ level [23]. However, the deletion of resA did not affect persister formation due to additional TA systems or other determinants involved in persister formation.

3.4. Plasmid Maintenance

Plasmid maintenance is a function of plasmid-encoded type II TA systems [15]. The two earliest identified TA systems, ccdAB and hok-sok, stabilise plasmids by PSK [1,64]. PSK systems are addiction systems that encode long-lived toxins and short-lived antitoxins, ensuring plasmid stability by killing plasmid-free daughter cells [15]. The functions of plasmid-encoded TA systems in P. aeruginosa are unclear. The PumAB TA system is encoded from the conjugative plasmid pUM505 of P. aeruginosa strain PUM503 [27]. pUM505 harbours a mercury resistance operon (merRTPFADE), chromate resistance operon (chrBAC), and ciprofloxacin resistance gene (crpP) [65]. Non-pumAB-carrying plasmids decreased in number after 216 generations and were undetected after 432 generations; by contrast, pumAB-carrying plasmids were not significantly affected [27]. Therefore, pumAB can confer post-segregational plasmid stability via PSK [65]. No mechanism other than PSK has been confirmed in P. aeruginosa. Additionally, the PumA toxin confers P. aeruginosa virulence and increased Caenorhabditis elegans and mouse mortality rates [66].

3.5. Phage Production

Prophages and satellite prophages are widely distributed among bacteria, in which they confer various phenotypic traits to their hosts, including pathogenicity [67], antibiotic tolerance [68], biofilm formation [41], and general stress [41]. Prophage-borne TA systems have been discovered in E. coli [69] and Shewanella oneidensis [70]. TA systems are also found to be responsible for controlling the production of prophages. The PfiTA TA system is encoded by the filamentous Pf4 prophage of the P. aeruginosa PAO1 strain [28]. The Pf4 prophage modulates physiology and virulence, affects biofilm matrix composition and structure, and enhances bacterial survival and antibiotic resistance [71,72]. Deletion of the toxin gene pfiT increased Pf4 phage production by approximately 100,000-fold compared with the wild type. In the ∆pfiT mutant, the phage replication initiation gene PA0727 was induced but not the phage repressor gene pf4r. Therefore, PfiT regulates Pf4 phage production by inducing the expression of the replication initiation gene PA0727 [28]. In addition, the toxin PfiT is involved in phage immunity and coordinates the phage repressor Pf4r in conferring immunity to Pf4 [28]. The chromosome-encoded CrlTA TA system in P. aeruginosa WK172 is involved in fighting against phage infection. In excess, the prophage Cro-like antitoxin CrlA inhibited infection by lytic Pseudomonas phages (PAP-L5, PAOP5, PAP8, and QDWS) because CrlA inhibited phage replication by binding to crlTA palindrome-like sequences in the phage genome [25]. The CrlTA complex confers phage resistance to lytic phages PAP8 and QDWS; however, the underlying molecular mechanism is unclear.

4. Potential Application of Type II TA Systems in P. aeruginosa

Bacterial TA systems have considerable biotechnological potential [2,73,74]. For instance, a novel antibiotic that targets specific pathogens can be developed based on toxin stability. Moreover, antitoxin or TA complexes could repress virulence gene expression, protect against stresses, and inhibit phage infection. TA systems have much potential for the development of novel antibacterials [73,74,75,76,77,78].
TA systems can be used as positive selectable markers for ensuring the stability of mobile genetic materials [2]. An example is TA systems as positive selection vectors. A toxin gene in such a vector will be inactivated by the insertion of a segment of foreign DNA; therefore, only bacteria harbouring the recombinant vector will proliferate [1]. Another example is plasmid stabilisation. The antitoxin gene, under the control of a constitutive promoter, is cloned into a plasmid. The toxin gene under the control of a promoter strongly repressed by the antitoxin protein is cloned into the bacterial chromosome. Therefore, a bacterial cell lacking the vector will not survive. This allows plasmid stabilisation without antibiotics [79].

5. Concluding Remarks

Although TA systems have been widely distributed and well studied in pathogenic bacteria for decades, in P. aeruginosa, the biological functions of TA systems have been focused on recently. Ten pairs of type II TA systems have been identified in P. aeruginosa. In P. aeruginosa, type II TA systems enhance fitness and have multiple biological functions. For instance, the HigBA TA system is associated with virulence and biofilm formation, persister formation, T3SS, and T6SS. Interestingly, the ParDE TA system confers protective or toxic effects depending on the toxin concentration, but the molecular basis remains unknown for the switch from protective effects at low concentrations to toxic effects at higher concentrations. The HicAB TA system has known functions in other bacteria, e.g., virulence, persister cell formation, and stress responses [80,81]; however, the HicAB system has no known function in P. aeruginosa [21].
The antitoxin can autoregulate TA operon transcription and also regulate other key regulatory genes. The antitoxin is a negative regulator and participates in repressing the expression of other key regulatory genes [82,83,84]. In P. aeruginosa, the antitoxin HigA regulates the transcriptional regulatory genes such as mvfR [44], exsA [34], and amrZ [34], which are involved in regulating virulence expression, T3SS, and T6SS, respectively. More novel type II TA systems are found in clinical isolates; however, their roles in clinical environments are unclear. Further investigations should focus on the medical applications of TA systems.

Author Contributions

All authors contributed to this review. Conceptualization, M.L. and Y.Z.; writing—original preparation and draft, M.L., T.W. and Y.Z.; literature researching, M.L., N.G., G.S., Y.H. and L.W.; writing—review and editing, M.L., T.W. and Y.Z.; supervision, T.W. and Y.Z.; funding acquisition, Y.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a grant from the Natural Science Foundation of Shaanxi Province of China (grant number 2021JZ-42).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ogura, T.; Hiraga, S. Mini-F plasmid genes that couple host cell division to plasmid proliferation. Proc. Natl. Acad. Sci. USA 1983, 80, 4784–4788. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Jurenas, D.; Fraikin, N.; Goormaghtigh, F.; Van Melderen, L. Biology and evolution of bacterial toxin-antitoxin systems. Nat. Rev. Microbiol. 2022, 20, 335–350. [Google Scholar] [CrossRef] [PubMed]
  3. Li, M.; Gong, L.; Cheng, F.; Yu, H.; Zhao, D.; Wang, R.; Wang, T.; Zhang, S.; Zhou, J.; Shmakov, S.A.; et al. Toxin-antitoxin RNA pairs safeguard CRISPR-Cas systems. Science 2021, 372, eabe5601. [Google Scholar] [CrossRef] [PubMed]
  4. Tam, J.E.; Kline, B.C. The F plasmid ccd autorepressor is a complex of CcdA and CcdB proteins. Mol. Genet. Genom. 1989, 219, 26–32. [Google Scholar] [CrossRef]
  5. Jurenas, D.; Van Melderen, L.; Garcia-Pino, A. Mechanism of regulation and neutralization of the AtaR-AtaT toxin-antitoxin system. Nat. Chem. Biol. 2019, 15, 285–294. [Google Scholar] [CrossRef] [Green Version]
  6. Fraikin, N.; Goormaghtigh, F.; Van Melderen, L. Type II Toxin-Antitoxin Systems: Evolution and Revolutions. J. Bacteriol. 2020, 202, e00763-19. [Google Scholar] [CrossRef] [Green Version]
  7. Jimmy, S.; Saha, C.K.; Kurata, T.; Stavropoulos, C.; Oliveira, S.R.A.; Koh, A.; Cepauskas, A.; Takada, H.; Rejman, D.; Tenson, T.; et al. A widespread toxin-antitoxin system exploiting growth control via alarmone signaling. Proc. Natl. Acad. Sci. USA 2020, 117, 10500–10510. [Google Scholar] [CrossRef]
  8. Jankevicius, G.; Ariza, A.; Ahel, M.; Ahel, I. The Toxin-Antitoxin System DarTG Catalyzes Reversible ADP-Ribosylation of DNA. Mol. Cell 2016, 64, 1109–1116. [Google Scholar] [CrossRef] [Green Version]
  9. Wang, X.; Lord, D.M.; Cheng, H.Y.; Osbourne, D.O.; Hong, S.H.; Sanchez-Torres, V.; Quiroga, C.; Zheng, K.; Herrmann, T.; Peti, W.; et al. A new type V toxin-antitoxin system where mRNA for toxin GhoT is cleaved by antitoxin GhoS. Nat. Chem. Biol. 2012, 8, 855–861. [Google Scholar] [CrossRef] [Green Version]
  10. Aakre, C.D.; Phung, T.N.; Huang, D.; Laub, M.T. A bacterial toxin inhibits DNA replication elongation through a direct interaction with the beta sliding clamp. Mol. Cell 2013, 52, 617–628. [Google Scholar] [CrossRef] [Green Version]
  11. Songailiene, I.; Juozapaitis, J.; Tamulaitiene, G.; Ruksenaite, A.; Sulcius, S.; Sasnauskas, G.; Venclovas, C.; Siksnys, V. HEPN-MNT Toxin-Antitoxin System: The HEPN Ribonuclease Is Neutralized by OligoAMPylation. Mol. Cell 2020, 80, 955–970.e957. [Google Scholar] [CrossRef] [PubMed]
  12. Yu, X.; Gao, X.; Zhu, K.; Yin, H.; Mao, X.; Wojdyla, J.A.; Qin, B.; Huang, H.; Wang, M.; Sun, Y.C.; et al. Characterization of a toxin-antitoxin system in Mycobacterium tuberculosis suggests neutralization by phosphorylation as the antitoxicity mechanism. Commun. Biol. 2020, 3, 216. [Google Scholar] [CrossRef] [PubMed]
  13. Harms, A.; Brodersen, D.E.; Mitarai, N.; Gerdes, K. Toxins, Targets, and Triggers: An Overview of Toxin-Antitoxin Biology. Mol. Cell 2018, 70, 768–784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Ni, S.; Li, B.; Tang, K.; Yao, J.; Wood, T.K.; Wang, P.; Wang, X. Conjugative plasmid-encoded toxin-antitoxin system PrpT/PrpA directly controls plasmid copy number. Proc. Natl. Acad. Sci. USA 2021, 118, e2011577118. [Google Scholar] [CrossRef] [PubMed]
  15. Kamruzzaman, M.; Wu, A.Y.; Iredell, J.R. Biological Functions of Type II Toxin-Antitoxin Systems in Bacteria. Microorganisms 2021, 9, 1276. [Google Scholar] [CrossRef] [PubMed]
  16. Qiu, J.; Zhai, Y.; Wei, M.; Zheng, C.; Jiao, X. Toxin-antitoxin systems: Classification, biological roles, and applications. Microbiol. Res. 2022, 264, 127159. [Google Scholar] [CrossRef]
  17. De Oliveira, D.M.P.; Forde, B.M.; Kidd, T.J.; Harris, P.N.A.; Schembri, M.A.; Beatson, S.A.; Paterson, D.L.; Walker, M.J. Antimicrobial Resistance in ESKAPE Pathogens. Clin. Microbiol. Rev. 2020, 33, e00181-19. [Google Scholar] [CrossRef]
  18. Dai, Z.; Wu, T.; Xu, S.; Zhou, L.; Tang, W.; Hu, E.; Zhan, L.; Chen, M.; Yu, G. Characterization of toxin-antitoxin systems from public sequencing data: A case study in Pseudomonas aeruginosa. Front. Microbiol. 2022, 13, 951774. [Google Scholar] [CrossRef]
  19. Wood, T.L.; Wood, T.K. The HigB/HigA toxin/antitoxin system of Pseudomonas aeruginosa influences the virulence factors pyochelin, pyocyanin, and biofilm formation. Microbiologyopen 2016, 5, 499–511. [Google Scholar] [CrossRef]
  20. Muthuramalingam, M.; White, J.C.; Murphy, T.; Ames, J.R.; Bourne, C.R. The toxin from a ParDE toxin-antitoxin system found in Pseudomonas aeruginosa offers protection to cells challenged with anti-gyrase antibiotics. Mol. Microbiol. 2019, 111, 441–454. [Google Scholar] [CrossRef] [Green Version]
  21. Li, G.; Shen, M.; Lu, S.; Le, S.; Tan, Y.; Wang, J.; Zhao, X.; Shen, W.; Guo, K.; Yang, Y.; et al. Identification and Characterization of the HicAB Toxin-Antitoxin System in the Opportunistic Pathogen Pseudomonas aeruginosa. Toxins 2016, 8, 113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Savari, M.; Rostami, S.; Ekrami, A.; Bahador, A. Characterization of Toxin-Antitoxin (TA) Systems in Pseudomonas aeruginosa Clinical Isolates in Iran. Jundishapur J. Microbiol. 2016, 9, e26627. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Zhou, J.; Li, S.; Li, H.; Jin, Y.; Bai, F.; Cheng, Z.; Wu, W. Identification of a Toxin-Antitoxin System That Contributes to Persister Formation by Reducing NAD in Pseudomonas aeruginosa. Microorganisms 2021, 9, 753. [Google Scholar] [CrossRef] [PubMed]
  24. Shmidov, E.; Lebenthal-Loinger, I.; Roth, S.; Karako-Lampert, S.; Zander, I.; Shoshani, S.; Danielli, A.; Banin, E. PrrT/A, a Pseudomonas aeruginosa Bacterial Encoded Toxin-Antitoxin System Involved in Prophage Regulation and Biofilm Formation. Microbiol. Spectr. 2022, 10, e01182-22. [Google Scholar] [CrossRef]
  25. Ni, M.; Lin, J.; Gu, J.; Lin, S.; He, M.; Guo, Y. Antitoxin CrlA of CrlTA Toxin-Antitoxin System in a Clinical Isolate Pseudomonas aeruginosa Inhibits Lytic Phage Infection. Front. Microbiol. 2022, 13, 892021. [Google Scholar] [CrossRef]
  26. Song, Y.; Zhang, S.; Ye, Z.; Song, Y.; Chen, L.; Tong, A.; He, Y.; Bao, R. The novel type II toxin-antitoxin PacTA modulates Pseudomonas aeruginosa iron homeostasis by obstructing the DNA-binding activity of Fur. Nucleic Acids Res. 2022, 50, 10586–10600. [Google Scholar] [CrossRef]
  27. Hernandez-Ramirez, K.C.; Chavez-Jacobo, V.M.; Valle-Maldonado, M.I.; Patino-Medina, J.A.; Diaz-Perez, S.P.; Jacome-Galarza, I.E.; Ortiz-Alvarado, R.; Meza-Carmen, V.; Ramirez-Diaz, M.I. Plasmid pUM505 encodes a Toxin-Antitoxin system conferring plasmid stability and increased Pseudomonas aeruginosa virulence. Microb. Pathog. 2017, 112, 259–268. [Google Scholar] [CrossRef]
  28. Li, Y.; Liu, X.; Tang, K.; Wang, W.; Guo, Y.; Wang, X. Prophage encoding toxin/antitoxin system PfiT/PfiA inhibits Pf4 production in Pseudomonas aeruginosa. Microb. Biotechnol. 2020, 13, 1132–1144. [Google Scholar] [CrossRef] [Green Version]
  29. Hayes, F.; Van Melderen, L. Toxins-antitoxins: Diversity, evolution and function. Crit. Rev. Biochem. Mol. Biol. 2011, 46, 386–408. [Google Scholar] [CrossRef]
  30. Moreno-Cordoba, I.; Chan, W.T.; Nieto, C.; Espinosa, M. Interactions of the Streptococcus pneumoniae Toxin-Antitoxin RelBE Proteins with Their Target DNA. Microorganisms 2021, 9, 851. [Google Scholar] [CrossRef]
  31. Moreno-Cordoba, I.; Diago-Navarro, E.; Barendregt, A.; Heck, A.J.; Alfonso, C.; Diaz-Orejas, R.; Nieto, C.; Espinosa, M. The toxin-antitoxin proteins relBE2Spn of Streptococcus pneumoniae: Characterization and association to their DNA target. Proteins 2012, 80, 1834–1846. [Google Scholar] [CrossRef] [PubMed]
  32. Garcia-Pino, A.; Balasubramanian, S.; Wyns, L.; Gazit, E.; De Greve, H.; Magnuson, R.D.; Charlier, D.; van Nuland, N.A.; Loris, R. Allostery and intrinsic disorder mediate transcription regulation by conditional cooperativity. Cell 2010, 142, 101–111. [Google Scholar] [CrossRef] [PubMed]
  33. Overgaard, M.; Borch, J.; Jorgensen, M.G.; Gerdes, K. Messenger RNA interferase RelE controls relBE transcription by conditional cooperativity. Mol. Microbiol. 2008, 69, 841–857. [Google Scholar] [CrossRef] [PubMed]
  34. Song, Y.; Zhang, S.; Luo, G.; Shen, Y.; Li, C.; Zhu, Y.; Huang, Q.; Mou, X.; Tang, X.; Liu, T.; et al. Type II Antitoxin HigA Is a Key Virulence Regulator in Pseudomonas aeruginosa. ACS Infect. Dis. 2021, 7, 2930–2940. [Google Scholar] [CrossRef]
  35. Zhang, Y.; Xia, B.; Li, M.; Shi, J.; Long, Y.; Jin, Y.; Bai, F.; Cheng, Z.; Jin, S.; Wu, W. HigB Reciprocally Controls Biofilm Formation and the Expression of Type III Secretion System Genes through Influencing the Intracellular c-di-GMP Level in Pseudomonas aeruginosa. Toxins 2018, 10, 424. [Google Scholar] [CrossRef] [Green Version]
  36. Mahmoudi, M.; Sadeghifard, N.; Maleki, A.; Yeo, C.C.; Ghafourian, S. relBE Toxin-antitoxin System as a Reliable Anti-biofilm Target in Pseudomonas aeruginosa. J. Appl. Microbiol. 2022, 133, 683–695. [Google Scholar] [CrossRef]
  37. Zander, I.; Shmidov, E.; Roth, S.; Ben-David, Y.; Shoval, I.; Shoshani, S.; Danielli, A.; Banin, E. Characterization of PfiT/PfiA toxin-antitoxin system of Pseudomonas aeruginosa that affects cell elongation and prophage induction. Environ. Microbiol. 2020, 22, 5048–5057. [Google Scholar] [CrossRef]
  38. Bjarnsholt, T. The role of bacterial biofilms in chronic infections. APMIS Suppl. 2013, 121, 1–58. [Google Scholar] [CrossRef]
  39. Lee, K.; Yoon, S.S. Pseudomonas aeruginosa Biofilm, a Programmed Bacterial Life for Fitness. J. Microbiol. Biotechnol. 2017, 27, 1053–1064. [Google Scholar] [CrossRef] [Green Version]
  40. Thi, M.T.T.; Wibowo, D.; Rehm, B.H.A. Pseudomonas aeruginosa Biofilms. Int. J. Mol. Sci. 2020, 21, 8671. [Google Scholar] [CrossRef]
  41. Wang, X.; Wood, T.K. Toxin-antitoxin systems influence biofilm and persister cell formation and the general stress response. Appl Environ. Microbiol. 2011, 77, 5577–5583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Ma, D.; Mandell, J.B.; Donegan, N.P.; Cheung, A.L.; Ma, W.; Rothenberger, S.; Shanks, R.M.Q.; Richardson, A.R.; Urish, K.L. The Toxin-Antitoxin MazEF Drives Staphylococcus aureus Biofilm Formation, Antibiotic Tolerance, and Chronic Infection. mBio 2019, 10, e01658-19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Li, M.; Long, Y.; Liu, Y.; Liu, Y.; Chen, R.; Shi, J.; Zhang, L.; Jin, Y.; Yang, L.; Bai, F.; et al. HigB of Pseudomonas aeruginosa Enhances Killing of Phagocytes by Up-Regulating the Type III Secretion System in Ciprofloxacin Induced Persister Cells. Front. Cell. Infect. Microbiol. 2016, 6, 125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Guo, Y.; Sun, C.; Li, Y.; Tang, K.; Ni, S.; Wang, X. Antitoxin HigA inhibits virulence gene mvfR expression in Pseudomonas aeruginosa. Environ. Microbiol. 2019, 21, 2707–2723. [Google Scholar] [CrossRef] [PubMed]
  45. Xiao, G.; Deziel, E.; He, J.; Lepine, F.; Lesic, B.; Castonguay, M.H.; Milot, S.; Tampakaki, A.P.; Stachel, S.E.; Rahme, L.G. MvfR, a key Pseudomonas aeruginosa pathogenicity LTTR-class regulatory protein, has dual ligands. Mol. Microbiol. 2006, 62, 1689–1699. [Google Scholar] [CrossRef] [PubMed]
  46. Brutinel, E.D.; Yahr, T.L. Control of gene expression by type III secretory activity. Curr. Opin. Microbiol. 2008, 11, 128–133. [Google Scholar] [CrossRef] [Green Version]
  47. Allsopp, L.P.; Wood, T.E.; Howard, S.A.; Maggiorelli, F.; Nolan, L.M.; Wettstadt, S.; Filloux, A. RsmA and AmrZ orchestrate the assembly of all three type VI secretion systems in Pseudomonas aeruginosa. Proc. Natl. Acad. Sci. USA 2017, 114, 7707–7712. [Google Scholar] [CrossRef] [Green Version]
  48. Gotfredsen, M.; Gerdes, K. The Escherichia coli relBE genes belong to a new toxin-antitoxin gene family. Mol. Microbiol. 1998, 29, 1065–1076. [Google Scholar] [CrossRef]
  49. Pandey, D.P.; Gerdes, K. Toxin-antitoxin loci are highly abundant in free-living but lost from host-associated prokaryotes. Nucleic Acids Res. 2005, 33, 966–976. [Google Scholar] [CrossRef]
  50. Coskun, U.S.S.; Cicek, A.C.; Kilinc, C.; Guckan, R.; Dagcioglu, Y.; Demir, O.; Sandalli, C. Effect of mazEF, higBA and relBE toxin-antitoxin systems on antibiotic resistance in Pseudomonas aeruginosa and Staphylococcus isolates. Malawi Med. J. 2018, 30, 67–72. [Google Scholar] [CrossRef] [Green Version]
  51. Chan, W.T.; Domenech, M.; Moreno-Cordoba, I.; Navarro-Martinez, V.; Nieto, C.; Moscoso, M.; Garcia, E.; Espinosa, M. The Streptococcus pneumoniae yefM-yoeB and relBE Toxin-Antitoxin Operons Participate in Oxidative Stress and Biofilm Formation. Toxins 2018, 10, 378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Golmoradi Zadeh, R.; Mirshekar, M.; Sadeghi Kalani, B.; Pourghader, J.; Barati, M.; Masjedian Jazi, F. The expression of type II TA system genes following persister cell formation in Pseudomonas aeruginosa isolates in the exponential and stationary phases. Arch. Microbiol. 2022, 204, 451. [Google Scholar] [CrossRef] [PubMed]
  53. Jones, C.J.; Newsom, D.; Kelly, B.; Irie, Y.; Jennings, L.K.; Xu, B.; Limoli, D.H.; Harrison, J.J.; Parsek, M.R.; White, P.; et al. ChIP-Seq and RNA-Seq reveal an AmrZ-mediated mechanism for cyclic di-GMP synthesis and biofilm development by Pseudomonas aeruginosa. PLoS Pathog. 2014, 10, e1003984. [Google Scholar] [CrossRef] [PubMed]
  54. Waligora, E.A.; Ramsey, D.M.; Pryor, E.E., Jr.; Lu, H.; Hollis, T.; Sloan, G.P.; Deora, R.; Wozniak, D.J. AmrZ beta-sheet residues are essential for DNA binding and transcriptional control of Pseudomonas aeruginosa virulence genes. J. Bacteriol. 2010, 192, 5390–5401. [Google Scholar] [CrossRef] [Green Version]
  55. Jones, C.J.; Ryder, C.R.; Mann, E.E.; Wozniak, D.J. AmrZ modulates Pseudomonas aeruginosa biofilm architecture by directly repressing transcription of the psl operon. J. Bacteriol. 2013, 195, 1637–1644. [Google Scholar] [CrossRef] [Green Version]
  56. Pasqua, M.; Visaggio, D.; Lo Sciuto, A.; Genah, S.; Banin, E.; Visca, P.; Imperi, F. Ferric Uptake Regulator Fur Is Conditionally Essential in Pseudomonas aeruginosa. J. Bacteriol. 2017, 199, e00472-17. [Google Scholar] [CrossRef] [Green Version]
  57. Sun, F.; Gao, H.; Zhang, Y.; Wang, L.; Fang, N.; Tan, Y.; Guo, Z.; Xia, P.; Zhou, D.; Yang, R. Fur is a repressor of biofilm formation in Yersinia pestis. PLoS ONE 2012, 7, e52392. [Google Scholar] [CrossRef] [Green Version]
  58. Garcia, C.A.; Alcaraz, E.S.; Franco, M.A.; De Rossi, B.N.P. Iron is a signal for Stenotrophomonas maltophilia biofilm formation, oxidative stress response, OMPs expression, and virulence. Front. Microbiol. 2015, 6, 926. [Google Scholar] [CrossRef] [Green Version]
  59. Roberts, R.C.; Strom, A.R.; Helinski, D.R. The parDE operon of the broad-host-range plasmid RK2 specifies growth inhibition associated with plasmid loss. J. Mol. Biol. 1994, 237, 35–51. [Google Scholar] [CrossRef]
  60. Kamruzzaman, M.; Iredell, J. A ParDE-family toxin antitoxin system in major resistance plasmids of Enterobacteriaceae confers antibiotic and heat tolerance. Sci. Rep. 2019, 9, 9872. [Google Scholar] [CrossRef] [Green Version]
  61. Ames, J.R.; Muthuramalingam, M.; Murphy, T.; Najar, F.Z.; Bourne, C.R. Expression of different ParE toxins results in conserved phenotypes with distinguishable classes of toxicity. Microbiologyopen 2019, 8, e902. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Fisher, R.A.; Gollan, B.; Helaine, S. Persistent bacterial infections and persister cells. Nat. Rev. Microbiol. 2017, 15, 453–464. [Google Scholar] [CrossRef] [PubMed]
  63. Harms, A.; Maisonneuve, E.; Gerdes, K. Mechanisms of bacterial persistence during stress and antibiotic exposure. Science 2016, 354, aaf4268. [Google Scholar] [CrossRef]
  64. Gerdes, K.; Rasmussen, P.B.; Molin, S. Unique type of plasmid maintenance function: Postsegregational killing of plasmid-free cells. Proc. Natl. Acad. Sci. USA 1986, 83, 3116–3120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Ramirez-Diaz, M.I.; Diaz-Magana, A.; Meza-Carmen, V.; Johnstone, L.; Cervantes, C.; Rensing, C. Nucleotide sequence of Pseudomonas aeruginosa conjugative plasmid pUM505 containing virulence and heavy-metal resistance genes. Plasmid 2011, 66, 7–18. [Google Scholar] [CrossRef] [PubMed]
  66. Hernandez-Ramirez, K.C.; Valerio-Arellano, B.; Valle-Maldonado, M.I.; Ruiz-Herrera, L.F.; Meza-Carmen, V.; Ramirez-Diaz, M.I. Virulence Conferred by PumA Toxin from the Plasmid-Encoded PumAB Toxin-Antitoxin System is Regulated by Quorum System. Curr. Microbiol. 2020, 77, 2535–2543. [Google Scholar] [CrossRef] [PubMed]
  67. Sweere, J.M.; Van Belleghem, J.D.; Ishak, H.; Bach, M.S.; Popescu, M.; Sunkari, V.; Kaber, G.; Manasherob, R.; Suh, G.A.; Cao, X.; et al. Bacteriophage trigger antiviral immunity and prevent clearance of bacterial infection. Science 2019, 363, eaat9691. [Google Scholar] [CrossRef]
  68. Wang, X.; Wood, T.K. Cryptic prophages as targets for drug development. Drug Resist. Updates 2016, 27, 30–38. [Google Scholar] [CrossRef] [Green Version]
  69. Guo, Y.; Quiroga, C.; Chen, Q.; McAnulty, M.J.; Benedik, M.J.; Wood, T.K.; Wang, X. RalR (a DNase) and RalA (a small RNA) form a type I toxin-antitoxin system in Escherichia coli. Nucleic Acids Res. 2014, 42, 6448–6462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Yao, J.; Guo, Y.; Wang, P.; Zeng, Z.; Li, B.; Tang, K.; Liu, X.; Wang, X. Type II toxin/antitoxin system ParESO /CopASO stabilizes prophage CP4So in Shewanella oneidensis. Environ. Microbiol. 2018, 20, 1224–1239. [Google Scholar] [CrossRef]
  71. Li, Y.; Liu, X.; Tang, K.; Wang, P.; Zeng, Z.; Guo, Y.; Wang, X. Excisionase in Pf filamentous prophage controls lysis-lysogeny decision-making in Pseudomonas aeruginosa. Mol. Microbiol. 2019, 111, 495–513. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Wettstadt, S. Protect thy host: Pf4 phages shield Pseudomonas aeruginosa from antibiotics. Environ. Microbiol. 2020, 22, 2461–2462. [Google Scholar] [CrossRef] [PubMed]
  73. Srivastava, A.; Pati, S.; Kaushik, H.; Singh, S.; Garg, L.C. Toxin-antitoxin systems and their medical applications: Current status and future perspective. Appl. Microbiol. Biotechnol. 2021, 105, 1803–1821. [Google Scholar] [CrossRef] [PubMed]
  74. Rownicki, M.; Lasek, R.; Trylska, J.; Bartosik, D. Targeting Type II Toxin-Antitoxin Systems as Antibacterial Strategies. Toxins 2020, 12, 568. [Google Scholar] [CrossRef]
  75. Culviner, P.H.; Laub, M.T. Global Analysis of the E. coli Toxin MazF Reveals Widespread Cleavage of mRNA and the Inhibition of rRNA Maturation and Ribosome Biogenesis. Mol. Cell 2018, 70, 868–880 e810. [Google Scholar] [CrossRef] [Green Version]
  76. Barth, V.C.; Woychik, N.A. The Sole Mycobacterium smegmatis MazF Toxin Targets tRNA(Lys) to Impart Highly Selective, Codon-Dependent Proteome Reprogramming. Front. Genet. 2019, 10, 1356. [Google Scholar] [CrossRef]
  77. Critchlow, S.E.; O’Dea, M.H.; Howells, A.J.; Couturier, M.; Gellert, M.; Maxwell, A. The interaction of the F plasmid killer protein, CcdB, with DNA gyrase: Induction of DNA cleavage and blocking of transcription. J. Mol. Biol. 1997, 273, 826–839. [Google Scholar] [CrossRef]
  78. Freire, D.M.; Gutierrez, C.; Garza-Garcia, A.; Grabowska, A.D.; Sala, A.J.; Ariyachaokun, K.; Panikova, T.; Beckham, K.S.H.; Colom, A.; Pogenberg, V.; et al. An NAD(+) Phosphorylase Toxin Triggers Mycobacterium tuberculosis Cell Death. Mol. Cell 2019, 73, 1282–1291 e1288. [Google Scholar] [CrossRef] [Green Version]
  79. Pecota, D.C.; Kim, C.S.; Wu, K.; Gerdes, K.; Wood, T.K. Combining the hok/sok, parDE, and pnd postsegregational killer loci to enhance plasmid stability. Appl. Environ. Microbiol. 1997, 63, 1917–1924. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Daimon, Y.; Narita, S.; Akiyama, Y. Activation of Toxin-Antitoxin System Toxins Suppresses Lethality Caused by the Loss of sigmaE in Escherichia coli. J. Bacteriol. 2015, 197, 2316–2324. [Google Scholar] [CrossRef] [Green Version]
  81. Butt, A.; Higman, V.A.; Williams, C.; Crump, M.P.; Hemsley, C.M.; Harmer, N.; Titball, R.W. The HicA toxin from Burkholderia pseudomallei has a role in persister cell formation. Biochem. J. 2014, 459, 333–344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Wang, X.; Kim, Y.; Hong, S.H.; Ma, Q.; Brown, B.L.; Pu, M.; Tarone, A.M.; Benedik, M.J.; Peti, W.; Page, R.; et al. Antitoxin MqsA helps mediate the bacterial general stress response. Nat. Chem. Biol. 2011, 7, 359–366. [Google Scholar] [CrossRef] [PubMed]
  83. Soo, V.W.; Wood, T.K. Antitoxin MqsA represses curli formation through the master biofilm regulator CsgD. Sci. Rep. 2013, 3, 3186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Hu, Y.; Benedik, M.J.; Wood, T.K. Antitoxin DinJ influences the general stress response through transcript stabilizer CspE. Environ. Microbiol. 2012, 14, 669–679. [Google Scholar] [CrossRef]
Figure 1. Model of type II toxin–antitoxin systems. The toxin and antitoxin are represented in purple and green, respectively. The antitoxin protein directly interacts with cognate toxin protein and inhibits its toxicity. The labile antitoxin is efficiently degraded when the production of protease, the stable toxin, leads to cell death or growth arrest. The antitoxin and/or toxin–antitoxin (TA) complex can negatively autoregulate the transcription of their operators by recognizing and binding to palindromic sequences.
Figure 1. Model of type II toxin–antitoxin systems. The toxin and antitoxin are represented in purple and green, respectively. The antitoxin protein directly interacts with cognate toxin protein and inhibits its toxicity. The labile antitoxin is efficiently degraded when the production of protease, the stable toxin, leads to cell death or growth arrest. The antitoxin and/or toxin–antitoxin (TA) complex can negatively autoregulate the transcription of their operators by recognizing and binding to palindromic sequences.
Toxins 15 00164 g001
Figure 2. Biological functions of type II TA systems in P. aeruginosa.
Figure 2. Biological functions of type II TA systems in P. aeruginosa.
Toxins 15 00164 g002
Figure 3. Regulatory pathways of HigB/HigA in control of virulence in P. aeruginosa. HigA negatively regulates the transcription of the TA operon by binding to the palindromic sequence and also negatively regulates the expression of exsA, armZ, and mvfR by binding to their promoter regions. HigB can repress biofilm formation and increase expression of the T3SS genes by negatively regulating the level of c-di-GMP and also negatively regulating motility and pyochelin. ‘→’ indicates induction, and ‘┤’ indicates repression.
Figure 3. Regulatory pathways of HigB/HigA in control of virulence in P. aeruginosa. HigA negatively regulates the transcription of the TA operon by binding to the palindromic sequence and also negatively regulates the expression of exsA, armZ, and mvfR by binding to their promoter regions. HigB can repress biofilm formation and increase expression of the T3SS genes by negatively regulating the level of c-di-GMP and also negatively regulating motility and pyochelin. ‘→’ indicates induction, and ‘┤’ indicates repression.
Toxins 15 00164 g003
Table 1. Type II toxin–antitoxin systems identified and characterised in P. aeruginosa.
Table 1. Type II toxin–antitoxin systems identified and characterised in P. aeruginosa.
TA SystemToxinAntitoxinLocalisationTargeted Cellular ProcessFunction
HigB/HigAHigBHigAchromosomeTranslationVirulence and biofilm formation
PumA/PumBPumAPumBplasmidUnknownVirulence and Plasmid maintenance
PfiA/PfiTPfiTPfiAprophageUnknownPhage production and immunity
ParD/ParEParEParDchromosomeReplicationProtection host against antibiotics
HicA/HicBHicAHicBchromosomeTranslationUnknown
RelB/RelERelERelBchromosomeTranslationPersistence and biofilm formation
ResA/XreAResAXreAchromosomeMetabolic stressPersistence
PrrA/PrrTPrrTPrrAchromosomeUnknownPhage production and biofilm formation
CrlA/CrlTCrlTCrlAchromosomeTranslationInhibits lytic phage infection
PacA/PacTPacTPacAchromosomeTranslationVirulence and biofilm formation
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, M.; Guo, N.; Song, G.; Huang, Y.; Wang, L.; Zhang, Y.; Wang, T. Type II Toxin–Antitoxin Systems in Pseudomonas aeruginosa. Toxins 2023, 15, 164. https://doi.org/10.3390/toxins15020164

AMA Style

Li M, Guo N, Song G, Huang Y, Wang L, Zhang Y, Wang T. Type II Toxin–Antitoxin Systems in Pseudomonas aeruginosa. Toxins. 2023; 15(2):164. https://doi.org/10.3390/toxins15020164

Chicago/Turabian Style

Li, Meng, Nannan Guo, Gaoyu Song, Yi Huang, Lecheng Wang, Yani Zhang, and Tietao Wang. 2023. "Type II Toxin–Antitoxin Systems in Pseudomonas aeruginosa" Toxins 15, no. 2: 164. https://doi.org/10.3390/toxins15020164

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