1932

Abstract

All bacteria must compete for growth niches and other limited environmental resources. These existential battles are waged at several levels, but one common strategy entails the transfer of growth-inhibitory protein toxins between competing cells. These antibacterial effectors are invariably encoded with immunity proteins that protect cells from intoxication by neighboring siblings. Several effector classes have been described, each designed to breach the cell envelope of target bacteria. Although effector architectures and export pathways tend to be clade specific, phylogenetically distant species often deploy closely related toxin domains. Thus, diverse competition systems are linked through a common reservoir of toxin-immunity pairs that is shared via horizontal gene transfer. These toxin-immunity protein pairs are extraordinarily diverse in sequence, and this polymorphism underpins an important mechanism of self/nonself discrimination in bacteria. This review focuses on the structures, functions, and delivery mechanisms of polymorphic toxin effectors that mediate bacterial competition.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-micro-020518-115638
2020-09-08
2024-04-26
Loading full text...

Full text loading...

/deliver/fulltext/micro/74/1/annurev-micro-020518-115638.html?itemId=/content/journals/10.1146/annurev-micro-020518-115638&mimeType=html&fmt=ahah

Literature Cited

  1. 1. 
    Ahmad S, Wang B, Walker MD, Tran HR, Stogios PJ et al. 2019. An interbacterial toxin inhibits target cell growth by synthesizing (p)ppApp. Nature 575:674–78
    [Google Scholar]
  2. 2. 
    Aoki SK, Diner EJ, de Roodenbeke CT, Burgess BR, Poole SJ et al. 2010. A widespread family of polymorphic contact-dependent toxin delivery systems in bacteria. Nature 468:439–42
    [Google Scholar]
  3. 3. 
    Aoki SK, Malinverni JC, Jacoby K, Thomas B, Pamma R et al. 2008. Contact-dependent growth inhibition requires the essential outer membrane protein BamA (YaeT) as the receptor and the inner membrane transport protein AcrB. Mol. Microbiol. 70:323–40
    [Google Scholar]
  4. 4. 
    Aoki SK, Pamma R, Hernday AD, Bickham JE, Braaten BA, Low DA 2005. Contact-dependent inhibition of growth in Escherichia coli. . Science 309:1245–48
    [Google Scholar]
  5. 5. 
    Aoki SK, Webb JS, Braaten BA, Low DA 2009. Contact-dependent growth inhibition causes reversible metabolic downregulation in Escherichia coli. J. Bacteriol 191:1777–86
    [Google Scholar]
  6. 6. 
    Arenas J, de Maat V, Caton L, Krekorian M, Herrero JC et al. 2015. Fratricide activity of MafB protein of N. meningitidis strain B16B6. BMC Microbiol 15:156
    [Google Scholar]
  7. 7. 
    Arenas J, Schipper K, van Ulsen P, van der Ende A, Tommassen J 2013. Domain exchange at the 3′ end of the gene encoding the fratricide meningococcal two-partner secretion protein A. BMC Genom 14:622
    [Google Scholar]
  8. 8. 
    Aschtgen MS, Bernard CS, De Bentzmann S, Lloubes R, Cascales E 2008. SciN is an outer membrane lipoprotein required for type VI secretion in enteroaggregative Escherichia coli. J. . Bacteriol 190:7523–31
    [Google Scholar]
  9. 9. 
    Bartelli NL, Sun S, Gucinski GC, Zhou H, Song K et al. 2019. The cytoplasm-entry domain of antibacterial CdiA is a dynamic α-helical bundle with disulfide-dependent structural features. J. Mol. Biol. 431:3203–16
    [Google Scholar]
  10. 10. 
    Basler M, Pilhofer M, Henderson GP, Jensen GJ, Mekalanos JJ 2012. Type VI secretion requires a dynamic contractile phage tail-like structure. Nature 483:182–86Showed the dynamic nature of T6SS assembly and contraction.
    [Google Scholar]
  11. 11. 
    Batot G, Michalska K, Ekberg G, Irimpan EM, Joachimiak G et al. 2017. The CDI toxin of Yersinia kristensenii is a novel bacterial member of the RNase A superfamily. Nucleic Acids Res 45:5013–25
    [Google Scholar]
  12. 12. 
    Baud C, Guerin J, Petit E, Lesne E, Dupre E et al. 2014. Translocation path of a substrate protein through its Omp85 transporter. Nat. Commun. 5:5271
    [Google Scholar]
  13. 13. 
    Beck CM, Morse RP, Cunningham DA, Iniguez A, Low DA et al. 2014. CdiA from Enterobacter cloacae delivers a toxic ribosomal RNase into target bacteria. Structure 22:707–18
    [Google Scholar]
  14. 14. 
    Beck CM, Willett JL, Cunningham DA, Kim JJ, Low DA, Hayes CS 2016. CdiA effectors from uropathogenic Escherichia coli use heterotrimeric osmoporins as receptors to recognize target bacteria. PLOS Pathog 12:e1005925
    [Google Scholar]
  15. 15. 
    Blondel CJ, Jimenez JC, Contreras I, Santiviago CA 2009. Comparative genomic analysis uncovers 3 novel loci encoding type six secretion systems differentially distributed in Salmonella serotypes. BMC Genom 10:354
    [Google Scholar]
  16. 16. 
    Bobrovskyy M, Willing SE, Schneewind O, Missiakas D 2018. EssH peptidoglycan hydrolase enables Staphylococcus aureus type VII secretion across the bacterial cell wall envelope. J. Bacteriol. 200:e00268-18
    [Google Scholar]
  17. 17. 
    Bondage DD, Lin JS, Ma LS, Kuo CH, Lai EM 2016. VgrG C terminus confers the type VI effector transport specificity and is required for binding with PAAR and adaptor-effector complex. PNAS 113:E3931–40
    [Google Scholar]
  18. 18. 
    Bonemann G, Pietrosiuk A, Diemand A, Zentgraf H, Mogk A 2009. Remodelling of VipA/VipB tubules by ClpV-mediated threading is crucial for type VI protein secretion. EMBO J 28:315–25
    [Google Scholar]
  19. 19. 
    Brackmann M, Wang J, Basler M 2018. Type VI secretion system sheath inter-subunit interactions modulate its contraction. EMBO Rep 19:225–33
    [Google Scholar]
  20. 20. 
    Brooks TM, Unterweger D, Bachmann V, Kostiuk B, Pukatzki S 2013. Lytic activity of the Vibrio cholerae type VI secretion toxin VgrG-3 is inhibited by the antitoxin TsaB. J. Biol. Chem. 288:7618–25
    [Google Scholar]
  21. 21. 
    Brunet YR, Zoued A, Boyer F, Douzi B, Cascales E 2015. The type VI secretion TssEFGK-VgrG phage-like baseplate is recruited to the TssJLM membrane complex via multiple contacts and serves as assembly platform for tail tube/sheath polymerization. PLOS Genet 11:e1005545
    [Google Scholar]
  22. 22. 
    Burkinshaw BJ, Liang X, Wong M, Le ANH, Lam L, Dong TG 2018. A type VI secretion system effector delivery mechanism dependent on PAAR and a chaperone-co-chaperone complex. Nat. Microbiol. 3:632–40
    [Google Scholar]
  23. 23. 
    Burts ML, Williams WA, DeBord K, Missiakas DM 2005. EsxA and EsxB are secreted by an ESAT-6-like system that is required for the pathogenesis of Staphylococcus aureus infections. PNAS 102:1169–74
    [Google Scholar]
  24. 24. 
    Busby JN, Panjikar S, Landsberg MJ, Hurst MR, Lott JS 2013. The BC component of ABC toxins is an RHS-repeat-containing protein encapsulation device. Nature 501:547–50
    [Google Scholar]
  25. 25. 
    Cao Z, Casabona MG, Kneuper H, Chalmers JD, Palmer T 2016. The type VII secretion system of Staphylococcus aureus secretes a nuclease toxin that targets competitor bacteria. Nat. Microbiol. 2:16183
    [Google Scholar]
  26. 26. 
    Carr S, Walker D, James R, Kleanthous C, Hemmings AM 2000. Inhibition of a ribosome-inactivating ribonuclease: the crystal structure of the cytotoxic domain of colicin E3 in complex with its immunity protein. Structure 8:949–60
    [Google Scholar]
  27. 27. 
    Cascales E, Buchanan SK, Duche D, Kleanthous C, Lloubes R et al. 2007. Colicin biology. Microbiol. Mol. Biol. Rev. 71:158–229
    [Google Scholar]
  28. 28. 
    Cascales E, Lloubes R, Sturgis JN 2001. The TolQ-TolR proteins energize TolA and share homologies with the flagellar motor proteins MotA-MotB. Mol. Microbiol. 42:795–807
    [Google Scholar]
  29. 29. 
    Chang YW, Rettberg LA, Ortega DR, Jensen GJ 2017. In vivo structures of an intact type VI secretion system revealed by electron cryotomography. EMBO Rep 18:1090–99
    [Google Scholar]
  30. 30. 
    Chauleau M, Mora L, Serba J, de Zamaroczy M 2011. FtsH-dependent processing of RNase colicins D and E3 means that only the cytotoxic domains are imported into the cytoplasm. J. Biol. Chem. 286:29397–407
    [Google Scholar]
  31. 31. 
    Chavira A, Belda-Ferre P, Kosciolek T, Ali F, Dorrestein PC, Knight R 2019. The microbiome and its potential for pharmacology. Handb. Exp. Pharmacol. 260:301–26
    [Google Scholar]
  32. 32. 
    Cherrak Y, Rapisarda C, Pellarin R, Bouvier G, Bardiaux B et al. 2018. Biogenesis and structure of a type VI secretion baseplate. Nat. Microbiol. 3:1404–16
    [Google Scholar]
  33. 33. 
    Cianfanelli FR, Alcoforado Diniz J, Guo M, De Cesare V, Trost M, Coulthurst SJ 2016. VgrG and PAAR proteins define distinct versions of a functional Type VI secretion system. PLOS Pathog 12:e1005735
    [Google Scholar]
  34. 34. 
    Cianfanelli FR, Monlezun L, Coulthurst SJ 2016. Aim, load, fire: the Type VI secretion system, a bacterial nanoweapon. Trends Microbiol 24:51–62
    [Google Scholar]
  35. 35. 
    Cramer WA, Sharma O, Zakharov SD 2018. On mechanisms of colicin import: the outer membrane quandary. Biochem. J. 475:3903–15
    [Google Scholar]
  36. 36. 
    Dalbey RE, Wang P, van Dijl JM 2012. Membrane proteases in the bacterial protein secretion and quality control pathway. Microbiol. Mol. Biol. Rev. 76:311–30
    [Google Scholar]
  37. 37. 
    De Maayer P, Venter SN, Kamber T, Duffy B, Coutinho TA, Smits TH 2011. Comparative genomics of the Type VI secretion systems of Pantoea and Erwinia species reveals the presence of putative effector islands that may be translocated by the VgrG and Hcp proteins. BMC Genom 12:576
    [Google Scholar]
  38. 38. 
    de Zamaroczy M, Mora L 2012. Hijacking cellular functions for processing and delivery of colicins E3 and D into the cytoplasm. Biochem. Soc. Trans. 40:1486–91
    [Google Scholar]
  39. 39. 
    Dekker N, Tommassen J, Verheij HM 1999. Bacteriocin release protein triggers dimerization of outer membrane phospholipase A in vivo. J. Bacteriol. 181:3281–83
    [Google Scholar]
  40. 40. 
    Dey A, Vassallo CN, Conklin AC, Pathak DT, Troselj V, Wall D 2016. Sibling rivalry in Myxococcusxanthus is mediated by kin recognition and a polyploid prophage. J. Bacteriol. 198:994–1004
    [Google Scholar]
  41. 41. 
    Dey A, Wall D. 2014. A genetic screen in Myxococcusxanthus identifies mutants that uncouple outer membrane exchange from a downstream cellular response. J. Bacteriol. 196:4324–32
    [Google Scholar]
  42. 42. 
    Diner EJ, Beck CM, Webb JS, Low DA, Hayes CS 2012. Identification of a target cell permissive factor required for contact-dependent growth inhibition (CDI). Genes Dev 26:515–25
    [Google Scholar]
  43. 43. 
    Durand E, Nguyen VS, Zoued A, Logger L, Pehau-Arnaudet G et al. 2015. Biogenesis and structure of a type VI secretion membrane core complex. Nature 523:555–60
    [Google Scholar]
  44. 44. 
    English G, Byron O, Cianfanelli FR, Prescott AR, Coulthurst SJ 2014. Biochemical analysis of TssK, a core component of the bacterial Type VI secretion system, reveals distinct oligomeric states of TssK and identifies a TssK-TssFG subcomplex. Biochem. J. 461:291–304
    [Google Scholar]
  45. 45. 
    Felisberto-Rodrigues C, Durand E, Aschtgen MS, Blangy S, Ortiz-Lombardia M et al. 2011. Towards a structural comprehension of bacterial Type VI secretion systems: characterization of the TssJ-TssM complex of an Escherichia coli pathovar. PLOS Pathog 7:e1002386
    [Google Scholar]
  46. 46. 
    Flaugnatti N, Le TT, Canaan S, Aschtgen MS, Nguyen VS et al. 2016. A phospholipase A1 antibacterial Type VI secretion effector interacts directly with the C-terminal domain of the VgrG spike protein for delivery. Mol. Microbiol. 99:1099–118
    [Google Scholar]
  47. 47. 
    Garufi G, Butler E, Missiakas D 2008. ESAT-6-like protein secretion in Bacillus anthracis.J. . Bacteriol 190:7004–11
    [Google Scholar]
  48. 48. 
    Ghequire MGK, Buchanan SK, De Mot R 2018. The ColM family, polymorphic toxins breaching the bacterial cell wall. mBio 9:e02267-17
    [Google Scholar]
  49. 49. 
    Gray TA, Clark RR, Boucher N, Lapierre P, Smith C, Derbyshire KM 2016. Intercellular communication and conjugation are mediated by ESX secretion systems in mycobacteria. Science 354:347–50
    [Google Scholar]
  50. 50. 
    Gucinski GC, Michalska K, Garza-Sanchez F, Eschenfeldt WH, Stols L et al. 2019. Convergent evolution of the Barnase/EndoU/Colicin/RelE (BECR) fold in antibacterial tRNase toxins. Structure 27:1660–74
    [Google Scholar]
  51. 51. 
    Guerin J, Bigot S, Schneider R, Buchanan SK, Jacob-Dubuisson F 2017. Two-partner secretion: combining efficiency and simplicity in the secretion of large proteins for bacteria-host and bacteria-bacteria interactions. Front. Cell Infect. Microbiol. 7:148
    [Google Scholar]
  52. 52. 
    Harms A, Brodersen DE, Mitarai N, Gerdes K 2018. Toxins, targets, and triggers: an overview of toxin-antitoxin biology. Mol. Cell 70:768–84
    [Google Scholar]
  53. 53. 
    Helbig S, Patzer SI, Schiene-Fischer C, Zeth K, Braun V 2011. Activation of colicin M by the FkpA prolyl cis-trans isomerase/chaperone. J. Biol. Chem. 286:6280–90
    [Google Scholar]
  54. 54. 
    Ho BT, Fu Y, Dong TG, Mekalanos JJ 2017. Vibrio cholerae type 6 secretion system effector trafficking in target bacterial cells. PNAS 114:9427–32
    [Google Scholar]
  55. 55. 
    Holberger LE, Garza-Sanchez F, Lamoureux J, Low DA, Hayes CS 2012. A novel family of toxin/antitoxin proteins in Bacillus species. FEBS Lett 586:132–36
    [Google Scholar]
  56. 56. 
    Hood RD, Singh P, Hsu F, Guvener T, Carl MA et al. 2010. A type VI secretion system of Pseudomonas aeruginosa targets a toxin to bacteria. Cell Host Microbe 7:25–37First demonstration of the antibacterial activity of T6SS.
    [Google Scholar]
  57. 57. 
    Housden NG, Hopper JT, Lukoyanova N, Rodriguez-Larrea D, Wojdyla JA et al. 2013. Intrinsically disordered protein threads through the bacterial outer-membrane porin OmpF. Science 340:1570–74Structural characterization of the quaternary colicin-translocon complex.
    [Google Scholar]
  58. 58. 
    Housden NG, Loftus SR, Moore GR, James R, Kleanthous C 2005. Cell entry mechanism of enzymatic bacterial colicins: porin recruitment and the thermodynamics of receptor binding. PNAS 102:13849–54
    [Google Scholar]
  59. 59. 
    Hu H, Zhang H, Gao Z, Wang D, Liu G et al. 2014. Structure of the type VI secretion phospholipase effector Tle1 provides insight into its hydrolysis and membrane targeting. Acta Crystallogr. D 70:2175–85
    [Google Scholar]
  60. 60. 
    Hullmann J, Patzer SI, Romer C, Hantke K, Braun V 2008. Periplasmic chaperone FkpA is essential for imported colicin M toxicity. Mol. Microbiol. 69:926–37
    [Google Scholar]
  61. 61. 
    Iyer LM, Zhang D, Rogozin IB, Aravind L 2011. Evolution of the deaminase fold and multiple origins of eukaryotic editing and mutagenic nucleic acid deaminases from bacterial toxin systems. Nucleic Acids Res 39:9473–97
    [Google Scholar]
  62. 62. 
    Jager F, Kneuper H, Palmer T 2018. EssC is a specificity determinant for Staphylococcus aureus type VII secretion. Microbiology 164:816–20
    [Google Scholar]
  63. 63. 
    Jakes KS, Finkelstein A. 2010. The colicin Ia receptor, Cir, is also the translocator for colicin Ia. Mol. Microbiol. 75:567–78
    [Google Scholar]
  64. 64. 
    Jamet A, Jousset AB, Euphrasie D, Mukorako P, Boucharlat A et al. 2015. A new family of secreted toxins in pathogenic Neisseria species. PLOS Pathog 11:e1004592
    [Google Scholar]
  65. 65. 
    Jana B, Fridman CM, Bosis E, Salomon D 2019. A modular effector with a DNase domain and a marker for T6SS substrates. Nat. Commun. 10:3595
    [Google Scholar]
  66. 66. 
    Johnson PM, Beck CM, Morse RP, Garza-Sanchez F, Low DA et al. 2016. Unraveling the essential role of CysK in CDI toxin activation. PNAS 113:9792–97
    [Google Scholar]
  67. 67. 
    Johnson PM, Gucinski GC, Garza-Sanchez F, Wong T, Hung LW et al. 2016. Functional diversity of cytotoxic tRNase/immunity protein complexes from Burkholderia pseudomallei. . J. Biol. Chem 291:19387–400
    [Google Scholar]
  68. 68. 
    Jones AM, Garza-Sanchez F, So J, Hayes CS, Low DA 2017. Activation of contact-dependent antibacterial tRNase toxins by translation elongation factors. PNAS 114:E1951–57
    [Google Scholar]
  69. 69. 
    Kajava AV, Cheng N, Cleaver R, Kessel M, Simon MN et al. 2001. Beta-helix model for the filamentous haemagglutinin adhesin of Bordetella pertussis and related bacterial secretory proteins. Mol. Microbiol. 42:279–92
    [Google Scholar]
  70. 70. 
    Kleanthous C, Kuhlmann UC, Pommer AJ, Ferguson N, Radford SE et al. 1999. Structural and mechanistic basis of immunity toward endonuclease colicins. Nat. Struct. Biol. 6:243–52
    [Google Scholar]
  71. 71. 
    Ko TP, Liao CC, Ku WY, Chak KF, Yuan HS 1999. The crystal structure of the DNase domain of colicin E7 in complex with its inhibitor Im7 protein. Structure 7:91–102
    [Google Scholar]
  72. 72. 
    Koskiniemi S, Lamoureux JG, Nikolakakis KC, t'Kint de Roodenbeke C, Kaplan MD et al. 2013. Rhs proteins from diverse bacteria mediate intercellular competition. PNAS 110:7032–37
    [Google Scholar]
  73. 73. 
    LaCourse KD, Peterson SB, Kulasekara HD, Radey MC, Kim J, Mougous JD 2018. Conditional toxicity and synergy drive diversity among antibacterial effectors. Nat. Microbiol. 3:440–46
    [Google Scholar]
  74. 74. 
    Leiman PG, Basler M, Ramagopal UA, Bonanno JB, Sauder JM et al. 2009. Type VI secretion apparatus and phage tail-associated protein complexes share a common evolutionary origin. PNAS 106:4154–59
    [Google Scholar]
  75. 75. 
    Li W, Keeble AH, Giffard C, James R, Moore GR, Kleanthous C 2004. Highly discriminating protein-protein interaction specificities in the context of a conserved binding energy hotspot. J. Mol. Biol. 337:743–59
    [Google Scholar]
  76. 76. 
    Liang X, Moore R, Wilton M, Wong MJ, Lam L, Dong TG 2015. Identification of divergent type VI secretion effectors using a conserved chaperone domain. PNAS 112:9106–11
    [Google Scholar]
  77. 77. 
    Logger L, Aschtgen MS, Guerin M, Cascales E, Durand E 2016. Molecular dissection of the interface between the Type VI secretion TssM cytoplasmic domain and the TssG baseplate component. J. Mol. Biol. 428:4424–37
    [Google Scholar]
  78. 78. 
    Lu FM, Chak KF. 1996. Two overlapping SOS-boxes in ColE operons are responsible for the viability of cells harboring the Col plasmid. Mol. Gen. Genet. 251:407–11
    [Google Scholar]
  79. 79. 
    Ma J, Pan Z, Huang J, Sun M, Lu C, Yao H 2017. The Hcp proteins fused with diverse extended-toxin domains represent a novel pattern of antibacterial effectors in type VI secretion systems. Virulence 8:1189–202
    [Google Scholar]
  80. 80. 
    Ma LS, Hachani A, Lin JS, Filloux A, Lai EM 2014. Agrobacterium tumefaciens deploys a superfamily of type VI secretion DNase effectors as weapons for interbacterial competition in planta. Cell Host Microbe 16:94–104
    [Google Scholar]
  81. 81. 
    Ma LS, Lin JS, Lai EM 2009. An IcmF family protein, ImpLM, is an integral inner membrane protein interacting with ImpKL, and its walker a motif is required for type VI secretion system-mediated Hcp secretion in Agrobacterium tumefaciens. J. . Bacteriol 191:4316–29
    [Google Scholar]
  82. 82. 
    MacIntyre DL, Miyata ST, Kitaoka M, Pukatzki S 2010. The Vibrio cholerae type VI secretion system displays antimicrobial properties. PNAS 107:19520–24First demonstration of the antibacterial activity of T6SS.
    [Google Scholar]
  83. 83. 
    Madgwick PG, Belcher LJ, Wolf JB 2019. Greenbeard genes: theory and reality. Trends Ecol. Evol. 34:1092–103
    [Google Scholar]
  84. 84. 
    Mariano G, Monlezun L, Coulthurst SJ 2018. Dual role for DsbA in attacking and targeted bacterial cells during Type VI secretion system-mediated competition. Cell Rep 22:774–85
    [Google Scholar]
  85. 85. 
    Mariano G, Trunk K, Williams DJ, Monlezun L, Strahl H et al. 2019. A family of Type VI secretion system effector proteins that form ion-selective pores. Nat. Commun. 10:5484
    [Google Scholar]
  86. 86. 
    Michalska K, Gucinski GC, Garza-Sanchez F, Johnson PM, Stols LM et al. 2017. Structure of a novel antibacterial toxin that exploits elongation factor Tu to cleave specific transfer RNAs. Nucleic Acids Res 45:10306–20
    [Google Scholar]
  87. 87. 
    Michalska K, Quan Nhan D, Willett JLE, Stols LM, Eschenfeldt WH et al. 2018. Functional plasticity of antibacterial EndoU toxins. Mol. Microbiol. 109:509–27
    [Google Scholar]
  88. 88. 
    Miyata ST, Kitaoka M, Brooks TM, McAuley SB, Pukatzki S 2011. Vibrio cholerae requires the type VI secretion system virulence factor VasX to kill Dictyostelium discoideum. . Infect. Immun 79:2941–49
    [Google Scholar]
  89. 89. 
    Miyata ST, Unterweger D, Rudko SP, Pukatzki S 2013. Dual expression profile of type VI secretion system immunity genes protects pandemic Vibrio cholerae. . PLOS Pathog 9:e1003752
    [Google Scholar]
  90. 90. 
    Mora L, Moncoq K, England P, Oberto J, de Zamaroczy M 2015. The stable interaction between signal peptidase LepB of Escherichia coli and nuclease bacteriocins promotes toxin entry into the cytoplasm. J. Biol. Chem. 290:30783–96
    [Google Scholar]
  91. 91. 
    Morse RP, Nikolakakis KC, Willett JL, Gerrick E, Low DA et al. 2012. Structural basis of toxicity and immunity in contact-dependent growth inhibition (CDI) systems. PNAS 109:21480–5
    [Google Scholar]
  92. 92. 
    Morse RP, Willett JL, Johnson PM, Zheng J, Credali A et al. 2015. Diversification of β-augmentation interactions between CDI toxin/immunity proteins. J. Mol. Biol. 427:3766–84
    [Google Scholar]
  93. 93. 
    Mosbahi K, Walker D, James R, Moore GR, Kleanthous C 2006. Global structural rearrangement of the cell penetrating ribonuclease colicin E3 on interaction with phospholipid membranes. Protein Sci 15:620–27
    [Google Scholar]
  94. 94. 
    Mosbahi K, Walker D, Lea E, Moore GR, James R, Kleanthous C 2004. Destabilization of the colicin E9 endonuclease domain by interaction with negatively charged phospholipids: implications for colicin translocation into bacteria. J. Biol. Chem. 279:22145–51
    [Google Scholar]
  95. 95. 
    Munoz-Dorado J, Marcos-Torres FJ, Garcia-Bravo E, Moraleda-Munoz A, Perez J 2016. Myxobacteria: moving, killing, feeding, and surviving together. Front. Microbiol. 7:781
    [Google Scholar]
  96. 96. 
    Ohr RJ, Anderson M, Shi M, Schneewind O, Missiakas D 2017. EssD, a nuclease effector of the Staphylococcus aureus ESS pathway. J. Bacteriol. 199:e00528-16
    [Google Scholar]
  97. 97. 
    Pallen MJ. 2002. The ESAT-6/WXG100 superfamily—and a new Gram-positive secretion system. Trends Microbiol 10:209–12
    [Google Scholar]
  98. 98. 
    Parker MW, Pattus F. 1993. Rendering a membrane protein soluble in water: a common packing motif in bacterial protein toxins. Trends Biochem. Sci. 18:391–95
    [Google Scholar]
  99. 99. 
    Parker MW, Pattus F, Tucker AD, Tsernoglou D 1989. Structure of the membrane-pore-forming fragment of colicin A. Nature 337:93–96
    [Google Scholar]
  100. 100. 
    Pathak DT, Wei X, Bucuvalas A, Haft DH, Gerloff DL, Wall D 2012. Cell contact-dependent outer membrane exchange in myxobacteria: genetic determinants and mechanism. PLOS Genet 8:e1002626
    [Google Scholar]
  101. 101. 
    Pathak DT, Wei X, Dey A, Wall D 2013. Molecular recognition by a polymorphic cell surface receptor governs cooperative behaviors in bacteria. PLOS Genet 9:e1003891
    [Google Scholar]
  102. 102. 
    Patzer SI, Albrecht R, Braun V, Zeth K 2012. Structural and mechanistic studies of pesticin, a bacterial homolog of phage lysozymes. J. Biol. Chem. 287:23381–96
    [Google Scholar]
  103. 103. 
    Pissaridou P, Allsopp LP, Wettstadt S, Howard SA, Mavridou DAI, Filloux A 2018. The Pseudomonas aeruginosa T6SS-VgrG1b spike is topped by a PAAR protein eliciting DNA damage to bacterial competitors. PNAS 115:12519–24
    [Google Scholar]
  104. 104. 
    Poole SJ, Diner EJ, Aoki SK, Braaten BA, T'Kint de Roodenbeke C et al. 2011. Identification of functional toxin/immunity genes linked to contact-dependent growth inhibition (CDI) and rearrangement hotspot (Rhs) systems. PLOS Genet 7:e1002217
    [Google Scholar]
  105. 105. 
    Pugsley AP, Schwartz M. 1983. Expression of a gene in a 400-base-pair fragment of colicin plasmid ColE2-P9 is sufficient to cause host cell lysis. J. Bacteriol. 156:109–14
    [Google Scholar]
  106. 106. 
    Pugsley AP, Schwartz M. 1984. Colicin E2 release: lysis, leakage or secretion? Possible role of a phospholipase. EMBO J 3:2393–97
    [Google Scholar]
  107. 107. 
    Pukatzki S, Ma AT, Revel AT, Sturtevant D, Mekalanos JJ 2007. Type VI secretion system translocates a phage tail spike-like protein into target cells where it cross-links actin. PNAS 104:15508–13
    [Google Scholar]
  108. 108. 
    Pukatzki S, Ma AT, Sturtevant D, Krastins B, Sarracino D et al. 2006. Identification of a conserved bacterial protein secretion system in Vibrio cholerae using the Dictyostelium host model system. PNAS 103:1528–33
    [Google Scholar]
  109. 109. 
    Quentin D, Ahmad S, Shanthamoorthy P, Mougous JD, Whitney JC, Raunser S 2018. Mechanism of loading and translocation of type VI secretion system effector Tse6. Nat. Microbiol. 3:1142–52
    [Google Scholar]
  110. 110. 
    Raaijmakers H, Vix O, Toro I, Golz S, Kemper B, Suck D 1999. X-ray structure of T4 endonuclease VII: a DNA junction resolvase with a novel fold and unusual domain-swapped dimer architecture. EMBO J 18:1447–58
    [Google Scholar]
  111. 111. 
    Riley MA. 1998. Molecular mechanisms of bacteriocin evolution. Annu. Rev. Genet. 32:255–78
    [Google Scholar]
  112. 112. 
    Ronneau S, Helaine S. 2019. Clarifying the link between toxin-antitoxin modules and bacterial persistence. J. Mol. Biol. 431:3462–71
    [Google Scholar]
  113. 113. 
    Ruhe ZC, Nguyen JY, Beck CM, Low DA, Hayes CS 2014. The proton-motive force is required for translocation of CDI toxins across the inner membrane of target bacteria. Mol. Microbiol. 94:466–81
    [Google Scholar]
  114. 114. 
    Ruhe ZC, Nguyen JY, Xiong J, Koskiniemi S, Beck CM et al. 2017. CdiA effectors use modular receptor-binding domains to recognize target bacteria. mBio 8:e00290-17
    [Google Scholar]
  115. 115. 
    Ruhe ZC, Subramanian P, Song K, Nguyen JY, Stevens TA et al. 2018. Programmed secretion arrest and receptor-triggered toxin export during antibacterial contact-dependent growth inhibition. Cell 175:921–33.e14Proposed that CdiA effectors build their own translocons in the outer membrane of target bacteria.
    [Google Scholar]
  116. 116. 
    Ruhe ZC, Townsley L, Wallace AB, King A, Van der Woude MW et al. 2015. CdiA promotes receptor-independent intercellular adhesion. Mol. Microbiol. 98:175–92
    [Google Scholar]
  117. 117. 
    Russell AB, Hood RD, Bui NK, LeRoux M, Vollmer W, Mougous JD 2011. Type VI secretion delivers bacteriolytic effectors to target cells. Nature 475:343–47
    [Google Scholar]
  118. 118. 
    Russell AB, LeRoux M, Hathazi K, Agnello DM, Ishikawa T et al. 2013. Diverse type VI secretion phospholipases are functionally plastic antibacterial effectors. Nature 496:508–12
    [Google Scholar]
  119. 119. 
    Russell AB, Singh P, Brittnacher M, Bui NK, Hood RD et al. 2012. A widespread bacterial type VI secretion effector superfamily identified using a heuristic approach. Cell Host Microbe 11:538–49
    [Google Scholar]
  120. 120. 
    Salomon D, Kinch LN, Trudgian DC, Guo X, Klimko JA et al. 2014. Marker for type VI secretion system effectors. PNAS 111:9271–76
    [Google Scholar]
  121. 121. 
    Salomon D, Klimko JA, Trudgian DC, Kinch LN, Grishin NV et al. 2015. Type VI secretion system toxins horizontally shared between marine bacteria. PLOS Pathog 11:e1005128
    [Google Scholar]
  122. 122. 
    Santin YG, Doan T, Journet L, Cascales E 2019. Cell width dictates type VI secretion tail length. Curr. Biol. 29:3707–13.e3
    [Google Scholar]
  123. 123. 
    Schaller K, Holtje JV, Braun V 1982. Colicin M is an inhibitor of murein biosynthesis. J. Bacteriol. 152:994–1000
    [Google Scholar]
  124. 124. 
    Shneider MM, Buth SA, Ho BT, Basler M, Mekalanos JJ, Leiman PG 2013. PAAR-repeat proteins sharpen and diversify the type VI secretion system spike. Nature 500:350–53
    [Google Scholar]
  125. 125. 
    Siegrist MS, Unnikrishnan M, McConnell MJ, Borowsky M, Cheng TY et al. 2009. Mycobacterial Esx-3 is required for mycobactin-mediated iron acquisition. PNAS 106:18792–97
    [Google Scholar]
  126. 126. 
    Silverman JM, Agnello DM, Zheng H, Andrews BT, Li M et al. 2013. Haemolysin coregulated protein is an exported receptor and chaperone of type VI secretion substrates. Mol. Cell 51:584–93
    [Google Scholar]
  127. 127. 
    Skjerning RB, Senissar M, Winther KS, Gerdes K, Brodersen DE 2019. The RES domain toxins of RES-Xre toxin-antitoxin modules induce cell stasis by degrading NAD+. Mol. Microbiol. 111:221–36
    [Google Scholar]
  128. 128. 
    Souza DP, Oka GU, Alvarez-Martinez CE, Bisson-Filho AW, Dunger G et al. 2015. Bacterial killing via a type IV secretion system. Nat. Commun. 6:6453
    [Google Scholar]
  129. 129. 
    Stanley SA, Raghavan S, Hwang WW, Cox JS 2003. Acute infection and macrophage subversion by Mycobacterium tuberculosis require a specialized secretion system. PNAS 100:13001–6
    [Google Scholar]
  130. 130. 
    Steczkiewicz K, Muszewska A, Knizewski L, Rychlewski L, Ginalski K 2012. Sequence, structure and functional diversity of PD-(D/E)XK phosphodiesterase superfamily. Nucleic Acids Res 40:7016–45
    [Google Scholar]
  131. 131. 
    Tang JY, Bullen NP, Ahmad S, Whitney JC 2018. Diverse NADase effector families mediate interbacterial antagonism via the type VI secretion system. J. Biol. Chem. 293:1504–14
    [Google Scholar]
  132. 132. 
    Ting SY, Bosch DE, Mangiameli SM, Radey MC, Huang S et al. 2018. Bifunctional immunity proteins protect bacteria against FtsZ-targeting ADP-ribosylating toxins. Cell 175:1380–92.e14
    [Google Scholar]
  133. 133. 
    Uchida K, Leiman PG, Arisaka F, Kanamaru S 2014. Structure and properties of the C-terminal β-helical domain of VgrG protein from Escherichia coli O157. J. Biochem. 155:173–82
    [Google Scholar]
  134. 134. 
    Unterweger D, Kostiuk B, Otjengerdes R, Wilton A, Diaz-Satizabal L, Pukatzki S 2015. Chimeric adaptor proteins translocate diverse type VI secretion system effectors in Vibrio cholerae. . EMBO J 34:2198–210
    [Google Scholar]
  135. 135. 
    Unterweger D, Miyata ST, Bachmann V, Brooks TM, Mullins T et al. 2014. The Vibrio cholerae type VI secretion system employs diverse effector modules for intraspecific competition. Nat. Commun. 5:3549
    [Google Scholar]
  136. 136. 
    van der Wal FJ, Luirink J, Oudega B 1995. Bacteriocin release proteins: mode of action, structure, and biotechnological application. FEMS Microbiol. Rev. 17:381–99
    [Google Scholar]
  137. 137. 
    Vassallo C, Pathak DT, Cao P, Zuckerman DM, Hoiczyk E, Wall D 2015. Cell rejuvenation and social behaviors promoted by LPS exchange in myxobacteria. PNAS 112:E2939–46
    [Google Scholar]
  138. 138. 
    Vassallo CN, Cao P, Conklin A, Finkelstein H, Hayes CS, Wall D 2017. Infectious polymorphic toxins delivered by outer membrane exchange discriminate kin in myxobacteria. eLife 6:e29397Discovered that SitA lipoproteins are novel polymorphic effectors delivered through OME.
    [Google Scholar]
  139. 139. 
    Vassallo CN, Wall D. 2019. Self-identity barcodes encoded by six expansive polymorphic toxin families discriminate kin in myxobacteria. PNAS 116:24808–18
    [Google Scholar]
  140. 140. 
    Vettiger A, Basler M. 2016. Type VI secretion system substrates are transferred and reused among sister cells. Cell 167:99–110.e12
    [Google Scholar]
  141. 141. 
    Virtanen P, Waneskog M, Koskiniemi S 2019. Class II contact-dependent growth inhibition (CDI) systems allow for broad-range cross-species toxin delivery within the Enterobacteriaceae family. Mol. Microbiol. 111:1109–25
    [Google Scholar]
  142. 142. 
    Vollmer W, Pilsl H, Hantke K, Holtje JV, Braun V 1997. Pesticin displays muramidase activity. J. Bacteriol. 179:1580–83
    [Google Scholar]
  143. 143. 
    Walker D, Mosbahi K, Vankemmelbeke M, James R, Kleanthous C 2007. The role of electrostatics in colicin nuclease domain translocation into bacterial cells. J. Biol. Chem. 282:31389–97
    [Google Scholar]
  144. 144. 
    Wang J, Brackmann M, Castano-Diez D, Kudryashev M, Goldie KN et al. 2017. Cryo-EM structure of the extended type VI secretion system sheath-tube complex. Nat. Microbiol. 2:1507–12
    [Google Scholar]
  145. 145. 
    Whitney JC, Peterson SB, Kim J, Pazos M, Verster AJ et al. 2017. A broadly distributed toxin family mediates contact-dependent antagonism between gram-positive bacteria. eLife 6:e26938
    [Google Scholar]
  146. 146. 
    Whitney JC, Quentin D, Sawai S, LeRoux M, Harding BN et al. 2015. An interbacterial NAD(P)+ glycohydrolase toxin requires elongation factor Tu for delivery to target cells. Cell 163:607–19
    [Google Scholar]
  147. 147. 
    Wiener M, Freymann D, Ghosh P, Stroud RM 1997. Crystal structure of colicin Ia. Nature 385:461–64
    [Google Scholar]
  148. 148. 
    Wiener MC. 2005. TonB-dependent outer membrane transport: going for Baroque. Curr. Opin. Struct. Biol. 15:394–400
    [Google Scholar]
  149. 149. 
    Willett JL, Gucinski GC, Fatherree JP, Low DA, Hayes CS 2015. Contact-dependent growth inhibition toxins exploit multiple independent cell-entry pathways. PNAS 112:11341–46
    [Google Scholar]
  150. 150. 
    Wood TE, Howard SA, Forster A, Nolan LM, Manoli E et al. 2019. The Pseudomonas aeruginosa T6SS delivers a periplasmic toxin that disrupts bacterial cell morphology. Cell Rep 29:187–201.e7
    [Google Scholar]
  151. 151. 
    Yang X, Clemens DL, Lee BY, Cui Y, Zhou ZH, Horwitz MA 2019. Atomic structure of the Francisella T6SS central spike reveals a unique α-helical lid and a putative cargo. Structure 27:1811–19.e6
    [Google Scholar]
  152. 152. 
    Zakharov SD, Eroukova VY, Rokitskaya TI, Zhalnina MV, Sharma O et al. 2004. Colicin occlusion of OmpF and TolC channels: outer membrane translocons for colicin import. Biophys. J. 87:3901–11
    [Google Scholar]
  153. 153. 
    Zakharov SD, Zhalnina MV, Sharma O, Cramer WA 2006. The colicin E3 outer membrane translocon: immunity protein release allows interaction of the cytotoxic domain with OmpF porin. Biochemistry 45:10199–207
    [Google Scholar]
  154. 154. 
    Zepeda-Rivera MA, Saak CC, Gibbs KA 2018. A proposed chaperone of the bacterial type VI secretion system functions to constrain a self-identity protein. J. Bacteriol. 200:e00688-17
    [Google Scholar]
  155. 155. 
    Zeth K, Romer C, Patzer SI, Braun V 2008. Crystal structure of colicin M, a novel phosphatase specifically imported by Escherichia coli. J. Biol. Chem 283:25324–31
    [Google Scholar]
  156. 156. 
    Zhang D, de Souza RF, Anantharaman V, Iyer LM, Aravind L 2012. Polymorphic toxin systems: comprehensive characterization of trafficking modes, processing, mechanisms of action, immunity and ecology using comparative genomics. Biol. Direct. 7:18Proposed that LXG/WXG, PrsW, and phage head morphogenesis proteins deliver toxins for bacterial competition.
    [Google Scholar]
  157. 157. 
    Zhang D, Iyer LM, Aravind L 2011. A novel immunity system for bacterial nucleic acid degrading toxins and its recruitment in various eukaryotic and DNA viral systems. Nucleic Acids Res 39:4532–52
    [Google Scholar]
  158. 158. 
    Zoued A, Cassaro CJ, Durand E, Douzi B, Espana AP et al. 2016. Structure-function analysis of the TssL cytoplasmic domain reveals a new interaction between the type VI secretion baseplate and membrane complexes. J. Mol. Biol. 428:4413–23
    [Google Scholar]
  159. 159. 
    Zoued A, Durand E, Brunet YR, Spinelli S, Douzi B et al. 2016. Priming and polymerization of a bacterial contractile tail structure. Nature 531:59–63
    [Google Scholar]
/content/journals/10.1146/annurev-micro-020518-115638
Loading
/content/journals/10.1146/annurev-micro-020518-115638
Loading

Data & Media loading...

  • Article Type: Review Article
This is a required field
Please enter a valid email address
Approval was a Success
Invalid data
An Error Occurred
Approval was partially successful, following selected items could not be processed due to error