Register      Login
Microbiology Australia Microbiology Australia Society
Microbiology Australia, bringing Microbiologists together
RESEARCH ARTICLE

Clostridium perfringens extracellular toxins and enzymes: 20 and counting

Sarah A Revitt-Mills A , Julian I Rood A and Vicki Adams A B
+ Author Affiliations
- Author Affiliations

A Monash University
19 Innovation Drive
Clayton, Vic. 3800, Australia
Tel: +61 3 9902 9139
Fax: +61 3 9902 2222

B Email: vicki.adams@monash.edu

Microbiology Australia 36(3) 114-117 https://doi.org/10.1071/MA15039
Published: 14 August 2015

Clostridium perfringens is a Gram-positive, anaerobic bacterium that is widely distributed in the environment; it is found in soil and commonly inhabits the gastrointestinal tract of humans and animals1,2. The ubiquitous nature of this bacterium has resulted in it becoming a major cause of histotoxic and enteric diseases3. The success of C. perfringens as both a pathogen and a commensal bacterium lies in its ability to produce a large number of potent toxins and extracellular enzymes4. This diverse toxin repertoire results in a broad range of diseases including gas gangrene, various enterotoxaemias, food poisoning and necrotic enteritis46. Since 2007, six new toxins have been identified, adding to the ever-increasing range of potential C. perfringens virulence determinants. This paper briefly reviews the plethora of toxins and extracellular enzymes produced by C. perfringens, highlighting their importance in disease and strain classification as well as introducing the latest additions to the ever increasing C. perfringens toxin family.


Toxinotype strain classification

Like many clostridial species, the virulence of C. perfringens is dependent on the production of toxins7. Not all toxins are produced by any one strain, instead individual isolates vary in toxin carriage and production8. This toxin expression profile forms the basis of a toxinotype classification scheme; in which strains are classified into five toxinotypes (A–E) based on the production of four different toxins (alpha, beta, epsilon and iota)4,9. This typing scheme is now very much outdated, but it has been useful for classification as the different toxinotypes are often associated with specific diseases4,10 (Table 1). For example, clostridial myonecrosis and human food poisoning are associated with type A strains, whereas type B, C and D strains are most strongly associated with enteric diseases of livestock4.


Table 1. Properties of Clostridium perfringens toxins.
Click to zoom


Toxins and toxin gene location

The number of characterised C. perfringens toxins is ever increasing; with more than 20 different toxins and enzymes classified to date, see Table 13,5,9,11. With a few important exceptions, these toxins are encoded on large conjugative plasmids4,10,1218, which allows for potential toxin gene transfer between different C. perfringens strains in the gastrointestinal tract and may prolong disease10. C. perfringens utilises chromosomally encoded toxins, such as alpha-toxin and perfringolysin O, during human histotoxic infections or human food poisoning (C. perfringens enterotoxin, CPE)3. However, for reasons that are probably related to disease epidemiology, plasmid-encoded toxins are critical for non-foodborne human gastrointestinal diseases, human enteritis necroticans and gastrointestinal diseases of animals3,10.


The toxin categories of C. perfringens

The toxins of C. perfringens can be functionally classified into four broad categories: membrane damaging enzymes, pore-forming toxins, intracellular toxins, and hydrolytic enzymes4. Membrane damaging toxins such as alpha-toxin are enzymes that damage target cell membranes through their ability to breakdown the constituents of the mammalian cell membrane19. The pore-forming toxins comprise the largest toxin category and function to disrupt membrane permeability and ion transport by inserting into the membrane and forming a permeable channel or pore20. This category includes toxins such as perfringolysin O, beta-toxin, CPE, NetB and epsilon-toxin. Intracellular toxins, such as TpeL, BEC and iota-toxin, are internalised into target host cells where they act to disrupt the cellular cytoskeleton21. Hydrolytic enzymes, such as sialidases and hyaluronidases, are secreted by C. perfringens and degrade surface associated glycans or glycoproteins4,22. These enzymes are not essential for disease; however, they may still contribute to the overall virulence of the bacterium23.


Recent toxin discoveries

In recent years, the number of characterised C. perfringens toxins has increased significantly. The latest editions to the C. perfringens arsenal are the six novel toxins or putative toxins: NetB, BEC, TpeL, NetE, NetF and NetG.

NetB is a beta-barrel pore-forming toxin and, like many other C. perfringens toxins, it is encoded on large conjugative plasmids24,25. Since its discovery, NetB-encoding plasmids have been identified in many avian necrotic enteritis isolates; and netB deletion studies have indicated that NetB toxin, is required for the development of necrotic enteritis in chickens26.

BEC is a novel binary toxin, composed of two distinct components, BECa and BECb and it appears to function in a similar fashion to iota-toxin15, with the BECa component having actin-specific ADP-ribosyltranferase activity. BEC was discovered after two unrelated food poisoning outbreaks that were caused by C. perfringens strains that did not encode CPE; the toxin typically associated with human food poisoning15. Further studies showed that these strains produced a novel binary toxin, designated as BEC, suggesting that this new toxin was responsible for the enteric symptoms observed during these outbreaks15.

TpeL is a member of the clostridial monoglycosyltransferase toxin family, which mediate cytotoxic effects through the glycosylation of host cell proteins, and is related to the large toxins, TcdA and TcdB, of Clostridium difficile27. There is no definitive evidence that TpeL is involved in disease; however, it is postulated that this toxin may make a contribution to virulence27.

The most recently characterised toxins are the cytotoxic pore-forming toxin NetF and the putative pore-forming toxins NetE and NetG, which were shown to be encoded on plasmids in isolates causing necrotizing gastroenteritis and necrotizing enterocolitis in foals and dogs28. There is a significant association between NetF-producing strains and enteric disease in these animals; however, the function of NetE and NetG in disease remains to be investigated28.


Conclusion

The production of toxins is essential for C. perfringens-mediated disease and this bacterium utilises an arsenal of different toxins and extracellular enzymes to cause a myriad of diseases. Note that the primary function of these toxins is most likely not to cause disease, but to provide nutrients for the growth of C. perfringens cells, which have limited capability to synthesise amino acids and essential co-factors. Many of the toxins that are crucial contributors to disease, for example NetB, are encoded on conjugative plasmids10, and therefore may be readily disseminated to other strains. Recent findings support the theory that C. perfringens strains have a tight association with the species in which they cause disease; for example, NetB-producing strains in birds and NetF-producers in foals and dogs. Just about everywhere we look, if there is an unclassified disease from which lots of C. perfringens cells can be isolated, the chances are good that another toxin is waiting to be discovered … stay tuned!



Acknowledgements

SAR-M is the recipient of an Australian Postgraduate Scholarship.


References

[1]  McClane, B. et al. (2013) Clostridium perfringens. In Food microbiology: fundamentals and frontiers, Fourth edn, pp. 465–489. ASM Press, Washington, DC.

[2]  McClane, B.A. (2014) Clostridium perfringens. In Encyclopedia of Toxicology, Third edn (Wexler, P., ed.), pp. 987–988. Academic Press.

[3]  Uzal, F.A. et al. (2014) Towards an understanding of the role of Clostridium perfringens toxins in human and animal disease. Future Microbiol. 9, 361–377.
Towards an understanding of the role of Clostridium perfringens toxins in human and animal disease.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2cXmslyitr4%3D&md5=cfbb05f5dc0e60c85755ea8faff71a87CAS | 24762309PubMed |

[4]  Petit, L. et al. (1999) Clostridium perfringens: toxinotype and genotype. Trends Microbiol. 7, 104–110.
Clostridium perfringens: toxinotype and genotype.Crossref | GoogleScholarGoogle Scholar | 1:STN:280:DyaK1M3is1aqtw%3D%3D&md5=0be2c88540ac78f87b0c3b24b2f642ccCAS | 10203838PubMed |

[5]  Uzal, F.A. et al. (2010) Clostridium perfringens toxins involved in mammalian veterinary diseases. Open Toxinology J. 2, 24–42.
Clostridium perfringens toxins involved in mammalian veterinary diseases.Crossref | GoogleScholarGoogle Scholar | 24511335PubMed |

[6]  Brynestad, S. et al. (2001) Enterotoxin plasmid from Clostridium perfringens is conjugative. Infect. Immun. 69, 3483–3487.
Enterotoxin plasmid from Clostridium perfringens is conjugative.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3MXjtV2qtL8%3D&md5=cb70860bec0090aea948283833f2474eCAS | 11292780PubMed |

[7]  Hatheway, C.L. (1990) Toxigenic clostridia. Clin. Microbiol. Rev. 3, 66–98.
| 1:STN:280:DyaK3c7isVWkug%3D%3D&md5=13876b1b86c3153c1e24f51578409d37CAS | 2404569PubMed |

[8]  Robertson, S. et al. (2014) Bacteria: Clostridium perfringens. In Encyclopedia of Food Safety (Motarjemi, Y., ed), pp. 395–402. Academic Press.

[9]  Rood, J.I. and Cole, S.T. (1991) Molecular genetics and pathogenesis of Clostridium perfringens. Microbiol. Rev. 55, 621–648.
| 1:CAS:528:DyaK38Xht1Wqtrw%3D&md5=e972ca31c1e78e3aed18c8ddd33a4fb4CAS | 1779929PubMed |

[10]  Li, J. et al. (2013) Toxin plasmids of Clostridium perfringens. Microbiol. Mol. Biol. Rev. 77, 208–233.
Toxin plasmids of Clostridium perfringens.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3sXpvVehsrk%3D&md5=6b7c401424c1fe6fbd0fb14d764402b5CAS | 23699255PubMed |

[11]  Rood, J.I. (1998) Virulence genes of Clostridium perfingens. Annu. Rev. Microbiol. 52, 333–360.
Virulence genes of Clostridium perfingens.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK1cXms1Giurc%3D&md5=28246a0c89754b676f3db6db1fbfd58aCAS | 9891801PubMed |

[12]  Brynestad, S. and Granum, P.E. (2002) Clostridium perfringens and foodborne infections. Int. J. Food Microbiol. 74, 195–202.
Clostridium perfringens and foodborne infections.Crossref | GoogleScholarGoogle Scholar | 11981970PubMed |

[13]  Bannam, T.L. et al. (2011) Necrotic enteritis-derived Clostridium perfringens strain with three closely related independently conjugative toxin and antibiotic resistance plasmids. MBio 2, e00190-11.
Necrotic enteritis-derived Clostridium perfringens strain with three closely related independently conjugative toxin and antibiotic resistance plasmids.Crossref | GoogleScholarGoogle Scholar | 21954306PubMed |

[14]  Hughes, M.L. et al. (2007) Epsilon-toxin plasmids of Clostridium perfringens type D are conjugative. J. Bacteriol. 189, 7531–7538.
Epsilon-toxin plasmids of Clostridium perfringens type D are conjugative.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2sXht1KnurbP&md5=783d8dddeabbb6eed01260a38f15247eCAS | 17720791PubMed |

[15]  Yonogi, S. et al. (2014) BEC, a novel enterotoxin of Clostridium perfringens found in human clinical isolates from acute gastroenteritis outbreaks. Infect. Immun. 82, 2390–2399.
BEC, a novel enterotoxin of Clostridium perfringens found in human clinical isolates from acute gastroenteritis outbreaks.Crossref | GoogleScholarGoogle Scholar | 24664508PubMed |

[16]  Nagahama, M. et al. (2015) Recent insights into Clostridium perfringens beta-toxin. Toxins (Basel) 7, 396–406.
Recent insights into Clostridium perfringens beta-toxin.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2MXjvFGjtrY%3D&md5=6b5f08ddad514f35f8530742e834afd8CAS | 25654787PubMed |

[17]  Katayama, S. et al. (1996) Genome mapping of Clostridium perfringens strains with I-CeuI shows many virulence genes to be plasmid-borne. Mol. Gen. Genet. 251, 720–726.
| 1:CAS:528:DyaK28XltFWgtL8%3D&md5=405c803d1eba5f8a5ac254d6d062a3aeCAS | 8757404PubMed |

[18]  Miyamoto, K. et al. (2006) Complete sequencing and diversity analysis of the enterotoxin-encoding plasmids in Clostridium perfringens type A non-food-borne human gastrointestinal disease isolates. J. Bacteriol. 188, 1585–1598.
Complete sequencing and diversity analysis of the enterotoxin-encoding plasmids in Clostridium perfringens type A non-food-borne human gastrointestinal disease isolates.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD28Xhs1GlsLk%3D&md5=010e10411026ad1f6c7cba8b77d47bc9CAS | 16452442PubMed |

[19]  Sakurai, J. et al. (2004) Clostridium perfringens alpha-toxin: characterization and mode of action. J. Biochem. 136, 569–574.
Clostridium perfringens alpha-toxin: characterization and mode of action.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2MXislehtLY%3D&md5=1df1bd6f9688fc3cfc855e9418827ae1CAS | 15632295PubMed |

[20]  Popoff, M.R. (2014) Clostridial pore-forming toxins: powerful virulence factors. Anaerobe 30, 220–238.
Clostridial pore-forming toxins: powerful virulence factors.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2cXhtFWitLvE&md5=1fef2b99d8a1546ff45f9f7ca7b17906CAS | 24952276PubMed |

[21]  Sakurai, J. et al. (2009) Clostridium perfringens iota-toxin: structure and function. Toxins (Basel) 1, 208–228.
Clostridium perfringens iota-toxin: structure and function.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3cXpsl2rtA%3D%3D&md5=26bd40c27f2b64c3df305aecb50adfb4CAS | 22069542PubMed |

[22]  Z̆ukaitė, V. and Biziulevičius, G.A. (2000) Physico-chemical and catalytic properties of Clostridium perfringens hyaluronidase: an update. Anaerobe 6, 347–355.
Physico-chemical and catalytic properties of Clostridium perfringens hyaluronidase: an update.Crossref | GoogleScholarGoogle Scholar |

[23]  Chiarezza, M. et al. (2009) The NanI and NanJ sialidases of Clostridium perfringens are not essential for virulence. Infect. Immun. 77, 4421–4428.
The NanI and NanJ sialidases of Clostridium perfringens are not essential for virulence.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1MXht1Ghu7rF&md5=e6621c13f66676cbb1195eac1130781aCAS | 19651873PubMed |

[24]  Keyburn, A.L. et al. (2008) NetB, a new toxin that is associated with avian necrotic enteritis caused by Clostridium perfingens. PLoS Pathog. 4, e26.
NetB, a new toxin that is associated with avian necrotic enteritis caused by Clostridium perfingens.Crossref | GoogleScholarGoogle Scholar | 18266469PubMed |

[25]  Parreira, V.R. et al. (2012) Sequence of two plasmids from Clostridium perfringens chicken necrotic enteritis isolates and comparison with C. perfringens conjugative plasmids. PLoS One 7, e49753.
Sequence of two plasmids from Clostridium perfringens chicken necrotic enteritis isolates and comparison with C. perfringens conjugative plasmids.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC38XhvVCns7bO&md5=651235fd3ef3d3588dc7686a462157daCAS | 23189158PubMed |

[26]  Keyburn, A.L. et al. (2010) Association between avian necrotic enteritis and Clostridium perfringens strains expressing NetB toxin. Vet. Res. 41, 21.
Association between avian necrotic enteritis and Clostridium perfringens strains expressing NetB toxin.Crossref | GoogleScholarGoogle Scholar | 19931005PubMed |

[27]  Chen, J. and McClane, B.A. (2015) Characterization of Clostridium perfringens TpeL toxin gene carriage, production, cytotoxic contributions, and trypsin sensitivity. Infect. Immun. 83, 2369–2381.
Characterization of Clostridium perfringens TpeL toxin gene carriage, production, cytotoxic contributions, and trypsin sensitivity.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2MXhtVSjsr3K&md5=2de0242e0199954a83429d9f37cb0b2fCAS | 25824828PubMed |

[28]  Gohari, I.M. et al. (2015) A novel pore-forming toxin in type A Clostridium perfringens is associated with both fatal canine hemorrhagic gastroenteritis and fatal foal necrotizing enterocolitis. PLoS One 10, e0122684.
A novel pore-forming toxin in type A Clostridium perfringens is associated with both fatal canine hemorrhagic gastroenteritis and fatal foal necrotizing enterocolitis.Crossref | GoogleScholarGoogle Scholar |


Biographies

Sarah A Revitt-Mills is a PhD student in the Department of Microbiology at Monash University and is studying conjugative toxin plasmid biology as part of her PhD studies.

Julian I Rood is a Professor of Microbiology at Monash University and has led the field in clostridial conjugation mechanisms for many years. He also studies bacterial pathogenesis, predominantly mechanisms utilised by the human and animal pathogen, Clostridium perfringens, and the sheep pathogen, Dichelobacter nodosus.

Vicki Adams is a Research Fellow in the Department of Microbiology at Monash University and has studied mobile genetic elements found in the clostridia, in addition to the biology of large conjugative toxin plasmids of Clostridium perfringens.