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

Clostridium difficile infection: the next big thing!

Michele M. Squire A and Thomas V. Riley A
+ Author Affiliations
- Author Affiliations

Microbiology & Immunology
School of Pathology & Laboratory Medicine
The University of Western Australia
Queen Elizabeth II Medical Centre
Nedlands 6009
Western Australia
Tel: 61-8-9346-3690
Fax: 61-8-9346-2912
E-mail: thomas.riley@uwa.edu.au

Microbiology Australia 33(4) 163-164 https://doi.org/10.1071/MA12163
Published: 1 November 2012

Abstract

Clostridium difficile causes infectious diarrhoea in humans and animals. It has been found in pigs, horses, and cattle, suggesting a potential reservoir for human infection, and in 20-40% of meat products in Canada and the USA, suggesting the possibility of food-borne transmission. It is likely that excessive antimicrobial exposure is driving the establishment of C. difficile in animals, in a manner analogous to human infection, rather than the organism just being normal flora of the animal gastrointestinal tract. Outside Australia, PCR ribotype 078 is the most common ribotype of C. difficile found in pigs (83% in one study in the USA) and cattle (up to 100%) and this ribotype is now the third most common ribotype of C. difficile found in humans in Europe. Human and pig strains of C. difficile are genetically identical in Europe confirming that a zoonosis exists. Rates of community-acquired C. difficile infection (CDI) are increasing world-wide, and a new community strain of unidentified origin has recently emerged in Australia. Environmental contamination may also play a role. C. difficile spores survive in treated piggery effluent, the by-products of which are used to irrigate crops and pasture and manufacture compost. There is abundant evidence that food products intended for human consumption contain toxigenic strains of C. difficile but food-borne transmission remains unproven. Thus there are four problems that require resolution: a human health issue, an animal health issue and the factors common to both these problems, environmental contamination and antimicrobial misuse.

Clostridium difficile is an anaerobic Gram positive spore-forming bacterium and the leading cause of infectious diarrhoea in hospitalised humans, usually after disruption of colonic flora by antimicrobials. The major virulence factors are two toxins, TcdA (an enterotoxin) and TcdB (a cytotoxin), while the role of a third “binary” toxin is unknown. Contamination of the environment with C. difficile spores plays a critical role in transmissibility. Spores are resistant to many commonly-used disinfectants and may remain in the environment for months1. There has been a suggestion that C. difficile is a threat to Australia’s biosecurity2. Now there is evidence that this warning has gone unheeded as rates of human CDI increase dramatically3 and C. difficile is recovered from Australian production animals4,5.


Human health

The emergence of strains causing C. difficile infection in the community (CA-CDI) is becoming more commonplace6. There is concern in Australia about the recent emergence of a novel highly virulent strain of C. difficile of community origin, the source of which has not been identified7-9[REMOVED HYPERLINK FIELD]. Unlike healthcare infections CA-CDI is associated with younger healthy people (particularly females), often without prior exposure to antimicrobials or contact with hospitalized patients10-13. Community strains differ from predominant hospital strains and are more heterogeneous, with many previously unidentified PCR ribotypes14. This suggests that other reservoirs of infection contribute to CA-CDI, possibly animals.


Animal and food sources of C. difficile

C. difficile is a recognized enteric pathogen in a variety of animals including companion animals (cats, dogs, horses) and food animals (cattle, sheep, goats, pigs)15,16. In Australia C. difficile has been isolated from piglets, sheep, lambs, horses, cats, dogs, and cattle, with the highest prevalence in neonatal animals due to a lack of established gut flora at birth. For this reason predisposing antibiotics may not be required for development of CDI in young animals although there is anecdotal evidence in Australia of routine use of extended-spectrum cephalosporins in production animals. This is particularly concerning in the pork industry where gross contamination of facilities with C. difficile spores is commonplace. C. difficile can be isolated from the faeces of piglets 1 hour after birth, presumably ingested from their environment. Within 48 hours 100% of piglets had acquired C. difficile of the same molecular type that was found in the piggery environment17. A 2011 Australian study showed contamination with toxigenic C. difficile increased from 0%-61% of sites within a swine facility only one month after occupation with pigs18[REMOVED HYPERLINK FIELD]. Airborne C. difficile spores can be found up to 20 metres from the pig facility19. The predominant genotype isolated from food production animals outside Australia is PCR ribotype 078, toxinotype V, NAP 7/8, REA group BK20. This ribotype is now the third most common European human ribotype21.

Overseas, meat products, seafood, ready-to-eat salads, salad leaves and vegetables are also contaminated with C. difficile, predominantly ribotype 078-like strains16,22-24. Contamination may occur through spillage of gut contents at slaughter or direct contamination by food handlers during processing or retailing. Environmental contamination may also play a role. C. difficile spores survive in treated piggery effluent, the by-products of which are then applied to agricultural land, used in retail compost manufacture, or recycled within the swine facility25.


Zoonotic transmission of C. difficile

Outside Australia, the increasing prevalence of PCR ribotype 078 in humans, food production animals and food products suggest potential zoonotic transmission. In the Netherlands, where infections with ribotype 078 increased more than four-fold from 2005 to 2008, patients infected with this ribotype were younger and acquired C. difficile in the community more frequently, particularly if they lived in rural pig producing areas26. In the USA, the prevalence of ribotype 078 infections in humans has increased from 0.02% to 1.3% (pre-2001 to 2006) and ribotype 078 is increasingly associated with CA-CDI. These strains are indistinguishable or very closely related to animal ribotype 078 strains by PFGE analysis27. Ribotype 078 strains from Dutch humans and pigs are indistinguishable by MLVA subtyping28. Transmission from humans to animals may also occur. C. difficile can be isolated from the faeces of hospital pet therapy dogs that had prior negative bacteriological cultures for C. difficile. These dogs were >2 times more likely to be colonised with C. difficile than dogs not visiting hospitals29.


Future

The Australian situation is less clear. Investigations are currently underway to identify the major strains of C. difficile causing community-onset infection and their origin. Interestingly, ribotype 078 has never been isolated from companion or production animals in Australia, and from very few human cases of CDI. Instead, we have found a variety of novel ribotypes and sequence types (STs) in animals, with some overlap of human strains. Many of these strains fall into clade 5, the clade of C. difficile that contains predominantly animal isolates. One, a pig strain, is genetically unique but MLST analysis shows it is similar to ribotype 078 (ST 11, clade 5)30. Much more work needs to be done, however it is likely that the use of certain antimicrobials in production animals is resulting in amplification of C. difficile and subsequent spillover into humans by an as yet undetermined mechanism. While the organism will be new to many proponents of “One Health” the concept will not and it will require a multi-disciplinary approach to prevent animal strains of C. difficile infiltrating human health systems.



References

[1]  Gerding, D. N. (2009) Clostridium difficile 30 years on: what has, or has not, changed and why? Int. J. Antimicrob. Agents 33, S2–8.
Clostridium difficile 30 years on: what has, or has not, changed and why?Crossref | GoogleScholarGoogle Scholar |

[2]  Riley, T. V. (2009) Is Clostridium difficile a threat to Australia's biosecurity? Med. J. Aust. 190, 661–662.

[3]  Western Australia, Department of Health, Public Health (2012) Healthcare Infection Surveillance Western Australia Quarterly Aggregate Report, 8. Retrieved from http://www.public.health.wa.gov.au/cproot/4496/2/q1-2012-agg-report.pdf.

[4]  Squire, M.M. et al. (2012) A novel molecular type of Clostridium difficile in neonatal pigs in Western Australia lacks tcdA and tcdC. (submitted for publication).

[5]  Knight, D.R. et al. (2012) Prevalence of Clostridium difficile carriage in Australian cattle of various ages. (submitted for publication).

[6]  Khanna, S. et al.. (2012) The epidemiology of community-acquired Clostridium difficile infection: a population-based study. Am. J. Gastroenterol. 107, 89–95.
The epidemiology of community-acquired Clostridium difficile infection: a population-based study.Crossref | GoogleScholarGoogle Scholar |

[7]  A. Gambotto-Burke, (2012, 28 July) Vigilance can pre-empt pandemics. In The Weekend Australian, News Corporation.

[8]  N. Bramwell, (2012, 6 July) Expert warns C. difficile deaths will rise. In Medical Observer, UBM Medica Australia Pty Ltd.

[9]  A. Bradley, (2012, 6 July) Rise in community C. difficile cases. In Australian Doctor, Reed Business.

[10]  Naggie, S. et al.. (2010) Community-associated Clostridium difficile infection: experience of a veteran affairs medical center in southeastern USA. Infection 38, 297–300.
Community-associated Clostridium difficile infection: experience of a veteran affairs medical center in southeastern USA.Crossref | GoogleScholarGoogle Scholar |

[11]  Kutty, P. K. et al.. (2010) Risk factors for and estimated incidence of community-associated Clostridium difficile infection, North Carolina, USA. Emerg. Infect. Dis. 16, 197–204.
Risk factors for and estimated incidence of community-associated Clostridium difficile infection, North Carolina, USA.Crossref | GoogleScholarGoogle Scholar |

[12]  Bauer, M. P. et al.. (2008) Community-onset Clostridium difficile-associated diarrhoea not associated with antibiotic usage--two case reports with review of the changing epidemiology of Clostridium difficile-associated diarrhoea. Neth. J. Med. 66, 207–211.

[13]  Wilcox, M. H. et al.. (2008) A case-control study of community-associated Clostridium difficile infection. J. Antimicrob. Chemother. 62, 388–396.
A case-control study of community-associated Clostridium difficile infection.Crossref | GoogleScholarGoogle Scholar |

[14]  Bauer, M. P. et al.. (2009) Clinical and microbiological characteristics of community-onset Clostridium difficile infection in The Netherlands. Clin. Microbiol. Infect. 15, 1087–1092.
Clinical and microbiological characteristics of community-onset Clostridium difficile infection in The Netherlands.Crossref | GoogleScholarGoogle Scholar |

[15]  Keel, M. K. and Songer, J. G. (2006) The comparative pathology of Clostridium difficile-associated disease. Vet. Pathol. 43, 225–240.
The comparative pathology of Clostridium difficile-associated disease.Crossref | GoogleScholarGoogle Scholar |

[16]  Rupnik, M. and Songer, J. G. (2010) Clostridium difficile: Its potential as a source of foodborne disease. Adv. Food Nutr. Res. 60C, 53–66.
Clostridium difficile: Its potential as a source of foodborne disease.Crossref | GoogleScholarGoogle Scholar |

[17]  Hopman, N. E. et al.. (2011) Acquisition of Clostridium difficile by piglets. Vet. Microbiol. 149, 186–192.
Acquisition of Clostridium difficile by piglets.Crossref | GoogleScholarGoogle Scholar |

[18]  Squire, M.M. and Riley, T.V. (2012) Clostridium difficile infection in humans and piglets: A 'One Health' Opportunity. Curr. Top. Microbiol. Immunol. Epub ahead of print 14 June.

[19]  Keessen, E. C. et al.. (2011) Aerial dissemination of Clostridium difficile on a pig farm and its environment. Environ. Res. 111, 1027–1032.
Aerial dissemination of Clostridium difficile on a pig farm and its environment.Crossref | GoogleScholarGoogle Scholar |

[20]  Songer, J. G. et al.. (2009) Clostridium difficile in retail meat products, USA, 2007. Emerg. Infect. Dis. 15, 819–821.
Clostridium difficile in retail meat products, USA, 2007.Crossref | GoogleScholarGoogle Scholar |

[21]  Bauer, M. P. et al.. (2011) Clostridium difficile infection in Europe: a hospital-based survey. Lancet 377, 63–73.
Clostridium difficile infection in Europe: a hospital-based survey.Crossref | GoogleScholarGoogle Scholar |

[22]  Bakri, M. M. et al.. (2009) Clostridium difficile in ready-to-eat salads, Scotland. Emerg. Infect. Dis. 15, 817–818.
Clostridium difficile in ready-to-eat salads, Scotland.Crossref | GoogleScholarGoogle Scholar |

[23]  Metcalf, D. et al.. (2011) Clostridium difficile in seafood and fish. Anaerobe 17, 85–86.
Clostridium difficile in seafood and fish.Crossref | GoogleScholarGoogle Scholar |

[24]  Metcalf, D. S. et al.. (2010) Clostridium difficile in vegetables, Canada. Lett. Appl. Microbiol. 51, 600–602.
Clostridium difficile in vegetables, Canada.Crossref | GoogleScholarGoogle Scholar |

[25]  M. M. Squire, et al., (2011) Detection of Clostridium difficile after treatment in a two-stage pond system. In Manipulating Pig Production (van Barneveld, R.J., ed), pp. 215, Australasian Pig Science Association.

[26]  Goorhuis, A. et al.. (2008) Emergence of Clostridium difficile infection due to a new hypervirulent strain, polymerase chain reaction ribotype 078. Clin. Infect. Dis. 47, 1162–1170.
Emergence of Clostridium difficile infection due to a new hypervirulent strain, polymerase chain reaction ribotype 078.Crossref | GoogleScholarGoogle Scholar |

[27]  Jhung, M. A. et al.. (2008) Toxinotype V Clostridium difficile in humans and food animals. Emerg. Infect. Dis. 14, 1039–1045.
Toxinotype V Clostridium difficile in humans and food animals.Crossref | GoogleScholarGoogle Scholar |

[28]  Debast, S. B. et al.. (2009) Clostridium difficile PCR ribotype 078 toxinotype V found in diarrhoeal pigs identical to isolates from affected humans. Environ. Microbiol. 11, 505–511.
Clostridium difficile PCR ribotype 078 toxinotype V found in diarrhoeal pigs identical to isolates from affected humans.Crossref | GoogleScholarGoogle Scholar |

[29]  Lefebvre, S. L. and Weese, J. S. (2009) Contamination of pet therapy dogs with MRSA and Clostridium difficile. J. Hosp. Infect. 72, 268–269.
Contamination of pet therapy dogs with MRSA and Clostridium difficile.Crossref | GoogleScholarGoogle Scholar |

[30]  Stabler, R. A. et al.. (2012) Macro and micro diversity of Clostridium difficile isolates from diverse sources and geographical locations. PLoS ONE 7, e31559.
Macro and micro diversity of Clostridium difficile isolates from diverse sources and geographical locations.Crossref | GoogleScholarGoogle Scholar |


Biographies

Michele Squire is completing a PhD in microbiology at The University of Western Australia, Nedlands. Her current research focuses on Clostridium difficile infection in neonatal piglets.

Tom Riley holds a Personal Chair at The University of Western Australia. He has had a long standing interest in healthcare-related infections, particularly the diagnosis, pathogenesis and epidemiology of Clostridium difficile infection.