Anda di halaman 1dari 12

Journal of Applied Microbiology ISSN 1364-5072

REVIEW ARTICLE

Dissemination of Clostridium difficile in food and the


environment: Significant sources of C. difficile
community-acquired infection?
K. Warriner1, C. Xu2, M. Habash3, S. Sultan3 and S.J. Weese4
1 Department of Food Science, University of Guelph, Guelph, ON, Canada
2 Shanghai Ocean University, Shanghai, China
3 School of Environmental Biology, University of Guelph, Guelph, ON, Canada
4 Pathobiology, University of Guelph, Guelph, ON, Canada

Keywords Summary
CDI, Clostridium difficile, Clostridium
perfringens, community-associated, Clostridium difficile is a significant pathogen with over 300 000 cases reported
endospores, food, germination, meat, in North America annually. Previously, it was thought that C. difficile was
microbiome. primarily a clinically associated infection. However, through the use of whole
genome sequencing it has been revealed that the majority of cases are
Correspondence
community acquired. The source of community-acquired C. difficile infections
Keith Warriner, Department of Food Science,
(CDI) is open to debate with foodborne being one route considered.
University of Guelph, Guelph, ON N1G 2W1,
Canada. Clostridium difficile fits the criteria of a foodborne pathogen with respect to
E-mail: kwarrine@uoguelph.ca being commonly encountered in a diverse range of foods that includes meat,
seafood and fresh produce. However, no foodborne illness outbreaks have been
2016/1640: received 27 July 2016, revised 25 directly linked to C. difficile there is also no conclusive evidence that its spores
October 2016 and accepted 27 October 2016 can germinate in food matrices. This does not exclude food as a potential
vehicle but it is likely that the pathogen is also acquired through zoonosis and
doi:10.1111/jam.13338
the environment. The most significant factor that defines susceptibility to CDI
is the host microbiome and functioning immune system. In this respect,
effective control can be exercised by reducing the environmental burden of C.
difficile along with boosting the host defences against the virulent enteric
pathogen.

source of C. difficile implicated in community-acquired


Introduction
CDI remains open to speculation although a foodborne
Clostridium difficile is responsible for 300 000 cases of link has been proposed given the faecaloral route of trans-
infection within North America each year and is attributed mission. Moreover, other clostridia such as Clostridium
to over 25 000 deaths. Although initially confined to North perfringens and Clostridium botulinum are well established
America, the pathogen has become distributed across the foodborne pathogens that would lend support to C. difficile
globe. Clostridium difficile was for a long time thought to following the same routes.
be restricted to clinical settings such as hospitals were it The classification of a foodborne pathogen is open to
continues to be the leading cause of antimicrobial-asso- interpretation although it has significant implications. Specif-
ciated diarrhoea (McFarland 2016). Yet, with increased ically, pathogens designated as foodborne need to be consid-
surveillance and the advent of whole genome sequencing ered in risk assessment and the means of prevention or
(WGS), the proportion of C. difficile infections (CDI) control to be then devised. The following provides a review
acquired from clinical settings is lower than 35% of the of our knowledge of C. difficile in relation to foodborne and
total cases reported (Knight et al. 2015). The remainder environmental sources. To provide context, the definition of
are placed under a large umbrella of community-acquired a foodborne pathogen will be A pathogen that can be
CDI that essentially encompasses cases that have no recent encountered or grow in foods at a sufficient level to cause ill-
contact with clinical settings or antibiotic use. The actual ness and has had historical association with foodborne illness

542 Journal of Applied Microbiology 122, 542--553 2016 The Society for Applied Microbiology
K. Warriner et al. Clostridium difficile in food and environment

outbreaks. This definition fits C. perfringens and C. botuli- would be too simplistic to suggest that these two ribo-
num both of which typically (but not exclusively) are types encompass all the virulent strains encountered.
encountered and grow in foods along with being implicated
in foodborne illness outbreaks. In this context, C. difficile can
Clostridium difficile mode-of-illness
be classified within the group of unspecified agents that are
recovered in food but the ability to cause illness via this route Clostridium difficile is an obligate anaerobe that follows
remains unknown (Scallen et al., 2011). the faecaloral route of infection. Endospores can be shed
in high levels by not only symptomatic patients but also
asymptomatic carriers, especially the young (Lees et al.
Background to Clostridium difficile
2016). The infectious dose of C. difficile is unknown but
When first isolated from an infants stool sample back in considered to be low (1001000 spores) and strongly
1935, C. difficile was named Bacillus difficile (Hall and dependent on host-related factors (Kansau et al. 2016).
OToole 1935). Later, the pathogen was transferred to clos- Once ingested, the spore passes through the stomach and
tridia on the basis that it was an obligate anaerobic spore then enters the duodenum where primary bile salts stim-
former. In some respects, C. difficile resembles C. perfrin- ulate the germination cascade (Bhattacharjee et al. 2016).
gens in terms of its primary habitat such as the gastroin- The cell undergoes outgrowth in the colon and caecum
testinal tract, association with meat and diarrhoea being where enterotoxins are produced. Virulent strains of
the major symptom of infection. However, on a genetic C. difficile express enterotoxin toxin A (TcdA) and the
level the two clostridia are distantly related to the extent cytotoxin toxin B (TcdB) (Lyras et al. 2009; Kuehne et al.
that it has been proposed to move C. difficile to the pep- 2010). Both toxins can disrupt the epithelial mucosal sur-
tostreptococcaceae family, thereby renaming the pathogen face and cause significant colonic inflammation (Badger
Peptoclostridium difficile (Yutin and Galperin 2013). More et al. 2012).
recently, reclassifying C. difficile as Clostridioides difficile Toxin A and toxin B are encoded by tcdA and tcdB
was proposed based on phenotypic, chemotaxonomic and genes that form part of the pathogenicity locus (PaLoc)
phylogenetic analysis (Lawson et al. 2016). Although along with the regulatory genes tcdR (positive regulator)
renaming C. difficile is unlikely it does underline distinct and tcdC (negative regulator) (Martin-Verstraete et al.
differences compared to other clostridia. 2016). In hypervirulent strains, there is a deletion in the
A further noteworthy feature of C. difficile is the large latter that results in increased toxin production although
genome size with a high proportion of mobile elements quorum sensing is also thought to be involved (Martin
that contributes to genetic plasticity (Knight et al. 2015). et al. 2013). Although nontoxigenic stains do not harbour
Phenotypically this has supported rapid diversification PaLoc, it can be acquired through horizontal transfer
and adaptability to different host environments (Knight from toxigenic C. difficile (Janoir 2016). Variations in the
et al. 2015). Genetically, C. difficile is comprised of a sequence of PaLoc are used to classify C. difficile into tox-
large number of strains divided between six phylogenetic inotypes of which there are currently 31 identified
clades. There has been a range of DNA typing techniques (Rupnik et al. 2009).
applied to differentiate C. difficile, with ribotyping emerg- Binary toxin is another virulence factor associated with
ing as the most widely applied. From a clinical point of hypervirulent strains that enhances the activity of Toxin
view, the most significant C. difficile are from Clade 2 A and B in a mechanism that is thought to disrupt the
and Clade 5 that contain ribotype 027 and 078 respec- immune response of the host (Janoir 2016). C. difficile
tively (Knight et al. 2015). However, the high rate of evo- strains harbouring the binary toxin but lacking Toxin A
lution within C. difficile gives rise to genetic lineages and B have also been recovered from CDI patients sug-
from ribotypes 027 and 078. With so many strains it is gesting that the latter two toxins are not necessarily
not unexpected to find that virulence and other pheno- required to cause symptoms of infection (Eckert et al.
typic characteristics vary significantly within the same 2015).
ribotype (Knight et al. 2015). For example, a number of
ribotypes that evolved from 027 and 078 have missing or
Antimicrobial resistance of Clostridium difficile
nonfunctional virulence genes, while others have alterna-
tive virulence factors (Badger et al. 2012). This is more The wide spread use of fluoroquinolone coincided with
so for ribotype 078 but has also been observed in dece- the emergence of hypervirulent C. difficile (Huang et al.
dents of 027 (Knetsch et al. 2011). Although WGS is clar- 2009). Resistance to antibiotics is encoded on trans-
ifying the phylogenetic relationships between C. difficile posons, some of which are encoded for by bacteriophages
strains, the majority of literature to date has focused on that facilitate widespread horizontal gene transfer (Roberts
ribotypes 027 and 078 due to clinical significance. Yet, it et al. 2014). Antimicrobial resistance of clinical C. difficile

Journal of Applied Microbiology 122, 542--553 2016 The Society for Applied Microbiology 543
Clostridium difficile in food and environment K. Warriner et al.

isolates have been found to vary between geographical proximity. However, this view was questioned with the
locations (Huang et al. 2009). For example, erythromycin WGS of C. difficile isolates from clinical settings (Eyre
resistance is more common in UK isolates with cefotax- et al. 2013). Eyre et al. (2013) sequenced over 1200 iso-
one resistance being more frequently encountered in lates collected over a period of 4 years at a University
North America (Huang et al. 2009). Different antibiotic- hospital. The key conclusions from the study were that
resistant profiles have also been observed in isolates only 35% of the cases could be attributed to direct
derived from animal production. For example, clin- patient-to-patient transmission. Over 45% of the isolates
damycin resistance is more commonly encountered in were genetically distinct and most likely originated out-
cattle C. difficile isolates compared to swine, with line- side the hospital environment (Eyre et al. 2013). This
zolid resistance being more frequently found in pigs finding again underlined the increased significance of
(Thitaram et al. 2016). Linezoild resistance has also been community-associated CDI that was for many years con-
encountered in cattle which is significant considering the sidered secondary to those of hospital-acquired
antibiotic is not commonly used in animal production infections.
(Rodriguez-Palacios et al. 2011). It has been speculated The risk factors for community-acquired CDI differ to
that linezolid-resistant strains could be derived from the those of hospital-acquired infections. Specifically, com-
environment although selective pressure based on munity-acquired CDI occurs in young persons without
production practices cannot be discounted (Rodriguez- contact with hospital environments or a recent history of
Palacios et al. 2011; Bandelj et al. 2016). In addition to antibiotic but frequent use of proton inhibitors (Freeman
resistance, genes for antibiotic biosynthesis have been et al. 2010).
identified in strains of C. difficile ribotype 078 that, if
expressed, could potentially confer a competitive advan-
Incidence of Clostridium difficile in food animals
tage within the gastrointestinal tract environment
(Knetsch et al. 2011). The carriage of C. difficile in animals and foods has been
subjected to significant attention in a bid to confirm that
community-acquired CDI has origins in a zoonotic and/
Hospital-acquired infection vs community-acquired
or foodborne source. Clostridium difficile is frequently
infections
encountered in a diverse range of animals that includes
Community-acquired infections are essentially those per- both pets and those destined for meat production
sons who have not been discharged from a healthcare (Table 1). Young animals typically have the highest
facility in the previous 12 weeks and a patient who had prevalence of C. difficile with higher carriage being associ-
symptom onset during the first 48 h after hospitalization ated with herds being provided medicated feed, or those
(Cohen et al. 2010). In theory at least, foodborne acquisi- carrying underlying infections such as mastitis in the case
tion can take place in both hospitals and the community of dairy farms (Bandelj et al. 2016).
given foods can become contaminated at any point from Although ribotype 027 has been recovered in animals
production through consumption. Yet, historically food- (pigs, cattle and poultry) the main ribotype 078 accounts
borne transmission is commonly linked to community- for 7590% of isolates (Hawken et al. 2013b). From
acquired infection.
Another point of differentiation between hospital- and Table 1 The prevalence of Clostridium difficile ribotype 078 in food
community-acquired infections relates to the implicated animals
genotype. Specifically, hospital-acquired infections are
Animal Prevalence (%) Country Source
commonly associated with the hypervirulent strain ribo-
type 027/North American pulsotype (NAP) 1/toxinotype Piglets 667 Belgium Rodriguez et al. (2012)
III (Cohen 2009). The success of ribotype 027 is consid- Piglets 100 Holland Hopman et al. (2011)
ered to be due to the hypertoxin production, antibiotic Piglets 944 Spain Pel
aez et al. (2013)
Pig 91 Canada Weese et al. (2010c)
resistance, along with high sporulation and germination
Pig 83 Canada Keel et al. (2007)
rates (Akerlund et al. 2008). The hypervirulent ribotype Pig 67 Canada Weese et al. (2011)
027 was thought to emerge under a background of heavy Pig 78 Holland Koene et al. (2012)
fluoroquinolone that imposed selective pressure to Pig 31 Holland Keessen et al. (2011)
increase antibiotic resistance and virulence. Through Calves 75 Belgium Rodriguez et al. (2012)
using typing techniques (MLST and ribotyping) along Calves 67 Canada Costa et al. (2011)
with epidemiology, it was demonstrated that the main Calves 94 Canada Keel et al. (2007)
Calves 17 Germany Schneeberg et al. (2013)
transmission route of ribotype 027 was via horizontal
Calves 94 United States Hammitt et al. (2008)
transmission from symptomatic patients to those in close

544 Journal of Applied Microbiology 122, 542--553 2016 The Society for Applied Microbiology
K. Warriner et al. Clostridium difficile in food and environment

DNA sequence studies it has been established that 078 manure samples taken from the holding area and 15% of
strains recovered in pigs show high similarity to those carcasses at postbleed, in addition to 15% at posteviscera-
isolated from clinical cases thereby confirming transmis- tion (Hawken et al. 2013a). In a similar study performed in
sion from animal-to-human but also human-to-animal the United States, C. difficile was recovered from only three
(Knight et al. 2015). The result could be interpreted as carcasses (post chill) from over 1500 tested samples despite
evidence that C. difficile populations associated with food up to 30% prevalence in pigs at the production level
animals have a wide spectrum of different hosts or may (Susick et al. 2012). Yet, the C. difficile recovered were
simply reflect lack of diversity within ribotype 078 (Bak- antibiotic resistant and toxigenic. In a comparable study in
ker et al. 2010; Knight et al. 2015). a Korean pork slaughter house, two carcasses from 659
sampled tested positive for C. difficile representing a 02%
prevalence (Cho et al. 2015). Both isolates recovered were
Incidence of Clostridium difficile in wild animals
identified as 078 and expressed antibiotic resistance.
Wild animals have been considered as a vector for dissemi- Sampling performed in a Belgium slaughter house
nating enteric pathogens between farms and urban centres. recovered toxigenic C. difficile from 8% of beef and 7%
It has been hypothesized that wild animals could act as vec- pork carcasses (Rodriguez et al. 2013). Through typing of
tors for C. difficile which is supported by the recovery of the isolates, 19 ribotypes were identified that included the
the pathogen from raccoons (8%, 4/52) and shrews (1/2) hypervirulent 078. More significantly, ribotypes recovered
(Jardine et al. 2013). Yet, no toxigenic C. difficile have been from carcasses could be matched to clinical cases within
recovered from deer mice, house mice, skunks, rats, voles, Belgium (Rodriguez et al. 2013).
opossums and groundhogs suggesting low carriage in wild From studies performed within slaughter houses, it can
animals although more data are required to support such a be concluded that carriage of C. difficile on carcasses is
conclusion (Jardine et al. 2013). low. This could be attributed to the lack of cross-contam-
ination events during the slaughter process or effective-
ness of carcass decontamination methods. Yet, the
Incidence in meat and other foods
carriage of C. difficile on carcasses is likely to reflect the
The high prevalence of C. difficile in animals would translate sanitary practices with meat processing facilities.
into carriage of the pathogen on meat through cross-con- Several studies have been performed to determine the
tamination events during the slaughter process. Hawken incidence of C. difficile in meat products, in addition to
et al. (2013a) undertook a sampling study within a pork other foods (Table 2). A common finding was the recov-
slaughter house, whereby samples were taken within the ery of toxigenic C. difficile in both ground and intact
holding area and from carcasses at each point of the slaugh- meat from various species (Table 2). The data presented
ter process. Clostridium difficile was recovered from 80% of cannot be considered as true prevalence given the limited

Table 2 Incidence of Clostridium difficile in meat, vegetables and seafood

% Positive for
Product Country C. difficile % Toxigenic Source

Ground beef Canada 61 800 Visser et al. (2012)


Ground beef Canada 122 100 Weese et al. (2009)
Ground veal USA 80 Not tested Houser et al. (2012)
Ground pork USA <1 03 Mooyottu et al. (2015)
Ground pork USA 95 100 Harvey et al. (2011)
Chicken meat USA 444 100 Songer et al. (2009)
Chicken meat Canada 125 100 Weese et al. (2010b)
Processed Canada 45 100 Metcalf et al. (2010)
vegetables
Raw vegetables UK 23 714 Bakri et al. (2009)
Leafy greens Iran 57 166 Yamoudy et al. (2015)
Lettuce USA 46 100 Rodriguez-Palacios et al. (2014)
Lettuce USA 138 100 Han (2016)
Lettuce France 29 100 Eckert et al. (2013)
Molluscs Italy 39 2 Troiano et al. (2015)
Shellfish USA 45 100 Norman et al. (2014)
Seafood (general) Canada 48 100 Metcalf et al. (2011)

Journal of Applied Microbiology 122, 542--553 2016 The Society for Applied Microbiology 545
Clostridium difficile in food and environment K. Warriner et al.

sampling size used, along with different methodologies constant input from land runoff or effluent derived from
and tendency for studies that failed to recover the patho- wastewater treatments along with the longevity of endo-
gen not being published. With regard to the latter, it was spores. In a study performed within Southern Ontario,
noted that in an extensive sampling trial that screened 92% of C. difficile recovered from river sediments were
over 1755 retail meat samples, no C. difficile was detected. found to be toxigenic (Xu et al. 2014). Of more concern
(Limbago et al. 2012). It should also be noted that in was the recovery of ribotype 078 from municipal water
those studies where C. difficile were recovered, there was samples taken from domestic housing that suggests the
a need to be enriched indicating low levels of the patho- C. difficile endospores can survive through drinking water
gen. Nevertheless, in fulfilling the requirements to be treatments (Bizaid 2013).
classed as a foodborne pathogen, it can be considered Studies to determine the incidence of C. difficile in soil
that C. difficile can be potentially carried on meat prod- have been restricted to high-risk areas where manure has
ucts. been applied directly or indirectly to land. For example,
Additional foods have also been screened for the pres- soil sampled within a stud farm recovered 14 isolates
ence of C. difficile with a focus on those prone to be con- from 132 samples screened (11% prevalence) compared
taminated with faecal contamination. For example, to 1% were only mature horses were kept (B averud et al.
bivalve shellfish are filter feeders that can potentially 2003). Clostridium difficile was recovered from 22% of
accumulate enteric pathogens when raised in polluted soil samples taken from a farmers market in Zimbabwe
waters. From the limited studies performed, C. difficile (Simango and Mwakurudza 2008). In the same study, C.
have been isolated from both shellfish and fish suggesting difficile was found in environmental surface samples taken
a potential source of toxigenic strains (Metcalf et al. from floors, tables, contact surfaces, windowsill, floor
2011). drain, feed bowl and around watering devices (Simango
Fresh produce is another food group that has been and Mwakurudza 2008).
linked to outbreaks caused by enteric pathogens. Again, Baverud et al. (2003) reported a 4% of soil samples
sampling studies were limited and primarily sampled at taken from rural area parks, playgrounds, gardens and
retail that raises questions with respect to the original cultivated fields tested positive for C. difficile. Clostridium
source of contamination. Yet, toxigenic, antibiotic-resis- difficile was also isolated from 21% (22/104) of the
tant C. difficile strains have been recovered in vegetables suburb soil samples from public parks, gardens, play-
including leafy greens (Metcalf et al. 2010; Eckert et al. grounds and fields of South Wales (Al-saif and Brazier
2013; Han 2016). 1996).
The origins of C. difficile associated with vegetables
remain to be elucidated but is likely from water, manure,
Prevalence of Clostridium difficile in domestic homes
compost and/or biosolids. This hypothesis is supported
and residences
by the finding that C. difficile can survive sewage treat-
ment and can be recovered in high prevalence in bioso- Within hospital environments there is significant person-
lids along with effluent discharge (Xu et al. 2014). In the to-person or contact surface-to-person transmission of C.
aforementioned study, ribotype 078 was recovered from difficile (Boyle et al. 2015). In principal, C. difficile could
19% of primary sludge (21/108), 8% of digested sludge also be acquired and transferred within enclosed spaces
(8/106), 35% of biosolids (15/43) and 60% of down- such as domestic homes and residences. Homes for the
stream river samples (15/25). When biosolids harbouring elderly are well acknowledged to have high risk of C. dif-
C. difficile was applied to land, the pathogen persisted ficile due to a concentration of highly susceptible popula-
over a 9-month period with no overall decrease in levels tions and challenges in sanitation (Shin et al. 2016). In a
(Xu et al. 2016b). Indeed, evidence was found to suggest small study performed within an elderly home a food
that C. difficile could grow within the subsurface of bio- sample was found to be positive for C. difficile ribotype
soild-amended soil (Xu et al. 2016b). Given the combina- 078 demonstrating how foods can be contaminated dur-
tion of discharge or deposition of C. difficile into the ing preparation and handling (Rodriguez et al. 2015).
environment, along with long-term persistence, it can be The incidence of C. difficile on domestic environments
suggested that crops would be exposed to the enteric (e.g. kitchens) has not been studied to any great extent
pathogen. despite the risk of contaminating foods. From studies
performed in private residences, C. difficile was isolated
from 32% of the surfaces sampled (Alam et al. 2014).
Incidence of Clostridium difficile in the environment
The highest prevalence of C. difficile was from shoes with
Toxigenic C. difficile are commonly encountered in the the pathogen also being recovered from bathrooms in
sediments from rivers and recreational lakes due to addition to floors. Of the 30 households visited, C.

546 Journal of Applied Microbiology 122, 542--553 2016 The Society for Applied Microbiology
K. Warriner et al. Clostridium difficile in food and environment

difficile was recovered from 25 illustrating the wide distri- and those hurdles applied to control either C. perfrin-
bution of the pathogen within the domestic setting (Alam gens and C. botulinum would also be effective against
et al. 2014). the enteric pathogen (Lund and Peck 2015).
The original source of C. difficile within domestic
homes remains unclear although domestic pets have been
Germination of Clostridium difficile
identified as a significant source. Studies have reported a
range of carriage levels of 5% for dogs and 3% for cats Given that C. difficile exists in the endospore form out-
although carriage depends on the pets age (Weese et al. side the host environment there is a need to initiate ger-
2010b). The source of C. difficile could be through envi- mination prior to growth. Like most spores, germination
ronmental exposure, pet food and water. However, the is enhanced through an initial heat shock applied for a
transfer from humans to pets is also thought to be signif- short period (10 min) between 60 and 100C (Ghosh
icant. This was underlined by the fact that dogs tested et al. 2009). It was naturally assumed that C. difficile
more frequently for C. difficile when living in a household would follow the same germination mechanisms as other
with immune-comprised persons who carried the patho- clostridia such as C. perfringens. That is, receptors on the
gen (Weese et al. 2010a). inner spore membrane are activated to induce a cascade
The incidence of C. difficile in high population envi- of degradation reactions leading to eventual outgrowth.
ronments such as schools, cruise ships, restaurants or In most spore formers the receptors sense the presence of
gymnasiums has not been studied. However, a study per- amino acids (L-alanine being a classic example) that initi-
formed to assess the sanitary status of hotel rooms has ates the germination cascade. However, C. difficile is dif-
been undertaken. Here, samples were taken from the liv- ferent in that germination is not triggered by amino acids
ing area and bathroom surfaces within hotels across alone but requires bile or bile breakdown products (Bhat-
Canada (Xu et al. 2015). Two toxigenic C. difficile isolates tacharjee et al. 2016). In evolutionary terms this would
were recovered from over 300 samples screened make sense given the presence of bile would be an envi-
representing <1% prevalence. Although low, the recovery ronmental trigger so that the C. difficile spores germi-
of C. difficile again underlines the wide distribution of nated within the gastrointestinal tract. By accident or
the enteric pathogen. design, the presence of bile in the colon would also sig-
nify a disrupted microflora as microbes within the gas-
trointestinal tract would ordinarily breakdown bile
Growth of Clostridium difficile in foods
(Allegretti et al. 2016). Specifically, with a disrupted
The growth of C. perfringens and C. botulinum in foods microflora (for example, in the event of administration
is commonly, but not exclusively, a prerequisite for of antibiotics) the bile acids, principally cholate, tauro-
both pathogens to cause foodborne illness. The extent cholate and glycocholate, are not metabolized thereby
to which this applies to C. difficile remains unknown, activating the germination of C. difficile that then under-
although it should be noted that spores rather than goes out-growth with associated toxin production (Alle-
vegetative cells need to be ingested to cause CDI. Nev- gretti et al. 2016). A further secondary bile by-product,
ertheless, there remain large knowledge gaps with deoxycholate, can also stimulate germination but inhibits
respect to growth ranges of C. difficile, especially in the outgrowth of the vegetative cell, thereby having a duel
foods. It has been reported that C. difficile can metabo- effect.
lize a broad range of carbohydrates and acids and Given that the C. difficile endospore has a relatively
hence is adaptable to utilize the energy sources within small time frame to germinate and grow in the lower gas-
the gastrointestinal tract (Scaria et al. 2015). A compar- trointestinal tract, it follows that those strains that germi-
ative study has been performed to determine the nate quickly are more successful in causing infection. In
growth limits of ribotype 078 and 027. Both ribotypes this regard, it is interesting to note that in hypervirulent
grew within the pH range of 79 although greatly stains the inclusion of glycine or histidine, along with bile
reduced growth rates at pH values <65 and salt con- salts, enhances the germination of spores and thereby the
centrations >4% was observed. No growth occurs at ability to grow and produce toxin (Carlson et al. 2015).
4C although it could at 22 with an optimum of Yet, it should be noted that strains within the same ribo-
37C (unpublished results). The growth of C. difficile is type exhibit different responses to germinating agents.
also inhibited by preservatives commonly applied to Specifically, in some strains the proportion of superdor-
control clostridia, namely nitrite, nitrate, sodium mancy is high. Superdormancy is a state where spores
metabisulphite and nisin (Le Lay et al. 2016; Lim et al. experience a lag period prior to initiating germination
2016). Therefore, from the current evidence available it despite germinating agents being present. The lag period
can be concluded that the growth of C. difficile is rare can range from hours to weeks in some cases and is

Journal of Applied Microbiology 122, 542--553 2016 The Society for Applied Microbiology 547
Clostridium difficile in food and environment K. Warriner et al.

significant in terms of returning false-negative results in 027 (Rodriguez-Palacios et al. 2016). In relative terms,
diagnostic tests (Rodriguez-Palacios and LeJeune 2011). endospores of C. difficile spores show comparable or lower
From an evolutionary standpoint, superdormancy allows resistance to C. perfringens although a high level of inter-
a proportion of the spores to remain in the dormant state strain variation exists (Nakamura et al. 1985).
in the event the environmental conditions become Like other endospores, those of C. difficile are resistant
adverse for the germinating cells. Superdormancy has to chemical antimicrobials with 5 log CFU reduction
been observed in C. difficile ribotype 078 but is strain requiring chlorine concentrations of over 3000 ppm for
dependent (Xu et al. 2016a). 15 min or 30% v/v hydrogen peroxide vapour (Perez
Given the essential need for bile salts and neutral to et al. 2005; Barbut et al. 2009). Therefore, C. difficile
alkaline pH for germination of C. difficile spores, it is spores are likely to survive standard sanitation regimes
unlikely that this would occur in food systems. Indeed, applied in the industry.
there have been no reports published in the literature on
C. difficile spore germination in food systems. In our own
Host-related factors increase susceptibility to Clostridium
studies, no germination of ribotype 027 and 078 occurred
difficile
in extracts of beef or fish without the addition of sodium
taurocholate bile salts (unpublished results). Yet, both The susceptibility of the host plays a significant role in
beef and fish extract did support the growth of both C. determining the outcomes of infection. This is most
difficile ribotypes when inoculated with vegetative cells. evident for Listeria monocytogenes infection and is espe-
cially relevant in the case of C. difficile. Those within
the population most susceptible to CDI are the elderly
Sporulation and endospore resistance in Clostridium
especially with most people suffering from CDI being
difficile
over the age of 65 although the young (<3 years old)
One of the notable features of sporulation in C. difficile is are also considered a high-risk group (Britton and
a network that regulates sporulation which also controls Young 2014).
solventogenesis, biofilm formation and toxin production The susceptibility towards C. difficile is dependent in
(Burns and Minton 2011; Zidaric and Rupnik 2016). several factors that likely interplay to lead to CDI. For
Therefore, individual cells would respond differently to example, the innate and humoral immune system is
the same signal with only a proportion entering into the thought to play a vital role in protecting against C. diffi-
sporulation cycle (Saujet et al. 2014). In hypervirulent C. cile along with neutrophiles and general gastrointestinal
difficile the yield of spores and faecal shedding is high, physiology, all of which weaken with age (Shin et al.
thereby increasing the probability to be transferred to the 2016). The microbiome is frequently identified as the
next host (Kansau et al. 2016). main factor on defining if C. difficile can germinate and
Although the germination and growth of C. difficile in go on to produce toxin. Specifically, the decrease in the
foods is unlikely its endospores are encountered in food proportion of bile acid degraders by exposure to antibi-
products (Table 2). Reports have indicated that C. difficile otics increases the concentration of germinants within the
spores inoculated into meat can survive recommended colon thereby supporting C. difficile outgrowth and toxin
cooking practices (i.e. internal temperature of 73C) production (Allegretti et al. 2016).
(Redondo-Solono et al. 2016). Yet, it should be noted that Members of the microbiome can also directly inhibit
endospores would be expected to survive such temperatures the growth and toxin production in C. difficile. These can
given the inherent thermal resistance (Setlow 2007; Rodri- be naturally present or administered through probiotic
guez-Palacios et al. 2010; Paredes-Sabja et al., 2014). The preparations or faecal transplants (McFarland 2016). Pro-
reported D values for C. difficile at 100C varies from 2.5 to biotics can also produce antimicrobial compounds to
33 min depending on the strain, suspending medium and inhibit C. difficile growth or inhibit toxin production
methods of recovery (Lund and Peck 2015). With regard to (McFarland 2016).
the latter, the apparent heat resistance of C. difficile can be In community-acquired CDI one of the most signifi-
increased by inclusion of lysozyme in the recovery media. cant risk factor is proton inhibitors that are taken to
Here, the lysozyme replaces the cortex-degrading enzyme reduce stomach acid (Clooney et al. 2016). The reduced
during spore germination in the event that the endogenous acidity remodel the microbiome to decrease bacteroide-
enzymes were degraded (Lund and Peck 2015). In terms of tes and increase firmicutes thereby increasing suscepti-
strain variation, there is no conclusive evidence that thermal bility to CDI (Clooney et al. 2016). The microbiome-
resistance correlates with virulence although it is interesting modulating effects of proton inhibitors are not due to
to note that endospores derived from ribotype 078 have changes in acidity but more with the types of nutrients
enhanced heat tolerance compared to other types including that reaches the lower gastrointestinal tract (Clooney

548 Journal of Applied Microbiology 122, 542--553 2016 The Society for Applied Microbiology
K. Warriner et al. Clostridium difficile in food and environment

Watercourse

Seafood

Sewage Treatment Facilities

Vegetables

Biosolids Spreading

Animal Production

Community Acquired CDI


Wild Animals Meat

Figure 1 The cyling and recycling of Clostridium difficile from zoonotic (- - - -), environmental (. . .. . .. . ..) or foodborne (______) sources implicated in
community-associated infections. The main incubators for C. difficile are animal production and general population. Clostridium difficile associ-
ated with animals can contaminate meat during processing and survive up to the point of consumption. Manure produced in animal production
can contaminate crops directly or leach into irrigation water or watersheds. Treated sewage effluent and biosolids disposed in the environment
can contaminate crops and seafood when released into estuaries. In addition, potential zoonotic transfer from animals to humans and vice versa
can occur to further disseminate C. difficile. [Colour figure can be viewed at wileyonlinelibrary.com]

et al. 2016). There is also evidence that proton inhibi- to the lack of ability to germinate in the absence of bile
tors disrupt the host immune system that would ordi- salts. A counter argument would be that there are numer-
narily reduce the risk of CDI (Larcombe et al. 2016). ous examples of foodborne pathogens that can cause ill-
ness without growing in foods. Enteric viruses, protozoa
along with virulent pathogens such as Campylobacter
Conclusions
jejuni are classic examples. Yet, there is a large knowledge
Community-acquired CDI is increasing and accounts for gap on what the infectious dose of C. difficile actually is.
65% of cases reported with actual numbers probably Regardless of this fact, the most significant piece of evi-
being underestimated. Through various lines of research, dence against C. difficile being classed as a foodborne
C. difficile satisfies a number of criteria that would desig- pathogen is that no outbreaks have been traced to foods
nate the pathogen as foodborne. Specifically, C. difficile is contaminated with the pathogen.
encountered in foods such as meat and can survive the In reality, it is likely that C. difficile is foodborne but
cooking process. The pathogen is also encountered in this is one of several routes that includes zoonotic, water-
fresh produce and seafood that are minimally processed borne, environmental and person-to-person (Fig. 1). It is
and thereby remain so up to the point of consumption. reasonable to suggest that we are exposed to toxigenic C.
However, in other ways C. difficile is not a typical food- difficile on a daily basis but only leads to CDI when a
borne pathogen given that it cannot grow in foods due combination of events occurs to disrupt the microbiome

Journal of Applied Microbiology 122, 542--553 2016 The Society for Applied Microbiology 549
Clostridium difficile in food and environment K. Warriner et al.

and compromise the functioning of the immune system. Bakker, D., Corver, J., Harmanus, C., Goorhuis, A., Keessen,
In addition to host-related factors, the virulence of C. dif- E.C., Fawley, W.N., Wilcox, M.H. and Kuijper, E.J. (2010)
ficile strains also contributes to the likelihood of acquir- Relatedness of human and animal Clostridium difficile PCR
ing CDI. In this respect, the rapid evolution and ribotype 078 isolates determined on the basis of
diversification of C. difficile means that the focus on ribo- multilocus variable-number tandem-repeat analysis and
type 027 and 078 is too restrictive and virulence markers tetracycline resistance. J Clin Microbiol 48, 37443749.
would be more appropriate (Janoir 2016). Bakri, M.M., Brown, D.J., Butcher, J.P. and Sutherland, A.D.
In terms of control, it would be unfeasible to inactivate (2009) Clostridium difficile in ready-to-eat salads, Scotland.
Emerg Infect Dis 15, 817818.
C. difficile in foods due to the inherent resistance of
Bandelj, P., Blagus, R., Briski, F., Frlic, O., Rataj, A.V., Rupnik,
endospores. A more productive approach is to minimize
M., Ocepek, M. and Vengust, M. (2016) Identification of
the environmental burden of C. difficile through a One
risk factors influencing Clostridium difficile prevalence in
Health approach. For example, it has been proposed to
middle-size dairy farms. Vet Res 47, 11.
administer vaccines against C. difficile to reduce carriage
Barbut, F., Menuet, D., Verachten, M. and Girou, E. (2009)
in animals (Rodriguez-Palacios et al. 2013; Ghose and Comparison of the efficacy of a hydrogen peroxide dry-
Kelly 2015). Other initiatives including composting of mist disinfection system and sodium hypochlorite solution
biosolids or thermophilic sludge digestion have proven to for eradication of Clostridium difficile spores. Infect Con
be effective interventions to reduce the carriage of C. dif- Hosp Epidemiol 30, 507514.
ficile in biosolids (Xu et al. 2016a,b). Increasing the resis- B
averud, V., Gustafsson, A., Franklin, A., Aspan, A. and
tance of the host to CDI by tailoring the microbiome Gunnarsson, A. (2003) Clostridium difficile: prevalence in
would also be an effective protective approach. horses and environment, and antimicrobial susceptibility.
Equine Vet J 35, 465471.
Bhattacharjee, D., Francis, M.B., Ding, X., McAllister, K.N.,
Acknowledgements
Shrestha, R. and Sorg, J.A. (2016) Reexamining the
We thank the Ontario Ministry of Agriculture, Food and germination phenotypes of several Clostridium difficile
Rural Affairs for funding our work and scientists around strains suggests another role for the CspC germinant
the globe for continuing to gain further insights into receptor. J Bacteriol 198, 777786.
Clostridium difficile. Bizaid, F (2013). Distribution and sources of Clostridium
difficile present in water sources. MSc University of Guelph.
https://atrium.lib.uoguelph.ca/xmlui/bitstream/handle/
Conflict of Interest 10214/5254/Bazaid_Fahad_201301_Msc.pdf?sequence=1
Boyle, M.L., Ruth-Sahd, L.A. and Zhou, Z. (2015) Fecal
None declared. microbiota transplant to treat recurrent Clostridium
difficile infections. Crit Care Nurse 35, 5164.
References Britton, R.A. and Young, V.B. (2014) Role of the intestinal
microbiota in resistance to colonization by Clostridium
Akerlund, T., Persson, I., Unemo, M., Noren, T., Svenungsson, difficile. Gastroenterology 146, 15471553.
B., Wullt, M. and Burman, L.G. (2008) Increased Burns, D.A. and Minton, N.P. (2011) Sporulation studies in
sporulation rate of epidemic Clostridium difficile Type 027/ Clostridium difficile. J Microbiol Meth 87, 133138.
NAP1. J Clin Microbiol 46, 15301533. Carlson, P.E., Kaiser, A.M., McColm, S.A., Bauer, J.M., Young,
Alam, M.J., Anu, A., Walk, S.T. and Garey, K.W. (2014) V.B., Aronoff, D.M. and Hanna, P.C. (2015) Variation in
Investigation of potentially pathogenic Clostridium difficile germination of Clostridium difficile clinical isolates
contamination in household environs. Anaerobe 27, 3133. correlates to disease severity. Anaerobe 33, 6470.
Allegretti, J.R., Kearney, S., Li, N., Bogart, E., Bullock, K., Cho, A., Byun, J.W., Kim, J.W., Oh, S.I., Lee, M.H. and Kim,
Gerber, G.K., Bry, L., Clish, C.B. et al. (2016) Recurrent H.Y. (2015) Low prevalence of Clostridium difficile in
Clostridium difficile infection associates with distinct bile slaughter pigs in Korea. J Food Prot 78, 10341036.
acid and microbiome profiles. Aliment Pharmacol Therap Clooney, A.G., Bernstein, C.N., Leslie, W.D., Vagianos, K.,
43, 11421153. Sargent, M., Laserna-Mendieta, E.J., Claesson, M.J. and
Al-saif, N. and Brazier, J.S. (1996) The distribution of Targownik, L.E. (2016) A comparison of the gut
Clostridium difficile in the environment of South Wales. microbiome between long-term users and non-users of
J Med Microbiol 45, 133137. proton pump inhibitors. Aliment Pharmacol Ther 43,
Badger, V.O., Ledeboer, N.A., Graham, M.B. and Edmiston, C.E. 974984.
Jr (2012) Clostridium difficile: epidemiology, pathogenesis, Cohen, M.B. (2009) Clostridium difficile infections: emerging
management, and prevention of a recalcitrant healthcare- epidemiology and new treatments. J Pediatr Gastroenterol
associated pathogen. JPEN 36, 645662. Nut 48, 6365.

550 Journal of Applied Microbiology 122, 542--553 2016 The Society for Applied Microbiology
K. Warriner et al. Clostridium difficile in food and environment

Cohen, S.H., Gerding, D.N., Johnson, S., Kelly, C.P., Loo, Hawken, P., Weese, J.S., Friendship, R. and Warriner, K.
V.G., McDonald, L.C., Pepin, J. and Wilcox, M.H. (2010) (2013b) Longitudinal study of Clostridium difficile and
Clinical practice guidelines for Clostridium difficile methicillin-resistant Staphylococcus aureus associated with
infection in adults: 2010 update by the Society for pigs from weaning through to the end of processing.
Healthcare Epidemiology of America (SHEA) and the J Food Prot 76, 624630.
Infectious Diseases Society of America (IDSA). Inf Con Hopman, N.E.M., Keessen, E.C., Harmanus, C., Sanders, I.,
Hosp Epidemiol 31, 431455. van Leengoed, L., Kuijper, E.J. and Lipman, L.J.A. (2011)
Costa, M.C., Stampfli, H.R., Arroyo, L.G., Pearl, D.L. and Acquisition of Clostridium difficile by piglets. Vet Microbiol
Weese, J.S. (2011) Epidemiology of Clostridium difficile on 149, 186192.
a veal farm: prevalence, molecular characterization and Houser, B.A., Soehnlen, M.K., Wolfgang, D.R., Lysczek, H.R.,
tetracycline resistance. Vet Microbiol 152, 379384. Burns, C.M. and Jayarao, B.M. (2012) Prevalence of
Eckert, C., Burghoffer, B. and Barbut, F. (2013) Clostridium difficile toxin genes in the feces of veal calves
Contamination of ready-to-eat raw vegetables with and incidence of ground veal contamination. Food Pathog
Clostridium difficile in France. J Med Microbiol 62, 1435 Dis 9, 3236.
1438. Huang, H.H., Weintraub, A., Fang, H. and Nord, C.E. (2009)
Eckert, C., Emirian, A., Le Monnier, A., Cathala, L., de Antimicrobial resistance in Clostridium difficile. Int J
Montclos, H., Goret, J., Berger, P., Petit, A. et al. (2015) Antimicrob Agents 34, 516522.
Prevalence and pathogenicity of binary toxin-positive Janoir, C. (2016) Virulence factors of Clostridium difficile and
Clostridium difficile strains that do not produce toxins A their role during infection. Anaerobe 37, 1324.
and B. New Micro New Infect 3, 1217. Jardine, C.M., Reid-Smith, R.J., Rousseau, J. and Weese, J.S.
Eyre, D.W., Cule, M.L., Wilson, D.J., Griffiths, D., Vaughan, (2013) Detection of Clostridium difficile in small and
A., OConnor, L. and Ip, C.L.C. (2013) Diverse Sources of medium-sized wild mammals in Southern Ontario,
C. difficile Infection Identified on Whole-Genome Canada. J Wildlife Dis 49, 418421.
Sequencing. N Engl J Med 369, 11951205. Kansau, I., Barketi-Klai, A., Monot, M., Hoys, S., Dupuy, B.,
Freeman, J., Bauer, M.P., Baines, S.D., Corver, J., Fawley, Janoir, C. and Collignon, A. (2016) Deciphering
W.N., Goorhuis, B., Kuijper, E.J. and Wilcox, M.H. (2010) adaptation strategies of the epidemic Clostridium difficile
The changing epidemiology of Clostridium difficile 027 strain during infection through in vivo transcriptional
infections. Clin Microbiol Rev 23, 529549. analysis. PLoS ONE 11, e0158204.
Ghose, C. and Kelly, C.P. (2015) The prospect for vaccines to Keel, K., Brazier, J.S., Post, K.W., Weese, S. and Songer, J.G.
prevent Clostridium difficile infection. Infect Dis Clin N (2007) Prevalence of PCR ribotypes among Clostridium
Am 29, 145162. difficile isolates from pigs, calves, and other species. J Clin
Ghosh, S., Zhang, P., Li, Y.Q. and Setlow, P. (2009) Microbiol 45, 19631964.
Superdormant spores of Bacillus species have elevated wet- Keessen, E.C., Van den Berkt, A.J., Haasjes, N.H., Hermanus,
heat resistance and temperature requirements for heat C., Kuijper, E.J. and Lipman, L.J.A. (2011) The relation
activation. J Bacteriol 191, 55845591. between farm specific factors and prevalence of
Hall, I.C. and OToole, E. (1935) Intestinal flora in new-born Clostridium difficile in slaughter pigs. Vet Microbiol 154,
infants: with a description of a new pathogenic anaerobe, 130134.
Bacillus difficilis. Am J Dis Child 49, 390402. Knetsch, C.W., Hensgens, M.P.M., Harmanus, C., van der Bijl,
Hammitt, M.C., Bueschel, D.M., Keel, M.K., Glock, R.D., M.W., Savelkoul, P.H.M., Kuijper, E.J., Corver, J. and van
Cuneo, P., DeYoung, D.W., Reggiardo, C., Trinh, H.T. Leeuwen, H.C. (2011) Genetic markers for Clostridium
et al. (2008) A possible role for Clostridium difficile in the difficile lineages linked to hypervirulence. Microbiology 157,
etiology of calf enteritis. Vet Microbiol 127, 343352. 31133123.
Han, Y. (2016) Detection of antibiotic resistant Clostridium Knight, D.R., Elliott, B., Chang, B.J., Perkins, T.T. and Riley,
difficile in lettuce. MSc thesis. Louisiana State University. T.V. (2015) Diversity and evolution in the genome of
http://etd.lsu.edu/docs/available/etd-07072016-015631/ Clostridium difficile. Clin Microbiol Rev 28, 721741.
unrestricted/YiHan.pdf (accessed October 2016). Koene, M.G.J., Mevius, D., Wagenaar, J.A., Harmanus, C.,
Harvey, R.B., Norman, K.N., Andrews, K., Norby, B., Hume, Hensgens, M.P.M., Meetsma, A.M., Putirulan, F.F., van
M.E., Scanlan, C.M., Hardin, M.D. and Scott, H.M. Bergen, M.A.P. et al. (2012) Clostridium difficile in Dutch
(2011) Clostridium difficile in retail meat and processing animals: their presence, characteristics and similarities
plants in Texas. J Vet Diagn Invest 23, 807811. with human isolates. Clin Microbiol Infect 18, 778784.
Hawken, P., Weese, J.S., Friendship, R. and Warriner, K. Kuehne, S.A., Cartman, S.T., Heap, J.T., Kelly, M.L.,
(2013a) Carriage and dissemination of Clostridium difficile Cockayne, A. and Minton, N.P. (2010) The role of toxin
and methicillin resistant Staphylococcus aureus in pork A and toxin B in Clostridium difficile infection. Nature
processing. Food Cont 31, 433437. 467, 711713.

Journal of Applied Microbiology 122, 542--553 2016 The Society for Applied Microbiology 551
Clostridium difficile in food and environment K. Warriner et al.

Larcombe, S., Hutton, M.L. and Lyras, D. (2016) Involvement spores of Clostridium difficile strains. Microbiol Immun 29,
of bacteria other than Clostridium difficile in antibiotic- 113118.
associated diarrhoea. Tren Microbiol 24, Norman, K.N., Harvey, R.B., Andrews, K., Hume, M.E.,
463476. Callaway, T.R., Anderson, R.C. and Nisbet, D.J. (2014)
Lawson, P.A., Citron, D.M., Tyrrell, K.L. and Finegold, S.M. Survey of Clostridium difficile in retail seafood in College
(2016) Reclassification of Clostridium difficile as Station, Texas. Food Addit Contam Part A Chem Anal
Clostridioides difficile (Hall and OToole 1935) Prevot Control Expo Risk Assess 31, 11271129.
1938. Anaerobe 40, 9599. Paredes-Sabja, D., Shen, A. and Sorg, J.A. (2014) Clostridium
Le Lay, C., Dridi, L., Bergeron, M.G., Ouellette, M. and Fliss, difficile spore biology: sporulation, germination, and spore
I. (2016) Nisin is an effective inhibitor of Clostridium structural proteins. Trends in Microbiology 22, 406416.
difficile vegetative cells and spore germination. J Med 
Pelaez, T., Alcala, L., Blanco, J.L., Alvarez-P
erez, S., Marn, M.,
Microbiol 65, 169175. Martn-L opez, A., Catalan, P., Reigadas, E. et al. (2013)
Lees, E.A., Miyajima, F., Pirmohamed, M. and Carrol, E.D. Characterization of swine isolates of Clostridium difficile in
(2016) The role of Clostridium difficile in the paediatric Spain: a potential source of epidemic multidrug resistant
and neonatal gut a narrative review. Eur J Clin Microbiol strains? Anaerobe 22, 4549.
Infect Dis 35, 10471057. Perez, J., Springthorpe, V.S. and Sattar, S.A. (2005) Activity of
Lim, S.C., Foster, N.F. and Riley, T.V. (2016) Susceptibility of selected oxidizing microbicides against the spores of
Clostridium difficile to the food preservatives sodium Clostridium difficile: relevance to environmental control.
nitrite, sodium nitrate and sodium metabisulphite. Am J Infect Control 33, 320325.
Anaerobe 37, 6771. Redondo-Solono, M., Burson, D.E. and Thippareddi, H.
Limbago, B., Thompson, A.D., Greene, S.A., MacCannell, D., (2016) Thermal resistance of Clostridium difficile spores in
MacGowan, C.E., Jolbitado, B., Hardin, H.D., Estes, S.R. peptone and pork meat. J Food Prot 79, 14681474.
et al. (2012) Development of a consensus method for Roberts, A.P., Allan, E. and Mullany, P. (2014) The impact of
culture of Clostridium difficile from meat and its use in a horizontal gene transfer on the biology of Clostridium
survey of US retail meats. Food Microbiol 32, 448451. difficile. In Adv Microb Physiol, Vol 65: Advances in
Lund, B.M. and Peck, M.W. (2015) A possible route for Bacterial Pathogen Biology ed. Poole, R.K., pp 6382.
foodborne transmission of Clostridium difficile? Foodborne London: Academic Press Ltd-Elsevier Science Ltd.
Pathog Dis 12, 177182. Rodriguez, C., Taminiau, B., van Broeck, J., Avesani, V.,
Lyras, D., OConnor, J.R., Howarth, P.M., Sambol, S.P., Delmee, M. and Daube, G. (2012) Clostridium difficile in
Carter, G.P., Phumoonna, T., Poon, R., Adams, V. et al. young farm animals and slaughter animals in Belgium.
(2009) Toxin B is essential for virulence of Clostridium Anaerobe 18, 621625.
difficile. Nature 458, 11761181. Rodriguez, C., Avesani, V., van Broeck, J., Taminiau, B.,
Martin, M.J., Clare, S., Goulding, D., Faulds-Pain, A., Barquist, Delmee, M. and Daube, G. (2013) Presence of Clostridium
L., Browne, H.P., Pettit, L., Dougan, G. et al. (2013) The difficile in pigs and cattle intestinal contents and carcass
agr locus regulates virulence and colonization genes in contamination at the slaughterhouse in Belgium. Int J
Clostridium difficile 027. J Bacteriol 195, 36723681. Food Microbiol 166, 256262.
Martin-Verstraete, I., Peltier, J. and Dupuy, B. (2016) The Rodriguez, C., Korsak, N., Taminiau, B., Avesani, V., van
regulatory networks that control Clostridium difficile toxin Broeck, J., Brach, P., Delmee, M. and Daube, G. (2015)
synthesis. Toxins 8, E153. Clostridium difficile from food and surface samples in a
McFarland, L.V. (2016) Therapies on the horizon for Belgian nursing home: an unlikely source of
Clostridium difficile infections. Expert Opin Invest Drugs contamination. Anaerobe 32, 8789.
25, 541555. Rodriguez-Palacios, A. and LeJeune, J.T. (2011) Moist-heat
Metcalf, D.S., Costa, M.C., Dew, W.M.V. and Weese, J.S. resistance, spore aging, and superdormancy in Clostridium
(2010) Clostridium difficile in vegetables, Canada. Lett Appl difficile. Appl Environ Microbiol 77, 30853091.
Microbiol 51, 600602. Rodriguez-Palacios, A., Reid-Smith, R.J., Staempfli, H.R. and
Metcalf, D., Avery, B.P., Janecko, N., Matic, N., Reid-Smith, Weese, J.S. (2010) Clostridium difficile survives minimal
R. and Weese, J.S. (2011) Clostridium difficile in seafood temperature recommended for cooking ground meats.
and fish. Anaerobe 17, 8586. Anaerobe 16, 540542.
Mooyottu, S., Flock, G., Kollanoor-Johny, A., Upadhyaya, I., Rodriguez-Palacios, A., Koohmaraie, M. and LeJeune, J.T.
Jayarao, B. and Venkitanarayanan, K. (2015) (2011) Prevalence, enumeration, and antimicrobial agent
Characterization of a multidrug resistant C. difficile meat resistance of Clostridium difficile in cattle at harvest in the
isolate. Int J Food Microbiol 192, 111116. United States. J Food Prot 74, 16181624.
Nakamura, S., Yamakawa, K., Izumi, J., Nakashio, S. and Rodriguez-Palacios, A., Borgmann, S., Kline, T.R. and Lejeune,
Nishida, S. (1985) Germinability and heat resistance of J.T. (2013) Clostridium difficile in foods and animals:

552 Journal of Applied Microbiology 122, 542--553 2016 The Society for Applied Microbiology
K. Warriner et al. Clostridium difficile in food and environment

history and measures to reduce exposure. Anim Health Res (2015) Toxigenic Clostridium difficile PCR ribotypes in
Rev 14, 2935. edible marine bivalve molluscs in Italy. Int J Food
Rodriguez-Palacios, A., Ilic, S. and LeJeune, J.T. (2014) Microbiol 208, 3034.
Clostridium difficile with moxifloxacin/clindamycin Visser, M., Sepehrim, S., Olson, N., Du, T., Mulvey, M.R. and
resistance in vegetables in Ohio, USA, and prevalence Alfa, M.J. (2012) Detection of Clostridium difficile in retail
meta-analysis. J Path 2014, 158601. ground meat products in Manitoba. Can J Inf Dis Med
Rodriguez-Palacios, A., Ilic, S. and LeJeune, J.T. (2016) Microbiol 23, 2830.
Subboiling moist heat favors the selection of enteric Weese, J.S., Avery, B.P., Rousseau, J. and Reid-Smith, R.J.
pathogen Clostridium difficile PCR ribotype 078 spores in (2009) Detection and enumeration of Clostridium difficile
food. Can J Infect Dis Med Microbiol 8, 1462405. spores in retail beef and pork. Appl Environ Microbiol 75,
Rupnik, M., Wilcox, M.H. and Gerding, D.N. (2009) Clostridium 50095011.
difficile infection: new developments in epidemiology and Weese, J.S., Finley, R., Reid-Smith, R.R., Janecko, N. and
pathogenesis. Nature Rev Microbiol 7, 526536. Rousseau, J. (2010a) Evaluation of Clostridium difficile in
Saujet, L., Pereira, F.C., Henriques, A.O. and Martin- dogs and the household environment. Epidemiol Infect
Verstraete, I. (2014) The regulatory network controlling 138, 11001104.
spore formation in Clostridium difficile. FEMS Microbiol Weese, J.S., Reid-Smith, R.J., Avery, B.P. and Rousseau, J.
Lett 358, 110. (2010b) Detection and characterization of Clostridium
Scallen, E., Griffin, P.M., Angulo, F.J., Tauxa, R.V. and difficile in retail chicken. Lett Appl Microbiol 50, 362365.
Hoekstra, M. (2011) Foodborne illness acquired in the Weese, J.S., Wakeford, T., Reid-Smith, R., Rousseau, J. and
United States- unspecified agents. Emerg Inf Dis 17, 1622. Friendship, R. (2010c) Longitudinal investigation of
Scaria, J., Suzuki, H., Ptak, C.P., Chen, J.-W., Zhu, Y., Guo, Clostridium difficile shedding in piglets. Anaerobe 16, 501
X.-K. and Chang, Y.-F. (2015) Comparative genomic and 504.
phenomic analysis of Clostridium difficile and Clostridium Weese, J.S., Rousseau, J., Deckert, A., Gow, S.l. and Reid-
sordellii, two related pathogens with differing host tissue Smith, R.J. (2011) Clostridium difficile and methicillin-
preference. BMC Genom 16, 448. resistant Staphylococcus aureus shedding by slaughter-age
Schneeberg, A., Neubauer, H., Schmoock, G., Grossmann, E. and pigs. BMC Vet Res 7, 17.
Seyboldt, C. (2013) Presence of Clostridium difficile PCR Xu, C., Weese, J.S., Flemming, C., Odumeru, J. and Warriner,
ribotype clusters related to 033, 078 and 045 in diarrhoeic K. (2014) Fate of Clostridium difficile during wastewater
calves in Germany. J Med Microbiol 62, 11901198. treatment and incidence in Southern Ontario watersheds.
Setlow, P. (2007) I will survive: DNA protection in bacterial J Appl Microbiol 117, 1219.
spores. Trends Microbiol 15, 172180. Xu, C., Weese, S.J., Namvar, A. and Warriner, K. (2015)
Shin, J.H., High, K.P. and Warren, C.A. (2016) Older is not Sanitary status and incidence of methicillin-resistant
wiser, immunologically speaking: effect of aging on host Staphylococcus aureus and Clostridium difficile within
response to Clostridium difficile infections. J Gerontol A Canadian hotel rooms. J Environ Health 77, 815.
Biol Sci Med Sci 71, 916922. Xu, C., Salsali, H., Weese, S. and Warriner, K. (2016a)
Simango, C. and Mwakurudza, S. (2008) Clostridium difficile Inactivation of Clostridium difficile in sewage sludge by
in broiler chickens sold at market places in Zimbabwe and anaerobic thermophilic digestion. Can J Microbiol 62, 1623.
their antimicrobial susceptibility. Int J Food Microbiol 124, Xu, C., Wang, D., Huber, A., Weese, S.J. and Warriner, K.
268270. (2016b) Persistence of Clostridium difficile in wastewater
Songer, J.G., Trinh, H.T., Killgore, G.E., Thompson, A.D., treatment-derived biosolids during land application or
McDonald, L.C. and Limbago, B.M. (2009) Clostridium windrow composting. J Appl Microbiol 120, 312320.
difficile in retail meat products, USA, 2007. Emerg Infect Yamoudy, M., Mirlohi, M., Isfahani, B.N., Jalali, M.,
Dis 15, 819821. Esfandiari, Z. and Hosseini, N.S. (2015) Isolation of
Susick, E.K., Putnam, M., Bermudez, D.M. and Thakur, S. toxigenic Clostridium difficile from ready-to-eat salads by
(2012) Longitudinal study comparing the dynamics of multiplex polymerase chain reaction in Isfahan, Iran. Adv
Clostridium difficile in conventional and antimicrobial free Biomed Res 4, 87.
pigs at farm and slaughter. Vet Microbiol 157, 172178. Yutin, N. and Galperin, M.Y. (2013) A genomic update on
Thitaram, S.N., Frank, J.F., Siragusa, G.R., Bailey, J.S., Dargatz, clostridial phylogeny: Gram-negative spore formers and other
D.A., Lombard, J.E., Haley, C.A., Lyon, S.A. et al. (2016) misplaced clostridia. Environ Microbiol 15, 26312641.
Antimicrobial susceptibility of Clostridium difficile isolated Zidaric, V. and Rupnik, M. (2016) Sporulation properties and
from food animals on farms. Int J Food Microbiol 227, 15. antimicrobial susceptibility in endemic and rare
Troiano, T., Harmanus, C., Sanders, I., Pasquale, V., Clostridium difficile PCR ribotypes. Anaerobe 39, 183188.
Dumontet, S., Capuano, F., Romano, V. and Kuijper, E.J.

Journal of Applied Microbiology 122, 542--553 2016 The Society for Applied Microbiology 553

Anda mungkin juga menyukai