Download Ecology and Epidemiology of Human Pathogen

Document related concepts

Foot-and-mouth disease wikipedia , lookup

Myxobolus cerebralis wikipedia , lookup

Fasciolosis wikipedia , lookup

Transcript
Ecology and Epidemiology of Human Pathogen Clostridium difficile in Foods, Food
Animals and Wildlife
DISSERTATION
Presented in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy
in the Graduate School of The Ohio State University
By
Alexander Rodriguez-Palacios
Graduate Program in Veterinary Preventive Medicine
The Ohio State University
2011
Dissertation Committee:
Dr. Jeffrey T. LeJeune, Advisor
Dr. Gireesh Rajashekara
Dr. Wondwossen A. Gebreyes
Dr. Song Liang
Copyrighted by
Alexander Rodriguez-Palacios
2011
Abstract
The world is experiencing an increase in illness and deaths associated with one of
the most common nosocomial pathogens, Clostridium difficile. However, the sources of
infection for people at risk, especially in the community which is also increasingly
affected, are often unknown. In 2006 a study reported new hypervirulent and multidrug
resistant epidemic C. difficile strains of relevance in human disease were present in
livestock and retail meats highlighting the potential for foodborne transmission. The main
goal of this PhD dissertation was to investigate and identify factors associated with C.
difficile shedding in food animals, food contamination at harvest and its relation with
wildlife and their environment with particular attention to inter/intra-species transmission
and dissemination potential. Knowing how epidemic strains of bacteria disseminate
among food animals and reach the food supply will help to identify strategies to mitigate
the risk of food contamination and therefore transmissibility to humans.
By using observational, experimental and applied microbiological studies this
dissertation documented: 1) the low prevalence of epidemic C. difficile in mature
livestock at harvest, 2) the potential emergence of multidrug resistant strains that are
unlikely to originate from food animal production systems, 3) fecal supershedding status
relevant for infectious disease dynamics and environmental contamination, and 4)
transient intestinal colonization and unaffected shedding dynamics despite enhanced use
ii
of antimicrobials in naturally infected cattle. Geographical clustering of captive whitetailed deer harboring emerging C. difficile PCR ribotype 078 was also documented.
Nontoxigenic C. difficile strains, which can prevent the virulent effect of toxigenic
strains, were unexpectedly prevalent in food animals at the time of harvest. Thermal
studies with recovered spores conclusively documented the actual food safety risk
highlighted in former publications, mechanistically showed that temperature affects cell
division but not spore germination, and that hypervirulent strains are significantly favored
with sublethal heating during cooking. Here it was also determined that wild birds are not
clinically affected by human epidemic C. difficile strains and that dissemination at a
regional scale is possible in anthropogenic and food animal ecosystems. The
identification of which food animal species and to what extent wild birds are active
disseminators of emerging C. difficile strains is relevant to advance our ecological
understanding of human diseases and to further develop strategies to reduce the risk of
food contamination. These studies are expected to help increase global public and
environmental health initiatives.
iii
To the ones I love, friends and family.
My Grandma, mama and tata.
To Cata who still forgives me with love and patience for captivating long working
sessions
And from the deepest of my liver …to my inspiring and cheering children
Gabi, Dani and beba Anna, my most precious non-repeatable experiments.
Thanks to them heartening butterflies still dance in celebration
iv
Acknowledgements
I am in great debt with all past and present laboratory members at Dr J. LeJeune
Laboratory for their support, technical assistance, laughter and sincere hugs of
appreciation in moments of joy. Especially, Rose Schelpii, Laura Harpster, Jennifer
Vanpelt (Mrs. Shrock), Pam Schlegel, Mark Morash, and Mike Kauffman. Special thanks
to my advisor Dr LeJeune for his continuous support and guidance and the other
members of my committee for being great professional examples. Thanks to our FAHRP
(Food Animal Health Research Program) family for the years of collaboration and
cooperation, especially Dr Y.M. Saif for advising me through small leadership enterprises
(OARDC scholars Association). I am also grateful to my former mentors in Canada,
especially Dr H Staemply and J.S. Weese for having kindly donated some of the C.
difficile strains used in some of the experiments here presented and for thoughtful
discussions. Thanks to the Public Health Preparedness for Infectious Diseases (PHPID)
program, Targeted Investment in Excellence, The Ohio State University for kindly
supporting me through their generous Graduate Award program. My dearest thanks to
Sanja Ilic, colleague and loving partner, and our little family for their permanent support;
without them and their help, half of the work here presented could have not been
v
seamlessly and happily accomplished. Thank you all, who gladly helped me personally
and professionally and that I have not listed.
vi
Vita
1989................................................................High - Colegio Nacional Nicolas Esguerra
1997................................................................D.V.M. Population Medicine,
........................................................................National University of Colombia
2004................................................................M.Sc. Microbiology, University of Guelph
2007 ...............................................................D.V.Sc. Internal Medicine-Epidemiology
........................................................................University of Guelph
2007 to present ..............................................Graduate Research Associate, Department
of Veterinary Preventive Medicine, The
Ohio State University
Publications
Journal Articles related to Research:
1. 2011 – Appl Env Microbiol. 77:3391-97. Rodriguez-Palacios, A, Pickworth C,
Loerch, S, JT LeJeune. Transient fecal shedding and limited animal-to-animal
transmission of Clostridium difficile in naturally infected feedlot cattle.
vii
2. 2011 – Appl Env Microbiol. 77:3085-91. Rodriguez-Palacios, A, JT LeJeune.
Moist-heat resistance, spore aging and superdormancy in Clostridium difficile.
[Google cites: 1 time].
3. 2010 – Anaerobe. 16:540-2. Rodriguez-Palacios, A, R Reid-Smith, H. Stämpfli, J.
Weese. Clostridium difficile survives minimal temperature recommended for
cooking ground meats. [Google cites: 4 times].
4. 2010 – Foodborne Path and Dis. 7:1031-7. French, E., Rodriguez-Palacios, A, and
J.T. LeJeune. Enteric bacterial pathogens with zoonotic potential isolated from farmraised deer.
5. 2009 – Emerg Infect Dis.. Rodriguez-Palacios, A, H. Stämpfli, R Reid-Smith, LC
McDonald, A. Thompson, Limbago B.,J. Weese. Possible seasonality of Clostridium
difficile in retail meats, Canada. [Google cites: 26 times].
6.
2009 - J Appl Microbiol. 106:393-401. Rodriguez-Palacios A, Staempfli HR,
Duffield T, Weese JS. Isolation of bovine intestinal Lactobacillus plantarum and
Pediococcus acidilactici with inhibitory activity against Escherichia coli O157 and
F5. [Google cites: 3 times].
7. 2008 - Toxicol Appl Pharmacol. 230:1-8. Epub Feb 12. You Q, Karrow NA, Cao H,
Rodriguez A, Mallard BA, Boermans HJ. Variation in the ovine cortisol response to
systemic bacterial endotoxin challenge is predominantly determined by signaling
within the hypothalamic-pituitary-adrenal axis. [Google cites: 2 times].
8. 2007 - Emerg Infect Dis. 13:485-7. Rodriguez-Palacios, A, H. Stämpfli, T.
Duffield, J. Weese. Isolation of Clostridium difficile from retail meat, Canada.
viii
[Google cites: 81 times, including the CDC in the MMRW weekly report, April 4,
2008 / 57(13);340-343, and Clin. Infect Dis. 2010 Sep 1;51(5):577-82].
9. 2007- Vet Microbiol. 124:166-72. Rodriguez-Palacios, A, H. Stämpfli, M. Satlker,
T. Duffield, J. Weese. Natural and experimental infection of neonatal calves with
Clostridium difficile. [Google cites: 10 times].
10. 2006 - Emerg Infect Dis. 12:1730-6. Rodriguez-Palacios, A., H. Stämpfli, T.
Duffield, A. Peregrine, L. Trotz-Williams, L. Arroyo, J. Brazier, J. Weese.
Clostridium difficile PCR-ribotypes in Calves (and Relatedness Between Bovine and
Human PCR Ribotypes), Canada. [Google cites: 78 times, including CDC -MMRW
weekly, April 4, 2008 / 57(13);340-343].
11. 2006 - BMC Vet Res. Nov 29;2:34. Cao H, Kabaroff LC, You Q, Rodriguez A,
Boermans HJ, Karrow NA. Characterization of ovine hepatic gene expression
profiles in response to Escherichia coli lipopolysaccharide using a bovine cDNA
microarray. [Google cites: 12 times].
12. 2006 - Vet Immunol Immunopathol. 113:113-24. Kabaroff LC, Rodriguez A,
Quinton M, Boermans H, Karrow NA. Assessment of the ovine acute phase response
and hepatic gene expression in response to Escherichia coli endotoxin. [Google
cites: 8 times]
13. 2004 - J Vet Intern Med 18:395-396. Rodriguez A., Weese, JS, Duffield, T,
Staempfli, HR. ACVIM Effect of oral probiotics on calf diarrhea: Clinical trial
published between 1973-2003. [Google cites: 2 times].
ix
Journal Articles related to Teaching:
1. 2007 - J Vet Intern Med. 21:336-9. Rodriguez-Palacios, A, Rolando Q., J. Baird,
M. Stalker, and K. McGurrin. Presumptive fluphenazine-induced hepatitis and
urticaria in a horse.
2. 2007 - Can Vet J. 48:420-2. J.B. Koenig, Rodriguez-Palacios, A; J.K.
Colquhoun, H. Stämpfli. Congenital colonic malformation in a 4-month-old
Standardbred foal.
3. 2006 - J. Vet Med Assoc. 229:104-10. Baird, J., Arroyo, L.G., Venust, M., Mc
Gurrin, M. K., Rodriguez-Palacios, A., Kenney, D., Aravagiri, M., Maylin, G.A.
Adverse extrapyramidal side effects in four horses given fluphenazine decanoate.
[Google cites: 2 times; key reference for new Chapter ―Fluphenazine toxicosis‖.
In: Current therapy in equine medicine. Robinson, E, Sprayberry, K, (eds).
Elsevier, 2009, the later cited 11 times].
4. 2006 - Can Vet J. 47:65-6. Fjordbakk, C.; Kenney, D.; Rodriguez A., Stalker, M;
Keller, S. Inflammatory aural polyp in a horse.
5. 1997 - Journal of the Colombian Vet Med association ACOVEZ, June. Oliver, O.,
Rodriguez A. et al. ―Chronic renal failure in a horse of Paso Colombiano: A Case
report‖.
Book Chapter
2008 - Weese, JS., Shariff, S. & Rodriguez-Palacios, A. Probiotics in Veterinary
Medicine. In Therapeutic Microbiology: Probiotics and Related Strategies, edited by
James Versalovic, Michael Wilson, July 2008. Publisher: ASM Press
x
Fields of Study
Major Field: Veterinary Preventive Medicine
xi
Table of Contents
Ecology and Epidemiology of Human Pathogen Clostridium difficile in Foods, Food
Animals and Wildlife .......................................................................................................... 1
DISSERTATION ................................................................................................................ 1
Abstract ............................................................................................................................... ii
Acknowledgements ............................................................................................................. v
Vita.................................................................................................................................... vii
Publications ....................................................................................................................... vii
Fields of Study ................................................................................................................... xi
Table of Contents .............................................................................................................. xii
List of Tables ................................................................................................................. xviii
List of Figures .................................................................................................................. xix
Chapter 1 ― PhD Program Project Description and Study Objectives .............................. 1
Abstract ........................................................................................................................... 1
Introduction ..................................................................................................................... 2
Hypothesis ....................................................................................................................... 3
Specific Aims .................................................................................................................. 3
Goals and Experimental Approach ................................................................................. 4
Goal 1 – Food Animals at harvest and food contamination ........................................ 4
Goal 2 – Ecology of C. difficile in wildlife and mechanisms of spore dispersion ...... 4
Goal 3 – Thermal resistance of C. difficile spores and spore superdormancy............. 5
Goal 4 – Risk of ingestion of C. difficile and food safety education ........................... 6
Chapter 2 ― Literature Review on Clostridium difficile in animals and retail foods:
Understanding and reducing risks ....................................................................................... 8
Summary ......................................................................................................................... 8
Introduction - why is this relevant today? ....................................................................... 9
History and Burden ......................................................................................................... 9
Major events in the history of C. difficile .................................................................... 9
More severity and cases of C. difficile infections internationally ............................. 13
The numbers .............................................................................................................. 13
xii
New patterns of disease — In brief ............................................................................... 15
Clostridium difficile has also changed as a pathogen ................................................ 16
Not everyone knows where or why they became sick with C. difficile..................... 17
Pathogen and facilitator factors ..................................................................................... 18
Strains today are more resistant to antibiotics than before 2000 ............................... 18
Clostridium difficile is an environmental-resistant microorganism .......................... 19
Fate of Clostridium difficile spores following ingestion ........................................... 19
If you get sick once, you can get sick twice, or more................................................ 21
Differentiating C.difficile from various potential origins.............................................. 22
Molecular fingerprinting (―bar-coding‖) of C. difficile ............................................. 22
There are several C. difficile strains (lineages) but often few are the bad ones ........ 23
About the Disease.......................................................................................................... 24
How does C. difficile ingestion result in intestinal disease?...................................... 24
New C. difficile strains can produce much more toxins - why? ................................ 25
C. difficile exposure vs. risk factors for CDI — Two different things ...................... 26
Where can we get both the spores and the risk for CDI? — Onset vs. Acquisition .. 27
About Animals, the Environment and Foods ................................................................ 28
Clostridium difficile in the environment and animals ................................................... 28
Companion Animals: Household pets - Dogs, cats and other creatures .................... 30
Companion Animals: Horses - Pleasure and working animals ................................. 31
Food Animals: Pigs ................................................................................................... 31
Food Animals: Cattle ................................................................................................. 33
Food Animals: Poultry............................................................................................... 34
Waters and the environment ...................................................................................... 35
C. difficile in Foods ................................................................................................... 36
Potential for control of C. difficile in foods and food animals ...................................... 38
References ..................................................................................................................... 45
Chapter 3 ― Transient Fecal Shedding and Limited Animal-to-Animal Transmission of
Clostridium difficile by Naturally Infected Finishing Feedlot Cattle ............................... 55
Abstract ......................................................................................................................... 55
Introduction ................................................................................................................... 56
Materials and Methods .................................................................................................. 58
Experimental unit and source of animals. ................................................................. 58
xiii
Animal allocation, sampling and antimicrobial therapy. ........................................... 58
Laboratory analysis.................................................................................................... 60
Molecular subtyping. ................................................................................................. 61
Verification of toxin production. ............................................................................... 62
Toxinotyping and antimicrobial susceptibility. ......................................................... 62
Statistical analysis...................................................................................................... 63
Results ........................................................................................................................... 64
Shedding at arrival. .................................................................................................... 64
Shedding over time and at harvest. ............................................................................ 64
Antimicrobial use and C. difficile shedding. ............................................................. 65
Reallocation and animal-to-animal transmission. ..................................................... 66
Molecular subtyping. ................................................................................................. 67
Toxin production, toxinotyping and antibiotic sensitivity......................................... 67
Discussion ..................................................................................................................... 68
Acknowledgements ....................................................................................................... 73
References ..................................................................................................................... 79
Chapter 4 ― Summer prevalence of Clostridium difficile in pigs, cattle, poultry, and
horses: finishing food animals are not naturally reservoirs .............................................. 84
Abstract ......................................................................................................................... 84
Introduction ................................................................................................................... 85
Materials and Methods .................................................................................................. 87
Food animal farms, processing plants, and horse racetrack. ..................................... 87
Sampling. ................................................................................................................... 88
C. difficile prevalence and molecular confirmation................................................... 89
Toxin ELISAs and dialysis cytotoxicity assay in Vero cells. .................................... 90
Statistics. .................................................................................................................... 90
Results ........................................................................................................................... 91
Discussion ..................................................................................................................... 92
Acknowledgements ....................................................................................................... 95
References ..................................................................................................................... 97
Chapter 5 ― Prevalence, enumeration and antimicrobial resistance in Clostridium
difficile in cattle at harvest, USA .................................................................................... 102
Abstract ....................................................................................................................... 102
Introduction ................................................................................................................. 103
Materials and Methods ................................................................................................ 104
xiv
Meat processing plants. ........................................................................................... 104
Sampling of animals. ............................................................................................... 105
Enrichment culture and molecular typing. .............................................................. 106
Toxin production. .................................................................................................... 107
Dialysis tubing and toxin cytotoxicity assay on Vero cells. .................................... 108
Antimicrobial susceptibility. ................................................................................... 109
Magnitude of fecal shedding. .................................................................................. 109
Statistics. .................................................................................................................. 110
Results ......................................................................................................................... 111
Prevalence. ............................................................................................................... 111
Genotypes. ............................................................................................................... 111
Antimicrobial susceptibility. ................................................................................... 112
Magnitude of shedding. ........................................................................................... 112
Discussion ................................................................................................................... 113
Acknowledgements ..................................................................................................... 118
References ................................................................................................................... 124
Chapter 6 ― Clostridium difficile in retail fresh vegetables, Ohio, USA ...................... 130
Abstract ....................................................................................................................... 130
Introduction ................................................................................................................. 130
Materials and Methods ................................................................................................ 131
Results and Discussion ................................................................................................ 132
Acknowledgements ..................................................................................................... 133
References ................................................................................................................... 135
Chapter 7 ― Enteric bacterial pathogens with zoonotic potential isolated from farmraised white-tailed deer ................................................................................................... 137
Abstract ....................................................................................................................... 137
Introduction ................................................................................................................. 138
Materials and Methods ................................................................................................ 139
Sampling frame........................................................................................................ 139
Microbiological and Molecular Analysis ................................................................ 139
Farm practices.......................................................................................................... 143
Statistics ................................................................................................................... 143
Results ......................................................................................................................... 143
Discussion ................................................................................................................... 145
Acknowledgements ..................................................................................................... 152
References ................................................................................................................... 157
xv
Chapter 8 ― Clostridium difficile in anthropogenic wild bird ecosystems and
dissemination potential ................................................................................................... 161
Abstract ....................................................................................................................... 161
Introduction ................................................................................................................. 162
The study ..................................................................................................................... 163
Retrospective spatial analysis of emerging C. difficile in cattle .............................. 163
Considerations about aerial dissemination .............................................................. 164
Clostridium spp. in wild birds over the past 40 years ............................................. 165
Prevalence of C. difficile in wild birds from anthropogenic and forest ecosystems 167
Verification of low shedding prevalence in small migratory birds ......................... 168
Experimental infection of wild birds with C difficile.............................................. 169
Telemetry study of bird movements and C. difficile typing .................................... 169
Discussion ................................................................................................................... 170
Acknowledgements ..................................................................................................... 172
References ................................................................................................................... 177
Chapter 9 ― Moist heat resistance, spore aging, and superdormancy in Clostridium
difficile ............................................................................................................................ 179
Abstract ....................................................................................................................... 179
Introduction ................................................................................................................. 180
Materials and Methods ................................................................................................ 182
Clostridium difficile strains ..................................................................................... 182
Spore production and aging ..................................................................................... 182
Thermal inhibition in liquid media .......................................................................... 183
Mechanism of thermal inhibition ............................................................................ 184
Thermal inhibition in ground beef and gravy .......................................................... 186
Statistics ................................................................................................................... 188
Results ......................................................................................................................... 188
Thermal inhibition in liquid media .......................................................................... 188
Heat treatment at 85oC affects cell division but not spore germination .................. 189
Thermal inhibition at 85oC can revert during incubation in broth .......................... 189
Thermal inhibition in ground beef and gravy .......................................................... 190
Discussion ................................................................................................................... 191
Acknowledgements ..................................................................................................... 195
References ................................................................................................................... 204
xvi
Chapter 10 ― Cooking and the Selection of Emerging Pathogens: Clostridium difficile
......................................................................................................................................... 210
Abstract ....................................................................................................................... 210
Introduction ................................................................................................................. 210
Materials and Methods ................................................................................................ 212
Results ......................................................................................................................... 213
Discussion ................................................................................................................... 213
Acknowledgements ..................................................................................................... 214
Supplementary Materials and Methods ....................................................................... 220
Rationale for temperatures tested and hypothesis ................................................... 220
Clostridium difficile strains ..................................................................................... 220
Experiment 1 – Individual thermal testing .............................................................. 221
Experiment 2 – Collective thermal testing of C. difficile mixture........................... 221
Systematic random sampling of survivor C. difficile .............................................. 222
Natural contamination in beef with C difficile and background microbial flora as
controls .................................................................................................................... 222
Statistical Analysis .................................................................................................. 223
References ................................................................................................................... 224
Chapter 11 ― Clostridium difficile and higher cooking temperatures for communities at
risk................................................................................................................................... 227
Acknowledgements ..................................................................................................... 232
References ................................................................................................................... 236
Chapter 12 ― Conclusions ............................................................................................. 237
All References ................................................................................................................. 239
xvii
List of Tables
Table 1. Clostridium difficile types with increasing relevance to humans and animals ... 44
Table 2.Toxin gene profiles of C. difficile PCR ribotypes from beef cattle in the feedlot.
........................................................................................................................................... 78
Table 3. Comparative summer prevalence of C. difficile fecal shedding in farm animals,
Ohio, USA......................................................................................................................... 96
Table 4.Estimated prevalence of cattle carrying C. difficile at harvest, USA, summer
2008.a,b ........................................................................................................................... 122
Table 5. Antimicrobial susceptibility of C. difficile from cattle at harvest.a .................. 123
Table 6. Retail fresh vegetables tested for Clostridium difficile in Ohio, USA, 2007.... 134
Table 7. Human enteric pathogens among 30 captive-deer operations in Ohio, November
2006-May 2007, USA. .................................................................................................... 155
Table 8. Culture recovery and toxin gene profiles of 37 C. difficile isolates from 11 deer
farms, USA, 2007a .......................................................................................................... 156
Table 9. Coefficient estimates of heat-shock enhancement of spore germination after 7h
of incubation in nutritious BHI broth.............................................................................. 202
Table 10. Minutes of heat needed to inhibit 90% of spores (D-values) estimated for a
1:1:1:1 mixture of C. difficile of PCR ribotypes 027, 078, 077 and ATCC 9689 strains in
three food matrices .......................................................................................................... 203
Table 11. Evidence that C. difficile PCR ribotype 078 is increasing internationally. .... 218
Table 12. Recommended safe minimum cooking temperatures and holding times to
reduce exposure to foodborne pathogens. ....................................................................... 219
Table 13. Minutes of heat needed to inhibit 90% of C. difficile spores at achievable
household cooking temperatures..................................................................................... 233
xviii
List of Figures
Figure 1. Increase of toxigenic C. difficile strains A-B+ in a multi-center study in
hospitals of Korea compared to A+B+ strains, 2000-2005 (Shin et al. 2008a). ................ 40
Figure 2. Study testing C. difficile toxins in fecal samples from patients visiting 40
hospitals and over 2,000 physicians in southern Germany. .............................................. 41
Figure 3. Possible routes of transmission of C. difficile spores. ....................................... 42
Figure 4. Clostridium difficile isolates from human patients that were molecularly
characterized with an international PCR ribotyping method in the Netherlands.............. 43
Figure 5. Clostridium difficile in steers at arrival to the feedlot. ...................................... 74
Figure 6. Longitudinal study ― Prevalence of C. difficile shedding in cattle in the feedlot,
and at harvest. ................................................................................................................... 75
Figure 7. Spatio-temporal distribution of steers shedding C. difficile at the feedlot ........ 76
Figure 8. PCR ribotypes of C. difficile from beef cattle in the feedlot. ............................ 77
Figure 9. Clostridium difficile in bovine feces collected from the colon of cattle.......... 119
Figure 10. Minimum inhibitory concentrations (MICs) of cattle-derived C. difficile
isolates to six antimicrobial classes of relevance for human C. difficile infections (CDI).
......................................................................................................................................... 120
Figure 11. Magnitude of C. difficile fecal shedding in cattle at harvest. ........................ 121
Figure 12. Distribution of 301 contacted premises with captive white-tailed deer, study
participants, and positive premises to C. difficile and other human enteric pathogens in
Ohio, USA....................................................................................................................... 153
Figure 13. Multiplex PCR and PCR-ribotyping. A) Multiplex PCR for tcdC, tcdE and
cdu-2 within the Pathogenicity Locus (PaLoc). .............................................................. 154
Figure 14. Retrospective spatial analysis of data from a study of C. difficile in calves,
Canada, 2004(Rodriguez-Palacios et al. 2006). Note the distribution of farms with PCR
xix
ribotype 017 (variant without gene for toxin A; tcdA-/tcdB+cdtAB-). PCR ribotype 017
is responsible for severe disease in people in many countries. ....................................... 173
Figure 15. Cross sectional study of fecal prevalence of C. difficile in wild birds in Ohio,
USA, 2007....................................................................................................................... 174
Figure 16. PCR ribotyping of isolates from wild waterfowl (geese or ducks), European
starlings and dairy cattle sampled within the same period and region. Strains subtypes AC. ..................................................................................................................................... 175
Figure 17. PCR ribotyping and telemetry data indicate that C. difficile (Subtype A in this
example) can be isolated from several birds (E. starlings) captured over time from farms
that had network connectivity through repetitive bird movements. ............................... 176
Figure 18. Effect of moist heat in phosphate buffered saline at 63 and 85oC on viability of
fresh (1-week-old) and aged (20-week-old) spores of C. difficile. ................................. 196
Figure 19. Effect of storage on cultivability of aged C. difficile spores during 18 days of
refrigeration at 4ºC. ......................................................................................................... 197
Figure 20. Germination and outgrowth of C. difficile .................................................... 198
Figure 21. Effect of heat shock on immediate cultivability and subsequent germination of
C. difficile spores determined by enumeration of ethanol-resistant cells before and after
7h of incubation in BHI. ................................................................................................. 199
Figure 22. Minutes of heat needed to inhibit a mixture of aged C. difficile spores heated
in three food matrices...................................................................................................... 200
Figure 23. Genetic diversity of representative isolates tested......................................... 201
Figure 24. Cooking favored the selection of C. difficile PCR ribotype 078 (and 027 to a
lesser extent) as time-temperatures increased. ................................................................ 215
Figure 25. Experiment 1 – Individual thermal testing of prototype C. difficile PCR
ribotypes in lean ground beef at the highest reported minimal cooking temperature
recommended for poultry meals in Canada. ................................................................... 216
Figure 26. PCR ribotyping (Bidet et al. 1999) was used to differentiate between the
inoculated, the naturally present, and the heat-survivor C. difficile strains recovered on
blood agar........................................................................................................................ 217
Figure 27. Box 1 - Risk and exposure factors for acquisition of C. difficile spores and
infection .......................................................................................................................... 234
xx
Figure 28. Box 2 - Expended recommendations to reduce the risk of acquiring C. difficile
spores .............................................................................................................................. 235
xxi
Chapter 1 ― PhD Program Project Description and Study Objectives
Abstract
Clostridium difficile has been estimated to be responsible for at least half million
cases of serious enteric disease in humans in the USA every year, which is just a fraction
of all probable cases. Although C. difficile was in the past considered a pathogen of
hospital environments, increasing numbers of cases occur now in the community. This
changing epidemiology has been partly attributed to the emergence and global
dissemination of more virulent strains. Since 2006, we know that such epidemic strains
can be isolated from neonatal dairy calves, piglets and most importantly, from up to 20%
of retail ground meat, which highlights foodborne transmission potential. At the time this
proposal was presented, it was unknown 1) to what extent food animals at harvest were a
source for food contamination, and 2) the ecological reasons for the rapid dissemination
of such virulent strains across distant agricultural regions in North America, and to 3)
what extent moist heat (i.e., routine cooking) affected the risk of ingesting foodborne
spores. Addressing these unknowns will have a profound impact on our understanding of
the ecology of this pathogen which is constantly re-emerging in the community.
1
Introduction
The frequency and severity of Clostridium difficile infections (CDI) have been
steadily increasing in North America since 2000. In Ohio, the number of annual cases
present only in health care centers is over 17,000 individuals, for a state of 11.5 million
inhabitants. Traditionally, CDI has been regarded as a disease of nosocomial origin, but
now it is well known that cases of more severe disease have emerged in the community,
with the inclusion of children and pregnant women who were previously deemed to be at
low risk. This increased severity has been largely attributed to emerging hyper-virulent
multi-drug resistant strains, especially of strains of an emerging genotype, designated
PCR-ribotype 027/North American Pulse-Field type 1. In 2006, I led a study reporting
for the first time the presence of emerging C. difficile strains from dairy calves raising
concerns about zoonotic transmissibility. A year later, I led another study reporting the
identification of emerging genotypes from retail ground beef indicating C. difficile might
induce foodborne infections.
Knowing how epidemic strains reach the food chain and how they disseminate
among food-producing animals will help to identify strategies to mitigate the ultimate
risk of food contamination and therefore transmissibility to humans. Today, it is unknown
1) if cattle and other food-producing animal species are sources of C. difficile at the time
of harvest, 2) to what extent intestinal carriage is related to carcass contamination, 3) if
the apparent dissemination of such strains among regions and farms is due in part to wild
birds, and 4) if cooking effectively inhibits epidemic C. difficile spores. The lack of
2
understanding of these basic aspects of disease ecology prevents us from studying and
implementing preharvest, harvest and postharvest strategies to reduce the risk of food and
environmental contamination.
Hypothesis
Our central hypothesis is that food animals at the time of harvest and wild birds
can be sources of toxigenic C. difficile of current epidemiological relevance for disease in
humans and be effective disseminators.
Specific Aims
The long-term goal of the research program in Dr LeJeune‘s laboratory and mine
as well is to identify and validate strategies to reduce the risk of foodborne
transmissibility of human pathogens from their source in animals or foods. Thus the
main objective of this project was to determine the role of food animals and wild birds
in the epidemiology of C. difficile from an ecological perspective, and to determine to
what extent moist heat is inhibitory when used as recommended.
The ‗Gaps in knowledge‘ we deemed necessary to address during this PhD
program to enhance global food safety against C. difficile included:

Understanding dissemination and reducing the risk of food contamination

Understanding the effect of cooking on spores‘ survival.

Promoting educational programs
3
Goals and Experimental Approach
Goal 1 – Food Animals at harvest and food contamination
To assess C. difficile shedding in food animals prior to harvest, as it is important
to infer the potential for carcass and food contamination. The working hypothesis of this
aim was that the intestinal carriage of C. difficile varies among livestock species, and that
in cattle such carriage represents a risk for carcass contamination at harvest.
The experimental approach:
Study 1. Single-cohort (longitudinal) study of C. difficile in a group of
beef steers from fattening to harvest, (Appl Env, Micro. Published; Chapter 3).
Study 2. Cross-sectional study of C. difficile prevalence in various food
animal species and horses in Ohio during summer. (For submission; Chapter 4).
Study 3. Two-group cross-sectional study of fecal prevalence of C. difficile
at harvest in large-scale processing plants for beef and culled dairy cattle in the
USA. (J of Food Prot. In Press Chapter 5).
Study 4. Prevalence study of C. difficile in retail fresh vegetables
(For submission; Chapter 6).
Goal 2 – Ecology of C. difficile in wildlife and mechanisms of spore dispersion
To understand the ecology of C. difficile and explore mechanisms of spore
dispersal, as it is important to identify critical control points to prevent spore
dissemination if applicable. The working hypothesis of this aim was that wild birds can
4
harbor and disseminate C. difficile strains with molecular virulence factors of relevance
for enteric disease.
The experimental approach:
Study 1. Cross-sectional study of C. difficile prevalence in captive white tailed
deer intended for food production and recreational hunting in designated reserves
in Ohio. (Foodborne Path and Dis, Published; Chapter 7).
Study 2a. Cross-sectional study of C. difficile prevalence in three wild bird
populations to identify common reservoirs, (For submission; Chapter 8).
Study 2b. Experimental infection study where selected wild birds were orally
inoculated with regular or hyper-virulent C. difficile strains to determine their
fecal shedding patterns, (For submission; Chapter 8).
Study 2c. Single-cohort (longitudinal) study of C. difficile from wild birds in a
public park (in rural Ohio) to determine the prevalence dynamics of C. difficile
overtime. (For submission; Chapter 8).
Goal 3 – Thermal resistance of C. difficile spores and spore superdormancy
To quantify the inhibitory effect of various time-temperature combinations and
determine the patterns of survivability and spore germination with recommended
minimum cooking temperatures. This is important because it will determine if current
guidelines are ineffective, and if revised recommendations to target educate susceptible
people are needed
5
The working hypothesis of this aim is that C. difficile survives and cannot be
inhibited at temperatures widely publicized in most international cooking guidelines.
The experimental approach:
Study 1. Experimental thermal studies to determine spore resistance to heat, and
mechanistic studies of spore resistance. (Appl. Env. Micro. Published; Chapter 9).
Study 2. Selection thermal studies to determine if cooking favors the
selection of emerging hypervirulent C. difficile strains (Submitted, Chapter 10).
Goal 4 – Risk of ingestion of C. difficile and food safety education
The final most comprehensive goal of this project was associated with the
production of two documents, one for general medical practitioners who can have access
to communities at risk (elderly, children, etc) and the other prepared for communities at
risk. This is important because revised food safety guidelines are more effective when
they target the populations at risk.
The working theory of this aim is that if susceptible individuals were targeted for
education, the risk of inadvertent ingestion of spores would decrease, and so theoretically
the incidence of CDI in the community.
The approach:
Perspective article. An manuscript written for a medical journal to target general
practitioners. It will inform medical and public health practitioners about the risk
of foodborne infection (as C. difficile can be isolated from foods) and the need to
―prescribe‘ recommendations to improve cooking and hygiene to minimize the
6
exposure to sources of C.difficile spores for communities at risk (Submitted;
Chapter 11).
White Paper. An encyclopedic white paper (literature review with high
readability index) to provide policy makers and average internet users with
information regaring C. difficile biology, infections and food safety research facts.
(For Submission; Chapter 1).
Invited Review. An editor-invited review paper is under preparation. This will
target the community of food and animal scientists
The identification of which food animal species and to what extent wild birds are
active disseminators of emerging C. difficile strains is relevant to further develop
strategies to increase our Nation‘s food safety and health. Results are expected to
contribute to the understanding of the disease epidemiology in humans, and to the
development of intervention strategies to reduce the risk of transmission in the future.
7
Chapter 2 ― Literature Review on Clostridium difficile in animals and retail foods:
Understanding and reducing risks
Written as an encyclopedic White Paper. Structured according to guidelines to achieve
the readability index of average internet users. (For Submission)
Summary
With no doubt, Clostridium difficile is the most important infectious agent
associated with intestinal diseases acquired in health-care centers, but now it is
increasingly linked with more cases in the community worldwide. C. difficile-associated
diseases represent a complex of enteric diseases ranging from transient diarrhea, sporadic
cases of bacteremia (bacteria in the blood), to often fatal cases of pseudo-membranous
colitis (severe inflammation of the colon). Although C. difficile have been traditionally
associated with hospital environments, recent molecular studies indicate that it is no
longer the case. Livestock and retail foods might be involved in the current changing
epidemiology of this international pathogenic bacterium. In 2006, the isolation of C.
difficile strains (responsible for disease in humas) in dairy calves and retail foods ntended
for humans, raised concern for its potential zoonotic (animal-human-animal) and
foodborne (food-human) transmissibility. Though this issue remains unsettled—due to
the absence of conclusive reports linking human disease with exposure to infected
animals or contaminated foods, the fact is that C. difficile can be found in ready-to-eat
8
foods. Thus, it is imperative to advance our understanding of how spores from C. difficile
of relevance for humans reach animals and foods, and/or vice versa. This paper highlights
disease facts to contextualize the relevance of potential health risks and future research
needs.
Introduction - why is this relevant today?
With the potential for inadvertent exposure of animals and humans to infective C.
difficile, there is need to enhance prevention and food safety measures at various levels.
Compared to other enteric pathogens—say the infamous Escherichia coli O157:H7, C.
difficile has been largely unexplored in livestock and food production.
Many different foods and animals can carry C. difficile. But it does not mean we
always get sick after every exposure. There is a chance of getting sick however. But the
magnitude of such chance is largely uncertain as it depends on interactions with other—
often unknown—risk factors. For example we have differences understanding concepts
related to such modifying factors and current research needs will allow citizens to make
informed/responsible decisions regarding prevention of C. difficile.
History and Burden
Major events in the history of C. difficile
The bacterium we now know as Clostridium difficile was first isolated from
healthy infants in 1935 (Bartlett 2008). Then, it was shown that those first isolates were
fatal to laboratory hamsters. However, despite that observation, no attention was directed
9
to study the potential health risks of C. difficile in older people until three decades later
(Bartlett 2008). During those years, it was assumed that C. difficile was a normal
inhabitant of the intestinal tract in children. In 1962, another study documented the
presence of C. difficile in various types of non-intestinal infections of adult people (Smith
and King 1962). That study examined eight C. difficile isolates obtained from human
patients, and found that such isolates were fatal to hamsters as shown in 1935. Despite
this finding, the authors‘ conclusion was that in those human cases “C. difficile was not
pathogenic for man‖. This bacterium continued being considered as a normal part of our
gut flora until it was associated with a life-threatening form of inflammation of the
human colon.
In the 1970s several studies in humans and hamsters finally establishd a
connecting link between C. difficile and a severe form of colitis (colon inflammation) in
people. Called pseudo-membranous colitis, that form of disease had been described since
1893 as a disease induced by antibiotic consumption (Bartlett 2008). Usually, pseudomembranous colitis would ocurr after the use of widely used clindamycin. In subsequent
years, other antibiotics increasingly prescribed to humans, mainly cephalosporins, were
also causing such pseudo-membranous colitis and other forms of C. difficile infections
(CDI) (Bartlett 2008). Currently, it is well accepted that almost every case of pseudomembranous colitis has as a culprit C. difficile. But not always C. difficile results in that
type of colitis. Infections were traditionally seen as sporadic cases, particularly among
people residing in hospitals or health-care centers. However, things changed in the 1990s
10
as highlighted in a medical journal article entitled “C. difficile: A pathogen of the
nineties” (Riley 1998; Pepin et al. 2004).
Compared to the early 1980s, the frequency of C. difficile infections in hospitals
in developed countries doubled during the 1990s. By the end of the 1990s, it was
considered that humans were the most important reservoirs for infection to other humans
(Riley 1998). At the same time, the first studies in domestic animals and the environment
demonstrated that C. difficile could also be found almost ―everywhere‖ including root
vegetables and household pets (al Saif and Brazier 1996). However, the increasing
number of cases in hospitals maintained the attention placed on the role of human-tohuman and to environment-to-human transmission (Kaatz et al. 1988). Infections
originated in the community, where patients acquire the diseases outside hospitals,
received little consideration for disease prevention.
Since the identification as a pathogen, C. difficile was thought to induce disease
via the concurrent production of two major toxins (A and B) inside the gut. Toxin A was
initially considered the initiating virulence factor. Toxin B was then thought as an
accompanying toxin worsening the injury path created by toxin A. Almost always, strains
capable fo causing disease carry both toxins, A+B+, and strains that cannot produce either
toxins (A-B-) are clinically deemed nontoxigenic and nonpathogenic. However, since
1999, new naturally-occuring C. difficile mutant strains with only toxin B (in short, A-B+)
have been identified as the cause of large outbreaks in hospitals worldwide (al-Barrak et
al. 1999; Loo et al. 2005). In 2009, a study showing that toxin B, and not toxin A, was
essential for disease induction in hamsters (Lyras et al. 2009) created controversy.
11
However, subsequent studies have shown that both toxins are important for disease and
can act independently, locally in the gut lumen, and systemically in various organs when
toxins reach the blood stream. In the past, C. difficile A-B+ mutants were uncommon
(often less than 5 percent of cases where C. difficile caused disease). Nowadays, A-B+
strains are becoming predominant in some regions, for instance, Korea (Figure 1) (Kim et
al. 2008; Shin et al. 2008; Shin et al. 2008). In aditiion to toxins A and B, a minority of
strains produce another toxin called binary toxin (CdtA, CdtB), but its role in disease
remains unclear (Geric et al. 2004; Geric et al. 2006; Stare et al. 2007). Some consider
that the production of binary toxins alone cannot explain C. difficile pathogenesis (Geric
et al. 2006), although it may be a contributing virulence factor (Terhes et al. 2004;
Schwan et al. 2009). Together, the three toxins through slightly different mechanisms
produce disruption of the cytoskeleton and cell death.
Since 2005, additional parallel discoveries indicated that unusual pathogenic
strains were associated with more outbreaks and disease severity in various countries.
Compared to isolates from before 2000, current C. difficile isolates from people are more
resistant to antibiotics (Muto et al. 2005). Further, some strains have been shown that can
produce up to 16-to-20 times more toxins A or B compared to regular strains (often
called hyper-virulent strains) (Loo et al. 2005; Warny et al. 2005).
Of public health relevance for communities at risk, identical hyper-virulent high
toxin-producing C. difficile strains associated with severe disease in humans have been
increasingly isolated from food animals and retail foods since 2006 (Rodriguez-Palacios
et al. 2006; Rodriguez-Palacios et al. 2007; Rodriguez-Palacios et al. 2009; Songer et al.
12
2009). Since the publication of those first reports, there is increased awareness about the
potential for animal-to-human (zoonotic) and food-to-human (foodborne) transmission
(Rupnik 2007). Although, there are no proven cases of zoonotic or foodborne-associated
C difficile infections, there are disease facts and research findings that highligh the
potential risk for transmission.
More severity and cases of C. difficile infections internationally
Testing of patients for C. difficile has increased since 2000, and with it, the
number of people diagnosed with C. difficile associated infections and the severity of
such infections have also increased (Dallal et al. 2002). Despite increased awareness of C
difficile infections and potential reporting bias, studies have confirmed that the increasing
prevalence of C. difficile in disease is real, and not simply a reflection of more testing.
For example, in Germany between 2000 and 2007, thousands of stool samples
from people in hospitals and the community tested for C. difficile confirmed a parallel
increase in the number of cases in hospitals and the community (Figure 2) (Borgmann et
al. 2008). At its peak in 2003, that study showed that the percentage of affected
outpatients (community) had increased almost 7 times compared to the year 2000—in
other words, went from one case every 40 outpatients in 2000 to one in every 6 in 2003.
Despite a subsequent decline observed, it was still one in every 10 outpatients by 2007.
The numbers
This increase frequency of CDI has also resulted in a financial burden to
individuals and national health systems. It is estimated that there are over 500,000 cases
of CDI every year in the USA in health care settings alone (Rupnik et al. 2009). This
13
estimate is based on data obtained during a state wide mandatory reporting system
adopted in Ohio during 2006 (Campbell et al. 2009). Few states in USA has adopted
mandatory report since, but that effort was critical to better grasp the magnitude of the
problem and raise awareness among health-care workers. During that reporting, health
care officials also kept track of recurrent cases of CDI. Considering that CDI reoccurrences is currently a growing problem (Pepin et al. 2005), and that CDI is
increasingly reported in the community (Limbago et al. 2009), the quoted national
estimates are likely low.
The cost for medical treatment for every CDI case hospitalized could be in excess
of $3,000 in North America (Dubberke and Wertheimer 2009). Overall, the cost to the
USA finances may represent over 3.2 billion dollars every year (O'Brien et al. 2007), the
amount equivalent to collecting yearly taxes from approximately 300,000 individuals in
USA (Bachman 2008).
With an increase in the number of human infections, there has also been an
increase of social concerns. Today there is more disease severity, with more patients
requiring surgical removal of the inflamed colon (one every ten CDI cases—10 percent),
higher mortality rates, and of relative concern more liability (Pepin et al. 2005; Sailhamer
et al. 2009). The catastrophic effect that such medical consequences have at the
individual level and their families cannot be quantified. Therein, infection control and
surveillance initiatives have been reinforced internationally to reduce the raising
incidence of C difficile infections (Pepin et al. 2004); unfortunately, in the majority of
cases with limited success.
14
Given that C. difficile also infects animals, the disease can have impact on
companion animals and livestock too. Financial loss estimates associated with C. difficile
infections in animals are not available. But in horses and other companion pets where C.
difficile causes enteric disease the costs associated with veterinary medical treatment are
assumed by the owners. In livestock production, no estimates are yet available either,
largely due to absent evidence that C. difficile causes disease and/or growth delays in
production animals. But the farmer will finally carry the financial burden for treatment
and loss of production (Songer 2004; Kiss and Bilkei 2005). Gaining further
understanding about the disease in animals would allow us to implement preventive
strategies to reduce this burden on agriculture and animal‘s owners.
New patterns of disease — In brief
Overall, C. difficile infections (CDI) can be classified based on the time of
acquisition and the onset of clinical signs as associated with health-care settings or with
the community when no attribution can be made to health-care settings. Often due to the
opportunistic nature of C. difficile and the variable number of days that pass between
exposure to a risk factor (for instance, antimicrobials or antacid use) and the recognition
of signs of disease, the distinction between hospital and community acquisition is often
difficult. Hence, definitions often vary from study to study. Health-care-facility onset
refers to cases when signs of disease start after 48 hours of admission, and communityonset applies when the complex definition of health-care onset/acquisition does not
apply.
15
Considering that elderly over 65 year-old are often affected in health-care centers,
two new patterns of C. difficile infections have internationally emerged since 2000:

More severe and frequent cases of CDI among people in health-care
centers (Gastmeier et al. 2009).

More severe and frequent cases of CDI among people in the community,
including those previously thought to be at low risk; specifically pregnant
women, children and young healthy adults (Limbago et al. 2009).
Comparatively, studies have highlighted the remarkable parallel in disease
occurrence between hospital and community-acquired infections (Dubberke et al. 2009).
In hospital settings, the admission of patients with community-acquired CDI represents
sources of new C. difficile strains for other patients. Thus, understanding the factors
associated with C. difficile infections in the community would in theory reduce the
incidence of C. difficile in hospitals (Figure 3).
Clostridium difficile has also changed as a pathogen
There is conclusive evidence that C. difficile has also genetically and
physiologically changed over the past decades. New C. difficile strains are often mutants,
produce more toxins and are more resistant to antimicrobials that would have otherwise
been effective to them before 2000. Recent microarray-based data indicate that food
animals might have been the original sources of epidemic strains of C. difficile,
particularly newly emerging C. difficile (Stabler et al. 2006; Goorhuis et al. 2008), but no
conclusive studies are yet available to determine if animal shedding or food
16
contamination are associated with changing patterns of disease in humans. Has the
pathogen moved from humans to animals and foods or viceversa? Given the spore
forming nature of C. difficile, it is possible that inter-species transmission occurs from
widespread environmental sources and that some level of host adaptation (Janvilisri et al.
2009) and clonality has also ensued in paralell over millions of years (He et al. ; Stabler
et al. 2006). Horizontal gene transfer and homologous recombinations (i.e., the exchange
of pieces of DNA among C. difficile strains) are very frequent processes in C. difficile
(He et al.).
Not everyone knows where or why they became sick with C. difficile
Only a fraction of all individuals affected with C. difficile can be linked to known
risk factors for disease. Some individuals are more likely than others to become ill with
C. difficile. Those include individuals that:

are over 65 years old

have direct or indirect contact with hospital/health-care centers, or

have consumed antibiotics or antacids.(Kuijper and van Dissel 2008)

among racial ethnicities, whites appear to be at higher risk (Redelings MD
et al. 2007)
However, a large proportion of (approximately one third) CDI patients do not
have any of the above mentioned risk factors, and the source of their infection cannot
often be traced (Wilcox et al. 2008). Less clear are the sources of C difficile spores for
people in the community that have no exposure to hospital settings. Moleculart studies
17
identifying pathogenic strains in rivers, animal manure and retail foods indicate those
could be source of spores which are needed for the induction of disease (RodriguezPalacios et al. 2006; Rodriguez-Palacios et al. 2007; Rodriguez-Palacios et al. 2009;
Songer et al. 2009; Von Abercron et al. 2009; Weese et al. 2009; Weese 2010; Zidaric et
al. 2010). However, no direct evidence is available confirming zoonotic or foodbore
transmission.
Pathogen and facilitator factors
Strains today are more resistant to antibiotics than before 2000
In addition to high amounts of toxin production in some genetically mutated
strains, C. difficile strains isolated today are more frequently resistant to antimicrobials
than before 2000. Today, the increase resistance to fluroquiniolones, metronidazole and
vancomycin has become an emerging global health issue (Spigaglia et al. 2008).
Although, the use of antimicrobial in animal production is often blamed for the increase
resistance among pathogens derived from infected humans, the same assumption is
difficult to justify in the case of emerging C. difficile. Fluoroquinolones, metronidazole
and vancomycin are prohibited drugs for use in food animals in most developed
countries, where C. difficile is a problem. More recently, high MIC values (intermediate
resistance) were identified in a C. difficile isolate derived from cattle at the time of
harvest (Rodriguez-Palacios et al. 2011). Linezolid is a new antimicrobial drug
(prototype of new drug class) that has not been reportedly used in the veterinary
literature. Horizontal gene transfer of antimicrobial resistance from naturally resistant
18
microorganisms and spontaneous mutations (Holzel et al. 2010) with natural selection are
possible options.
Clostridium difficile is an environmental-resistant microorganism
Clostridium difficile is a bacterium that grows in the absence of oxygen
(anaerobic environments), including the intestinal tract of animals and humans. When
ideal growing conditions turn into adverse scenarios the bacteria produce spores―this is,
cells that are dormant and very resistant to environmental temperature changes and the
presence of sanitizers (Lawley et al. 2009). When the conditions become more
favourable, these dormant spores awake and germinate to produce a new offspring of
actively dividing bacterial cells. These growing cells have the capacity to produce
virulent toxins and escape the action of antimicrobials to become opportunistic pathogens
of the gut.
In susceptible animals and people, antimicrobials (among other factors) apart
from facilitating the induction of enteric disease, they also stimulate the rapid production
of spores by growing C. difficile in the intestinal tract which favours the survival of
antimicrobial-resistant strains (Baverud 2004; Rupnik et al. 2009). In the environment,
spores from infected animals can survive for months or even years in feces or soil
(Baverud et al. 2003) and could be dispersed via air, waters and animals, at least locally.
There, spores can remain dormant until they are ingested by a new host.
Fate of Clostridium difficile spores following ingestion
Like other enteric pathogens, activation of C. difficile starts with inadvertent
ingestion of spores. Ingested spores may have two fates, remain dormant and be expelled
19
in the feces or wake up, produce offspring, and induce intestinal disease. The latter only
results in disease if the immune status of the individual and the integrity of his/her
microbial gut flora are impaired. However, since the mid 2000s, more cases of disease
seem to be occurring in otherwise healthy individuals (Rupnik et al. 2009).
Ingested spores or insulated delicate vegetative cells can survive gastric acids and
germinate and/or proliferate in the gut owing to its adequate nutritional conditions,
optimal host temperature and low oxygen. Under ideal laboratory conditions, a spore can
awake from dormancy to produce active C. difficile cells within one hour. But how long
it takes inside the gut of humans or animals is largely unknown.
The likelihood of becoming sick following ingestion of spores is uncertain to
date. Due to ethical limitations and the opportunistic multi-factorial nature of CDI it is
difficult to determine how many spores are needed to cause disease (i.e., the infective
dose) in people. Experimental infection studies in neonatal calves (Rodriguez-Palacios et
al. 2007) and foals (Arroyo et al. 2004) have shown that spore ingestion alone is not
sufficient to induce disease. Predisposing factors are almost always needed. Treatment
with antimicrobials is one of these factors.
In farm animals, C. difficile may persist in the intestinal tract of animals (Norman
et al. 2009) or in the soil (Baverud et al. 2003) at the farm for months if not years. One
study, conducted in a single swine (pig) operation indicates that at least a strain of human
relevance can be prevalent among animals through most of the production cycle,
although it becomes less common close to the time of harvest compared to newborn
piglets (Norman et al. 2009). However, the ability of C. difficile to complete a life cycle
20
outside animated hosts has not been elucidated. The ecological component of C. difficile
in the patterns of human and animal disease remains largely unexplored. The relevance of
such understanding lies in the fact that susceptible individuals often acquire infective
spores from environments formerly inhabited by shedding individuals (Figure 4).
If you get sick once, you can get sick twice, or more
Of additional concern is that following recovery from a C. difficile infection, reinfections in the same individual are occurring with increasing frequency. Although it is
happening more often, not everyone suffers C. difficile re-infections. There are however
overly sensitive individuals that can be seriously affected; the reasons of such
susceptibility is currently under intense investigation. Low antibody titters effective
against the C. difficile toxins (Wilcox 2004), and disrupted intestinal flora due to
antimicrobials (Rupnik et al. 2009) are among the risk factors that may enhance
susceptibility to re-infections.
Reported rates of re-infections vary between 15 and 30 percent, with recurrences
being more common in elderly patients (over 65 years old)(Kyne 2010). Further, the
more recurrences a person has, the more likely he/she is to have a recurrence again. In
most cases, subsequent infections are caused by a new C. difficile strain that is
molecularly different from that of the first CDI episode.
Prevention measures in re-infected individuals are particularly important
considering that most CDI that require medical attention are treated with antibiotics that
disrupt the gut flora beyond the time of recovery from the CDI symptoms.(Rupnik et al.
2009) Unfortunately, preventive medicine reports are not yet available. Understanding
21
the interaction dynamics between the antimicrobials, their impact on the gut flora, and the
windows of opportunity for C. difficile to proliferate and re-induce disease might reduce
the risk of exposure to infective spores during periods of higher risk for CDI.
Differentiating C.difficile from various potential origins
Molecular fingerprinting (―bar-coding‖) of C. difficile
To differentiate C. difficile strains (or lineage) and thus to study their possible
origin, scientists use molecular methods to create genetic fingerprints. These fingerprints
are largely unique and identical amomg highly related isolates, but different among
different distanly related strains. Genetic fingerprints allow tracking routes of
contamination and study the potential for pathogen transmission, and the vehicles of
dissemination that could be involved. To enhance accuracy, scientists use a variety of
molecular tests to classify strains of C. difficile, often in combination. The most common
methods for studying C. difficile isolates include:

PCR ribotyping

Pulse field gel electrophoresis (PFGE)

Restricted fragment length polymorphism (widely known as toxinotyping)

Multilocus variable tandem repeat analysis (MLVA)
The most common typing methods cited in current studies are PCR-ribotyping
(mostly used in Europe), followed by Toxinotyping (international), and Pulse Field Gel
Electrophoresis (PFGE) and Restriction Endonuclease typing (REA) mostly used in the
22
USA. A search of studies about C. difficile published in PubMed search in 2009 showed
that PCR ribotyping was more used (48 studies) compared to PFGE and REA (7 and 6
studies, respectively) (Rodriguez-Palacios, A, unpublished data). PCR ribotyping is
widely used because it allows for the contextualization of results at a global scale.
Toxinotyping is widely used because it helps group the strains by their virulence profile
(i.e., toxins A and B gene molecular characteristics).
There are several C. difficile strains (lineages) but often few are the bad ones
Molecular typing methods help understand how strains are genetically related and
to source-track and infer potential routes of transmissibility locally or across large
geographical distances. However, each technique has pros and cons that are considered
depending on the purpose of the typing (Tanner et al. 2010).
Depending on the typing method used, the number of C. difficile strain types can
vary largely. According to an internationally cited Reference Laboratory in the United
Kingdom—where the strain collection begun in the early 1990s (O'Neill et al. 1996),
there are now over 300 C. difficile strain types (i.e., molecularly distinct with a
fingerprinting method)(Balassiano et al. 2009).
To date, many studies are using the UK PCR-ribotyping nomenclature for
international comparisons (Rodriguez-Palacios et al. 2006; Balassiano et al. 2009). The
nomenclature designation for that system is numeric and started with PCR-ribotype 001
in the 1980s and continued in order as new strains were identified. The use of such
international naming system has allowed scientist to compare isolates across landscapes
and over time, studying the dynamics and emergence of epidemic C. difficile strains
23
(Figure 4). Today, PCR ribotypes 001, 014, 017, 027, 077 and 078 are mong the most
common strains in certain regions. Of public health relevance, all these strains have been
isolated from animals and foods (Rodriguez-Palacios et al. 2006; Rodriguez-Palacios et
al. 2007).
About the Disease
How does C. difficile ingestion result in intestinal disease?
Whether an individual who is exposed to C. difficile spores becomes sick depends
upon many factors. First, it depends on the age of the host that is exposed and their
immune status (elderly patients with cancer, or debilitating medical conditions are at
higher risk), the concurrent use of medications that are disruptive of the intestinal flora
(i.e., antimicrobials), and the ability of the spores to grow and produce damaging toxins
in the gut. Ingested C. difficile spores are not capable of inducing disease, they need to
―wake up‖ from dormancy (germinate) and actively produce rapidly multiplying daughter
cells which produce the toxins responsible for intestinal lesions and disease. With few
exceptions, C. difficile strains that induce disease need to be able to produce at least one
of three currently recognized virulent toxins—A, B, and CDT. Strains that cannot
produce toxins do not cause disease, and rather may play a protective role (Songer et al.
2007).
Toxins A and B may act independently in vitro; however, a recent study showed
than toxin B alone is essential to induce disease and mortality in experimental animal
models. Although most strains carry both toxins A and B, more cases of strains producing
24
only toxin B have been associated with disease worldwide since the first outbreak of
serious disease in hospitalized people was identified in Canada in 1999 (Alfa et al. 2000).
The role of the third toxin, referred as CDT, is under investigation.
Functionally, the toxins induce cell roundening and death, which creates
microscopic gaps or craters on the lining that cover the intestinal tract. Disruption of this
protective layer is responsible for the clinical signs observed in humans and animals
when disease occurs. Inflammation in the colon, with destruction of the gut barrier,
renders hosts susceptible to other opportunistic bacteria and complications associated
with initial intestinal damage. Bacterial invasion of deeper tissues and blood originating
from the intestine can occasionally occur in some individuals (Smith and King 1962;
Gérard et al. 1989 ; Thomas et al. 2011). This complication however seem to occur in the
absence of intestinal signs of disease and in immunocompetent patients (Cid et al. 1998;
Libby and Bearman 2009).
While the effect of toxigenic strains has been well studied, the effect of ingesting
non-toxigenic strains (i.e., strains unable to produce toxins due to lack of functional toxin
genes; A-B-CDT-) on intestinal health is largely unclear. Preliminary studies in animals
and humans indicate however that exposure to non-toxigenic strains might have a
protective role against disease by preventing the proliferation of toxigenic strains if
subsequently ingested. Ongoing clinical trials are underway to test related hypothesis.
New C. difficile strains can produce much more toxins - why?
While humans required years to reach reproductive age, bacteria—including C.
difficile—can have a new offspring every 20-60 minutes (Goulding et al. 2009). With
rapid generation turnover, genetic mistakes (mutations) can result during bacterial
25
replication. Often those genetic mistakes can be fatal to the new daughter cells, but at
times, mutations may boost the chances for a bacterial cell to thrive in challenging
environments facilitating adaptation.
The production of toxins A and B is controlled by protein TcdC, which is a
product of the regulatory gene tcdC. Natural mutations in this gene have resulted in C.
difficile strains with increased toxin production. In 2005, a study of the common strain
(epidemic PCR ribotype 027/NAP1) from human origin showed that genetic machinery
that control toxin production in that strain was disrupted increasing the amount of toxin
production. Strains since about the early 2000s are likely to produce more toxins
compared to regular non-epidemic strains (Warny et al. 2005). Although some studies
indicate that such strains have increased the incidence of disease and associated mortality
outside health care centers (Kuijper and van Dissel 2008), few studies have reported no
direct associations with disease severity in hospitals (Morgan et al. 2008). Irrespecive of
the finding, all studies demonstrate that this and other emerging mutated strains are on
the rise as responsible for disease in humans.
In animals, recent studies have shown this and other types of mutations in the
same toxin regulatory mechanism, including strains designated PCR ribotype
078/Toxinotype V. This strain increasingly isolated in humans, is a major predominant
strain in livestock, namely young cattle and piglets.
C. difficile exposure vs. risk factors for CDI — Two different things
One thing is identifying the sources where spores can be ingested from, but the
other is the settings where risk factors are likely to make subjects prone to developing
enteric disease. Risk factors such as antibiotic or antacid ingestion can occur at home and
26
in health care centers, where spores may or may not be present in the environment. In
humans and animals, CDI is a diagnosis primarily based on the combined presence of
clinical signs of enteric disease and the identification of toxins in the stools
(feces).(Russmann et al. 2007) In practice, the simultaneous association of both—the risk
factor and the source of spores—largely depend on diagnostic criteria used to confirm
CDI and the clarity of definitions used to call something a risk.
Where can we get both the spores and the risk for CDI? — Onset vs. Acquisition
To date most studies have divided all forms of CDI in two majors groups, but the
epidemiology of it is more complicated than that. One thing is the onset (when clinical
signs start) and the other is acquisition (where the spores were ingested). To enhance our
understanding in this area, scientists have long collaborated to achieve global scientific
standards. Now, there is international consensus definitions for community associated
cases:(Kuijper and van Dissel 2008)
“The European Centre for Disease Prevention and Control and the US
Centers for Disease Control and Prevention have proposed similar definitions for
community-acquired C. difficile infection:
[A] the onset of symptoms occurred while the patient was outside a health
care facility and the patient had not been discharged from a health care facility within 12
weeks before symptom onset (community onset, community acquired);
[B] the onset of symptoms occurred within 48 hours after admission to a health
care facility and the patient had no prior stay in a health care facility within the 12 weeks
before symptom onset (health-care-facility onset, community acquired”
27
“Community-acquired C. difficile infection is almost certainly under diagnosed”
(Kuijper and van Dissel 2008)
In the past, CDI were a disease of mostly elderly over 65 years of age. Now things
appear to be changing, affecting younger populations. The unanswered question is where
these affected individuals are picking up the virulent C. difficile strains. Among many
possible alternatives, the following excerpt further illustrates the absence of traditionally
known risk factors among people that get sick in the community:
“Of 241 patients, 36 percent had no history of antibiotic use within 3 months
before symptom onset, and 25percent had no underlying medical condition or recent
hospital admission and, moreover, were younger than 45. ...”
About Animals, the Environment and Foods
In animals, the first studies reporting the isolation of C. difficile from fecal
samples companion animals and pigs were published in the early 1980s. However, it was
early in the 2000s when association with enteric disease started to be confirmed in
animals. Molecular (i.e., genetic) fingerprinting studies comparing animal and human
isolates began to be reported in 2005.(Arroyo et al. 2005) These started to indicate the
potential for identical strains to share human and animal habitats. With it the risk of
cross-transmission began to be considered.
Clostridium difficile in the environment and animals
28
As an environmental microorganism, it is possible to think that an obvious source
of infective spores for people is unnoticed contaminated environments. However, such
observations been documented since the 1980s by the same groups of scientists who
started the UK collection of C. difficile that is now reference landmark for molecular
epidemiology studies internationally (Borriello et al. 1983).
Clostridium difficile was then (1982) isolated by Borriello et al. from rivers
draining a city, and nearby ocean waters, soils and a few root vegetables. Concurrently
studies in companion animals showed that C. difficile was present in dogs, especially
puppies but no linking studies were conclusive to establish a strong connection. Other
species assessed included horses, in which C. difficile has been now recognized as an
established cause of serious enteric diseases in adult horses (Weese et al. 2006). In young
foals the situation is not completely clear, but antimicrobials are deemed a major
predisposing factor (Arroyo et al. 2004). Although the possibility of animals being
reservoirs of disease was suggested by the researchers at that time, no subsequent studies
were conducted to validate such observation.
In 2005, a study from Staempfli and Weese‘s group, at the University of Guelph,
Canada, reported the first molecular studies comparing human and strains derived from
companion animals indicated that no strains were common (Arroyo et al. 2005). However
in 2006, during studies conducted with calves in an attempt to determine the role of C.
difficile in enteric disease in young calves the majority isolates matched now known as
epidemic human PCR genotypes, namely PCR ribotypes 078, 027, 014 and 017
(Rodriguez-Palacios et al. 2006).
29
Companion Animals: Household pets - Dogs, cats and other creatures
Dogs and cats can carry toxigenic strains of C. difficile in their feces. This was
first identified in the 1980s (Borriello et al. 1983). But in the majority of cases, C.
difficile does not cause disease (Weese et al. 2010). If risk factors, that parallel that of
humans (antimicrobials) animals may have diarrhea (Weese et al. 2001). No
pseudomembranous colitis or bacteraemia have being documented in dogs.
Screening studies indicated recently that up to 10 percent of household pets may
carry C. difficile in the community representing a risk for owners (Weese et al. 2010).
Although no direct transmissibility from pets has been documented to humans, reports of
the virulent strains of C. difficile (including PCR ribotype 027) in visitation dogs indicate
that hospital visiting animals might carry strains within and outside health-care facilities
(Lefebvre et al. 2006). Of epidemiological relevance, pets owned by an inmmunecompromissed person are more likely to be colonized by C. difficile (Weese et al. 2010).
In veterinary hospitals, outbreaks of severe diarrhea associated C. difficile have
been reported in small animal clinics (Weese and Armstrong 2003), therefore pets that
visit animal hospitals may be inadvertent carriers of C. difficile spores. In households, a
report in 2009 found C. difficile in dogs of immune-compromised people and in the
environment of a few households, but molecular typing (fingerprinting) of the strains did
not support that dogs were the source for household contamination (Weese et al. 2010).
Those dogs, however, had strains identical to the ones that had affected people in the
region of the study a year earlier. No studies have assessed the potential of dogs and cats
to be vehicles of C. difficile strains in food or livestock production systems.
30
Companion Animals: Horses - Pleasure and working animals
Clostridium difficile has been studied in horses since the mid 1980s (Ehrich et al.
1984). Today, it is known that horses can carry C. difficile, but the proportion of animal
shedding the pathogen varies across studies as it depends on culture methods and the
animals‘ age. Adult horses are less likely to carry the bacterium compared to neonatal
foals. Overall, between 2 and 30 percent of horses could be found carrying spores at any
given time without showing signs of disease (Baverud et al. 2003). But like other species,
horses also can develop diarrhea and other forms of serious disease. In equine hospitals,
strict isolation and infection control measures are widely recommended to avoid
outbreaks.(Baverud 2004). As in human hospitals, antimicrobials increase the risk of
horses being affected with CDI (Weese et al. 2006).
Currently, horses provide pleasure for some, but historically, horses have been
companions for work purposes since the time of wars. And although they have been large
displaced by modern mechanized equipment, some cultures still rely on horse power to
work on their agricultural lands to produce foods. Indeed, regulated production and
slaughter of horse meat sustain at least partially local economies in some regions (USDA
1997). Since horse manure can contain living C. difficile spores for years (Baverud et al.
2003) and it has been used as fertilizer for decades (Randall 2003), the potential for crop
contamination (Pell 1997) is an additional alternative to currently documented food
contamination with C. difficile. In this regard, no studies are yet available in horses to
further address the potential risk for zoonotic and foodborne transmissibility.
Food Animals: Pigs
31
As in horses, the first studies reporting the presence of C. difficile in pigs date to
the early 1980s (Jones and Hunter 1983). Since then, over 50 published studies have
served to now recognize C. difficile as an enteric pathogen in this domesticated species.
And among pigs, the major clinical relevance resides in piglets (Post et al. 2002).
Mortality and morbidity rates are largely uncertain, but some estimates indicate that up
to100 percent of litters and individual piglets can be affected in infected farrowing
facilities (Songer 2004). In non-fatal cases, weaning weights of diseased pigs can be 10
percent below the expected average weight (Songer 2004). In older animals, C. difficile
has been associated with increased mortality in sows that received antimicrobial
treatment against a post-parturient syndrome characterized by lack of milk production,
and uterine inflammation (Kiss and Bilkei 2005). Sows that received antimicrobial
treatment were more likely to die if they concurrently had C. difficile compared to sows
that were not treated. Apart from this report in an outdoor production system in Croatia
(Kiss and Bilkei 2005), no other studies have described the syndrome in swine.
On the other hand, C. difficile can also be isolated from up to 4 percent of healthy
pigs close to the harvest time, which has further raised concerns about potential for
carcass contamination.(Norman et al. 2009) As with other species, in the harvest stages,
C. difficile infections goes unnoticed as no obvious signs of disease differentiate shedders
from non-shedders.
Molecularly, swine-derived C. difficile isolates have called the largest attention
from public health officers as among all possible PCR ribotypes, PCR ribotype 078—the
increasingly documented emerging human ribotype in the community—have accounted
32
for up to 80 percent of all swine isolates in most studies conducted involving pigs in
North America and Europe (Keel et al. 2007; Debast et al. 2009; Songer 2009).
Regarding the type of production system, a study in Europe found no difference between
organic and conventional swine operations (Keessen et al. 2011).
Food Animals: Cattle
Despite all efforts to document the presence of C. difficile in animals since the
1980s, little attention has been directed to study ruminants and bovine animals in
particular. The first study that described C. difficile in calves indicated the presence of C.
difficile in veal calves in 2002 (Porter MC et al. 2002). However, the first peer-reviewed
study assessing the impact of C. difficile in the bovine industry aimed to determine—in
2004—the role of the pathogen as a cause of diarrhea in young calves (RodriguezPalacios et al. 2006). In that study, Rodriguez-Palacios and his colleagues conducted a
case–control study of calves less than 28 days of age—from 102 dairy farms in Canada—
and showed that significantly more diarrheic calves than non-diarrheic controls were
positive to C. difficile toxins, suggesting its association with disease (Rodriguez-Palacios
et al. 2006). However, in further experimental studies the same group could not induce
disease when calves fed colostrums were given orally high numbers of C. difficile
(Rodriguez-Palacios et al. 2007). Further, C. difficile spores were more common in
healthy calves, clearly showing the potential for inadvertent dissemination. Subsequent
studies conducted in calf ranches conducted by others indicated that the lesions observed
are compatible with C. difficile toxin action (Hammitt et al. 2008). Other unpublished
33
studies indicate that veal calves can increase the rate of C. difficile shedding as they are
treated with antibiotics upon entry to finishing operations (Weese 2010).
In older cattle, C. difficile shedding has been observed to decrease overtime
during the finishing period. However, C. difficile can be found in healthy feedlot steers
and culled dairy cattle at the time of harvest highlighting the risk for carcass
contamination.
Regardless of its association with enteric disease, C. difficile isolates derived from
cattle were the first to call the attention for the potential for foodborne transmissibility
involving current epidemic human strains.(Rodriguez-Palacios et al. 2006) Calf isolates
that were identified in 2006 include strains of PCR ribotypes that occur in humans
worldwide and that have been later found in retail foods, particularly PCR ribotypes 017,
027, 077, 014 and 078 (Rodriguez-Palacios et al. 2006; Keel et al. 2007; Hammitt et al.
2008; Rodriguez-Palacios et al. 2009). Together, the molecular characteristics of bovine
and swine C. difficile isolates represent an epidemiological link with human disease that
is under intense research. However, much more need to be done to enhance prevention.
This area of science is relatively new.
Food Animals: Poultry
This is probably the livestock species that has been studied the least. But perhaps
the study that highlight the potential relevance of this species are from Africa (Simango
2006; Simango and Mwakurudza 2008) and Slovenia (Zidaric et al. 2008). In Zimbawe,
Simango and colleagues have showed that up to 30 percent of free range chickens carried
toxigenic C. difficile strains and antimicrobial resistance patterns of relevance for humans
34
(Simango 2006; Simango and Mwakurudza 2008). Although the presence of C. difficile
in cultures where marketing free range living animals is well rooted, results from a study
in confined poultry operations in Slovenia indicate that in confined farming environments
the percentage of birds colonized could be higher. In that study, over 60 percent of
animals carried C. difficile, but it was less frequent as animals approached harvest time.
In a study conducted in operations in intensive raring facilities in Ohio, USA, the
percentage of contamination approached zero (Rodriguez-Palacios, et al. 2011). In a
published report from Austria, 5 percent of poultry tested had C. difficile (Indra et al.
2009). Clearly much more work needs to be done in this field also.
Waters and the environment
In general, spore-forming bacteria including clostridia can be considered long
lasting microorganisms inhabiting the environment. With very few exceptions, C. difficile
produce spores that can survive for months in the environment (Baverud et al. 2003), but
not many publications are available in this regard. Perhaps the most relevant of them is a
British publication that in 1996 described the presence of toxigenic C. difficile in
different environmental samples (al Saif and Brazier 1996). In that study, scientists found
C. difficile in soils, well and recreational waters, and veterinary clinics and in households;
they even suggested that those places could be potential sources of infection (al Saif and
Brazier 1996). Another relevant study in Africa showed that soils in rural areas of
Zimbawe had toxigenic spores (Simango 2006). However, despite these publications,
little attention has been placed in this very relevant aspect of the disease epidemiology—
this is the environment as a source of infectious spores but also its mediating effect in
35
spore dissemination. C. difficile spores have been shown to be able to disseminate via air
in indoor environments (Roberts et al. 2008). At the farm level, this area of research
remains largely unexplored. But sewage should be expected to be a way how farming
may be impacting the environment.
C. difficile in Foods
Great efforts have been accomplished by various governments around the world
to monitor the patterns of antimicrobial resistance in bacterial isolates from foods which
cause disease in humans. Although countries in North America and Europe are constantly
monitoring trends, such surveillance programs are important initiatives (Jones and
Masterton 2001) that have assisted in the study of food products as carriers of C. difficile.
Before summarizing this interesting growing body of literature, it is important to
mention that as of August 16, 2010 there were no reports confirming that people have
gotten C. difficile infections via consumption of contaminated foods. Also it is important
to highlight that at least three reports published in the 1980s warning about the potential
foodborne (Gurian et al. 1982; Borriello et al. 1983) and zoonotic (Borriello et al. 1983)
transmissibility of C. difficile remain largely overlooked until the mid 2000s.(Rupnik
2007) Now is the time to further enhance our knowledge in this regard.
Raw meat products
Raw ground beef and pork were the first products to be found contaminated with
C. difficile, and as usual with major discoveries it was an incidental finding. In 1994,
Broda and her colleagues studying which microorganisms caused gas blown-packed
spoilage in meats (indeed they were ready-to-eat products) found C. difficile, which did
36
not produce gas. So the relevance of such findings seemed as disregarded by food
scientists in the field that could have read the article. The next study, unaware of Broda‘s
findings—and therefore truly unbiased—was conducted by veterinarians in Raw meat
products for dogs also found C. difficile but little was thought about the validity of that
observation in human health by others in veterinary medicine or related fields that might
have also read the study in both human and veterinary medicine-in retrospect- also
overlooked that observation.
In 2006, Rodriguez-Palacios and his colleagues, for the first time, found identical
strains of C. difficile in ground meats in Canada that matched strains of relevance for
humans locally but also in the UK. Subsequent studies have confirmed that this pathogen
can be found elsewhere. Now, scientific reports describe toxigenic C. difficile in meats in
several countries. Although, the percentages of meat packages that have showed
contamination with C. difficile have ranged from 3 to 42 percent, the overall expected
real prevalence of C. difficile contamination under natural conditions might be around 6
percent. Poultry has been the type of meats least studied; at least one published study
found no C. difficile in retail poultry,(Indra et al. 2009) but a recent review indicate that
poultry meats can also carry toxigenic strains at about 7% frequency (Weese 2010)
Ready-to-eat foods
In modern times when life style allows us a few spare minutes to buy, cook and
eat our own foods, convenient ready-to-eat products are gaining market share. And with
it food scientists have been accompanying the change with processing methods that seek
to eliminate most unwanted microorganisms, while marinating the flavoury nutritional
37
value. Among this group, processed meats and ready to eat fruits and vegetables are
common ingredients recommended for healthy lives. However, like other infamous
bacteria- including E. coli O157:H7, C. difficile has also been found in these products.
But this is not something new; we just missed its relevance in the past.
In part due to lack of molecular studies which have now help to evaluate the
potential risk by showing that strains derived from ready to eat salads in Scotland were
identical to the most common PCR ribotypes affecting people (Bakri et al. 2009). In the
past; we overlooked initial reports. C. difficile was first isolated in the UK in 1996, from
root vegetables (al Saif and Brazier 1996). Their finding allowed scientists then to warn
about the potential food safety concern might have been a potential problem. However,
only recently studies have confirmed that between 1and 6 percent of vegetables can be
sources of C. difficile, including PCR ribotype 078 (Metcalf et al. 2010).
Potential for control of C. difficile in foods and food animals
Because several aspects of ecology and epidemiology are still unknown regarding
the ability of this pathogen to survive in different environments and to effectively
amplify, it is necessary to determine survival and persistence factors at preharvest and
postharvest level. The main objective of the present thesis was to determine factors
associated with fecal shedding dynamics, seasonality and prevalence in food producing
animals and wildlife, and to determine factors of thermal resistance that would be
relevant for the enhancement of food safety through cooking and education. This is the
particular objective of this White Paper. It will be modified and adjusted for free online
38
publication. Appropriate quantitative data is also needed to assist in the implementation
of risk assessment programs, food safety policies and recommendations, and to propose
intervention strategies for the future.
39
Annual prevalence rate of toxigenic C. difficile in Korea
100
% of all isolates tested
93.3
80
73.8
64.9
60
53.9
50.3
46.9
A+B+
40
35.2
20
6.7
5
5.4
2001
2002
A–B+
27.2
13.2
0
2000
2003
2004
2005
Figure 1. Increase of toxigenic C. difficile strains A-B+ in a multi-center study in
hospitals of Korea compared to A+B+ strains, 2000-2005 (Shin et al. 2008a).
40
Study of C. difficile toxin testing, human patients, Southe rn Germany, 2000-2007
Number of C. difficilepositive patients
Outpatients (“Community“)
6000
Inpatients (“Hospitals“)
5000
4000
3000
2000
1000
900
Outpatients
750
Inpatients
Percentage of patients positive to C. difficile
20
% of positive samples
7000
Number of positive sampels
600
450
300
150
2006
All patients
15
Inpatients
Outpatients
10
5
0
2007
2005
2004
2003
2002
2006
2007
2005
2004
2003
2002
2001
2000
0
2001
0
2000
Number of samples tested/year
Number of patients
examined
2000
2001
2002
2003
2004
2005
2006
2007
Year
Figure 2. Study testing C. difficile toxins in fecal samples from patients visiting 40
hospitals and over 2,000 physicians in southern Germany.
Despite the increased number of samples submitted for C. difficile testing, these
illustration depicts the rising percentage of people affected by this pathogen over the
years, in both hospitals and the community. Similar trends occur in other regions of
Europe and North America. (Archibald et al. 2004; Ricciardi et al. 2007; Kuijper et al.
2008; Limbago et al. 2009) (Reproduced with permission from Borgmann et
al.(Borgmann et al. 2008) Copyright Eurosurveillance, 2008)
41
Possible routes of transmission of C. difficile spores
IN HEALTH-CARE FACILITIES
IN THE COMMUNITY
Health-care environment
Public or household
environments?
Patient-to-patient
Health-care personnel
Therapy animals or food?
Person-to-person?
Animals or food?
Figure 3. Possible routes of transmission of C. difficile spores.
Representation of confirmed and unconfirmed routes of transmission of C. difficile spores
to susceptible individuals. Elucidating new routes will help with prevention.
42
Other types (n=1,236 )
Type 002 (n=125)
Type 014 (n=358)
Number of isolates
300
Type 001 (n=351)
Type 078 (n=301)
Type 027 (n=417)
250
200
150
100
50
0
2005
2006
2007
2008
2009
Quarter of the year
Figure 4. Clostridium difficile isolates from human patients that were molecularly
characterized with an international PCR ribotyping method in the Netherlands.
Note the increased number of samples obtained and tested since 2005. Most importantly
note that only 5—of all possible 180 PCR ribotypes known—account for almost half of
all human cases. Those PCR ribotypes are very common among diseased people and have
been isolated from foods and food animals (see Table 1). This study represents the first
report where the emerging PCR ribotype 027 started to decrease; this highlights the
changing epidemiology of this pathogen. (Reproduced with permission, Copyright
Eurosurveillance, 2009) (Hensgens et al. 2009)
43
Table 1. Clostridium difficile types with increasing relevance to humans and animals
PCR/PFGE
Toxinotype
Toxins
Toxin
Regulator
International
Foods/animals
027/NAP1
III
A B CDT
Mutated-18bp
NA, Europe
Yes/yes
078/NAP7-8
V
A B CDT
Mutated-39bp
NA, Europe
Yes/yes
001/NAP
0
A B CDT
Normal
Europe
Yes/unknown
014/NAP
A B CDT
Normal
NA, Europe
Yes/yes
077/NAP
AB
Normal
NA, Europe
Yes/yes
017/NAP
B CDT
Normal
NA, Europe
Yes/yes
Other strains of PFGE-NAP1 differ from PCR-type 027 but have identical virulent genes.
PCR, PCR ribotype; PFGE, pulse field gel electrophoresis nomenclature in North
America; AB and CDT, names of toxins; NA, North America.
44
References
al-Barrak, A., J. Embil, et al. (1999). "An outbreak of toxin A negative, toxin B positive
Clostridium difficile-associated diarrhea in a Canadian tertiary-care hospital." Can
Commun Dis Rep 25(7): 65-9.
al Saif, N. and J. S. Brazier (1996). "The distribution of Clostridium difficile in the
environment of South Wales." J Med Microbiol 45(2): 133-7.
Alfa, M. J., A. Kabani, et al. (2000). "Characterization of a Toxin A-Negative, Toxin BPositive Strain of Clostridium difficile Responsible for a Nosocomial Outbreak of
Clostridium difficile-Associated Diarrhea." J. Clin. Microbiol. 38(7): 2706-2714.
Archibald, L. K., S. N. Banerjee, et al. (2004). "Secular trends in hospital-acquired
Clostridium difficile disease in the United States, 1987-2001." J Infect Dis 189(9):
1585-9.
Arroyo, L. G., S. A. Kruth, et al. (2005). "PCR ribotyping of Clostridium difficile isolates
originating from human and animal sources." J Med Microbiol 54(Pt 2): 163-6.
Arroyo, L. G., J. S. Weese, et al. (2004). "Experimental Clostridium difficile enterocolitis
in foals." J Vet Intern Med 18(5): 734-8.
Bachman, J. (2008, May 30th, 2008). "Visualizing One Billion Dollars
" Retrieved January 15, 2010, from <http://www.wallstats.com/blog/visualizing-onebillion-dollars/>.
Bakri, M. M., D. J. Brown, et al. (2009). "Clostridium difficile in ready-to-eat salads,
Scotland." Emerging Infectious Diseases 15(5): 817-818.
Balassiano, I. T., K. R. Miranda, et al. (2009). "Characterization of Clostridium difficile
strains isolated from immunosuppressed inpatients in a hospital in Rio de Janeiro,
Brazil." Anaerobe 15(3): 61-4.
45
Bartlett, J. G. (2008). "Historical perspectives on studies of Clostridium difficile and C.
difficile infection." Clin Infect Dis 46 Suppl 1: S4-11.
Baverud, V. (2004). "Clostridium difficile diarrhea: infection control in horses." Vet Clin
North Am Equine Pract 20(3): 615-30.
Baverud, V., A. Gustafsson, et al. (2003). "Clostridium difficile: prevalence in horses and
environment, and antimicrobial susceptibility." Equine Vet J 35(5): 465-71.
Borgmann, S., M. Kist, et al. (2008). "Increased number of Clostridium difficile
infections and prevalence of Clostridium difficile PCR ribotype 001 in southern
Germany." Euro Surveill 2008;13(49):pii=19057.
(http://www.eurosurveillance.org/ViewArticle.aspx?ArticleId=19057).
Borriello, S. P., P. Honour, et al. (1983). "Foodborne transmission of Clostridium
difficile." Gastroenterology 84(1): 201.
Borriello, S. P., P. Honour, et al. (1983). "Household pets as a potential reservoir for
Clostridium difficile infection." J Clin Pathol 36(1): 84-7.
Campbell, Robert J. P., L. M. P. H. Giljahn, et al. (2009). "Clostridium difficile Infection
in Ohio Hospitals and Nursing Homes During 2006." Infection Control and
Hospital Epidemiology 30(6): 526-533.
Cid, A., A. R. Juncal, et al. (1998). "Clostridium difficile Bacteremia in an
Immunocompetent Child." J. Clin. Microbiol. 36(4): 1167-1168.
Dallal, R. M., B. G. Harbrecht, et al. (2002). "Fulminant Clostridium difficile: an
underappreciated and increasing cause of death and complications." Ann Surg
235(3): 363-72.
Debast, S. B., L. A. van Leengoed, et al. (2009). "Clostridium difficile PCR ribotype 078
toxinotype V found in diarrhoeal pigs identical to isolates from affected humans."
Environ Microbiol 11(2): 505-511.
Dubberke, Erik R., Anne M. Butler, et al. (2009). "Multicenter Study of the Impact of
Community-Onset Clostridium difficile Infection on Surveillance for C. difficile
Infection." Infection Control and Hospital Epidemiology 30(6): 518-525.
46
Dubberke, E. R. and A. I. Wertheimer (2009). "Review of current literature on the
economic burden of Clostridium difficile infection." Infect Control Hosp
Epidemiol 30(1): 57-66.
Ehrich, M., B. D. Perry, et al. (1984). "Acute diarrhea in horses of the Potomac River
area: examination for clostridial toxins." J Am Vet Med Assoc 185(4): 433-5.
Gastmeier, P., D. Weitzel-Kage, et al. (2009). "Surveillance of Clostridium difficileassociated diarrhoea with the German nosocomial infection surveillance system
KISS (CDAD-KISS)." International journal of antimicrobial agents
33(Supplement 1): S19-S23.
Gérard, M., N. Defresne, et al. (1989 ). "Polymicrobial septicemia with Clostridium
difficile in acute diverticulitis." Eur J Clin Microbiol Infect Dis 8(4): 300-302.
Geric, B., R. J. Carman, et al. (2006). "Binary toxin-producing, large clostridial toxinnegative Clostridium difficile strains are enterotoxic but do not cause disease in
hamsters." J Infect Dis 193(8): 1143-50.
Geric, B., M. Rupnik, et al. (2004). "Distribution of Clostridium difficile variant
toxinotypes and strains with binary toxin genes among clinical isolates in an
American hospital." J Med Microbiol 53(Pt 9): 887-94.
Goorhuis, A., D. Bakker, et al. (2008). "Emergence of Clostridium difficile infection due
to a new hypervirulent strain, polymerase chain reaction ribotype 078." Clin
Infect Dis 47(9): 1162-1170.
Goulding, D., H. Thompson, et al. (2009). "Distinctive profiles of infection and
pathology in hamsters infected with Clostridium difficile strains 630 and B1."
Infect Immun 77(12): 5478-85.
Gurian, L., T. T. Ward, et al. (1982). "Possible foodborne transmission in a case of
pseudomembranous colitis due to Clostridium difficile: influence of
gastrointestinal secretions on Clostridium difficile infection." Gastroenterology
83(2): 465-9.
Hammitt, M. C., D. M. Bueschel, et al. (2008). "A possible role for Clostridium difficile
in the etiology of calf enteritis." Vet Microbiol 127(3-4): 343-52.
47
He, M., M. Sebaihia, et al. "Evolutionary dynamics of Clostridium difficile over short and
long time scales." Proceedings of the National Academy of Sciences 107(16):
7527-7532.
Hensgens, M. P., A. Goorhuis, et al. (2009). "Decrease of hypervirulent Clostridium
difficile PCR ribotype 027 in the Netherlands." Euro Surveill 14(45).
Holzel, C. S., K. S. Harms, et al. (2010). "Resistance to Linezolid in a Porcine
Clostridium perfringens Strain Carrying a Mutation in the rplD Gene Encoding
the Ribosomal Protein L4." Antimicrobial Agents and Chemotherapy 54(3):
1351-1353.
Indra, A., H. Lassnig, et al. (2009). "Clostridium difficile: a new zoonotic agent?" Wien
Klin Wochenschr 121(3-4): 91-5.
Janvilisri, T., J. Scaria, et al. (2009). "Microarray identification of Clostridium difficile
core components and divergent regions associated with host origin." J Bacteriol
191(12): 3881-91.
Jones, M. A. and D. Hunter (1983). "Isolation of Clostridium difficile from pigs." Vet
Rec 112(11): 253.
Jones, R. N. and R. Masterton (2001). "Determining the value of antimicrobial
surveillance programs." Diagn Microbiol Infect Dis 41(4): 171-5.
Kaatz, G. W., S. D. Gitlin, et al. (1988). "Acquisition of Clostridium difficile from the
hospital environment." Am J Epidemiol 127(6): 1289-94.
Keel, K., J. S. Brazier, et al. (2007). "Prevalence of PCR ribotypes among Clostridium
difficile isolates from pigs, calves, and other species." J Clin Microbiol 45(6):
1963-1964.
Keessen, E. C., A. J. van den Berkt, et al. (2011). "The relation between farm specific
factors and prevalence of Clostridium difficile in slaughter pigs." Vet Microbiol.
Kim, H., T. V. Riley, et al. (2008). "Increasing prevalence of toxin A-negative, toxin Bpositive isolates of Clostridium difficile in Korea: impact on laboratory
diagnosis." J Clin Microbiol 46(3): 1116-7.
48
Kiss, D. and G. Bilkei (2005). "A new periparturient disease in Eastern Europe,
Clostridium difficile causes postparturient sow losses." Theriogenology 63(1): 1723.
Kuijper, E. J., F. Barbut, et al. (2008). "Update of Clostridium difficile infection due to
PCR ribotype 027 in Europe, 2008." Euro Surveill 13(31).
Kuijper, E. J. and J. T. van Dissel (2008). "Spectrum of Clostridium difficile infections
outside health care facilities." CMAJ 179(8): 747-8.
Kyne, L. (2010). "Clostridium difficile--beyond antibiotics." N Engl J Med 362(3): 264-5.
Lawley, T. D., S. Clare, et al. (2009). "Antibiotic Treatment of Clostridium difficile
Carrier Mice Triggers a Supershedder State, Spore-Mediated Transmission, and
Severe Disease in Immunocompromised Hosts." Infect. Immun. 77(9): 36613669.
Lefebvre, S. L., L. G. Arroyo, et al. (2006). "Epidemic Clostridium difficile strain in
hospital visitation dog." Emerg Infect Dis 12(6): 1036-7.
Libby, D. B. and G. Bearman (2009). "Bacteremia due to Clostridium difficile--review of
the literature." International Journal of Infectious Diseases 13(5): e305-e309.
Limbago, B., C. Long, et al. (2009). "Clostridium difficile Strains from Community-1 Associated
Infections." J of Clinical Micro Published online ahead of print July 1, 2009.
Loo, V. G., L. Poirier, et al. (2005). "A predominantly clonal multi-institutional outbreak
of Clostridium difficile-associated diarrhea with high morbidity and mortality." N
Engl J Med 353(23): 2442-9.
Lyras, D., J. R. O'Connor, et al. (2009). "Toxin B is essential for virulence of Clostridium
difficile." Nature 458(7242): 1176-9.
Metcalf, D. S., M. C. Costa, et al. (2010). "Clostridium difficile in vegetables, Canada."
Lett Appl Microbiol 51(5): 600-602.
Morgan, O. W., B. Rodrigues, et al. (2008). "Clinical severity of Clostridium difficile
PCR ribotype 027: a case-case study." PLoS One 3(3): e1812.
49
Muto, C. A., M. Pokrywka, et al. (2005). "A large outbreak of Clostridium difficileassociated disease with an unexpected proportion of deaths and colectomies at a
teaching hospital following increased fluoroquinolone use." Infect Control Hosp
Epidemiol 26(3): 273-80.
Norman, K. N., R. B. Harvey, et al. (2009). "Varied prevalence of Clostridium difficile in
an integrated swine operation." Anaerobe 15(6): 256-260.
O'Brien, J. A., B. J. Lahue, et al. (2007). "The emerging infectious challenge of
Clostridium difficile-associated disease in Massachusetts hospitals: clinical and
economic consequences." Infect Control Hosp Epidemiol 28(11): 1219-27.
O'Neill, G. L., F. T. Ogunsola, et al. (1996). "Modification of a PCR Ribotyping Method
for Application as a Routine Typing Scheme forClostridium difficile." Anaerobe
2(4): 205-209.
Pell, A. N. (1997). "Manure and microbes: public and animal health problem?" J Dairy
Sci 80(10): 2673-81.
Pepin, J., M. E. Alary, et al. (2005). "Increasing risk of relapse after treatment of
Clostridium difficile colitis in Quebec, Canada." Clin Infect Dis 40(11): 15911597.
Pepin, J., L. Valiquette, et al. (2004). "Clostridium difficile-associated diarrhea in a
region of Quebec from 1991 to 2003: a changing pattern of disease severity."
CMAJ 171(5): 466-72.
Porter MC, Reggiardo C, et al. (2002). Association of Clostridium difficile with bovine
neonatal diarrhea [abstract]. In: Abstracts of the Proceedings of the American
Association of Veterinary Laboratory Diagosticians. 45th annual meeting. Saint
Louis; 2002 Oct 19-21. Davis (CA): American Association of Veterinary
Laboratory Diagosticians.
Post, K. W., B. H. Jost, et al. (2002). "Evaluation of a test for Clostridium difficile toxins
A and B for the diagnosis of neonatal swine enteritis." J Vet Diagn Invest 14(3):
258-9.
Randall, E. J. (2003) "Horse Manure Management: The Nitrogen Enhancement System."
Redelings MD, Sorvillo F, et al. (2007). "Increase in Clostridium difficile-related
mortality rates, United States, 1999-2004." Emerg Infect Dis 13(9): 1417-1419.
50
Ricciardi, R., D. A. Rothenberger, et al. (2007). "Increasing prevalence and severity of
Clostridium difficile colitis in hospitalized patients in the United States." Arch
Surg 142(7): 624-31; discussion 631.
Riley, T. V. (1998). "Clostridium difficile: a pathogen of the nineties." Eur J Clin
Microbiol Infect Dis 17(3): 137-41.
Roberts, K., C. F. Smith, et al. (2008). "Aerial dissemination of Clostridium difficile
spores." BMC Infect Dis 8: 7.
Rodriguez-Palacios, A., M. Koohmaraie, et al. (2011). "Prevalence, Enumeration, and
Antimicrobial Agent Resistance in Clostridium difficile in Cattle at Harvest in the
United States." J Food Prot October, In press.
Rodriguez-Palacios, A., R. J. Reid-Smith, et al. (2009). "Possible seasonality of
Clostridium difficile in retail meat, Canada." Emerg Infect Dis 15(5): 802-805.
Rodriguez-Palacios, A., H. R. Staempfli, et al. (2007). "Clostridium difficile in retail
ground meat, Canada." Emerg Infect Dis 13(3): 485-7.
Rodriguez-Palacios, A., H. R. Stampfli, et al. (2006). "Clostridium difficile PCR
ribotypes in calves, Canada." Emerg Infect Dis 12(11): 1730-6.
Rodriguez-Palacios, A., H. R. Stampfli, et al. (2007). "Natural and experimental infection
of neonatal calves with Clostridium difficile." Vet Microbiol 124(1-2): 166-72.
Rodriguez-Palacios, A., H. R. Stampfli, et al. (2009). "Possible seasonality of
Clostridium difficile in retail meat, Canada." Emerg Infect Dis: 831-834.
Rupnik, M. (2007). "Is Clostridium difficile-associated infection a potentially zoonotic
and foodborne disease?" Clin Microbiol Infect 13(5): 457-459.
Rupnik, M., M. H. Wilcox, et al. (2009). "Clostridium difficile infection: new
developments in epidemiology and pathogenesis." Nat Rev Microbiol 7(7): 52636.
Russmann, H., K. Panthel, et al. (2007). "Evaluation of three rapid assays for detection of
Clostridium difficile toxin A and toxin B in stool specimens." Eur J Clin
Microbiol Infect Dis 26(2): 115-9.
51
Sailhamer, E. A., K. Carson, et al. (2009). "Fulminant Clostridium difficile colitis:
patterns of care and predictors of mortality." Arch Surg 144(5): 433-9; discussion
439-40.
Schwan, C., B. r. Stecher, et al. (2009). "<italic>Clostridium difficile</italic> Toxin CDT
Induces Formation of Microtubule-Based Protrusions and Increases Adherence of
Bacteria." PLoS Pathog 5(10): e1000626.
Shin, B. M., E. Y. Kuak, et al. (2008). "Multicentre study of the prevalence of toxigenic
Clostridium difficile in Korea: results of a retrospective study 2000-2005." J Med
Microbiol 57(Pt 6): 697-701.
Shin, B. M., E. Y. Kuak, et al. (2008). "Emerging toxin A-B+ variant strain of
Clostridium difficile responsible for pseudomembranous colitis at a tertiary care
hospital in Korea." Diagn Microbiol Infect Dis 60(4): 333-7.
Simango, C. (2006). "Prevalence of Clostridium difficile in the environment in a rural
community in Zimbabwe." Trans R Soc Trop Med Hyg 100(12): 1146-50.
Simango, C. and S. Mwakurudza (2008). "Clostridium difficile in broiler chickens sold at
market places in Zimbabwe and their antimicrobial susceptibility." Int J Food
Microbiol 124(3): 268-70.
Smith, L. D. and E. O. King (1962). "Occurrence of Clostridium difficile in infections of
man." J Bacteriol 84: 65-7.
Songer, J. G. (2004). "The emergence of Clostridium difficile as a pathogen of food
animals." Anim Health Res Rev 5(2): 321-6.
Songer, J. G. (2009). "Clostridia as agents of zoonotic disease." Vet Microbiol.
Songer, J. G., R. Jones, et al. (2007). "Prevention of porcine Clostridium difficileassociated disease by competitive exclusion with nontoxigenic organisms." Vet
Microbiol 124(3-4): 358-361.
Songer, J. G., H. T. Trinh, et al. (2009). "Clostridium difficile in retail meat products,
USA, 2007." Emerg Infect Dis 15(5): 819-821.
Spigaglia, P., F. Barbanti, et al. (2008). "Fluoroquinolone resistance in Clostridium
difficile isolates from a prospective study of C. difficile infections in Europe." J
Med Microbiol 57(Pt 6): 784-9.
52
Stabler, R. A., D. N. Gerding, et al. (2006). "Comparative phylogenomics of Clostridium
difficile reveals clade specificity and microevolution of hypervirulent strains." J
Bacteriol 188(20): 7297-305.
Stare, B. G., M. Delmee, et al. (2007). "Variant forms of the binary toxin CDT locus and
tcdC gene in Clostridium difficile strains." J Med Microbiol 56(Pt 3): 329-35.
Tanner, H. E., K. J. Hardy, et al. (2010). "Co-existence of multiple MLVA sub-types of
Clostridium difficile PCR-Ribotype 027 strains within faecal specimens." J Clin
Microbiol.
Terhes, G., E. Urban, et al. (2004). "Community-Acquired Clostridium difficile Diarrhea
Caused by Binary Toxin, Toxin A, and Toxin B Gene-Positive Isolates in
Hungary." J. Clin. Microbiol. 42(9): 4316-4318.
Thomas, J. A., K. C. Newman, et al. (2011). "Bacteraemia from an unrecognized source
(occult bacteraemia) occurring during Clostridium difficile infection."
Scandinavian Journal of Infectious Diseases 43(4): 269-274.
USDA, F. S. a. I. S. (1997) "U.S.D.A. Promotes Horse & Goat Meat."
Von Abercron, S. M., F. Karlsson, et al. (2009). "Low occurrence of Clostridium difficile
in retail ground meat in Sweden." J Food Prot 72(8): 1732-4.
Warny, M., J. Pepin, et al. (2005). "Toxin production by an emerging strain of
Clostridium difficile associated with outbreaks of severe disease in North America
and Europe." Lancet 366(9491): 1079-1084.
Weese, J. S. (2010). "Clostridium difficile in food--innocent bystander or serious threat?"
Clin Microbiol Infect 16(1): 3-10.
Weese, J. S. and J. Armstrong (2003). "Outbreak of Clostridium difficile-associated
disease in a small animal veterinary teaching hospital." J Vet Intern Med 17(6):
813-6.
Weese, J. S., B. P. Avery, et al. (2009). "Detection and enumeration of Clostridium
difficile in retail beef and pork." Appl Environ Microbiol 75(15): 5009-5011.
Weese, J. S., R. Finley, et al. (2010). "Evaluation of Clostridium difficile in dogs and the
household environment." Epidemiol Infect: 1-5.
53
Weese, J. S., L. Toxopeus, et al. (2006). "Clostridium difficile associated diarrhoea in
horses within the community: predictors, clinical presentation and outcome."
Equine Vet J 38(2): 185-8.
Weese, J. S., H. E. Weese, et al. (2001). "Suspected Clostridium difficile-associated
diarrhea in two cats." J Am Vet Med Assoc 218(9): 1436-9, 1421.
Wilcox, M. H. (2004). "Descriptive study of intravenous immunoglobulin for the
treatment of recurrent Clostridium difficile diarrhoea." J Antimicrob Chemother
53(5): 882-4.
Wilcox, M. H., L. Mooney, et al. (2008). "A case-control study of community-associated
Clostridium difficile infection." J Antimicrob Chemother 62(2): 388-96.
Zidaric, V., S. Beigot, et al. (2010). "The occurrence and high diversity of Clostridium
difficile genotypes in rivers." Anaerobe.
Zidaric, V., M. Zemljic, et al. (2008). "High diversity of Clostridium difficile genotypes
isolated from a single poultry farm producing replacement laying hens." Anaerobe
14(6): 325-327.
54
Chapter 3 ― Transient Fecal Shedding and Limited Animal-to-Animal
Transmission of Clostridium difficile by Naturally Infected Finishing Feedlot Cattle
As Published in :
Rodriguez-Palacios, A., C. Pickworth, S. Loerch, and LeJeune. (2011b). "Transient Fecal
Shedding and Limited Animal-to-Animal Transmission of Clostridium difficile by
Naturally Infected Finishing Feedlot Cattle." Appl Environ Microbiol 77(10):
3391-3397.
Abstract
To longitudinally assess fecal shedding and animal-to-animal transmission of
Clostridium difficile among finishing feedlot cattle as a risk for beef carcass
contamination, we tested 186±12 steers (1,369 samples) in an experimental feedlot
facility during the finishing period and at harvest. Clostridium difficile was isolated from
12.9% of steers on arrival (24/186; 0 to 33% among 5 suppliers). Shedding decreased to
undetectable levels a week later (0%, P < 0.001); and remained low (<3.6%) until
immediately prior to shipment for harvest (1.2%). Antimicrobial use did not increase
fecal shedding despite treatment of 53% of animals for signs of respiratory disease.
Animals shedding C. difficile on arrival, however, had 4.6 times higher odds of receiving
antimicrobials for respiratory signs than nonshedders (95% CI OR = 1.4, 14.8; P = 0.01).
Most isolates (39/42) were deemed nontoxigenic based on two complementary multiplex
55
PCRs and ELISA testing. Two linezolid- and clindamycin-resistant PCR ribotype 078
(tcdA+tcdB+cdtB+/39bp-type deletion in tcdC) isolates were identified from two steers (at
arrival and on week 20), but these did not become endemic. The other toxigenic isolate
(tcdA+tcdB+cdtB+/classic tcdC; PCR ribotype 078-like) was identified in the cecum of
one steer at harvest. Spatio-temporal analysis indicated transient shedding and failed to
identify direct animal-to-animal transmission. The association between C. difficile
shedding upon arrival and the subsequent need of antimicrobials for respiratory disease
might indicate common predisposing factors. The isolation of toxigenic C. difficile from
bovine intestines at harvest highlights the potential for food contamination in meat
processing plants.
Introduction
Clostridium difficile is the leading cause of infectious diarrhea in hospitals
worldwide, with more than double the incidence observed today compared to the early
2000s (McFarland 2008). In the United States, C. difficile is estimated to be associated
with over 300,000 clinical cases each year resulting in a major financial burden to the
health care system (Dubberke and Wertheimer 2009). Moreover, the incidence of severe
infections and deaths in the community, particularly in children and pregnant women, is
also increasing (Centers for Disease Control and Prevention (CDC) 2008). This change in
the epidemiology of C. difficile is coincidental with the emergence of hypervirulent
antimicrobial resistant strains, namely, epidemic PCR ribotypes 027 and 078 (Jhung et al.
2008). The presence of these strains in livestock and up to 42% of retail meats (Rupnik
56
2007; Songer et al. 2009) indicates that livestock may serve as a reservoir for
contamination of the food supply, making foods a plausible source for communityacquired C. difficile infections. Resistance to fluoroquinolones and clindamycin,
antibiotics used in human medicine, is also common among C. difficile isolates of animal
and food origin (Rodriguez-Palacios et al. 2006).
Most pathogens that contaminate meat originate from the bacterial flora harbored
by the live animal (Gill 2007). As such, preharvest interventions to control foodborne
pathogens on the farm are predicted to decrease the frequency of carcass contamination
therefore enhancing food safety (Oliver et al. 2009). Clostridium difficile spores of
epidemic genotypes have been isolated from retail meats (Rodriguez-Palacios et al. 2007;
Rodriguez-Palacios et al. 2009; Songer et al. 2009) and vegetables (al Saif and Brazier
1996; Bakri et al. 2009; Metcalf et al. 2010), and from neonatal and young cattle and pigs
(Rodriguez-Palacios et al. 2006; Rupnik et al. 2008; Norman et al. 2009). However, the
patterns of fecal C. difficile shedding among cattle during the finishing period, and most
importantly, information regarding the frequency of bovine carcass contamination at the
time of harvest is limited.
A better understanding of the epidemiology of C. difficile in food-producing
animal populations will help to elucidate the sources of food contamination and
determine if preharvest interventions for this pathogen may be beneficial. The goal of
this study was to determine the prevalence of C. difficile in cattle upon entry to a feedlot,
the incidence of new infections (assessed as new fecal shedders) and the spatiotemporal
57
distribution of C. difficile subtypes in a group of beef cattle up until and at the time of
harvest.
Materials and Methods
Experimental unit and source of animals.
This study was conducted as part of a study evaluating body weight and dietary
supplementation in a finishing feedlot facility at The Ohio State University, Wooster,
Ohio, USA. Purchased Angus-cross steers were maintained as a closed population;
animals entered the feedlot in early fall (October, 2007) and were harvested between 7
and 8.5 months later in spring. Prior to animal arrival, disinfection of the facility
included pressure washing of floors, feeders, waterers, and fresh painting of pen divisions
and fences. The waterers were stainless-steel, automatically refillable, and shared
between two pens. The pens used had concrete slatted floors and an underground pit to
accumulate fecal matter.
The steers originated from four university-owned operations and one commercial
livestock auction. University-owned cattle originated from cow-calf farms that had the
steers (~7-8 months old) alongside their dams in permanent pasture with no transitional
grain based diets offered prior to shipment. Animals from the auction had unknown
origin and management history. Multiple trucks were used for transporting animals from
the source to the feedlot facility. Farms A, B, C, D and the auction market were located at
151, 283, 70, 155, and 265 km away, respectively (transport time 1-4 hours).
Animal allocation, sampling and antimicrobial therapy.
58
Upon arrival to the feedlot, fecal samples were collected from the rectum using
individual plastic sleeves (occasionally, some animals had no feces available for testing).
Animals were also weighed, ear tagged, and systematically assigned to 24 pens based on
order of arrival (8-10 animals/pen). Fecal samples were also collected from each animal
on weeks 1, 4, 12, 20, and the day prior to shipment to the meat processing plant. During
the first month, the steers remained in the originally assigned pens while they were fed a
transitional diet. This diet consisted of a stepwise increase from corn silage, soy hulls and
high moisture corn (45, 20, and 10% of the diet, respectively) to a finishing diet of mostly
high moisture corn (65%) and corn silage (10%). The remaining 25% of the diet was a
pelleted supplement containing protein (soybean meal and urea), vitamins, and minerals
to meet the requirements of growing cattle (Council 2000). No grass or hay was offered
to the animals during the study.
Signs of bovine respiratory disease complex (ranging from fever to fatal
pneumonia) are common in beef steers entering feedlots. A protocol based on
antiinflamatory and antimicrobial drugs approved for use in cattle in the USA was used to
treat affected animals. Flunixin meglumine (1.1 mg/kg, IV, single dose, Banamine,
Schering-Plough) and tulathromycin (2.5 mg/kg, subcutaneous, one dose, Draxxin,
Pfizer) were initially administered to animals with fever (>39.7oC), anorexia, and
depression. If no clinical response was observed within 24-48 hours indicated by
persistently elevated rectal temperature, florfenicol was given (40 mg/kg, subcutaneous,
one dose, Nuflor, Schering-Plough). Ceftiofur (1mg/kg, subcutaneous, one dose a day,
for 3 days, Excenel, Pharmacia & Upjohn) and oxytetracycline (20 mg/kg, subcutaneous,
59
one dose, Liquamycin LA-200, Pfizer) were the third and last antimicrobial options,
respectively, used in nonresponsive cases. Due to removals associated with respiratory
disease, sudden death, or injury, animals were systematically reallocated on weeks 4 and
6 to ensure a final animal density of 7 steers per pen (of comparable body weights)
providing the space needed at the end of the finishing period.
Animals were harvested at two commercial meat processing plants in groups of
26-28 animals (4 pens per week) over a period of 6 weeks. Following evisceration, two
additional samples were collected: cecal content (2-4 ml by transmural puncture of cecal
apex with 16 gauge needle), and a composite external carcass swab using a single
Hydrated-Sponge (3M™, HS10BPW, 10 ml of buffered peptone water broth) per animal.
Swabbing occurred after carcass trimming on both front quarters and the thoracic area
(40 x 40 cm).
Laboratory analysis.
Eight samples per animal (six fecal samples at the feedlot, and one cecal sample
and one carcass swab at harvest) were cultured for C. difficile. Bacterial isolation was
based on a described enrichment protocol (Arroyo et al. 2005; Rodriguez-Palacios et al.
2007). Briefly, one gram of feces, cecal contents, and the carcass swabs were inoculated
into 10, 10, and 35 ml of broth, respectively. This broth was prepared and supplemented
with sodium thaurocholate (0.1%, Sigma Aldrich, St Louis, MO, USA) and
antimicrobials D-cycloserine (250 mg/L) and cefoxitin (8 mg/L) (SR0096, Oxoid, MD,
USA) as previously described (Arroyo et al. 2005; Rodriguez-Palacios et al. 2007). The
broths were anaerobically incubated (CO2/H2/N2, 10/10/80%) at 37oC for 10 days and
60
then centrifuged (7,000 x g, 10 min). The sediments were treated with 96% ethanol (1:1,
v/v, 30 min) to reduce contaminants, centrifuged again and stored at -80oC without
supernatant. Thawed sediments were inoculated onto pre-reduced cycloserine cefoxitin
fructose C. difficile agar supplemented with 7% defibrinated horse blood. Agar plates
were inspected after 5 days of anaerobic incubation at 37oC for suspect colonies (Arroyo
et al. 2005). Up to five C. difficile-suspect (i.e., morphology and 365-nm fluorescence)
colonies from each plate were subcultured onto tryptic soy agar (Acumedia, Lansing, MI)
supplemented with 5% defibrinated sheep blood. Nonhemolytic colonies were
biochemically confirmed as C. difficile via L-proline aminopeptidase activity (Fedorko
and Williams 1997). Five days later, spores were harvested and stored in selective broth
with 30% glycerol at -80oC. All samples were re-coded and blindly processed in
composite batches. Positive controls included C. difficile strain ATCC 9569.
Molecular subtyping.
Thawed C. difficile isolates were revived on blood agar for crude DNA extraction.
Single colonies resuspended in 200 µL of distilled water were boiled for 15 min and
centrifuged (16,000 x g, 1 min) to harvest the supernatant which was stored at -20oC.
DNA concentrations ranged between 17 and 70 ng · µl-1 (mean, 40.7; Nanodrop ND1000, Wilmington, DE). Recovered isolates were molecularly assessed using three
multiplex PCRs (Spigaglia and Mastrantonio 2002; Lemee et al. 2004; Persson et al.
2008). The gene combinations tested were tpi/tcdA/tcdB for multiplex 1, 16S
rDNA/tcdA/tcdB/cdtA/cdtB for multiplex 2, and tcdE/tcdC for multiplex 3 (Spigaglia and
Mastrantonio 2002; Lemee et al. 2004; Persson et al. 2008). PCR ribotyping was also
61
conducted to assess strain similarity (Bidet et al. 1999). Cluster analysis of fingerprint
patterns was conducted using the unweighted pair group method with median and
Pearson coefficients (Bionumerics, v5.1, Applied Maths Inc, Austin, TX). International
PCR ribotype designations were given when available; otherwise, descriptive
nomenclature for the first representative isolate was used for its respective cluster.
Verification of toxin production.
Two commercial ELISAs were used to verify toxin production in the first two
isolates recovered from each PCR ribotype clusters (Thonnard et al. 1996; Garcia et al.
2000) using 72-hour-old colonies grown on blood agar. One ELISA detected toxins A
and B (Wampole-C. difficile TOX A/B II, TechLab, Blacksburg, VA) and the other
detected toxin A only (Clearview C.diff A, Inverness Medical Innovations, Princeton,
NJ). ELISA testing was performed in duplicate. A strain was defined: i) A-B- if no toxins
were detected with the ELISA TOX A/B kit, ii) A-B+ if the ELISA TOX A/B was
positive but the ELISA C diff A was negative, and iii) A+B+ if both ELISAs were
positive. Clostridium difficile ATCC 9569 was used as control. Binary toxin production
was not evaluated.
Toxinotyping and antimicrobial susceptibility.
Restriction fragment length polymorphism of the pathogenicity locus
(toxinotyping) was conducted on select isolates (one per major phylogenetic cluster) as
described by Rupnik (toxin gene fragments A3 and B1) (http://www.mf.unimb.si/mikro/tox/). Minimum inhibitory concentrations against six antimicrobial classes
relevant for treatment or induction of C. difficile infections in humans were also
62
determined. Metronidazole, vancomycin, moxifloxacin, clindamycin, linezolid (first
compound of oxazolidinones) and tigecycline (first compound of glycylcyclines) were
assessed using the E-test method as described in a previous study (AB biodisk,
bioMérieux, France) using Brucella agar (Acumedia, Lansing, MI) (Rodriguez-Palacios
et al. 2009). Reference breakpoints from the Clinical and Laboratory Standards Institute
or reported ranges for vancomycin, linezolid and tigecycline were used for interpretation
(Ackermann et al. 2003; Hecht and Citron 2007; Jones et al. 2009; Nagy and Dowzicky
2010).
Statistical analysis.
Based on reported C. difficile prevalences in calves (Rodriguez-Palacios et al.
2006), a one-sample size estimation indicated that 171 animals would be sufficient to test
a hypothetical shedding prevalence of 7% (alternate 13%, power=0.8, α=0.05). As a
closed population cohort study, variables were assessed using a risk based cumulative
incidence analysis (Dohoo et al. 2003). Binary and continuous data on arrival were
analyzed using chi-square, Fisher‘s exact, t-test or ANOVA statistics, or their
nonparametric options, accordingly (Petrie and Watson 1999). Adjusting for events with
probability close to zero, 95% confidence intervals were calculated using binomial exact
statistics (Clopper and Pearson 1934; Brown et al. 2001). Analyses were conducted using
STATA software package (v 10.1, College Station, TX). To quantify the risk of animalto-animal transmission for C. difficile, the effective reproduction number (‗R(t)‘, number
of secondary shedders produced by each primary shedder during the first week following
arrival) was estimated (Chowell et al. 2009). No reproduction numbers were calculated
63
after week one due to larger sampling intervals (3-8 weeks). We also compared the
cumulative risk of becoming a shedder when exposed to shedders and nonshedders
(Dohoo et al. 2003). For binary data at pen level, Join-Count analysis using the Rook‘s
case was computed as an index of spatial autocorrelation (Bell et al. 2008) to differentiate
clustered from disperse and random distributions of C. difficile-positive pens.
Results
In total, 197 steers arrived to the feedlot, 116 directly from cow-calf farms and 81
from the livestock auction. Upon arrival, fecal samples were obtained from 186 animals
(95.4% of 197; 105 from farms and 81 from auction; 11 steers had no feces in the rectum
at arrival). A total of 1,369 samples were tested for C. difficile.
Shedding at arrival.
Clostridium difficile was isolated from 12.9% of steers on arrival (24/186; 95% CI
= 8.4, 18.5). At the source level, the prevalence of C. difficile ranged from 0 to 33% (Fig.
5A), but there was no statistical difference between animals from farms and the livestock
auction (Chi-square P > 0.5). Univariate and logistic regression analysis controlling by
farm effect showed no association between body weight and C. difficile shedding at
arrival (P > 0.2; odds ratio = 1); C. difficile was isolated from steers of all body weight
ranges (Fig. 5B) indicating no effect of body weight (and possibly age) on shedding in
young finishing steers.
Shedding over time and at harvest.
64
The proportion of animals shedding C. difficile at the feedlot facility decreased
significantly from 12.9% (24/186) at arrival to 0% (0/176; 95% CI = 0, 2.1) a week later
and to 1.7% (3/176; 95% CI = 0.4, 4.9) by week four (Fisher‘s exact P < 0.0001).
Clostridium difficile was not detected on week 12 (0%; 0/168; 95% CI = 0, 2.2) but the
prevalence at week 20 was 3.6% (5/168; 95% CI = 1.3, 7.6). Clostridium difficile was
also isolated from the feces of two animals (1.2%, 2/167; 95% CI = 0.1, 4.2) prior to
shipment for harvest, and from the intestinal content of two other animals at the time of
harvest (cecum, 1.2%; 2/168; 95% CI = 0.1, 4.2). One of these two steers that was C.
difficile positive prior to shipment for harvest was also positive when it entered the
feedlot. Likewise, one of the two steers that had C. difficile at harvest was positive on
week four. Neither of these two animals shed the same strain twice. All carcass swabs
were negative (0/168; 95% CI = 0, 2.2; Fig. 6).
Antimicrobial use and C. difficile shedding.
Clinical signs of respiratory disease (i.e. fever and increased respiratory rate)
prompted the use of antimicrobials in 53% (99/186) of animals; most of them were
treated between weeks one and five. One, two, three and four courses of the antibiotic
choices were necessary in 63, 24, 8, and 7% of the 99 animals treated, respectively.
Tulathromycin was the most commonly used antibiotic, followed by ceftiofur, florfenicol
and oxytetracycline (60, 22, 17 and 1%, of 184 total single doses, respectively). The use
of antimicrobials during the first month of the study (3 single doses on day 6, plus 156
doses between days 8 and 31) did not enhance C. difficile shedding (Fig. 6). By the end of
the study, cumulative data analysis of the 12 steers that became new shedders at the
65
feedlot indicated that antimicrobial use remained not associated with C. difficile shedding
(7/12 treated vs. 5/12 nontreated, P > 0.1). Although no signs of enteric disease (i.e.,
watery diarrhea) were observed, logistic regression analysis controlling for farm effect
indicated that animals shedding C. difficile on arrival had higher odds of subsequently
receiving antimicrobials at the feedlot (odds ratio 4.6; 95% CI = 1.4, 14.8; P = 0.01)
compared to nonshedders. Compared to the source of origin, this effect was driven by
farm B (P = 0.003) whose steers also had the longest transport distance to the feedlot
facility (283 km, ~3.2 h, directly from cow-calf farm).
Reallocation and animal-to-animal transmission.
A total of 29 steers were removed (25 had moderate respiratory signs, 2 had leg
injuries, and 2 died) early in the study leaving 168 animals by week 6. The body weight
of animals removed was not different from that of the remaining animals (t-test P =
0.132). Three of the 29 removed steers (10.3%) were positive for C. difficile on arrival.
Adjusting for withdrawals associated with disease, injuries, and to balance body weights
in each pen; animals were reallocated after having completed the dietary adjustment, on
weeks 4 and 6. At the animal level, spatial statistics showed that the ordered allocation of
incoming steers resulted in a random distribution of shedders within pens (Join-Count, P
> 0.4) (Fig. 7). Regarding animal-to-animal transmission, the reproduction number
derived from data obtained within one week of arrival was zero (i.e., no new shedders
arose from contact with shedding animals detected at arrival). No spatial statistics were
done at individual sampling points after week one due to limited number of shedders
detected at each point (n<5).
66
Molecular subtyping.
In total, 42 isolates recovered from 34 steers were examined. Molecular typing
with one of the multiplex PCR methods (Lemee et al. 2004) showed most strains were
tcdA-tcdB+ but the second multiplex method used (Persson et al. 2008) indicated they
were tcdA-tcdB- (Table 2). A variant genotype (tcdA+tcdB+cdtB+/39bp-type deletion in
tcdC) identical to human and animal epidemic PCR ribotype 078 was also identified in
two steers (arrival, n=1; week 20, n=1) using the two multiplex methods. At harvest, one
of the two cecal isolates identified was a toxigenic tcdA+tcdB+cdtB+/classic tcdC strain.
PCR ribotyping clearly resulted in distinguishable ribotypes recovered from the same
animal on different dates. In addition, some animals shed distinct ribotypes on the same
date (Fig. 8). The two animals that shed PCR ribotype 078 in this feedlot never had direct
nose-to-nose contact, shared a common pen, or have indirect contact via their pen mates.
Toxin production, toxinotyping and antibiotic sensitivity.
ELISA testing of 20 isolates confirmed that tcdA-tcdB-cdtB- strains produced no
toxins and that tcdA+tcdB+cdtB+ toxigenic strains produced both toxins. The multiplexdiscrepant (tcdA-tcdB+; tcdA-tcdB-) strains had no detectable toxin. Cytotoxicity
neutralization assays were not conducted. Toxinotyping and antimicrobial susceptibility
testing showed that the two PCR-ribotype 078 strains, identified at arrival and on week
20, were toxinotype V and resistant to clindamycin and linezolid (MICs were 256 µg/ml
and 8-12 µg/ml, respectively). All isolates were susceptible to metronidazole,
vancomycin, tigecycline and moxifloxacin.
67
Discussion
This study provides new insights to the on-farm epidemiology of C. difficile in
food-producing animals, particularly during the final stages of beef production as well as
the associated potential for food contamination. Although young steers entering the
feedlot harbored C. difficile, fecal carriage of this bacterium, irrespective of toxigenic
genotype, was transient and the transmission to other animals within the feedlot was
neither temporally nor spatially clustered. The prevalence of C. difficile in purchased
steers arriving from the suppliers (i.e., cow-calf farms) was comparable to the reported
prevalence for young calves (i.e., 10-13%), however most isolates were nontoxigenic
(Rodriguez-Palacios et al. 2006). The reasons for the discrepancy observed between the
multiplex PCRs performance regarding the detection of gene tcdB, which made the
difference between the preliminary identification of some isolates as toxigenic by
multiplex 1 and as nontoxigenic by multiplex 2, are unclear but may be in part due to
differences in primer design and specificity.
Irrespective of toxin genotype, we did not observe molecular or temporal
evidence to indicate that shedding persisted over time, either when animals were sampled
one week apart after arrival to the feedlot or when testing occurred 24 hours apart prior to
harvest. These two short sampling intervals were considered critical to assess shedding
persistence upon entry and prior to leaving the feedlot, but sparser sampling was chosen
during the rest of the study (every 4 and 8 weeks) to reflect transition periods typical of
feedlots (i.e., dietary adjustment, susceptibility to bovine respiratory disease, and
68
reallocations). More frequent sampling could provide information about C. difficile
shedding at a more precise timescale, but may yield limited information with regards to
food safety risks.
In the feedlot, the cause of the marked reduction of C. difficile during the first
week was not determined but may be attributed to dietary changes (i.e., animals shifted
from grass and milk based diets to grain based diets in the feedlot). In humans, diet
appears to modulate the growth of C. difficile in the intestinal tract (Iizuka et al. 2004),
but similar effects are unknown for food-producing animals. Alternatively, the move of
cattle to the feedlot may have eliminated exposure to potential sources of C. difficile
spores present on the farms of origin. At the feedlot, the slatted floors could have also
reduced the number of spores to which the animals were exposed.
In humans and other animal species, antimicrobials enhance shedding,
transmission of C. difficile, and the induction of C. difficile associated disease (Clooten et
al. 2008; McFarland 2008). For example, under experimental conditions, antimicrobials
given to asymptomatic shedding mice resulted in ―super-shedder‖ states and increased
transmission to previously C. difficile-negative immune-competent co-housed cohorts
(Lawley et al. 2009). In contrast, in this study, we conclude that antimicrobial use was
not the cause of the decrease in C. difficile shedding because most animals were treated
only after the significant shedding reduction to 0.6% was identified on week one. Only
1.6% of all antimicrobial doses were given (i.e., two animals) prior to sampling the
feedlot on week one.
69
The effect of transportation as a source of stress and contamination is also
unknown. In feedlot steers, bovine respiratory disease complex often develops as
indicator of increased stress-related immune suppression following transportation and
entry to feedlots (Yates 1982). In our study, the association of C. difficile shedding at
arrival (primarily nontoxigenic isolates) with higher odds of receiving antimicrobials for
respiratory signs in the feedlot indicates that C. difficile shedding and the bovine
respiratory disease complex may share common predisposing factors at the farm of
origin.
In neonatal calves, C. difficile shedding has been documented to last for at least 6
days (Rodriguez-Palacios et al. 2007). Fecal excretion of C. difficile was detected on only
one occasion in most shedding steers, and those animals with two positive samples were
shedding distinct C. difficile genotypes at non-consecutive sampling points. The fecal
shedding pattern was therefore considered transient. It is uncertain if detectable C.
difficile shedding resulted from short-term successful bacterial colonization and
proliferation or from intestinal passage of dormant spores ingested from the environment
in food or water.
Despite the low incidence of new shedders in this study, we identified two
animals with C. difficile immediately prior to shipment for slaughter, and two different
animals carrying C. difficile in the cecum at harvest. One of the cecal isolates was fully
toxigenic (tcdA+tcdB+cdtB+/classic tcdC) and had 83% fingerprint pattern similarity to
PCR ribotype 078, but it was clearly different as it had additional bands of lower
molecular weight (Fig. 4). Epidemic PCR ribotype 078 (tcdA+tcdB+cdtB+/39bp deletion
70
in putative toxin negative regulator tcdC/toxinotype V) was identified on two occasions
in this cattle population (one animal at arrival and one on week 20) but did not become
endemic. PCR ribotype 078/toxinotype V is emerging as a cause of disease in
hospitalized people (Debast et al. 2009) and the community(Limbago et al. 2009), and is
a predominant type in swine operations (Keel et al. 2007), retail meats (Songer et al.
2009), and young dairy cattle (Rodriguez-Palacios et al. 2006). It was also recently
isolated from clustered white-tailed deer farms in Ohio, USA (French et al. 2010), in
geographical proximity to where this study was conducted.
The presence of toxigenic C. difficile in feces and intestinal contents of cattle at
the time of harvest indicates that contamination of transport vehicles, lariage areas, and
meat processing plants is possible. For other foodborne pathogens, for instance
Escherichia coli O157:H7, the prevalence of carcass contamination is strongly associated
with the prevalence and load of fecal shedding by live animals at the time of harvest
(Elder et al. 2000). The extent of such an association for C. difficile remains unknown. In
this study, C difficile was not identified in sponges of carcass swabs. To date, there are no
conclusive reports of foodborne C. difficile infections. However, the increasing isolation
of human epidemic strains from retail meats and food animals, including finishing beef
cattle in the present study, and the thermal resistance of C. difficile to minimum
recommended cooking temperatures (Rodriguez-Palacios et al. 2010) underscore the
potential for foodborne transmission.
In contrast to recent studies where PCR ribotype 078 was clearly prevalent in
young food animals (up to 94% in calves) and retail meats (73% of identified genotypes)
71
(Keel et al. 2007; Rupnik et al. 2008; Debast et al. 2009; Songer et al. 2009), this
longitudinal study showed that such epidemic genotype did not become endemic in our
facility. The transient shedding observed in this feedlot indicates that cattle are unlikely
persistent reservoirs of C. difficile during the finishing period. However, the isolation of
one tcdA+tcdB+cdtB+ C. difficile strain from the cecum at harvest (although all carcass
swabs were negative) demonstrates that meat contamination with clinically relevant
isolates could occur at harvest (43, 47). In addition, as with other clostridia (Turcs et al.
2001; Vengust et al. 2003; Sathish and Swaminathan 2008; Rodriguez-Palacios et al.
2009), spore intestinal translocation to deeper muscle tissues could also occur in vivo.
Because bacterial proliferation during processing or in the final products is plausible,
identifying and applying multiple barrier approaches including those that emphasize
control of cross-contamination, and decontamination during processing and food
preparation, could reduce the risk of food contamination with this emerging pathogen.
72
Acknowledgements
This study was supported by an Ohio Agricultural Research and Development
Center (OARDC) SEED grant, The Ohio State University and state and federal funds
allocated to the OARDC. ARP is a research fellow of the Public Health Preparedness for
Infectious Diseases (Targeted Investments in Excellence) Program, The Ohio State
University.Authors are grateful to Mike Kauffman, Rose Schleppi, Jennifer Schrock,
Laura Harpster, and Pamela Schlegel for technical assistance during sample collection,
and Dr. Michele Williams for thorough comments during final preparation of the
manuscript. Special thanks to Dr. J. Scott Wesse (Univ. of Guelph, Canada) for kindly
donating reference strains, and to the reviewers and editors for helpful suggestions.
73
B
Steers with C. difficile - %
40
Chi-square,
farms Vs. auction,
P > 0.5
30
20
10
Body weight - kg
A
F-test, body weight means,
P < 0.001
350
C. difficile:
Negative
Positive
300
250
200
150
0
A (54)
B (27)
C (3)
Auction (81)
D (21)
B
Farm of origin (steers tested, n =)
D
Auction
C
A
Farm of origin
Figure 5. Clostridium difficile in steers at arrival to the feedlot.
A) Prevalence of fecal shedding of C. difficile by source. The combined prevalence was
12.9%. B) C. difficile shedding by body weight. Controlling for farm, there was no
association between shedding status and body weight (F-test, P > 0.1). Lines with solid
squares represent group means.
74
Steers with C. difficile - %
14
12
12
10
%
Fisher’s exact , arrival Vs.
others, P < 0.001
Period of
antimicrobial use
88
6
44
At harvest
2
00
Figure 6. Longitudinal study ― Prevalence of C. difficile shedding in cattle in the feedlot,
and at harvest.
Cecum and carcass samples were collected at harvest, 24 hours after the last on-farm
sampling (shipment).
75
C. difficile at arrival
Solid Wall
Pen 13
Week 1
Reallocation
by week 6
Week 4
Pen 12
Week 12
Week 20
Cumulative:
New shedders
Preharvest/
harvest
A
1
A,C
1
2
1
1
Auction
A
B
2
1
1
D
a
b
2
a
Pen 24
Pen 1
Removed =
Removed =
Positive upon arrival
Positive upon arrival, but negative at sampling
Pen that holds ,or held, at least one shedder
a
Isolate was PCR-type 078, linezolid and
clindamycin resistant
Positive while in the feedlot, but negative at sampling b
Isolate was PCR-type 078-similar,
Positive at harvest – cecal contents
A+B+cdtA/B+
New positive shedder detected while in the feedlot
Figure 7. Spatio-temporal distribution of steers shedding C. difficile at the feedlot
over time (n = 1,369 samples tested). Panels represent two separate rows of 12
contiguous pens (6-7 animals/pen).
Spatial analysis indicated that pens holding incoming shedders followed a random
distribution (Join Count, P > 0.4). Note the absence of new infections in association with
PCR ribotype 078. Arrows and capital letters in left panel indicate order of arrival and
farm of origin.
76
Pearson correlation (Opt:7.00%) [0.0%-100.0%]
Farm - pen
ELISA Type
PCR-078
-
Pos
D
Arrival
86
Au - 1
Pos
D
20
187
A - 23
Pos
D
Harvest
250
D - 2
-
D1
Arrival
117
Au - 5
-
E
Arrival
86
Au - 1
-
E
Arrival
87
Au - 1
Neg
E
Arrival
238
B - 19
Neg
E
Arrival
169
A - 11
Neg
E
Arrival
161
C - 11
Neg
E
Arrival
225
B - 25
-
E
Arrival
204
A - 15
-
A
Arrival
206
A - 15
-
A
Arrival
198
A - 15
-
A
Arrival
169
A - 11
Neg
A
Arrival
160
Au - 10
-
A
Arrival
186
A - 13
-
A
Arrival
123
A - 5
-
A
Arrival
248
B - 21
Neg
A1
Arrival
212
A - 16
Neg
A
20
95
Au - 16
-
A2
Arrival
213
A - 16
Neg
B
Arrival
125
Au - 6
-
I
Arrival
231
B - 19
Neg
J
20
124
Au - 12
Neg
F
20
118
Au - 20
Neg
F1
Arrival
103
Au - 3
Neg
F
Arrival
127
Au - 6
-
G
04
259
D - 22
Neg
H
Arrival
104
Au - 3
-
C
20
94
Au - 12
Neg
C1
04
169
A - 11
-
K
04
169
A - 11
-
L
B
380-week 20
Steer
Control
680-Cecum
Week
177-arrival
bp
Control
350.00
350
300.00
300
250.00
250
500.00
500
600.00
600
800.00
800
700.00
700
100
100
1000
1000
80
80
200.00
200
bp
PCR-Bidet
PCR type-078
A
60
60
40
40
PCR-Bidet
1500bp
1000bp
500bp
100bp
Figure 8. PCR ribotypes of C. difficile from beef cattle in the feedlot.
A) Cluster analysis of representative isolates with at least one toxin gene reaction
positive. Note similarity at arrival. The upper shaded rectangle highlights the similarity of
three toxigenic isolates to PCR ribotype 078. ELISA tested for toxins A and B. -, not
tested. Type, arbitrary letter designation; subscripts indicate the extra bands were
observed compared to other isolates in cluster. See Table 2 for toxin gene profile of the
PCR types (similarity >80%). Steers 86 and 169 carried two different strains.
B) PCR ribotypes highlighted in cluster type D illustrating similarity to PCR ribotype
078.
77
Table 2.Toxin gene profiles of C. difficile PCR ribotypes from beef cattle in the feedlot.
PCR ribotype
Multiplex 1
Multiplex 2
tcdE/tcdC
ELISA
A, A1, A2
tpi+ tcdA- tcdB+
16S+ tcdA-tcdB-cdtA-cdtB-
-/-
Negative
B
tpi+ tcdA- tcdB+
16S+ tcdA- tcdB- cdtA-cdtB-
-/-
Negative
C, C1
tpi+tcdA-tcdB+
16S+ tcdA- tcdB- cdtA- cdtB-
-/-
Negative
D (PCR 078)
tpi+ tcdA+ tcdB+
16S+ tcdA+ tcdB+ cdtA+ cdtB+
+/39bp del
A+B+
D1
tpi+ tcdA+ tcdB+
16S+ tcdA+ tcdB+ cdtA+ cdtB+
+/classic
ND
E
tpi+ tcdA- tcdB+
16S+tcdA-tcdB-cdtA-cdtB-
-/-
Negative
F-L
tpi+ tcdA- tcdB+
16S+ tcdA-tcdB-cdtA-cdtB-
-/-
Negative
Others
tpi+ tcdA- tcdB-
16S+ tcdA-tcdB-cdtA-cdtB-
ND
ND
Subscripts in PCR ribotype letters indicate extra bands were observed compared to
representative isolates in cluster (Fig 4). Isolates were tested with two complementary
PCR multiplex methods for C. difficile (Lemee et al. 2004; Persson et al. 2008). The
discrepancy between tcdB in both multiplexes for some PCR ribotypes (e.g., PCR
ribotype A) indicates differential performance for tcdB gene. Isolate type D1 was not
available for further testing. ND, not determined. Nontoxigenic C. difficile are unlikely to
induce enteric disease.
78
References
Ackermann, G., D. Adler and A. C. Rodloff (2003). "In vitro activity of linezolid against
Clostridium difficile." J Antimicrob Chemother 51(3): 743-745.
al Saif, N. and J. S. Brazier (1996). "The distribution of Clostridium difficile in the
environment of South Wales." J Med Microbiol 45(2): 133-137.
Arroyo, L. G., J. Rousseau, B. M. Willey, et al. (2005). "Use of a selective enrichment
broth to recover Clostridium difficile from stool swabs stored under different
conditions." J Clin Microbiol 43(10): 5341-5343.
Bakri, M. M., D. J. Brown, J. P. Butcher, et al. (2009). "Clostridium difficile in ready-toeat salads, Scotland." Emerging Infectious Diseases 15(5): 817-818.
Bell, N., N. Schuurman and S. M. Hameed (2008). "Are injuries spatially related?
Joinâ€―count spatial autocorrelation for small-area injury analysis." Injury
Prevention 14(6): 346-353.
Bidet, P., F. Barbut, V. Lalande, et al. (1999). "Development of a new PCR-ribotyping
method for Clostridium difficile based on ribosomal RNA gene sequencing."
FEMS Microbiology Letters 175(2): 261-266.
Brown, L. D., T. T. Cai and A. DasGupta (2001). "Interval Estimation for a Binomial
Proportion." Statistical Science 16(2): 101-117.
Centers for Disease Control and Prevention (CDC), U. S. (2008). "Surveillance for
community-associated Clostridium difficile--Connecticut, 2006." MMWR Morb
Mortal Wkly Rep 57(13): 340-343.
Chowell, G., J. M. Hyman, L. M. A. Bettencourt, et al. (2009). The Effective
Reproduction Number as a Prelude to Statistical Estimation of Time-Dependent
Epidemic Trends. Mathematical and Statistical Estimation Approaches in
Epidemiology, Springer Netherlands: 103-121.
79
Clooten, J., S. Kruth, L. Arroyo, et al. (2008). "Prevalence and risk factors for
Clostridium difficile colonization in dogs and cats hospitalized in an intensive
care unit." Vet Microbiol 129(1-2): 209-214.
Clopper, C. J. and E. S. Pearson (1934). "The use of confidence or fiducial limits
illustrated in the case of the binomial " Biometrika 26: 404-413
Council, N. R., Ed. (2000). Nutrient requirements of beef cattle. Washington, DC., Natl.
Acad. Press.
Debast, S. B., L. A. van Leengoed, A. Goorhuis, et al. (2009). "Clostridium difficile PCR
ribotype 078 toxinotype V found in diarrhoeal pigs identical to isolates from
affected humans." Environ Microbiol 11(2): 505-511.
Dohoo, I., W. Martin and H. Stryhn (2003). Cohort studies. Veterinary Epidemiologic
Research. S. M. McPike. Charlottetown, Prince Edward Island, Canada., AVC,
Inc.: 69, 151-162
Dubberke, E. R. and A. I. Wertheimer (2009). "Review of current literature on the
economic burden of Clostridium difficile infection." Infect Control Hosp
Epidemiol 30(1): 57-66.
Elder, R. O., J. E. Keen, G. R. Siragusa, et al. (2000). "Correlation of enterohemorrhagic
Escherichia coli O157 prevalence in feces, hides, and carcasses of beef cattle
during processing." Proc Natl Acad Sci U S A 97(7): 2999-3003.
Fedorko, D. P. and E. C. Williams (1997). "Use of cycloserine-cefoxitin-fructose agar
and L-proline-aminopeptidase (PRO Discs) in the rapid identification of
Clostridium difficile." J Clin Microbiol 35(5): 1258-1259.
French, E., A. Rodriguez-Palacios and J. T. LeJeune (2010). "Enteric bacterial pathogens
with zoonotic potential isolated from farm-raised deer." Foodborne Pathog Dis
7(9): 1031-1037.
Garcia, A., J. Perez, A. Pulido, et al. (2000). "Evaluation of four rapid methods for the
investigation of the toxigenic capacity of Clostridium difficile strains isolated in a
selective medium." Enferm Infecc Microbiol Clin 18(3): 109-112.
Gill, C. O. (2007). "Microbiological conditions of meats from large game animals and
birds." Meat Science 77(2): 149-160.
80
Hecht, D. W. and D. Citron (2007). "Clinical Laboratory Standards Institute. M11-A7Methods for antimicrobial susceptibility testing of anaerobic bacteria - Seven
Edition." 27(2): 64p.
Iizuka, M., H. Itou, S. Konno, et al. (2004). "Elemental diet modulates the growth of
Clostridium difficile in the gut flora." Aliment Pharmacol Ther 20 Suppl 1: 151157.
Jhung, M. A., A. D. Thompson, G. E. Killgore, et al. (2008). "Toxinotype V Clostridium
difficile in humans and food animals." Emerg Infect Dis 14(7): 1039-1045.
Jones, R. N., S. Kohno, Y. Ono, et al. (2009). "ZAAPS International Surveillance
Program (2007) for linezolid resistance: results from 5591 Gram-positive clinical
isolates in 23 countries." Diagn Microbiol Infect Dis 64(2): 191-201.
Keel, K., J. S. Brazier, K. W. Post, et al. (2007). "Prevalence of PCR ribotypes among
Clostridium difficile isolates from pigs, calves, and other species." J Clin
Microbiol 45(6): 1963-1964.
Lawley, T. D., S. Clare, A. W. Walker, et al. (2009). "Antibiotic Treatment of
Clostridium difficile Carrier Mice Triggers a Supershedder State, Spore-Mediated
Transmission, and Severe Disease in Immunocompromised Hosts." Infect.
Immun. 77(9): 3661-3669.
Lemee, L., A. Dhalluin, S. Testelin, et al. (2004). "Multiplex PCR targeting tpi (triose
phosphate isomerase), tcdA (Toxin A), and tcdB (Toxin B) genes for toxigenic
culture of Clostridium difficile." J Clin Microbiol 42(12): 5710-5714.
Limbago, B. M., C. M. Long, A. D. Thompson, et al. (2009). "Clostridium difficile
strains from community-associated infections." J Clin Microbiol 47(9): 30043007.
McFarland, L. V. (2008). "Update on the changing epidemiology of Clostridium difficileassociated disease." Nat Clin Pract Gastroenterol Hepatol 5(1): 40-48.
Metcalf, D. S., M. C. Costa, W. M. Dew, et al. (2010). "Clostridium difficile in
vegetables, Canada." Lett Appl Microbiol 51(5): 600-602.
Nagy, E. and M. J. Dowzicky (2010). "In vitro activity of tigecycline and comparators
against a European compilation of anaerobes collected as part of the Tigecycline
Evaluation and Surveillance Trial (TEST)." Scand J Infect Dis 42(1): 33-38.
81
Norman, K. N., R. B. Harvey, H. M. Scott, et al. (2009). "Varied prevalence of
Clostridium difficile in an integrated swine operation." Anaerobe 15(6): 256-260.
Oliver, S. P., D. A. Patel, T. R. Callaway, et al. (2009). "ASAS Centennial Paper:
Developments and future outlook for preharvest food safety." J. Anim Sci. 87(1):
419-437.
Persson, S., M. Torpdahl and K. E. Olsen (2008). "New multiplex PCR method for the
detection of Clostridium difficile toxin A (tcdA) and toxin B (tcdB) and the binary
toxin (cdtA/cdtB) genes applied to a Danish strain collection." Clin Microbiol
Infect 14(11): 1057-1064.
Petrie, A. and P. Watson (1999). Statistics for Veterinary and Animal Sciences. Oxford,
UK., Blackwell Science.
Rodriguez-Palacios, A., R. J. Reid-Smith, H. R. Staempfli, et al. (2009). "Possible
seasonality of Clostridium difficile in retail meat, Canada." Emerg Infect Dis
15(5): 802-805.
Rodriguez-Palacios, A., R. J. Reid-Smith, H. R. Staempfli, et al. (2010). "Clostridium
difficile survives minimal temperature recommended for cooking ground meats."
Anaerobe 16: 540-542.
Rodriguez-Palacios, A., H. R. Staempfli, T. Duffield, et al. (2007a). "Clostridium difficile
in retail ground meat, Canada." Emerg Infect Dis 13(3): 485-487.
Rodriguez-Palacios, A., H. R. Stampfli, T. Duffield, et al. (2006). "Clostridium difficile
PCR ribotypes in calves, Canada." Emerg Infect Dis 12(11): 1730-1736.
Rodriguez-Palacios, A., H. R. Stampfli, M. Stalker, et al. (2007b). "Natural and
experimental infection of neonatal calves with Clostridium difficile." Vet
Microbiol 124(1-2): 166-172.
Rupnik, M. (2007). "Is Clostridium difficile-associated infection a potentially zoonotic
and foodborne disease?" Clin Microbiol Infect 13(5): 457-459.
Rupnik, M., A. Widmer, O. Zimmermann, et al. (2008). "Clostridium difficile toxinotype
V, ribotype 078, in animals and humans." J Clin Microbiol 46(6): 2146.
82
Sathish, S. and K. Swaminathan (2008). "Molecular characterization of the diversity of
Clostridium chauvoei isolates collected from two bovine slaughterhouses:
Analysis of cross-contamination." Anaerobe 14(3): 190-199.
Songer, J. G., H. T. Trinh, G. E. Killgore, et al. (2009). "Clostridium difficile in retail
meat products, USA, 2007." Emerg Infect Dis 15(5): 819-821.
Spigaglia, P. and P. Mastrantonio (2002). "Molecular analysis of the pathogenicity locus
and polymorphism in the putative negative regulator of toxin production (TcdC)
among Clostridium difficile clinical isolates." J Clin Microbiol 40(9): 3470-3475.
Thonnard, J., F. Carreer, V. Avesani, et al. (1996). "Toxin A detection on Clostridium
difficile colonies from 24-h cultures." Clin Microbiol Infect 2(1): 50-54.
Turcs, aacute, J. n, et al. (2001). "Occurrence of Anaerobic Bacterial, Clostridial, and
Clostridium perfringens Spores in Raw Goose Livers from a Poultry Processing
Plant in Hungary." Journal of food protection 64: 1252-1254.
Vengust, M., L. G. Arroyo, J. S. Weese, et al. (2003). "Preliminary evidence for dormant
clostridial spores in equine skeletal muscle." Equine Vet J 35(5): 514-516.
Yates, W. D. (1982). "A review of infectious bovine rhinotracheitis, shipping fever
pneumonia and viral-bacterial synergism in respiratory disease of cattle." Can J
Comp Med 46(3): 225-263.
83
Chapter 4 ― Summer prevalence of Clostridium difficile in pigs, cattle, poultry, and
horses: finishing food animals are not naturally reservoirs
(For Submission)
Abstract
Enteric human pathogen Clostridium difficile has been isolated from food and
companion animals and from retail foods raising concerns about foodborne transmission.
To comparatively assess the potential for food or environmental contamination at the
time of animal harvest, we conducted a cross-sectional fecal prevalence study in animals
intensively raised for food production (swine, beef and dairy cattle, poultry and turkey)
within 24 hours of harvest, and compared it to the prevalence of a group of intensively
managed horses at a racetrack. Animals were sampled in the same region (Ohio, USA)
during one month period in summer. Clostridium difficile was isolated from 5 out of 875
(0.6% prevalence) animals. One isolate was from a commercial beef steer at a feedlot
(1/180), while the remaining isolates were from racing horses (7.2%; 4/55). Hence,
horses had the highest prevalence of C. difficile compared to the food animal groups (2 x
6, chi-square, P < 0.001). The low prevalence of C. difficile in finishing food animals
during summer, in relation to the reproducible prevalence of C. difficile in horses,
indicate commercial finishing food animals are not naturally successful reservoirs for this
84
enteric pathogen, and that the risk of food contamination at harvest is relatively low in the
warmth of summer. With this low level of C. difficile prevalence (summer baseline), it is
now necessary to determine the shedding dynamics in winter, when C difficile in people,
animals and foods seem to occur with higher frequency.
Introduction
The enteric pathogen Clostridium difficile associated with infections (CDI) in
humans has been repeatedly isolated from retail foods since 2006, raising concerns about
foodborne transmission to communities at risk (Rodriguez-Palacios et al. 2006;
Rodriguez-Palacios et al. 2007). However, little is known about how food gets
contaminated. Initial studies documented the presence of emerging hypervirulent strains
of C difficile in foods and young livestock (Rodriguez-Palacios et al. 2006; Keel et al.
2007; Pirs et al. 2008; Norman et al. 2009), but there is limited research regarding the
presence of such strains in food animas at the time of harvest, and when they enter the
food chain. Some studies have been conducted at single commercial operations and
largely on single animal species. The resistant nature of this pathogen indicates that
contamination of various animal species within the same region could be possible. This is
particularly possible, especially because there are reports of long survival (years) of
spores in soil and horse manure (Baverud et al. 2003), presence of toxigenic spores in
bodies of water (Zidaric et al. 2010), and indications of aerial dissemination of spores
within commercial swine operations and towards their surrounding environment
(Keessen et al. 2010) in Europe.
85
Independent studies indicate that regional predominance of C. difficile
hypervirulent subtypes is possible. A recent study on C. difficile conducted on farmed
deer, indentified the presence of C. difficile PCR ribotype 078 in a cluster of farms in a
relatively small region of Ohio, USA (French et al. 2010). The lack of apparent social or
food/water network connectivity between those farms indicated that other mechanisms
might facilitate spore dissemination at a local scale (French et al. 2010). In the same
region, the PCR ribotype 078 was concurrently isolated from a feedlot cattle operation,
although the pathogen was not persistently shed until the time of harvest (RodriguezPalacios et al. 2011).
In other regions of North America, C. difficile have been isolated from pleasure
healthy horses (Weese et al. 2006), but little is known about the prevalence in horses in
Ohio, and no studies are available for poultry in the USA. In Africa and Europe, single
studies indicate that the prevalence of C. difficile in poultry can be high (18-29%)
towards the end of the production system (Simango 2006; Simango and Mwakurudza
2008), but the production conditions differ from the ones encountered in commercial
operations in the USA.
Considering that fecal shedding at harvest is a major risk factor for carcass
contamination with foodborne pathogens (Elder et al. 2000), understanding if the risk of
carriage of a given pathogen is higher in any given farm animal species would be
important to target intervention efforts to reduce the risk of food contamination at the
time of harvest and to prevent the potential for environmental contamination via animal
excrements, which may vary with the type of operation.
86
Because the fecal shedding of C. difficile in young livestock (at their farms) has
been reported to be high in winter and low in summer in calves and piglets (RodriguezPalacios et al. 2006; Norman et al. 2009), the present study was designed to quantify the
lowest baseline prevalence of fecal shedding of C. difficile in farm animals (mature, or
ready for harvest) during the least favorable reported season for fecal shedding at the
farm, summer. In a cross sectional study we determined if common farm animal species,
particularly food animals and horses, which have different management production
systems, have the same probability of carrying C. difficile in a relatively well defined
geographical area and temporal frame (Ohio, USA), i.e., during a one month sampling in
the warmest months of summer.
Materials and Methods
Food animal farms, processing plants, and horse racetrack.
We conducted a cross-sectional study of C. difficile in fecal or hind gut intestinal
samples from finishing swine, dairy calves and mature dairy and feedlot cattle, chicken
and turkey, and stabled horses at the end of a racing season in Ohio, USA. The facilities
visited and sampled were located within a radius of 290 ± 32 kilometers from our
research center. Establishments sampled corresponded to commercial farms intended for
food production, with the exception of the racing horses. Samples were collected during
one month in summer (August–September), 2007. Farms were sampled based on
convenience, but it corresponded to producers with moderate or large market share in the
region and who kindly volunteered to participate in the study. Importantly, it was
87
considered that all animals and operations here sampled had in common that they were
intensively managed, have relatively high animal production densities, and were assumed
to have an average high level of stress for the type of production system (i.e., prior to or
at harvest for food animals, racing conditions for horses).
Sampling.
Within each farm or production unit, the median number of animals sampled was
thirty. Animals were sampled using a systematic sampling approach. This is the sample
size divided by the number of finishing lots or animals of interest (i.e., immediately prior
to harvest as in the case of feedlot cattle or finishing pigs, or at harvest as in the case of
poultry). Animals were sampled at regular intervals to minimize potential clustering
associated with animal groupings or lots. The intervals were calculated by dividing the
number of animals available by the estimated sample size. For every production system
five farms were sampled for a total of 150 animals per production system. Horses were
one exception; 55 animals (~25% of all possible horses) were sampled from one single
available racetrack facility, where animal traffic on a weekly basis was inherently high.
None of the facilities simultaneously housed more than one animal species of interest.
Cattle samples were collected from the rectum while swine and horse samples
were collected from the pen floor. For poultry, an inspector directed the collection of the
viscera and internal organs from selected animals at the time of evisceration; the viscera
was packed in individual plastic bags, refrigerated and sent to the laboratory for sample
collection (composite two grams of cecal and rectal-cloacal content). In all cases, fecal
88
and intestinal samples were transported refrigerated to the Ohio State University for C.
difficile analysis and processed within 24 hours of collection.
C. difficile prevalence and molecular confirmation.
We used an enrichment protocol for C. difficile as in previous studies (Arroyo et
al. 2005; Rodriguez-Palacios et al. 2006; Rodriguez-Palacios et al. 2007). In brief,
homogenized fecal or intestinal samples (1 gram) were enriched in 15 ml of selective C.
difficile broth (Arroyo et al. 2005) supplemented with 0.1% sodium taurocholate (Sigma
Aldrich, St Louis, MO) and C. difficile supplement SR0096 (Oxoid, Columbia, MD,
cycloserine and cefoxitin). The following steps included incubation (37oC, 7 days),
centrifugation of the broths and treatment of sediments with 96% ethanol, and plating
onto C. difficile fructose agar supplemented with 7% defibrinated horse blood (Quad
Five, Ryegate, MT) and the cycloserine/cefoxitin supplement. After 5 days of anaerobic
incubation, suspect nonhemolitic fluorescent (UV365) colonies on tryptic soy (plus 5%
defibrinated sheep blood) agar were biochemically (L-proline aminopeptidase activity;
Pro Disc, Remel, Lenexa, KS, USA) and molecularly confirmed (Fedorko and Williams
1997). These steps are described in more detail in one of our recent enumeration studies
(Rodriguez-Palacios et al. 2011).
Pure single colonies revived on blood agar (48 hours of anaerobic incubation)
were suspended in double-deionized sterile water and boiled for 8 minutes to obtain
crude DNA, which was used as template in the PCR reactions. Isolates were confirmed
for the presence of the triose-phosphate isomerase (tpi) gene; toxin genes tcdA and tcdB
were also assessed (Lemee et al. 2004). PCR ribotyping was conducted as described
89
(Bidet et al. 1999) to determine strain diversity. Isolates were deemed toxigenic if they
had one of the toxin genes tcdA, tcdB, cdtA/B and toxins (Rupnik et al. 2005).
Toxin ELISAs and dialysis cytotoxicity assay in Vero cells.
Representative isolates were tested with two complementary commercial ELISAs
(C. difficile TOX A/B II, TechLab, Blacksburg, VA, and Clearview C. diff A (Wampole
Laboratories, Inverness Medical Professional Diagnostics, Princeton, NJ) as described
(Thonnard et al. 1996; Garcia et al. 2000) using 72-hour culture colonies. Toxigenic
strain (tcdA+/tcdB+/cdtAB-/toxinotype 0) ATCC 9689 was used as a positive control.
Binary toxin production was not determined. Dialysis tubing and citotoxicity assay was
conducted on ELISA and toxin-gene negative isolates to verify the lack of toxin
production as described (Rodriguez-Palacios et al. 2011).
Statistics.
In a previous study, we estimated based on simple-random sampling calculations
a sample size of 120 animals to verify a hypothetical prevalence of C. difficile of 5.6 ±
6% in cattle at harvest (one-sample, α = 0.05, power = 0.8). Our target here was to
sample 5 farms for each animal species based on convenience. To empirically control for
clustering due to animal-animal transmissibility (within farms previously shown to be
very low in mature finishing cattle) (Rodriguez-Palacios et al. 2011), we arbitrarily
increase the sample size to 150 individuals per animal species. Based on a previous study
where farm level prevalence was documented to be 23% for cattle based on testing of 2-6
calves per farm (Rodriguez-Palacios et al. 2006), it was deemed sufficient to sample 30
animals per farm to detect at least one clustered group of finishing animals (farm)
90
exposed to C. difficile, as recently shown for pig farms (Thakur et al. 2010). Culture
binary data was analyzed using Fisher‘s exact or chi-square statistics (Petrie and Watson
1999). Confidence intervals were estimated using exact statistics (Clopper and Pearson
1934). (STATA, v10.1, College Station, TX). Although this was not a clinical trial, we
considered the ‗REFLECT‘ guidelines for study reporting (Sargeant et al. 2010).
Results
Clostridium difficile was isolated from 5 out of 875 (0.57% prevalence) animals
in summer. One isolate was derived from one commercial beef steer at a feedlot (0.55%;
1/180), while the remaining were racing horses (7.3%; 4/55). In this study, the population
of clinically healthy race horses (with no evidence of diarrhea at the individual basis not
at the facility as a collective; visited by A.R.P) was the one with the highest rate of C.
difficile isolation (2 x 6, chi-square, P < 0.001). Despite the isolation of C. difficile from
one cattle, the shedding prevalences were comparably low across the food animal
operations (2 x 5, chi-square, P < 0.469; Table 3).
Available ELISA and molecular tests identified that the PCR ribotype found in
the horse population was indistinguishable by PCR ribotyping to the most common
nontoxigenic strain identified in harvested cattle a year later. Dialysis citotoxicity assay
on vero cells confirmed that the nontoxigenic isolate did not produce toxins even after 72
hours of incubation and despite heavy noticeable production of cell biomass. No toxin
testing is yet available for the remaining isolates.
91
Discussion
As the incidence of Clostridium difficile infections (CDI) in humans have
increased, the prevalence and clinical relevance of this enteric pathogen in domestic
animals has also been increasingly documented since the mid 2000s. Comprehensive
studies in humans and food animals have indicated that there is a seasonal component
which modifies the prevalence (and incidence) dynamics of this pathogen, with the
lowest estimates observed during summer (Rodriguez-Palacios et al. 2006; Norman et al.
2009). This observation, in addition to the identification of emerging hypervirulent
strains from young livestock (Rodriguez-Palacios et al. 2006; Goorhuis et al. 2008),
indicated that there is great similarity between the epidemiology and seasonal dynamics
of C difficile in animals and humans. Weather conditions are an obvious factor that can
independently produce seasonal transformation in animals and humans (e.g., disease
susceptibility to debilitating diseases, pathogen persistence). However, the documented
evidence for seasonality of food contamination with C. difficile indicates that tainted
foods could theoretically serve as a complement link between the two species seasonality
patterns (Rodriguez-Palacios et al. 2009). Here we determined the shedding of C. difficile
in food animals ready to enter the food chain during summer when the probability of
shedding is the lowest to determine if they could be successful natural reservoirs, and to
infer the risk for food contamination at harvest.
By sampling various farm animal species we confirmed a very low prevalence of
C. difficile in food animals in a region where C. difficile is a growing problem for the
92
health care system since the 2000s (Campbell et al. 2009). This low shedding prevalence
indicates that the potential for food contamination during the first steps of food
production in summer is low, compared to the prevalence of pathogens such as
Campylobacter spp (2-8%, cattle and pigs) (Boes et al. 2005; Dodson and LeJeune 2005).
However, we have shown that even a small number of positive shedders might have
enough spores in the feces to consider environmental contamination of processing
facilities a risk for post-harvest contamination of the food supply (Rodriguez-Palacios et
al. 2011). Especially, because regular disinfection eliminate most vegetative fecal
contaminants, but spores are resistant to most commercial disinfection products (Wilcox
and Fawley 2000).
Here we recovered C. difficile from 7.2% of racing horses at the racetrack
facilities. Compared to food animals, horse shed at a higher proportion, but the proportion
observed was reproducible as it was comparable to reported values in horses from other
racetracks (5-8%) (Medina-Torres et al. 2011). Thus, the isolation of C. difficile from
horses documents the appropriate performance of our culture method, which has been
recently shown to be adequate to comparatively screen for this pathogen in beef and dairy
cattle, with some limitations for pigs (Thitaram et al. 2011). Although no methods have
been validated for poultry, the isolation principles of the method used here (Arroyo et al.
2005) are also comparable to methods used in previous poultry studies. Comparative
considerations regarding culture methodology thus indicate that food animals are indeed
uncommon C. difficile shedders, at least during summer. A study of pigs, cattle and
broiler sampled at abattoirs in Austria (March-July, 2008) found prevalence of C. difficile
93
between 3.3 and 5% (total n=187) (Indra et al. 2009), suggesting that weather conditions,
and/or regional differences, definitely play a substantial role in the epidemiology of C.
difficile in food animals.
The low shedding prevalence observed in our finishing food animals is in
agreement with the low prevalence observed in longitudinal studies conducted
independently for finishing cattle, pigs and poultry in diverse regions (Zidaric et al. 2008;
Norman et al. 2009; Rodriguez-Palacios et al. 2011; Rodriguez-Palacios et al. 2011).
Regarding poultry species, no studies are yet available characterizing the shedding
dynamics of this pathogen in intensive production systems in the USA. As a referent,
studies in Slovenia found in a single farm an age-dependet reduction of C. difficile
shedding in laying hens, with decreasing shedding prevalence from 100% in chicks to a
final 30 % in mature laying hens (Zidaric et al. 2008). Here we did not sample hens, but
the high prevalence (7%) of C. difficile in retail chicken meats in Canada (Weese et al.
2010) indicate that there must be very relevant management factors that could modulate
the dynamics of the pathogen in broilers in intensive systems. If our results were
applicable to other farms, alternatively, there could be factors at harvest or postharvest to
explain the rise in the rate of retail chicken contamination (7%) with C. difficile. Of food
safety reassurance, none of the commercially raised poultry and the finishing swine tested
(450 animals), which are part of a single commercial network (farms span a radius of
about 40 kilometers), had C. difficile at the time of harvest.
Taken together, the low prevalence of C. difficile in finishing food animals, in
relation to the reproducible prevalence of C. difficile shedding in intensively managed
94
racing horses during summer, indicate that commercially raised food animals are not
naturally successful reservoirs for this enteric pathogen, and that the risk of food
contamination at harvest is relatively low in summer. With this low baseline prevalence,
it is now necessary to explore the dynamics and prevalence of shedding in winter months,
in developed temperate nations when reportedly higher prevalence of C difficile in
people, animals and foods seem to occur.
Acknowledgements
This study was funded in part by state and federal funds allocated to the Ohio
Agricultural Research and Development Center, The Ohio State University, and an
internal funding SEEDs program. ARP is a research fellow of the Public Health
Preparedness for Infectious Diseases Program, The Ohio State University. Thanks to
Laura Harpster and Pamela Schlegel for invaluable technical assistance. Special thanks
go to the participating meat processing plants, farms, and horse owners. Sincere thanks to
Dr Ramiro Toribio for guidance and comments during sampling phase.
95
Table 3. Comparative summer prevalence of C. difficile fecal shedding in farm animals,
Ohio, USA.
Animals/
operation
(Operations,
n)
C. difficile
prevalence
95% CI of
prevalence
Grain-based
30 (6)
0.6%
(1/180)
0.14 – 3%
Milk Production
Total mixed
ratio
30 (5)
0% (0/150)
0 – 2%
Swine, finishing
Meat -Prior harvest
Concentrate
30 (5)
0% (0/150)
0 – 2%
Broilers, finishing
Chickens, local
Turkeys, finishing
Meat - At harvest
Meat - At market
Meat - At harvest
None - End of
season
Concentrate
Unspecified
Concentrate
Hay,
concentrate
30 (5)
40 (1)
30 (5)
0% (0/150)
0% (0/40)
0% (0/150)
7.2%
(4/55)*
0 – 2%
0 – 9%
0 – 2%
n/a
n/a
Animal
production
Cattle, beef
feedlot
Cattle, mature
dairy
Horses, racing
Total
Relation to food
chain/sampling
Primary
Diet
Meat -Prior harvest
55 (1)
875 (28)
*Statistically different from the rest, Chi square, P < 0.001.
96
2 – 17%
References
Arroyo, L. G., J. Rousseau, B. M. Willey, et al. (2005b). "Use of a selective enrichment
broth to recover Clostridium difficile from stool swabs stored under different
conditions." J Clin Microbiol 43(10): 5341-5343.
Baverud, V., A. Gustafsson, A. Franklin, et al. (2003). "Clostridium difficile: prevalence
in horses and environment, and antimicrobial susceptibility." Equine Vet J 35(5):
465-471.
Bidet, P., F. Barbut, V. Lalande, et al. (1999). "Development of a new PCR-ribotyping
method for Clostridium difficile based on ribosomal RNA gene sequencing."
FEMS Microbiology Letters 175(2): 261-266.
Boes, J., L. Nersting, E. M. Nielsen, et al. (2005). "Prevalence and diversity of
Campylobacter jejuni in pig herds on farms with and without cattle or poultry."
Journal of Food Protection 68(4): 722-727.
Campbell, Robert J. P., L. M. P. H. Giljahn, K. M. P. H. Machesky, et al. (2009).
"Clostridium difficile Infection in Ohio Hospitals and Nursing Homes During
2006." Infection Control and Hospital Epidemiology 30(6): 526-533.
Clopper, C. J. and E. S. Pearson (1934). "The use of confidence or fiducial limits
illustrated in the case of the binomial " Biometrika 26: 404-413
Dodson, K. and J. LeJeune (2005). "Escherichia coli O157:H7, Campylobacter jejuni,
and Salmonella Prevalence in cull dairy cows marketed in northeastern Ohio."
Journal of Food Protection 68(5): 927-931.
97
Elder, R. O., J. E. Keen, G. R. Siragusa, et al. (2000). "Correlation of enterohemorrhagic
Escherichia coli O157 prevalence in feces, hides, and carcasses of beef cattle
during processing." Proc Natl Acad Sci U S A 97(7): 2999-3003.
Fedorko, D. P. and E. C. Williams (1997). "Use of cycloserine-cefoxitin-fructose agar
and L-proline-aminopeptidase (PRO Discs) in the rapid identification of
Clostridium difficile." J Clin Microbiol 35(5): 1258-1259.
French, E., A. Rodriguez-Palacios and J. T. LeJeune "Enteric bacterial pathogens with
zoonotic potential isolated from farm-raised deer." Foodborne Pathog Dis 7(9):
1031-1037.
Garcia, A., J. Perez, A. Pulido, et al. (2000). "Evaluation of four rapid methods for the
investigation of the toxigenic capacity of Clostridium difficile strains isolated in a
selective medium." Enferm Infecc Microbiol Clin 18(3): 109-112.
Goorhuis, A., D. Bakker, J. Corver, et al. (2008). "Emergence of Clostridium difficile
infection due to a new hypervirulent strain, polymerase chain reaction ribotype
078." Clin Infect Dis 47(9): 1162-1170.
Indra, A., H. Lassnig, N. Baliko, et al. (2009). "Clostridium difficile: a new zoonotic
agent?" Wien Klin Wochenschr 121(3-4): 91-95.
Keel, K., J. S. Brazier, K. W. Post, et al. (2007). "Prevalence of PCR ribotypes among
Clostridium difficile isolates from pigs, calves, and other species." J Clin
Microbiol 45(6): 1963-1964.
Keessen, E. C., C. J. Donswijk, S. P. Hol, et al. (2010) "Aerial dissemination of
Clostridium difficile on a pig farm and its environment. Available from:
http://igitur-archive.library.uu.nl/student-theses/2011-0518-200326/artikel.doc.
Accessed July 21, 2011."
Lemee, L., A. Dhalluin, S. Testelin, et al. (2004). "Multiplex PCR targeting tpi (triose
phosphate isomerase), tcdA (Toxin A), and tcdB (Toxin B) genes for toxigenic
culture of Clostridium difficile." J Clin Microbiol 42(12): 5710-5714.
98
Medina-Torres, C. E., J. S. Weese and H. R. Staempfli (2011). "Prevalence of
Clostridium difficile in horses." Veterinary Microbiology.
Norman, K. N., R. B. Harvey, H. M. Scott, et al. (2009). "Varied prevalence of
Clostridium difficile in an integrated swine operation." Anaerobe 15(6): 256-260.
Petrie, A. and P. Watson (1999). Statistics for Veterinary and Animal Sciences. Oxford,
UK., Blackwell Science.
Pirs, T., M. Ocepek and M. Rupnik (2008). "Isolation of Clostridium difficile from food
animals in Slovenia." J Med Microbiol 57(Pt 6): 790-792.
Rodriguez-Palacios, A., M. Koohmaraie and J. T. LeJeune (2011a). "Prevalence,
Enumeration, and Antimicrobial Agent Resistance in Clostridium difficile in
Cattle at Harvest in the United States." J Food Prot October, In press.
Rodriguez-Palacios, A., C. Pickworth, S. Loerch, et al. (2011b). "Transient Fecal
Shedding and Limited Animal-to-Animal Transmission of Clostridium difficile by
Naturally Infected Finishing Feedlot Cattle." Appl Environ Microbiol 77(10):
3391-3397.
Rodriguez-Palacios, A., R. J. Reid-Smith, H. R. Staempfli, et al. (2009a). "Possible
seasonality of Clostridium difficile in retail meat, Canada." Emerg Infect Dis
15(5): 802-805.
Rodriguez-Palacios, A., H. R. Staempfli, T. Duffield, et al. (2007a). "Clostridium difficile
in retail ground meat, Canada." Emerg Infect Dis 13(3): 485-487.
Rodriguez-Palacios, A., H. R. Stampfli, T. Duffield, et al. (2006). "Clostridium difficile
PCR ribotypes in calves, Canada." Emerg Infect Dis 12(11): 1730-1736.
Rodriguez-Palacios, A., H. R. Stampfli, M. Stalker, et al. (2007c). "Natural and
experimental infection of neonatal calves with Clostridium difficile." Vet
Microbiol 124(1-2): 166-172.
99
Rupnik, M., B. Dupuy, N. F. Fairweather, et al. (2005). "Revised nomenclature of
Clostridium difficile toxins and associated genes." J Med Microbiol 54(Pt 2): 113117.
Sargeant, J. M., A. M. O‘Connor, I. A. Gardner, et al. (2010). "The REFLECT Statement:
Reporting Guidelines for Randomized Controlled Trials in Livestock and Food
Safety: Explanation and Elaboration†." Zoonoses and Public Health 57(2): 105136.
Simango, C. (2006b). "Prevalence of Clostridium difficile in the environment in a rural
community in Zimbabwe." Transactions of the Royal Society of Tropical
Medicine and Hygiene 100(12): 1146-1150.
Simango, C. and S. Mwakurudza (2008). "Clostridium difficile in broiler chickens sold at
market places in Zimbabwe and their antimicrobial susceptibility." Int J Food
Microbiol 124(3): 268-270.
Thakur, S., M. Putnam, P. R. Fry, et al. (2010). "Prevalence of antimicrobial resistance
and association with toxin genes in Clostridium difficile in commercial swine."
American Journal of Veterinary Research 71(10): 1189-1194.
Thitaram, S. N., J. F. Frank, S. A. Lyon, et al. (2011). "Clostridium difficile from Healthy
Food Animals: Optimized Isolation and Prevalence." J Food Prot 74(1): 130-133.
Thonnard, J., F. Carreer, V. Avesani, et al. (1996). "Toxin A detection on Clostridium
difficile colonies from 24-h cultures." Clin Microbiol Infect 2(1): 50-54.
Weese, J. S., R. J. Reid-Smith, B. P. Avery, et al. (2010). "Detection and characterization
of Clostridium difficile in retail chicken." Lett Appl Microbiol 50(4): 362-365.
Weese, J. S., L. Toxopeus and L. Arroyo (2006). "Clostridium difficile associated
diarrhoea in horses within the community: predictors, clinical presentation and
outcome." Equine Vet J 38(2): 185-188.
Wilcox, M. H. and W. N. Fawley (2000). "Hospital disinfectants and spore formation by
Clostridium difficile." Lancet 356(9238): 1324.
100
Zidaric, V., S. Beigot, S. Lapajne, et al. (2010). "The occurrence and high diversity of
Clostridium difficile genotypes in rivers." Anaerobe.
Zidaric, V., M. Zemljic, S. Janezic, et al. (2008b). "High diversity of Clostridium difficile
genotypes isolated from a single poultry farm producing replacement laying
hens." Anaerobe 14(6): 325-327.
101
Chapter 5 ― Prevalence, enumeration and antimicrobial resistance in Clostridium
difficile in cattle at harvest, USA
As Accepted in:
Rodriguez-Palacios, A., M. Koohmaraie, et al. (2011a). "Prevalence, Enumeration, and
Antimicrobial Agent Resistance in Clostridium difficile in Cattle at Harvest in the
United States." J Food Prot In press for October Issue.
Abstract
To assess the potential for food contamination with Clostridium difficile from
food animals, we conducted a cross-sectional fecal prevalence study in 944 randomly
selected cattle harvested at seven commercial meat processing plants, representing four
distant regions (median distance 1,500 km) of the United States. In total 944 animals
were sampled in summer 2008. C. difficile was isolated from 1.8% (17/944) of cattle,
with median fecal shedding concentration of 2.2 log10 CFU·g-1 (range = 1.6 to 4.8; 95%
CI = 1.6, 4.3). Toxigenic C. difficile isolates were recovered from only four cattle (0.4%).
One of these isolates was emerging PCR ribotype 078/toxinotype V. The remaining
toxigenic isolates were toxinotype 0, one of which was an isolate with resistance to
linezolid, clindamycin and moxifloxacin (E-test). All isolates were susceptible to
vancomycin, metronidazole and tigecycline, but the minimum inhibitory concentrations
against linezolid were as high as the highest reported values for human-derived isolates.
102
The source of the linezolid-clindamycin-moxifloxacin resistance in a toxigenic C difficile
isolate from cattle is uncertain. However, since the use of these three antimicrobials in
cattle is not allowed in North America, it is possible that resistance originated from an
environmental source, from other species where those antimicrobials are used, or
transferred from other intestinal bacteria. This study confirms that commercial cattle can
carry epidemiologically relevant C. difficile strains at the time of harvest, but the
prevalence at the time they enter the food chain is low.
Introduction
Clostridium difficile has been isolated from food animals and retail meats
(including ground beef) since 2006, raising concerns about potential zoonotic and
foodborne transmission of this pathogen (Rodriguez-Palacios et al. 2006; RodriguezPalacios et al. 2007). Despite the increasing number of published studies on the subject,
most research about C. difficile in food producing animals has focused on determining the
prevalence of fecal shedding in very young stock (Rodriguez-Palacios et al. 2006; Keel et
al. 2007; Pirs et al. 2008; Norman et al. 2009), an animal population that contributes
minimally (~0.5%) to the millions of pounds of meat processed for human consumption
in the United States every year (USDA 2010).
During processing of ground beef, carcass trimmings from mature animals are
mixed to achieve target fat contents between 3 and 30% in final products. While fat
trimmings often originate from cattle intentionally raised for beef production (hereinafter
referred to as fed cattle), the sources of leaner trimmings are frequently mature cattle
103
primarily used for breeding purposes or milk production (hereinafter referred to as mature
cattle) which can be domestic or imported (NCBA 2002). Both types of cattle represent
distinct production systems, may have different risks and prevalence of pathogens, and
therefore could differentially contribute to microbial contamination of ground beef
products.
At the time of harvest, fecal contamination of the hides and processing facilities
contaminate carcasses (Elder et al. 2000). Establishing the fecal prevalence of C. difficile
in cattle at harvest is thus a valuable step to assess the potential risk for food
contamination (carcass or meat products) in the early stages of food processing. This
study determined the prevalence and magnitude of fecal shedding of C. difficile in
commercial fed and mature cattle at harvest in the USA. Furthermore, we assessed the
public health relevance of recovered isolates by characterizing their toxin genes and
antimicrobial resistance patterns to six drugs of relevance in human medicine.
Materials and Methods
Meat processing plants.
We conducted a cross-sectional study of C. difficile in fed and mature cattle at the
time of harvest at seven commercial beef processing plants in four regions of the United
States. Establishments inspected by the United States Department of Agriculture that
were harvesting between 400 and 1,500 cattle per day were visited and sampled once
during summer (July–September) in 2008. Five of these participants specialized in
harvesting either fed or mature cattle, whereas the other two processed both animal types.
104
Sampling of animals.
Animals sampled were part of commercial lots scheduled for harvest the day of
sampling and were sampled during the normal harvest process. The geographical
distribution of the plants was deemed adequate to represent the diversity of cattle
production systems that ship animals for slaughter in the USA, including eastern, central,
western and southern regions. Three of the seven meat processing plants indicated that
animals arriving from Canada were also being harvested on the day of sampling. In the
processing line, sampling was conducted by dedicated personal working at the conveyor
belt immediately after each evisceration station. In plants where fed and mature cattle
were both harvested, the plant supervisor guided the sampling by indicating the
respective viscera to the sampling personnel. Based on systematic sampling principles,
animals were sampled at regular animal intervals to minimize potential clustering
associated with cattle groupings or lots. The intervals were calculated by dividing the
number of animals available for harvest the day of sampling over the estimated sample
size.
Starting with the first harvested animals, and continuing for 5-6 consecutive hours
(50-80% of required time to process all animals), 10 grams of fecal contents were
aseptically obtained through a 10-15 cm incision made on the dorsal wall of the rectum
about 20 cm from the anus. Fecal samples were stored in WhirlPak bags (Nasco, Fort
Atkinson, WI) and shipped refrigerated to The Ohio State University for C. difficile
testing. Within 24 h of collection, samples were coded and blindly processed. Aliquots
105
from the first two processing plants (west region) were frozen at -80oC to determine the
magnitude of fecal shedding (i.e., CFU enumeration) at a later date.
Enrichment culture and molecular typing.
Clostridium difficile prevalence was based on an enrichment protocol used in
previous studies (Arroyo et al. 2005; Rodriguez-Palacios et al. 2006; Rodriguez-Palacios
et al. 2007). Briefly, one gram of manually homogenized feces was enriched in 15 ml of
selective C. difficile broth (Arroyo et al. 2005) supplemented with 0.1% sodium
taurocholate (Sigma Aldrich, St Louis, MO) and Clostridium difficile supplement
SR0096 (Oxoid, Columbia, MD; broth concentrations of D-cycloserine and cefoxitin are
250 and 8 mg · L-1, respectively). Following anaerobic incubation at 37oC for 7 days
(Anaerobic chamber; CO2:H:N gas ratio of 10:10:80) and centrifugation of the broths
(7,000 x g, 10 min), the sediments were treated with 96% ethanol (1:1, v/v, 30 min),
centrifuged again to remove the supernatants, and plated onto 3-hour prereduced C.
difficile fructose agar (CM0601, Oxoid, Columbia, MD) supplemented with 7%
defibrinated horse blood (Quad Five, Ryegate, MT) and the cycloserine/cefoxitin
supplement. After 5 days of anaerobic incubation at 37oC, up to three suspect yellowgreen fluorescent (UV365) colonies were subcultured onto tryptic soy agar (Acumedia,
Lansing, MI) supplemented with 5% defibrinated sheep blood (Quad Five, Ryegate, MT).
Nonhemolytic isolates, positive for L-proline aminopeptidase activity (Pro Disc, Remel,
Lenexa, KS, USA) (Fedorko and Williams 1997) were stored at -80oC for molecular
confirmation.
106
Frozen isolates were revived on blood agar and after 48 hours of anaerobic
incubation single colonies were processed for DNA extraction using a commercial chelex
resin (InstaGene Matrix, Bio-Rad, Hercules, CA). Clostridium difficile was molecularly
characterized by using three multiplex PCR reactions. One multiplex confirmed C.
difficile by detecting the triose-phosphate isomerase (tpi) gene and toxin genes tcdA and
tcdB (Lemee et al. 2004), the second by detecting a typical 16S rRNA gene amplicon and
toxin genes tcdA, tcdB, cdtA and cdtB (Persson et al. 2008); the third multiplex assessed
the presence of the putative toxin regulator gene tcdC using tcdE as a referent amplicon
(Spigaglia and Mastrantonio 2002; French et al. 2010). PCR ribotyping and visual
comparison of resulting fingerprint patterns were used to assess strain similarity among
cattle types and regions (Bidet et al. 1999). In addition, toxinotyping (i.e., restriction
fragment length polymorphism of the pathogenicity locus) was conducted on toxin gene
fragments A3 and B1 as described by Rupnik (http://www.mf.uni-mb.si/mikro/tox/).
Isolates were deemed toxigenic if they had at least one of the tcdA, tcdB, cdtA/B toxin
genes and associated toxins (Rupnik et al. 2005).
Toxin production.
In addition to identifying the toxin genes, representative isolates were tested with
two complementary commercial ELISAs to assess their toxigenic potential as previously
described (Thonnard et al. 1996; Garcia et al. 2000) using 72-hour culture colonies.
Toxin gene testing was conducted at least in duplicate using reference toxigenic strain
ATCC 9569 (tcdA+/tcdB+/cdtAB-/toxinotype 0) as positive control. The production of C.
difficile toxins A and B was confirmed by interpreting in parallel ELISA results from
107
commercial kits C. difficile TOX A/B II (TechLab, Blacksburg, VA) and Clearview C. diff
A (Wampole Laboratories, Inverness Medical Professional Diagnostics, Princeton, NJ).
Binary toxin production was not evaluated.
Dialysis tubing and toxin cytotoxicity assay on Vero cells.
To confirm the lack of toxin production by strains identified as toxin-negative by
PCR and ELISA, selected isolates were tested in dialysis assays (Sullivan et al. 1982)
using toxigenic strain ATCC 9569 and nontoxigenic C difficile PCR ribotype M26
(Rodriguez-Palacios et al. 2007) as controls. In brief, 48h C difficile cultures, grown on
blood agar were anaerobically resuspended in antibiotic-free brain heart infusion (BHI)
broth supplemented with 0.5% yeast extract (Acumedia, Lansing, MI) at an OD600 of 0.3.
Ten milliliters of those suspensions were placed inside 6 mm diameter pretreated dialysis
tubing (Spectra/Por, 12-14,000 MWCO, Spectrum Laboratories, Rancho Dominguez,
CA), which were submerged in 500 ml of BHI-yeast broth and anaerobically incubated at
37oC for 72 hours. At the end of the assay, the broths within the tubing were harvested,
centrifuged (10,000 x g, 5 minutes) and filtered (0.2 µm) for sterile testing on 90%
confluent monolayer cultures of Vero cells line. Vero cells were grown using D-MEM
high glucose medium (Gibco, Carlsbad, CA) supplemented with penicillin and
streptomycin (100U and 0.1mg ml-1, respectively; Sigma Aldrich, St Louis, MO) and
10% fetal bovine serum (Atlas Biologicals, Fort Collins, CO), but tested for cytotoxin
activity only with the medium (D-MEM with antibiotics; 80 µl), 1% fetal bovine serum
and 20 µL of the toxin filtrate. Cytotoxicity (cell roundening) of two-fold serial dilutions
was quantified 24h after toxin inoculation in duplicate using 96-well microtiter plates
108
(Corning Life Sciences, USA). A virologist blindly assessed the plates using a scoring
system with determined kappa agreement between ratings greater than 0.95.
Antimicrobial susceptibility.
Commercial E-test strips (AB Biodisk, Solna, Sweden) were used to determine
the minimum inhibitory concentrations (MICs) against six antimicrobial classes of
clinical relevance in humans: metronidazole and vancomycin (first choice treatments for
C. difficile infections, CDI) (Cohen et al.), moxifloxacin and clindamycin (widely
associated with CDI induction) (Biller et al. 2007; Cohen et al.), and linezolid and
tigecycline (newer therapeutic classes against CDI) (Rello 2005; Jones et al. 2009).
Single isolates from 7 fed and 7 mature cattle representing all PCR ribotypes were tested.
Briefly, 24 hour old C. difficile colonies grown on blood agar and resuspended (inside
anaerobic chamber) in Brucella broth (Oxoid, Columbia, MD) were used at an OD600
0.23 ± 0.01 to lawn 6-hour prereduced Brucella agar, where the six E-test strips were
seeded. Plates were read after 48 hours of anaerobic incubation at 37°C following the
manufacturer‘s recommendations. Minimal inhibitory concentrations and interpretive
categories for metronidazole, clindamycin, and moxifloxacin derived from the Clinical
and Laboratory Standards Institute, while the values for vancomycin, linezolid and
tigecycline were from previously published studies (Rodriguez-Palacios et al. 2006;
Hecht and Citron 2007; Jones et al. 2009; Nagy and Dowzicky 2010).
Magnitude of fecal shedding.
The quantification of C. difficile shedding (log10 CFU·g-1 of feces) was
determined using the standard spread plate method. Six available frozen samples that
109
were C. difficile positive by the enrichment method and three negative samples were
enumerated. In brief, duplicate 2 g aliquots of feces homogenized with phosphate
buffered saline (PBS, 1:1, vol/vol) were split in two subsamples, one for direct plating
and one for ethanol shock. Ethanol was used to inactivate vegetative fecal contaminants
to facilitate enumeration of C. difficile spores (Arroyo et al. 2005). Then, 100 to 10-6
dilutions were spread plated onto blood and C. difficile selective agars. After 72 h of
incubation, C. difficile-like colonies were enumerated, subcultured onto blood agar, and
confirmed with a multiplex PCR (Persson et al. 2008) and PCR ribotyping (Bidet et al.
1999). A reference strain of C. difficile PCR ribotype 078 was used as control to estimate
the detection thresholds of our enrichment and spread plate methods. For this purpose,
duplicates from three samples were spiked with spores of PCR ribotype 078 (3.4 log10
CFU·g-1) that had been aged for four weeks in PBS.
Statistics.
Assuming simple random sampling, 120 animals were deemed adequate to test a
hypothesized C. difficile prevalence of 5.5 (±6%) for each cattle type at harvest (onesample, α = 0.05, power = 0.8). Clostridium difficile culture binary data was analyzed
using logistic regression with cattle type (fed vs. mature) and region as predictors.
Fisher‘s exact statistics tested proportions with zero counts (Petrie and Watson 1999).
Confidence intervals of binomials were estimated using exact statistics (Clopper and
Pearson 1934). Nonparametric statistics and significance values for log10 CFU·g-1 data
were adjusted for ties (Minitab-15, State College, PA; STATA-10.1, College Station,
110
TX). Geometrical means of MIC values (Wilcox et al. 2000) were used for descriptive
statistics.
Results
Prevalence.
In total, single fecal samples from 944 animals were tested for C. difficile
(118±12 in each processing plant). Using the enrichment method, 29 C. difficile isolates
were recovered from 17 fecal samples (1.8%); 10 from fed cattle and 7 from mature cattle
(Fig. 9). The proportion of C. difficile shedders was comparable for all plants and regions,
and no differences were observed between fed and mature cattle controlling for region
(odds ratio of shedding in mature cattle = 0.8; 95% CI = 0.03, 2.1; Logistic P = 0.596).
Toxigenic C. difficile was recovered from four (24%) of the 17 shedding animals;
although these four animals were all fed cattle, there was not statistically difference when
compared to the shedding prevalence in mature cattle (Fisher‘s exact, P = 0.125; Table
4).
Genotypes.
PCR ribotyping identified most isolates had unique fingerprints except for a tcdA/tcdB-/cdtAB- PCR ribotype (toxins A and B not detectable) that was common in the west
and central regions (superscript a in Fig. 9). Dialysis tubing assay confirmed this strain
was nontoxigenic (control ATCC strain cytotoxic titer was 1:1024). Emerging PCR
ribotype 078/toxinotype V was identified in the east in one of the four fed cattle shedding
toxigenic C. difficile. The other three toxigenic isolates were toxinotype 0, each had
unique PCR ribotype fingerprints. Toxin fragments A3 and B1 did not amplified in the
111
remaining isolates which were tcdA-/tcdB-/cdtAB-. Simultaneous shedding of two C.
difficile PCR ribotypes was identified in 3 of 17 shedders (18%). Two of these three
animals shed simultaneously two PCR ribotypes which were negative to the toxins and
toxin genes tested. In the remaining animal that shed two PCR ribotypes, one isolate was
toxigenic, the other had no toxin genes or toxin detectable.
Antimicrobial susceptibility.
All isolates tested were susceptible to metronidazole and vancomycin but manyfold more susceptible to tigecycline (Fig. 10, Table 5). In contrast, 93% of isolates had
intermediate or resistant MICs for clindamycin. Only one isolate (7%) was resistant to
moxifloxacin. The MICs for linezolid were comparable to moxifloxacin and
metronidazole but were among the highest values reported for C. difficile isolates of
human origin (Ackermann et al. 2003; Jones et al. 2009). One of the toxigenic isolates
(toxinotype 0) was multidrug resistant and had the highest measurable MICs for
clindamycin and moxifloxacin and among the highest MICs reported for linezolid (Fig.
10). Using the actual MIC values, the cut offs for resistance, and the geometrical means
as previously reported (Wilcox et al. 2000), C. difficile isolates from fed and mature
cattle had comparable levels of susceptibility to the drugs tested (P > 0.05). However, C.
difficile from fed cattle had more variability, due to high MICs in toxigenic isolates
(Bartlett‘s test controlling for cattle type and strain toxinotype, P < 0.001; Fig. 10).
Magnitude of shedding.
Analysis of enumerated plate series where C. difficile was confirmed (n = 28)
indicated that the median of shedding of C. difficile at the time of harvest was 2.2 log10
112
CFU·g-1 of feces, but in some cases it reached almost 5 log10 units (range = 1.6 to 4.8;
95% CI = 1.6, 4.3). When duplicate samples were spiked with spores from strain PCR
ribotype 078, our spread plate-serial dilution method recovered a slightly lower number
of spores than inoculated (median = 0.45 log10 CFU·g-1 lower; 95.7% CI = 0.076, 1.283;
Mann-Whitney, P = 0.083; Fig. 11). In contrast, direct plating of PBS suspensions
containing PCR ribotype 078 yielded 1.2 log10 units more C. difficile than matched
samples treated with ethanol shock (95% CI = 0.9-1.5; Mann-Whitney, P = 0.011; Fig.
11).
Discussion
This enrichment-based study indicate that the prevalence of fecal shedding of
toxigenic C. difficile in cattle at harvest (0.4%; 4/944) is in agreement with a recent
longitudinal study conducted in a naturally infected feedlot (0.6% at harvest) (RodriguezPalacios et al. 2011), but it is much lower than the prevalence reported in calves at the
farm (7-13%) and upon entry to veal calf operations (33-36%) (Costa et al. 2009). It is
however important to note that the present study was conducted in summer when C.
difficile shedding in neonatal calves (and possible the extent and susceptibility to
environmental contamination) is the lowest (Rodriguez-Palacios et al. 2006).
The toxigenic C. difficile isolates here identified correspond to important
toxinotypes associated with CDI in humans, toxinotype 0 and V. PCR ribotype
078/toxinotype V is an emerging C. difficile genotype increasing its participation in
human disease and its presence in food animals and foods (Keel et al. 2007; Rupnik et al.
113
2008; Songer et al. 2009). The isolation of emerging PCR ribotype 078 from one of our
fed cattle in one of the meat processing plants in the east is of epidemiological interest.
This genotype has also been isolated from dairy cattle in Ontario, Canada (RodriguezPalacios et al. 2006), and from finishing steers (Rodriguez-Palacios et al. 2011) and
farmed deer in Ohio, USA (French et al. 2010), indicating that PCR ribotype 078 might
have the potential for geographical distribution across farming operations in the east. Of
complementary interest, the most common genotype identified in meat plants of the west
and central regions was a nontoxigenic PCR ribotype (superscript ‗a‘ in Fig 1A) that has
not been reported in our studies on farm animals in the east. Understanding the factors
associated with the dissemination of C. difficile in livestock and the shedding prevalence
(toxigenic:nontoxigenic), which can vary along the production system, is important
because nontoxigenic C. difficile has been proposed as a probiotic strategy to prevent C.
difficile colonization in animals and people (Shim et al. 1998; Songer et al. 2007).
The magnitude of C. difficile shedding ranged from about 1 to 5 log10 CFU·g-1 of
feces. This shedding variability could represent actual temporal variation in vivo, local
heterogeneity of bacterial concentration within the fecal samples (Pearce et al. 2004), or
simply the variability of enumeration methods. In our study, ethanol shock of control
strain PCR ribotype 078 spores yielded fewer spores than with direct plating. This
indicates that our enumeration of spores in fecal samples is likely an underestimation of
the real C. difficile shedding in cattle. Recently, a study of culture methods by Thitaram
et al. (2011) comparing single versus double alcohol shock treatments as part of an
enrichment protocol similar to ours demonstrated differential performance across food
114
animal production systems, and that single alcohol shock (used in our study) was best to
concurrently assess the prevalence of C. difficile in dairy and beef cattle feces (Thitaram
et al. 2011). Comparing the observed prevalence of C. difficile in both studies for single
alcohol shock and plating onto TCCFA agar results, there were no differences for mature
cattle (7/462 [1.5%] vs. 19/1,325 [1.4%]; Chi square P = 0.9). However, C. difficile was
significantly more prevalent in beef cattle in Thitaram‘s study (10/482 [2.1%] vs.
157/2965 [5.3%]; Chi square P = 0.002). Collectively, both studies indicate that fecal
shedding of C. difficile in beef feedlot cattle may be more subject to significant regional
or production variability than in mature cattle.
It is uncertain if fecal shedding of large numbers of C. difficile persists over time
or if it is an oscillating pattern (Chase-Topping et al. 2008). Periods of fecal shedding of
greater than 3 or 4 log10 units have been referred by some as ―supershedding‖ status, a
determinant for infectiousness (Chase-Topping et al. 2008). Supershedding has recently
been described for C. difficile in response to antimicrobial therapy in mice (Lawley et al.
2009), but the factors associated with the magnitude of fecal shedding in cattle and its
causal role in food contamination are unknown. The identification of up to 4-5 log10
CFU·g-1 of feces indicates that small amounts of feces could be sufficient to contaminate
meat processing facilities, equipment, beef carcasses, and processed foods.
There is an important contrast between this and other studies conducted at harvest
(Rodriguez-Palacios et al. 2011) and studies with similar culture methods conducted on
retail meat products: the low prevalence of C. difficile shedding at harvest (<1.2%) seems
remarkably different from the much higher proportion of C. difficile contamination
115
observed in retail products (6-42%) (Rodriguez-Palacios et al. 2009; Songer et al. 2009).
Identifying post-harvest factors that could contribute to product contamination, and
proliferation, of C. difficile within foods is important, especially because emerging strains
can survive minimum cooking recommendations for ground meats (Rodriguez-Palacios
et al. 2010; Rodriguez-Palacios and LeJeune 2011).
In contrast to other reports of cattle- and meat-derived C difficile isolates where
100% of fluoroquinolone resistance was observed (Rodriguez-Palacios et al. 2006;
Rodriguez-Palacios et al. 2007), all our isolates, except one, were susceptible to
moxifloxacin. The regulations in the USA restrict fluoroquinolone usage in food animals
(FDA 2009; FDA 2009; FARAD 2011). Enrofloxacin, the only fluoroquinolone
permitted, is restricted to treatment of nonresponsive respiratory disease in i) beef cattle
and ii) nonlactating replacement dairy heifers under 20 months of age. Explicitly,
enrofloxacin is prohibited for dairy cattle or calves intended for production of beef (FDA
2002), although noncompliance might occur (FDA 2009; FDA 2009).
Purposely, we chose tigecycline and linezolid, which are newer drug classes with
emerging potential for treatment of CDI, to contribute to the general understanding of
susceptibility in C. difficile. These drugs exhibit marked inhibitory activity against
bacterial strains resistant to other antimicrobial classes, including fluoroquinolones.
Because these antimicrobials are relatively new, and not used in cattle in North America,
it would be unlikely that resistance in cattle-derived isolates could be attributable to the
use of such medications on farms. The isolates recovered in this study displayed the same
high susceptibility documented for tigecycline in global surveillance programs (Jones et
116
al. 2007; Nagy and Dowzicky 2010). But high MICs of linezolid were required to inhibit
most C. difficile isolates. The MICs found here are comparable to the highest MICs
reported for human-derived isolates (Wise et al. 1998). Of public health relevance, the
isolate that exhibited the highest MIC for linezolid (8 µg · ml-1) was also highly resistant
to clindamycin and moxifloxacin and was toxigenic. Although clindamycin is not
approved for use in food animals (FDA 2007), most isolates had intermediate or resistant
MICs to clindamycin.
The source of the linezolid-clindamycin-moxifloxacin resistance in the
abovementioned toxigenic C difficile isolate is uncertain. However, since the use of these
antimicrobials, particularly linezolid, in commercial cattle for harvest is unlikely in North
America, it is possible that this resistant strain originated from an environmental source,
or indirectly from species where those antimicrobials are more commonly used, including
humans. Transfer of resistance from other enteric bacteria (e.g., Enterococcus spp.)
where intrinsic resistance may occur (particularly linezolid) (Fluckey et al. 2009; Zhang
et al. 2010) is another possibility.
In summary, this study indicates that live cattle may be a source of C. difficile
contamination of meat processing facilities at the time of harvest, but the prevalence of
cattle harboring this pathogen at the time they enter the food chain is low. Of
epidemiological concern, two common toxigenic genotypes of C. difficile (toxinotype 0
that in addition was linezolid-clindamycin-moxifloxacin resistant and emerging
toxinotype V/PCR ribotype 078 which has been increasingly isolated from animals and
foods) were identified in cattle at harvest. Understanding how nontoxigenic C. difficile
117
modulate fecal shedding of toxigenic strains in various steps of food production might
provide insight into prevention strategies to reduce the risk of food contamination.
Acknowledgements
This study was funded by the Beef Checkoff. ARP and JTL were supported by
state and federal funds allocated to the Ohio Agricultural Research and Development
Center, Ohio State University, Wooster, OH. ARP is a research fellow of the Public
Health Preparedness for Infectious Diseases (Targeted Investments in Excellence)
Program, The Ohio State University.
Authors are grateful to Mike Kauffman, Jennifer Schrock, Laura Harpster, and
Pamela Schlegel for invaluable technical assistance during sample collection, and
Michele Williams for editorial assistance during final preparation of the manuscript.
Special thanks go to the participating meat processing plants and Maria Murgia for her
splendid collaboration during dialysis and cytotoxicity assays.
118
A
Clostridium difficile in cattle at harvest
Region
Plants
1
41
61
81
101
a
121
21
41
61
81
101
121
1
0
e
1
0
c
g
0.8%
1/120
East
d
f
4.4%
4/90
Central
1
1
0
21
1
21
41
61
81
101
121
1
South
aa
1
*
21
41
61
a, 81
*
101
1
0
1
1
21
41
81
101
**
21
41
61
0.8%
1/122
West
121
2.5%
3/120
Central
0
1
0
0.7%
1/130
1.5% (7/462)
West
1
10
19
28
37
46
55
64
73
82
91
100
109
118
127
Clostridium difficile
(positive = 1; negative = 0)
c
a
Mature cattle
b
1
0
aa a
2.1% (10/482)
1
0
Fed beef cattle
a
121
2.5%
3/120
East
61
81
101
121
2.5%
3/120
South
1
21
41
61
81
101
121
0.8%
1/122
Sample collection order
B
Driving distance between plants - km
West
Central
East
South
West
320
-
Central
2,600
50
-
East
4,500
1,900
0
-
South
3,500
1,500
2,700
450
1
Figure 9. Clostridium difficile in bovine feces collected from the colon of cattle
processed at meat processing/packing plants, USA. A) Sequential distribution of positive
animals (vertical bars) at harvest. Only isolates sharing identical PCR ribotypes are
indicated (superscript letter a), the remaining isolates had unique ribotypes. Asterisks
indicate toxigenic C. difficile. B) Driving distances between plants among and within
(black cells) regions.
119
● Fed beef cattle
3
Toxinotype 0
PCR-type I-574
○ Mature cattle
421
00.5 0.25 0.125 -1
0.064 0.032 0.016 -2 0.008
Inter. Res.
128
264 32 16 18
b
Data
a
Me_log
Va-Log
Susceptible
Antimicrobial in E-test strips (µg·ml-1)
256 -
Liz_log
tig_1
Metronidazole Vancomicyin Linezolid Tigecycline
Common treatment
New classes
Cli_1
mox_1
scale_log
Clindamycin Moxifloxacin
Associated with CDI
Figure 10. Minimum inhibitory concentrations (MICs) of cattle-derived C. difficile
isolates to six antimicrobial classes of relevance for human C. difficile infections (CDI).
Susceptibility zones within antibiotic ranges tested are illustrated. Vertical bars indicate
published linezolid MICs ranges for human isolates; superscript ‗a‘ for ref (Ackermann et
al. 2003) and ‗b‘ for ref (Wise et al. 1998).
120
Individual Value Plot of negative, positive/g, spiked/g, inocula/g
95% Bonferroni CI for the Mean
M-W, P < 0.001
5
M-W, P = 0.083
· g -1
logLog
10 cfu
10 CFU
4
3
2
1
0
Negative
negative
n=3
Positive
positive/g
n=6
Original f rozen samples
Spiked
spiked/g
n=3
Inoculum
inocula/g
PCR-078
Controls
Figure 11. Magnitude of C. difficile fecal shedding in cattle at harvest.
Bars and 95% Bonferroni confidence intervals of the means; M-W, Mann-Whitney
statistics.
121
Table 4.Estimated prevalence of cattle carrying C. difficile at harvest, USA, summer
2008.a,b
95% CI of estimated shedding prevalence by strain type (n)
Cattle tested
Toxigenic strains
‘Nontoxigenic’ strains
Totalc
Mature
n=462
0 to 0.8% (0)
0.6 to 3.1% (7)
0.6 to 3.1% (7)
Fed
n=482
0.2 to 2.1% (4)
0.6 to 3.0% (7)
1.0 to 3.8% (10)
Total
n=944
0.1 to 1.1% (4)
0.8 to 2.5% (14)
1.1 to 2.9% (17)
P = 0.12
P = 1.0
P = 0.6
Fed vs. Mature
a
Two-sided 95% exact confidence intervals (CI); one-sided 97.5% CI reported for zero
counts.
b
Relevance: toxigenic strains cause CDI in humans and animals (Rodriguez-Palacios et
al. 2006; Hammitt et al. 2008; Rupnik et al. 2008), but nontoxigenic strains might
prevent it (Shim et al. 1998; Songer et al. 2007).
c
One fed animal simultaneously shed a toxigenic and a toxin-gene negative strain.
122
Table 5. Antimicrobial susceptibility of C. difficile from cattle at harvest.a
Antibiotic
MIC50
MIC90
Maximum
Metronidazole
1.0
1.5
1.5
Vancomycin
0.75
0.75
1.5
Linezolid
1.5
2.0
8
Tigecycline
0.036
0.064
0.5
Clindamycin
6
12
256
Moxifloxacin
1.0
1.5
32
a
determined with the E-test method (µg ∙ ml-1)
123
References
Ackermann, G., D. Adler and A. C. Rodloff (2003). "In vitro activity of linezolid against
Clostridium difficile." J Antimicrob Chemother 51(3): 743-745.
Arroyo, L. G., J. Rousseau, B. M. Willey, et al. (2005b). "Use of a selective enrichment
broth to recover Clostridium difficile from stool swabs stored under different
conditions." J Clin Microbiol 43(10): 5341-5343.
Bidet, P., F. Barbut, V. Lalande, et al. (1999). "Development of a new PCR-ribotyping
method for Clostridium difficile based on ribosomal RNA gene sequencing."
FEMS Microbiology Letters 175(2): 261-266.
Biller, P., B. Shank, L. Lind, et al. (2007). "Moxifloxacin therapy as a risk factor for
Clostridium difficile-associated disease during an outbreak: attempts to control a
new epidemic strain." Infect Control Hosp Epidemiol 28(2): 198-201.
Chase-Topping, M., D. Gally, C. Low, et al. (2008). "Super-shedding and the link
between human infection and livestock carriage of Escherichia coli O157." Nat
Rev Micro 6(12): 904-912.
Clopper, C. J. and E. S. Pearson (1934). "The use of confidence or fiducial limits
illustrated in the case of the binomial " Biometrika 26: 404-413
Cohen, S. H., D. N. Gerding, S. Johnson, et al. (2010). "Clinical practice guidelines for
Clostridium difficile infection in adults: 2010 update by the society for healthcare
epidemiology of America (SHEA) and the infectious diseases society of America
(IDSA)." Infect Control Hosp Epidemiol 31(5): 431-455.
Costa, M. C., L. Arroyo, D. L. Pearl, et al. (2009). "Incidence of Clostridium difficile in
feces of male Holstein veal calves. American College of Veterinary Internal
Medicine Forum. http://www.acvim.org/websites/acvim/File/Costa.pdf."
Retrieved 15 March 2011.
124
Elder, R. O., J. E. Keen, G. R. Siragusa, et al. (2000). "Correlation of enterohemorrhagic
Escherichia coli O157 prevalence in feces, hides, and carcasses of beef cattle
during processing." Proc Natl Acad Sci U S A 97(7): 2999-3003.
FARAD. (2011). "Food Animal Residue Avoidance and Depletion Program. Prohibited
Drug List " Retrieved 15 March 2011, from
http://www.farad.org/eldu/prohibit.html.
FDA. (2002). "September 19. Reminder - Extra-label use of fluoroquinolones prohibited.
United States Food and Drug Administration." Retrieved 31 August 2010, from
http://www.fda.gov/AnimalVeterinary/NewsEvents/CVMUpdates/ucm129354.ht
m.
FDA. (2007). "March 19. Clindamycin hydrochloride oral drops (for use in dogs and
cats). United States Food and Drug Administration. ANADA 200-398 original
approval." May 20 2011, from
http://www.fda.gov/downloads/AnimalVeterinary/Products/ApprovedAnimalDru
gProducts/FOIADrugSummaries/ucm061348.pdf.
FDA. (2009a). "November 18. United States Food and Drug Administration takes action
against dairy farm and owner - Farm selling animals with illegal drug residues in
violation of federal law." FDA NEWS RELEASE, May 20 2011, from
http://www.fda.gov/NewsEvents/Newsroom/PressAnnouncements/2009/ucm1912
32.htm.
FDA. (2009b). "November 19. United States Food and Drug Administration takes action
against Maryland veal calf dealer - Man allegedly sold animals for human
consumption that contained illegal drug residues." May 20 2011, from
http://www.fda.gov/NewsEvents/Newsroom/PressAnnouncements/2009/ucm1914
46.htm.
Fedorko, D. P. and E. C. Williams (1997). "Use of cycloserine-cefoxitin-fructose agar
and L-proline-aminopeptidase (PRO Discs) in the rapid identification of
Clostridium difficile." J Clin Microbiol 35(5): 1258-1259.
Fluckey, W. M., G. H. Loneragan, R. D. Warner, et al. (2009). "Diversity and
Susceptibility of Enterococcus Isolated from Cattle before and after Harvest."
Journal of Food Protection 72: 766-774.
125
French, E., A. Rodriguez-Palacios and J. T. LeJeune (2010). "Enteric bacterial pathogens
with zoonotic potential isolated from farm-raised deer." Foodborne Pathog Dis
7(9): 1031-1037.
Garcia, A., J. Perez, A. Pulido, et al. (2000). "Evaluation of four rapid methods for the
investigation of the toxigenic capacity of Clostridium difficile strains isolated in a
selective medium." Enferm Infecc Microbiol Clin 18(3): 109-112.
Hammitt, M. C., D. M. Bueschel, M. K. Keel, et al. (2008). "A possible role for
Clostridium difficile in the etiology of calf enteritis." Vet Microbiol 127(3-4):
343-352.
Hecht, D. W. and D. Citron (2007). "Clinical Laboratory Standards Institute. M11-A7Methods for antimicrobial susceptibility testing of anaerobic bacteria - Seven
Edition." 27(2): 64p.
Jones, R. N., T. R. Fritsche, H. S. Sader, et al. (2007). "LEADER surveillance program
results for 2006: an activity and spectrum analysis of linezolid using clinical
isolates from the United States (50 medical centers)." Diagn Microbiol Infect Dis
59(3): 309-317.
Jones, R. N., S. Kohno, Y. Ono, et al. (2009). "ZAAPS International Surveillance
Program (2007) for linezolid resistance: results from 5591 Gram-positive clinical
isolates in 23 countries." Diagn Microbiol Infect Dis 64(2): 191-201.
Keel, K., J. S. Brazier, K. W. Post, et al. (2007). "Prevalence of PCR ribotypes among
Clostridium difficile isolates from pigs, calves, and other species." J Clin
Microbiol 45(6): 1963-1964.
Lawley, T. D., S. Clare, A. W. Walker, et al. (2009). "Antibiotic Treatment of
Clostridium difficile Carrier Mice Triggers a Supershedder State, Spore-Mediated
Transmission, and Severe Disease in Immunocompromised Hosts." Infect.
Immun. 77(9): 3661-3669.
Lemee, L., A. Dhalluin, S. Testelin, et al. (2004). "Multiplex PCR targeting tpi (triose
phosphate isomerase), tcdA (Toxin A), and tcdB (Toxin B) genes for toxigenic
culture of Clostridium difficile." J Clin Microbiol 42(12): 5710-5714.
126
Nagy, E. and M. J. Dowzicky (2010). "In vitro activity of tigecycline and comparators
against a European compilation of anaerobes collected as part of the Tigecycline
Evaluation and Surveillance Trial (TEST)." Scand J Infect Dis 42(1): 33-38.
NCBA (2002). "April 18. National Cattlemen's Beef Association. News archive:
Increased Export of U.S. Trimmings Could Benefit McDonald‘s, Help Improve
Value for U.S. Producers.
http://www.beefusa.org/NEWSIncreasedExportofUSTrimmingsCouldBenefitMc
DonaldsHelpImproveValueforUSProducers4927.aspx." September
14(September 14).
Norman, K. N., R. B. Harvey, H. M. Scott, et al. (2009). "Varied prevalence of
Clostridium difficile in an integrated swine operation." Anaerobe 15(6): 256-260.
Pearce, M. C., D. Fenlon, J. C. Low, et al. (2004). "Distribution of Escherichia coli O157
in Bovine Fecal Pats and Its Impact on Estimates of the Prevalence of Fecal
Shedding." Appl. Environ. Microbiol. 70(10): 5737-5743.
Persson, S., M. Torpdahl and K. E. Olsen (2008). "New multiplex PCR method for the
detection of Clostridium difficile toxin A (tcdA) and toxin B (tcdB) and the binary
toxin (cdtA/cdtB) genes applied to a Danish strain collection." Clin Microbiol
Infect 14(11): 1057-1064.
Petrie, A. and P. Watson (1999). Statistics for Veterinary and Animal Sciences. Oxford,
UK., Blackwell Science.
Pirs, T., M. Ocepek and M. Rupnik (2008). "Isolation of Clostridium difficile from food
animals in Slovenia." J Med Microbiol 57(Pt 6): 790-792.
Rello, J. (2005). "Pharmacokinetics, pharmacodynamics, safety and tolerability of
tigecycline." J Chemother 17 Suppl 1: 12-22.
Rodriguez-Palacios, A. and J. T. LeJeune (2011). "Moist-heat resistance, spore aging,
and superdormancy in Clostridium difficile." Appl. Environ. Microbiol. 77(9):
3085-3091.
127
Rodriguez-Palacios, A., C. Pickworth, S. Loerch, et al. (2011b). "Transient Fecal
Shedding and Limited Animal-to-Animal Transmission of Clostridium difficile by
Naturally Infected Finishing Feedlot Cattle." Appl Environ Microbiol 77(10):
3391-3397.
Rodriguez-Palacios, A., R. J. Reid-Smith, H. R. Staempfli, et al. (2009a). "Possible
seasonality of Clostridium difficile in retail meat, Canada." Emerg Infect Dis
15(5): 802-805.
Rodriguez-Palacios, A., R. J. Reid-Smith, H. R. Staempfli, et al. (2010). "Clostridium
difficile survives minimal temperature recommended for cooking ground meats."
Anaerobe 16: 540-542.
Rodriguez-Palacios, A., H. R. Staempfli, T. Duffield, et al. (2007a). "Clostridium difficile
in retail ground meat, Canada." Emerg Infect Dis 13(3): 485-487.
Rodriguez-Palacios, A., H. R. Stampfli, T. Duffield, et al. (2006). "Clostridium difficile
PCR ribotypes in calves, Canada." Emerg Infect Dis 12(11): 1730-1736.
Rodriguez-Palacios, A., H. R. Stampfli, M. Stalker, et al. (2007c). "Natural and
experimental infection of neonatal calves with Clostridium difficile." Vet
Microbiol 124(1-2): 166-172.
Rupnik, M., B. Dupuy, N. F. Fairweather, et al. (2005). "Revised nomenclature of
Clostridium difficile toxins and associated genes." J Med Microbiol 54(Pt 2): 113117.
Rupnik, M., A. Widmer, O. Zimmermann, et al. (2008). "Clostridium difficile toxinotype
V, ribotype 078, in animals and humans." J Clin Microbiol 46(6): 2146.
Shim, J. K., S. Johnson, M. H. Samore, et al. (1998). "Primary symptomless colonisation
by Clostridium difficile and decreased risk of subsequent diarrhoea." Lancet
351(9103): 633-636.
Songer, J. G., R. Jones, M. A. Anderson, et al. (2007). "Prevention of porcine
Clostridium difficile-associated disease by competitive exclusion with
nontoxigenic organisms." Vet Microbiol 124(3-4): 358-361.
128
Songer, J. G., H. T. Trinh, G. E. Killgore, et al. (2009). "Clostridium difficile in retail
meat products, USA, 2007." Emerg Infect Dis 15(5): 819-821.
Spigaglia, P. and P. Mastrantonio (2002). "Molecular analysis of the pathogenicity locus
and polymorphism in the putative negative regulator of toxin production (TcdC)
among Clostridium difficile clinical isolates." J Clin Microbiol 40(9): 3470-3475.
Sullivan, N. M., S. Pellett and T. D. Wilkins (1982). "Purification and characterization of
toxins A and B of Clostridium difficile." Infect. Immun. 35(3): 1032-1040.
Thitaram, S. N., J. F. Frank, S. A. Lyon, et al. (2011). "Clostridium difficile from Healthy
Food Animals: Optimized Isolation and Prevalence." J Food Prot 74(1): 130-133.
Thonnard, J., F. Carreer, V. Avesani, et al. (1996). "Toxin A detection on Clostridium
difficile colonies from 24-h cultures." Clin Microbiol Infect 2(1): 50-54.
USDA. (2010, September 14). "Estimated red meat production. CattleFax and United
States Department of Agriculture reports: Livestock slaughter and meat
production." Retrieved May 14 2010, from http://www.cattlefax.com/quicklinksdata/usda-reports/.
Wilcox, M. H., W. Fawley, J. Freeman, et al. (2000). "In vitro activity of new generation
fluoroquinolones against genotypically distinct and indistinguishable Clostridium
difficile isolates." J Antimicrob Chemother 46(4): 551-556.
Wise, R., J. M. Andrews, F. J. Boswell, et al. (1998). "The in-vitro activity of linezolid
(U-100766) and tentative breakpoints." J Antimicrob Chemother 42(6): 721-728.
Zhang, J., S. K. Wall, L. Xu, et al. (2010). "Contamination Rates and Antimicrobial
Resistance in Bacteria Isolated from Grass-Fed Labeled Beef Products."
Foodborne Pathogens and Disease 7(11): 1331-1336.
129
Chapter 6 ― Clostridium difficile in retail fresh vegetables, Ohio, USA
Abstract
A prevalence study of C. difficile in retail fresh vegetables was conducted in Ohio, USA
to determine the prevalence of toxigenic C. difficile. Retail products were sampled from
various retailers during summer. A toxigenic strain of Clostridium difficile was isolated
from the outer leaves of a head of lettuce.
Introduction
More frequent outbreaks of severe Clostridium difficile infections and the
appearance of more virulent strains have emerged since 2000 as a public health problem
worldwide, yet the major sources of community acquired infections remain enigmatic
(McFarland 2008). Since the isolation of hyper-virulent C. difficile strains from livestock
in 2006 (Rodriguez-Palacios et al. 2006) and retail meats in 2007 (Rodriguez-Palacios et
al. 2007), concerns about food contamination and foodborne transmissibility are now
being increasingly documented (Bakri et al. 2009; Rodriguez-Palacios et al. 2009;
Songer et al. 2009).
The ability of infectious C. difficile spores to survive on indoor surfaces, in feces
and soil (Baverud et al. 2003) for several years, and the current use of livestock manure
130
as fertilizer during vegetable production with increasingly green technologies, led us to
determine the period prevalence of C. difficile contamination in a variety of retail fresh
vegetables. We hypothesized that fresh vegetables including root products (onions,
potatoes, etc.) could be contaminated with C. difficile serving as a source of infection as
recently reported for ready-to-eat salads in Scotland (Bakri et al. 2009).
Materials and Methods
Aiming to achieve sufficient study power (p=0.8), in 2007, we collected 125
various vegetable products from 4 different retailers (2 chain, 1 auction, and 1 in-farm
market) in Ohio, USA, during July and August. Purchased products corresponded to
single units systematically selected for most classes of vegetables available. Assuming
the possible presence of toxigenic C. difficile in surface soils, as previously reported in
the United Kingdom (al Saif and Brazier 1996) and Zimbabwe (Simango 2006), we
purposively included product units having soil material on the surface. All products were
from US (CA, PA, OH, MI, WI, NY and FL) and Mexican origin, but some produce
lacked origin description.
In the laboratory, 10-15 g portions of edible parts of the produce, including
surface cuts (<3mm thick) and/or the outermost leaves, were aseptically cultured for C.
difficile by enrichment-based culture methods. The enrichment (50 mL of selective broth
with cycloserine and cefoxitin) and subsequent isolation steps were performed as
previously described for retail meats (Rodriguez-Palacios et al. 2007; Rodriguez-Palacios
et al. 2009). UV-fluorescence was used to screen selective agar plates for suspect
131
colonies. Biochemical confirmation was based on L-proline aminopeptidase activity (Pro
Disc, Remel, Lenexa, KS, USA). Identification bias was minimized by re-coding and
blindly processing all samples. Positive and negative controls included C. difficile strain
ATCC 9689. For quality control purposes, the countertop used for sample processing was
swabbed at regular intervals and cultured for C. difficile to rule out cross-contamination.
Results and Discussion
Clostridium difficile was isolated from the outer leaves of a single conventionallygrown iceberg lettuce head, for which the origin could not be determined (Table 1). In
contrast to our expectations, none of the other vegetables tested positive for C. difficile.
The overall prevalence of contamination of all vegetables tested was 0.8% (1 of 125); but
the relative proportion of contamination for the leafy vegetable category was 2.4% (1 of
41; table 1). With a single isolate, no differences were detected when comparing aerial(1/65) and the remaining soil/root-associated (0/60) vegetables (Fisher‘s exact p>0.1).
No molecular typing of the C. difficile isolate was possible as it could not be revived after
storage.
Since the product was offered for sale wrapped in plastic, contamination might
have occurred anywhere from production to post-harvest processing and handling. The
identification of C difficile from this retail lettuce head, from minimally processed readyto-eat salads in Scotland (Bakri et al. 2009), and also from bagged spinach in the USA
(PCR-ribotype 078; per. com. Dr J. G. Songer, University of Arizona, Tucson, USA,
unpublished data), indicate that there are possible critical points during the farm-to-fork
132
continuum that have resulted in the contamination of some retail vegetables with C.
difficile.
To date, Clostridium difficile has been isolated from retail leafy (0.8-7.5%) and root
(2.3%) vegetables, and from retail meats (6.1-42%)(al Saif and Brazier 1996; RodriguezPalacios et al. 2007; Bakri et al. 2009; Songer et al. 2009); but the epidemiological role
of foods as sources of infection in the community remains uncertain. C. difficile is an
opportunistic pathogen that especially affects elderly and immune-compromised people;
however, other individuals taking antimicrobial and/or antacid medications are also at
risk (McFarland 2008). Food contamination thus raises concern as it may serve as a direct
or indirect source of infectious spores.
Although C. difficile has traditionally been considered to be ubiquitous, its
distribution across food production and processing environments is unlikely to be
homogeneous. Therefore, studies are needed to identify factors and intervening strategies
to minimize vegetable contamination, especially of freshly-eaten leafy green vegetables
and herbs, which have large-scale global markets and a complex array of production and
post-harvest processing systems that greatly vary around the world (Organization 2008).
Acknowledgements
This study was conducted with federal funds allocated to the Ohio Agricultural Research
and Development Center, The Ohio State University. All authors declare no financial
conflict of interest; JTL currently has an active grant to conduct research on microbial
hazards of vegetables.
133
Table 6. Retail fresh vegetables tested for Clostridium difficile in Ohio, USA, 2007
Arbitrary
Category
WHO/FAO
groups
n=†
Organic‡
C.
Priority*
difficile
Aerial parts of
Leafy (lettuce/spinach/chard/herbs)
1
41
4
1§
plant
Tomato/pepper
2
13
3
0
(n=65)
Berries
2
5
-
0
Broccoli
-
3
-
0
Green beans
-
3
-
0
Common
Melons
2
3
-
0
contact with
Cucurbits
3
10
3
0
soil surface
(n=13)
Root
Onions
2-3
20
-
0
vegetables
Roots
3
19
3
0
(n=39)
(carrots/potatoes/beets/parsnip)
Other
Sprouts
2
2
-
0
(n=8)
Mushrooms
-
6
1
0
-
125
14
1 (0.8%)
Total
*
Priority levels in terms of fresh produce safety assigned by the Food and Agriculture
Organization of the United Nations and the World Health Organization, 2008
(Organization 2008) . Leafy green vegetables are global top priority; priorities 2 and 3
depend on regional relevance.
†
Total of samples tested per category; retailers sampled (n=4) were located at 0, 26, 174
and 201 km from research center.
‡
Products with organic claims or from local communities that use manure as fertilizer.
Farm practices were unknown for most products.
§
Toxigenic C difficile was isolated from conventional iceberg lettuce
134
References
al Saif, N. and J. S. Brazier (1996). "The distribution of Clostridium difficile in the
environment of South Wales." J Med Microbiol 45(2): 133-137.
Bakri, M. M., D. J. Brown, J. P. Butcher, et al. (2009). "Clostridium difficile in ready-toeat salads, Scotland." Emerg Infect Dis 15(5): 817-818.
Baverud, V., A. Gustafsson, A. Franklin, et al. (2003). "Clostridium difficile: prevalence
in horses and environment, and antimicrobial susceptibility." Equine Vet J 35(5):
465-471.
McFarland, L. V. (2008). "Update on the changing epidemiology of Clostridium difficileassociated disease." Nat Clin Pract Gastroenterol Hepatol 5(1): 40-48.
Organization, F. a. A. O. o. t. U. N. a. W. H. (2008). Microbiological hazards in fresh
leafy vegetables and herbs: Meeting Report. Microbiological Risk Assessment
Series. , FAO/WHO.
Rodriguez-Palacios, A., H. R. Staempfli, T. Duffield, et al. (2007). "Clostridium difficile
in retail ground meat, Canada." Emerg Infect Dis 13(3): 485-487.
Rodriguez-Palacios, A., H. R. Stampfli, T. Duffield, et al. (2006). "Clostridium difficile
PCR ribotypes in calves, Canada." Emerg Infect Dis 12(11): 1730-1736.
Rodriguez-Palacios, A., H. R. Stampfli and J. S. Weese (2009). "Possible seasonality of
Clostridium difficile in retail meat, Canada." Emerg Infect Dis: 831-834.
Simango, C. (2006). "Prevalence of Clostridium difficile in the environment in a rural
community in Zimbabwe." Trans R Soc Trop Med Hyg 100(12): 1146-1150.
135
Songer, J. G., H. T. Trinh, G. E. Killgore, et al. (2009). "Clostridium difficile in retail
meat products, USA, 2007." Emerg Infect Dis 15(5): 819-821.
136
Chapter 7 ― Enteric bacterial pathogens with zoonotic potential isolated from
farm-raised white-tailed deer
As Published in :
French, E., A. Rodriguez-Palacios, and LeJeune. (2010) "Enteric bacterial pathogens with
zoonotic potential isolated from farm-raised deer." Foodborne Pathog Dis 7(9):
1031-1037. First and second authors were equal contributors.
Abstract
The raising of captive white-tailed deer (Odocoileus virginianus) is a growing
agricultural industry in Ohio as it is in several other areas of the United States and around
the world. Pooled fecal samples were collected from 30 white-tailed deer confinement
facilities. Samples were cultured for five enteric bacterial pathogens. Premise
prevalence rates were as follows: Escherichia coli O157, 3.3%; Listeria monocytogenes,
3.3%; Salmonella enterica, 0 %; and Yersinia enterocolitica, 30%; Clostridium difficile,
36.7%. The ail virulence gene could not be amplified from any of the Y. enterocolitica
isolates recovered. Toxigenic strains of C. difficile PCR-ribotype 078, an emerging C.
difficile genotype of humans and food animals, were recovered from four (36.4%) of 11
C. difficile positive deer farms. Venison from farm-raised deer might become
contaminated with foodborne pathogens, deer farmers may have occupational exposure to
these zoonotic agents, and farm raised deer could be a reservoir from which the
137
environment and other livestock may become contaminated with a number of pathogenic
bacteria.
Introduction
The farming of white-tailed deer is a growing industry of agriculture across the
United States. According to the National Agricultural Statistical Service (US), there were
approximately 5,654 farms raising a collective total of 269,537 deer, up 20% from the
previous census. Ohio reported a total of 303 farms in 2007, 4900 deer farms in the US,
with Ohio reporting 224 deer farms in 2002 (Service 2009). Commercial farming of
white-tailed deer is conducted for the production of venison, harvest of antlers, urine
collection, and for the population of hunting preserves. As ruminants, these animals are
subject to colonization of many of the same bacterial agents as domestic livestock. It is
estimated that annually the industry of rearing cervids contributes $2.3 billion dollars into
the US economy through the purchase of feeds, veterinary services, fuel and other
purchases (Anderson et al. 2007).
As in other ruminants, bacterial infections in deer can cause illness or remain as
asymptomatic infections. Several zoonotic pathogens, notably E. coli O157:H7, Listeria
monocytogenes, Yersinia enterocolitica and Salmonella enterica have been reported in
wild deer populations (Branham et al. 2005,Lillehaug et al. 2005,Renter et al.
2006,Garcia-Sanchez et al. 2007,Lyautey et al. 2007). The prevalence of Clostridium
difficile in white-tailed deer has not been reported. The presence of these zoonotic
organisms poses food and environmental safety risks if carcass contamination occurs at
138
the time of harvest or pathogens are released into the environment during the pre-harvest
stages of production. The objective of this study was to determine the farm-level
prevalence of the aforementioned five pathogens among confined and intensively
managed white-tailed deer in Ohio, USA.
Materials and Methods
Sampling frame
All 301 individuals licensed through Ohio Department of Natural Resources,
Division of Wildlife in 2006 to commercially and non-commercially propagate deer in
Ohio were sent an invitation to participate in this study (Figure 1). The mailing contained
a letter describing the study, sampling materials, and a postage-prepaid response
envelope. Reminder letters were sent to non-respondents. In addition, non-respondents
farms, in geographic proximity to the laboratory were visited by investigators as an
additional recruitment strategy. Using the clean latex gloves provided, participants were
requested to collect one to two tablespoons of freshly deposited feces from 5 or 6
different locations in the confinement area. Samples were submitted in the sterile
containers provided. No financial incentives were offered. Upon receipt in the
laboratory, samples were stored refrigerated for up to 5 days so that they could be
processed in batches. All samples were collected and assayed between November 2006
and May 2007.
Microbiological and Molecular Analysis
139
For bacterial isolation, 20g of each composite sample was homogenized with 180
ml of Universal Pre-enrichment broth (UPB, Acumedia 7510A, Lansing, Michigan) using
a peristaltic-type mixer (120 rpm, 2 min). Two 15 ml aliquots of this homogenate was
removed and used for Clostridium difficile detection (see details below). The remainder
of the sample was incubated for 14-18h at 35oC. Testing for Escherichia coli O157,
Listeria monocytogenes, Yersinia enterocolitica and Salmonella enterica was conducted
on aliquots of the pre-enrichment broth as previously described (Dodson and LeJeune
2005,Bowman et al. 2007,Kersting et al. 2009) and outlined below.
Escherichia coli O157:H7 was screened in single 1 ml aliquots of UPB using
immunomagnetic separation and a commercial automated bead retrieval protocol
(BeadRetriever, Dynal, Oslo, Norway). Briefly, retrieved beads were then plated on
Sorbitol MacConkey agar plates containing cefixime (50g/ml) and tellurite (100g/ml).
Up to five suspect sorbitol-negative colonies from each plate were further screened for
lactose fermentation, absence of D-umbeliferyl-Beta-glucoronide cleavage, and latex
agglutination for O157 antigens (E. coli O157 Latex, Oxoid Ltd, Basingstoke, UK).
Isolates were confirmed as E. coli O157:H7 using a multiplex PCR protocol (Hu et al.
1999)
For Listeria monocytogenes screening, one ml of the UPB enrichment was
transferred to 9 ml of Fraser broth and incubated for 48h at 35°C. Samples that darkened
the broth were streaked onto polymyxin acrifflavin LiCl ceftazadime esculine manitol
agar (PALCAM, Oxoid) and incubated for 14-18h at 35°C. Suspect Listeria colonies
were tested for β-hemolysis on Columbia Blood agar and chromogenic characteristics on
140
Rapid‘L mono agar (Bio-Rad Laboratories, Hercules, CA). Isolates were confirmed using
a multiplex PCR assay (Zhang and Knabel 2005).
For Salmonella enterica, 10 μl of the UPB enrichment was incubated into 10 ml
of Rappaport-Vassiliadi broth at 37oC. After 14-18h of incubation, a loopful of this
enrichment was streaked onto Xylose-lysine-tergitol-4 (XLT4, Neogen Acumedia, MI)
agar and incubated 18 hours at 37oC. From each agar plate, up to three suspect colonies
were selected and plated on Triple Sugar Iron, Citrate, and Urea agars (Neogen
Acumedia, MI). One colony from each fecal sample that produced an acid butt and
alkaline slant with H2S production on Triple Sugar Iron, urea-negative, and citratepositive isolates were serogrouped using group specific antisera (Becton-Dickinson,
Sparks, MD).
For Yersinia enterocoltica, one ml of the UPB enrichment was incubated into 9
ml peptone sorbitol bile broth at 4°C (Neogen Acumedia). After 10 days of incubation, 1
ml of this broth was mixed with 9 ml of a 0.5%-KOH 0.5%-NaCl solution; from this
KOH/NaCl mixture 100 µl was streaked onto celfsulodin-irgasan-novobiocin agar (CIN,
Hardy Diagnostics, Santa Maria, CA) and incubated overnight at 30°C. Suspect colonies
(clear zone surrounding a red center) were transferred to lysine-arginine-iron, urea, and
bile-esculin agar slants and incubated for 48 h at 30 °C (Neogen Acumedia). Colonies
that produced a purple color change with no gas on lysine-arginine-iron, pink color
change on urea, and blackening of bile-esculin agars were transferred to 2 tubes with of
methyl red-Vogues Proskauer broth. One tube was incubated overnight at 30 °C and the
other for 48 h at 37 °C. Isolates that exhibited turbidity at 30 °C and agglutination at 37
141
°C were considered presumptive Y. enterocolitica and were stored at -80°C for PCR
analysis. Confirmation was based on detection of a 16s DNA amplicon and the ail
virulence gene (Bowman et al. 2007).
For Clostridium difficile, duplicate analysis was conducted on sediments of the 15
ml UPB/fecal aliquots (8200 x g, 10 min). Sediments were re-suspended with 15ml of an
enrichment broth prepared with the ingredients from a commercial agar base (Oxoid Ltd)
as previously described (Rodriguez-Palacios et al. 2007). One aliquot was supplemented
with commercial cycloserine/cefoxitine (Oxoid, Ltd) and the other with
moxalactam/norfloxacin (CDMN, Oxoid, Ltd). Broths were anaerobically incubated for
7-10 days at 37oC to favor sporulation. To reduce fecal contaminants and select for
spores, 2 ml of these homogenized broths were then mixed with 96% ethanol (1:1, v/v,
50min), centrifuged, and streaked onto C. difficile agar supplemented with the
corresponding antibiotics (cycloserine/cefoxitine or CDMN). Plates were examined for
up to 5 days under anaerobic incubation for suspect colonies (non-hemolytic, swarming,
p-cresol smell). Up to 3 colonies per plate were sub-cultured onto 5% sheep blood agar.
Preliminary confirmation was based on L-proline aminopeptidase activity. Crudely
extracted DNA (boiling, 10 min) was used for PCR confirmation (housekeeping tpi gene
detection), toxin gene profiling (tcdA, tcdB, cdtA, cdtB, tcdC deletions, tcdE, cdu) and
fingerprinting (PCR-ribotyping) of isolates as performed in previous studies (RodriguezPalacios et al. 2006 & 2009). Toxin production was confirmed in bacterial suspensions
(72 hour old colonies, and PBS) for up to two isolates for each toxin profile using
142
commercial ELISA Wampole Tox A/B™ (TechLab, Blacksburg, Va) following the
manufacturer‘s recommendations.
Farm practices
Questionnaires (see supplementary material) were completed by participants to
ascertain commercial purpose, herd demographic factors (presence of other animals), and
management factors (density, grazing area, health, etc).
Statistics
Proportions of binary data were compared with Chi-square or Fishers‘ exact
statistics. Adjusting for rare events, all reported 95% confidence intervals for proportions
used exact binomial Clopper-Pearson statistics (Clopper and Pearson 1934). Univariate
analysis of continuous data was conducted using 2-sample t-test or Mann-Whitney
statistics where appropriate (Sheskin 2000). STATA was the statistical software used
(v10, College Station, TX).
Results
Thirty of 301 (10%) possible responders collected samples within the time frame
specified (Figure 12). Twenty samples were mailed in and an additional 10 samples were
collected during farm visits by the investigators. Twenty-two of 30 (73%) of the
questionnaires were fully completed. Regarding animal facility, participants reported to
have between 4 and 364 deer (median 31; mean 57.9±86.4), and between 0.8 and 40
hectares (median 2; mean 4.9±8.6) available for grazing at the time of sampling.
Calculated stocking density ranged from 0.5 to 11.5 deer/hectare (median 5; mean
143
4.7±2.7); which was similar for pathogen-positive and pathogen-negative farms (t-test P
= 0.94).
Among the 30 farms, pathogens were isolated from 15 (50%) of the premises; 8
(27%) and 7 (23%) farms yielded one and two pathogens, respectively. Five of the farms,
three from which pathogens were isolated, had free access to open waters. Sharing
pasture with other livestock or free-range poultry was reported in 5 (22.7%) of the 22
farms submitting fully-answered questionnaires. Analysis of the 22 survey responders,
indicate that sharing pastures with other livestock was more common among pathogenpositive farms (4/7) compared to pathogen-negative (1/15) farms (Fisher‘s exact P =
0.021). None of the respondents reported recent problems with infections or disease in
their animals.
Escherichia coli O157 was isolated from one of the 30 premises (3.3%) (Table 7).
It was noted that the farm from which this sample was collected allowed co-grazing of
deer and other animals (cattle, bison, horses, small ruminants, and poultry). Listeria
monocytogenes was recovered from one specimen. Salmonella enterica was not
recovered from any specimen in this study.
A total of 14 (46.7%) of all 30 specimens yielded bacterial colonies that were
phenotypically indistinguishable and antigenically cross-reactive with the anti-Yersinia
antisera used. However, the 16S rDNA fragment specific to Yersinia enterocolitica could
be amplified from only nine (9/30, 30%) specimens, including the specimen from which
E. coli O157 was recovered. The gene ail, could not be amplified from any of these
144
isolates despite positive control assays for this gene consistently demonstrating the assay
was functioning properly.
Clostridium difficile was isolated from 11/30 (36.7%) of farms. Seven of the 11
(63.6%) locations yielded isolates with one or more of the toxin genes assayed (Table 8).
Among these toxigenic isolates there was a common variant isolates (tcdA+tcdB+cdtB+)
with a 39-bp type A deletion in putative negative toxin regulator tcdC (n=4) (Fig. 13A).
PCR-ribotyping, an international genomic fingerprinting based on 16S-23S rRNA gene
intergenic spacer regions, identified all four strains (i.e., tcdC 39-bp deletion) as human
epidemic PCR-ribotype 078 (57% of toxigenic C. difficile-positive farms, Fig. 13B, Table
1). Specimens with PCR-ribotype 078 were collected in March 2007 from farms that
were clustered within 30 km of distance (Fig. 12). No spatial statistics were conducted to
differentiate clustered from random data due to limited computational power, i.e., only 4
farms were PCR ribotype 078.
Discussion
In this study, one or more bacterial pathogens were isolated from 50% of farmed
white-tailed deer facilities following a single farm visit. This is the first report of
Clostridium difficile in deer, captive or wild. Although the specific within-farm animal
prevalence was not determined, this study demonstrates that enteric pathogens are present
among operations that raise captive cervids and that there is the potential to produce food
products that are contaminated with dangerous microorganisms, entails possible
occupational exposure to zoonosis, can contribute to pathogen loading in the
145
environment, and may be involved in the cycling of pathogenic organisms between
wildlife and domestic animals.
All the human enteric pathogens studied herein, except C. difficile, have
previously been reported among healthy and ill wild white-tailed deer populations
(Lillehaug et al. 2005,Kemper et al. 2006,Renter et al. 2006,Garcia-Sanchez et al.
2007,Lyautey et al. 2007). Although deer may become colonized with Escherichia coli
O157 and shed the organism in their feces for periods similar to cattle (Fischer et al.
2001), the prevalence of the organism is relatively low in the wild deer populations, with
occasional sharing of indistinguishable molecular subtypes between cattle and deer (Rice
et al. 1995,Renter et al. 2001). A longitudinal study of a single captive deer population in
the state of Louisiana, USA, identified one positive animal out of 226 samples tested
(Dunn et al. 2004). In contrast, a very small study of 10 captive deer in the UK
demonstrated 3/10 positive for E. coli O157 (Chapman and Ackroyd 1997). Like
commercial livestock production, the factors that influence the occurrence of E. coli
O157 on farms and the prevalence within a herd of deer are not fully understood. The
single farm testing positive in this study also raised cattle in the deer pen. Samples were
not collected from the cattle nor the environment outside the deer pen to determine if
other animals shared similar subtypes of E. coli O157 or if other environmental niches
were at this location contaminated with indistinguishable E. coli O157 strains.
Unlike E. coli O157, Salmonella enterica in addition to being asymptomatic may
cause disease in cervids (Robinson et al. 1970,Sato et al. 2000,Clark et al. 2004). Several
studies have been conducted to assess the prevalence of Salmonella enterica in wild and
146
semi-domesticated deer populations with the prevalence of Salmonella spp. generally
found to be extremely low, if detectable at all (Kemper et al. 2006,Renter et al. 2006).
However, a prevalence of 7% was reported in deer grazing rangeland shared by cattle
(Branham et al. 2005). It has been hypothesized that deer are somewhat refractory to
long-term carriage of Salmonellae because of their lack of a gall bladder (Henderson and
Hemmingsen 1983). In support of this theory, a recent study in humans and mice showed
gallstones playing a major role in prolonged fecal shedding due to enhanced colonization
of gallbladder tissue and bile with Salmonella enterica spp. compared to E. coli
infections (Crawford et al. 2010). From a public health perspective, prevalence-based
cross sectional studies indicate that deer are seemingly less frequent carriers of
Salmonella spp. compared to other animal species. Nevertheless, assumptions regarding a
lower prevalence for salmonella should not overrule the need to exercise preventive
measures due to the frequent presence of other relevant pathogens.
Listeria monocytogenes is a cause of meningo-ecephalitis and septicemia in
cervids (Kemenes et al. 1983,Eriksen et al. 1988,Tham et al. 1999), but it has been
recently isolated from feces of clinically healthy deer (6%) and moose (5%) in the wild
(Lyautey et al. 2007). Likewise, Yersinia enterocolitica has been reported in both wild
and captive deer populations (Henderson 1983,Henderson 1984). In addition to other
virulence genes, most pathogenic Yersinia enterocolitica strains harbour the
chromosomal ail marker (Miller et al. 1989,Zheng et al. 2008).Given that the majority of
isolates obtained in this study were likely of low pathogencity and minimal risk for food
147
or environmental safety, additional serological or molecular analyses were not performed
these isolates.
This is the first report documenting the presence of C. difficile in deer, and
specifically, a geographic cluster of farms with highly-relevant human emerging C.
difficile PCR ribotype 078 (Figure 1). Considering that there are more than 160
recognized C. difficile PCR ribotypes to date (Rodriguez-Palacios et al. 2006), the
molecular characteristics of the C. difficile strain 078 identified in four (36.4%) of 11 C.
difficile positive farms provides molecular evidence of a cluster. Although no spatial
statistics were computed due to low power to distinguish clustered from random
distribution, the joint probability of the combined events (culture-yes/no vs. molecular vs.
spatial localization-3 of 4 farms were located with 30 km) indicate that the cluster
observed occurring due to chance alone is very unlikely. Since the first identification of
this human C. difficile genotype in cattle in 2004 (Rodriguez-Palacios et al. 2006), and
the documented presence of emerging C. difficile strains in retail meat products intended
for human consumption since 2005 (Rodriguez-Palacios et al. 2006 & 2009, Songer et al.
2009), it has become increasingly evident that PCR ribotype 078 is predominant among
calves and swine, and is an emerging strain among people in hospital and community
settings in North America and Europe (Rupnik et al. 2008, Debast et al. 2009, Limbago et
al. 2009). This region on Ohio where the PCR ribotypes 078 were identified is home to a
high density of animal agriculture, particularly dairy production (Bender 2001). Although
it is uncertain whether the PCR ribotype 078 strains isolated from deer in this study
represent inter-species or only deer-adapted strains, the presence of this genotype,
148
previously isolated from swine, cattle, and humans indicate the potential of regional
dissemination of pathogens among farming operations.
Venison derived from farm raised deer is used for personal consumption, and sold
commercially in restaurants in metropolitan areas. Like other cervids, the frequency that
hair is shed during post-slaughter handling is of particular concern as it may contaminate
the carcass (Gracey et al. 1999). Slaughter hygiene, in addition to pathogen prevalence
of incoming animals may impact final microbial loads (Gill 2007). Incidence of
microbial contamination of carcasses at approved deer slaughter facilities in New
Zealand reported plant-to-plant and lot-to-lot variability in contamination of carcasses
with coliforms and E. coli (Sumner et al. 1977b, Sumner et al. 1977a). Generally, the
microbial counts were lower on carcasses from farm raised deer, but 16/30 samples
collected from one plant yielded Salmonella St. Paul (Sumner et al. 1977a).
Oftentimes the removal of viscera is performed by individuals who may not be
aware of food safety threats posed by contamination events that are not grossly obvious,
and performed in the field, lacking appropriate hygienic and temperature controls. One
might speculate that the rate on contamination under such more primitive conditions is
considerably higher. Subsequent processing of contaminated primal cuts in the
processing environment may serve as a source of contamination to other products. These
deficiencies can lead to food contamination and public health threats. Of particular
concern and direct relevance to food safety is the fact that the slaughter, dressing and
processing of deer and other game meats in the US falls outside the regulatory
jurisdiction of any federal food safety authority. Current regulations simply require the
149
processing of game meats to be separated from the processing of meat and poultry. In the
US, there is no federal requirement for subsequent professional food safety inspection,
although some states may have specific regulations governing this activity. Processing
oversight falls outside of the United States Department of Agriculture (USDA) Food
Safety and Inspection preview. When intended for human consumption, venison will fall
under the charge of the Food and Drug Administration (FDA) only when it enters interstate commerce. FDA does not have regulatory authority if the venison is in intra-state
commerce, but the USDA Agricultural Marketing Act of 1946 does provide service for a
voluntary USDA inspection (fee for service) of game animals at slaughter.
In addition to food safety hazards, there is a zoonosis hazard via occupational
exposure to the approximately 30,000 individuals that are employed in the deer farming
industry (Anderson et al. 2007). Many of these individuals may not be aware of the
potential hazards to which they are exposed or the possibility that they may bring
pathogens from farm sites back to their home or alternate work site (Kersting et al. 2009).
Contact between a child with an elevated risk of opportunistic infections and a deer was
followed by the child developing diarrhea, septicemia, and multiple organ abcessation
(Grigull et al. 2005).
Another concern of production agriculture that has implications to pathogen
carriage among farm raised deer is waste management. The bacterial pathogens evaluated
in this study, particularly C. difficile (al Saif and Brazier 1996), can survive for extended
periods and contribute to pathogen loading in the environment, waterways, and serve as a
reservoir for bi-directional cycling of pathogens to other wildlife, domestic livestock (Pell
150
1997, Vercauteren et al. 2006) and even humans. However, it is not possible from this
study to determine the extent that deer farms might be a source reservoir for pathogen
transfer to livestock, or if the deer were incidental secondary host as a result of the
intensive livestock production in the surrounding area. In our study, management was
such that deer on 5 of the 22 farms had free access to open waters. More than one
pathogen was isolated from 3 of these five farms.
In conclusion, this study demonstrates that fecal samples of deer raised in
confinement can harbor enteric human pathogens. Notably, the emerging epidemic C.
difficile PCR-ribotype 078 was identified in farmed deer. Although this study was based
on voluntary participation of deer framers and therefore there is potential for biased
estimates (Table 9), the presence of the pathogens tested in farm-raised deer indicates
there are threats to food safety, direct contact zoonoses, and environmental
contamination. As culture-based diagnostic methods often yield false negative results
(i.e., sensitivity is not 100%), it is possible that the actual prevalence of such pathogens
among deer farms might also be slightly higher. As deer farming continues to grow, more
individuals will become involved in farming of deer and the processing of venison. It is
critical that those individuals currently working in the industry, as well as those who start
in the future, are made aware of these hazards and educated in a way to minimize
environmental, carcass and direct-contact transmission of these organisms.
151
Acknowledgements
All authors were supported by state and federal funds allocated to the Ohio
Agricultural Research and Development Center, Wooster, Ohio, USA. This work was
also financed in part by a grant to The College of Wooster from the Howard Hughes
Medical Institute through the Undergraduate Science Education Program to support coauthor EF. A special thanks to Dr. Amber McCoig, FDA, for valuable comments on our
interpretation of FDA regulations and to the anonymous reviewers for their thoughtful
comments and time.
152
Figure 12. Distribution of 301 contacted premises with captive white-tailed deer, study
participants, and positive premises to C. difficile and other human enteric pathogens in
Ohio, USA.
Small grey irregular areas represent urban centers. Note the close-up map of clustered C.
difficile positive; four of the C. difficile farms shown had epidemic PCR-ribotype 078
(solid black stars). No water sources were common to all farms. Geographical coordinate
system, NAD, datum 1927. Geocoding accuracy was modified for illustration.
153
A.
B.
Figure 13. Multiplex PCR and PCR-ribotyping. A) Multiplex PCR for tcdC, tcdE and
cdu-2 within the Pathogenicity Locus (PaLoc).
C. difficile isolates from 3 white-tailed deer farms had deletion type A (39bp) in tcdC
gene. Note that cdu-2, a downstream flanking gene in PaLoc did not amplify with this
method indicating a variant gene in PCR-ribotype 078 strains. B) PCR-ribotyping using
isolates from duplicate testing in 3 farms have identical pattern to the first PCR-ribotype
078 isolate identified in food animals, Ontario, Canada, 2004 (label 078 in both gels) .
OH561, control strain verifying typing reproducibility; L, 100-bp molecular marker.
154
Table 7. Human enteric pathogens among 30 captive-deer operations in Ohio, November
2006-May 2007, USA.
Pathogen combination
Farms (%)
Estimated 95%CI
0/30 (0)
0, 11.6
E. coli and Y. enterocolitica
1/30 (3.3)
0.08,17.2
L. monocytogenes & C. difficile
1/30 (3.3)
0.08, 17.2
C difficile & Y. enterocolitica
5/30 (16.7)
5.6, 34.7
Total Y. enterocoliticaa
9/30 (30.3)
14.7, 49.4
Total C. difficileb
11/30 (36.7)
19.9, 56.1
15/30 (50)
31.3, 68.7
S. enterica
Total (Any pathogen)
a
None of the 9 Y. enterocolitica farms had the ail virulence marker.
b
Seven of the 11 C. difficile –positive farms had toxigenic strains. Four (57%) of these 7
farms had epidemic C. difficile PCR-ribotype 078 (Table 2). CI, confidence interval.
155
Table 8. Culture recovery and toxin gene profiles of 37 C. difficile isolates from 11 deer
farms, USA, 2007a
Toxin gene profile
Isolates tested,
Farms, n= (ID)
n=
tcdA+tcdB+cdtB+; tcdC with
13
(isolates,n=)
4 (12, 23, 29, 30)
39-bp deletionb
tcdA-tcdB+cdtB-; tcdC and
9
6 (14,16,20,23,24,28)
8
5 (12,13,20,24,28)
6
2 (16, 26)
CCFB (4), CDMN
(2)
1
1 (26)
(type B) deletion
a
CCFB (7), CDMN
(1)
deletion (classical)
tcdA+tcdB+cdtB+; tcdC 18-bp
CCFB (7), CDMN
(2)
amplify
tcdA+tcdB+cdtB-; tcdC had no
CCFB (4), CDMN
(9)
tcdA did not amplify
tcdA-tcdB-cdtB-; tcdC did not
Culture media
CCFB (0), CDMN
(1)
All culture and molecular analysis were conducted in a blinded fashion. C. difficile was
confirmed with tpi gene. Toxin production was confirmed with commercial ELISA
Wampole Tox A/B™ (TechLab, Blacksburg, Va).
b
PCR-ribotype 078 (n=3).
156
References
al Saif, N and JS Brazier. The distribution of Clostridium difficile in the environment of
South Wales. J Med Microbiol 1996; 45: 133-7.
Anderson, D, B Frosch and J Outlaw. Economic impact of the United States Cervid
Farming Industry, Agricultural & Food Policy Center, Texas A & M
University.2007
Bender, M. An economic comparison of traditional and conventional agricultural systems
at a county level. American journal of alternative agriculture 2001; 16: 2-15.
Bowman, AS, C Glendening, TE Wittum, et al. Prevalence of Yersinia enterocolitica in
different phases of production on swine farms. J Food Prot 2007; 70: 11-6.
Branham, LA, MA Carr, CB Scott, et al. E. coli O157 and Salmonella spp. in white-tailed
deer and livestock. Curr Issues Intest Microbiol 2005; 6: 25-9.
Chapman, PA and HJ Ackroyd. Farmed deer as a potential source of verocytotoxinproducing Escherichia coli O157. Vet Rec 1997; 141: 314-5.
Clark, RG, SG Fenwick, CM Nicol, et al. Salmonella brandenburg - emergence of a new
strain affecting stock and humans in the South Island of New Zealand. N Z Vet J
2004; 52: 26-36.
Clopper, CJ and ES Pearson. The use of confidence or fiducial limits illustrated in the
case of the binomial. Biometrika 1934; 26: 404-413.
Crawford, RW., R Rosales-Reyes, MD. Ramirez-Aguilar, et al. Gallstones play a
significant role in Salmonella spp. gallbladder colonization and carriage. Proc
Natl Acad Sci U S A. 2010; 107:4353-4358.
Debast, SB, LA van Leengoed, A Goorhuis, et al. Clostridium difficile PCR ribotype 078
toxinotype V found in diarrhoeal pigs identical to isolates from affected humans.
Env Micro 2009; 11: 505-511
157
Dodson, K and J LeJeune. Escherichia coli O157:H7, Campylobacter jejuni, and
Salmonella prevalence in cull dairy cows marketed in northeastern Ohio. J Food
Prot 2005; 68: 927-31.
Dunn, JR, JE Keen, D Moreland, et al. Prevalence of Escherichia coli O157:H7 in whitetailed deer from Louisiana. J Wildl Dis 2004; 40: 361-5.
Eriksen, L, HE Larsen, T Christiansen, et al. An outbreak of meningo-encephalitis in
fallow deer caused by Listeria monocytogenes. Vet Rec 1988; 122: 274-6.
Fischer, JR, T Zhao, MP Doyle, et al. Experimental and field studies of Escherichia coli
O157:H7 in white-tailed deer. Appl Environ Microbiol 2001; 67: 1218-24.
Garcia-Sanchez, A, S Sanchez, R Rubio, et al. Presence of Shiga toxin-producing E. coli
O157:H7 in a survey of wild artiodactyls. Vet Microbiol 2007; 121: 373-7.
Gill, C. Microbiological conditions of meats from large game animals and birds. Meat
Science 2007; 77: 149-160.
Gracey, J, D Collins and R Huey (1999). Meat Hygiene. London, Harcourt Brace and
Company.
Grigull, L, C Linderkamp, A Sander, et al. Multiple spleen and liver abscesses due to
Yersinia enterocolitica septicemia in a child with congenital sideroblastic anemia.
J Pediatr Hematol Oncol 2005; 27: 624-6.
Henderson, TG. Yersiniosis in deer from the Otago-Southland region of New Zealand. N
Z Vet J 1983; 31: 221-4.
Henderson, TG. The isolation of Yersinia sp. from feral and farmed deer faeces. N Z Vet
J 1984; 32: 88-90.
Henderson, TG and P Hemmingsen. Faecal survey of deer for Yersinia
pseudotuberculosis and Salmonella sp. N Z Vet J 1983; 31: 225-6.
Hu, Y, Q Zhang and JC Meitzler. Rapid and sensitive detection of Escherichia coli
O157:H7 in bovine faeces by a multiplex PCR. J Appl Microbiol 1999; 87: 86776.
Kemenes, F, R Glavits, E Ivanics, et al. Listeriosis of roe-deer in Hungary. Zentralbl
Veterinarmed B 1983; 30: 258-65.
Kemper, N, A Aschfalk and C Holler. Campylobacter spp., Enterococcus spp.,
Escherichia coli, Salmonella spp., Yersinia spp., and Cryptosporidium oocysts in
semi-domesticated reindeer (Rangifer tarandus tarandus) in Northern Finland and
Norway. Acta Vet Scand 2006; 48: 7.
158
Kersting, AL, LC Medeiros and JT Lejeune. Differences in Listeria monocytogenes
Contamination of Rural Ohio Residences With and Without Livestock. Foodborne
Pathog Dis 2009.
Limbago, BM, CM Long, AD Thompson, et al. Clostridium difficile Strains from
Community-Associated Infections. J of Clin Microbiol 2009; 44: 304-7
Lillehaug, A, B Bergsjo, J Schau, et al. Campylobacter spp., Salmonella spp.,
verocytotoxic Escherichia coli, and antibiotic resistance in indicator organisms in
wild cervids. Acta Vet Scand 2005; 46: 23-32.
Lyautey, E, A Hartmann, F Pagotto, et al. Characteristics and frequency of detection of
fecal Listeria monocytogenes shed by livestock, wildlife, and humans. Can J
Microbiol 2007; 53: 1158-67.
Miller, VL, JJ Farmer, 3rd, WE Hill, et al. The ail locus is found uniquely in Yersinia
enterocolitica serotypes commonly associated with disease. Infect Immun 1989;
57: 121-31.
Pell, AN. Manure and microbes: Public and animal health problem? Journal of Dairy
Science 1997; 80: 2673-2681.
Renter, DG, DP Gnad, JM Sargeant, et al. Prevalence and serovars of Salmonella in the
feces of free-ranging white-tailed deer (Odocoileus virginianus) in Nebraska. J
Wildl Dis 2006; 42: 699-703.
Renter, DG, JM Sargeant, SE Hygnstorm, et al. Escherichia coli O157:H7 in free-ranging
deer in Nebraska. J Wildl Dis 2001; 37: 755-60.
Rice, DH, DD Hancock and TE Besser. Verotoxigenic E coli O157 colonisation of wild
deer and range cattle. Vet Rec 1995; 137: 524.
Robinson, RM, RJ Hidalgo, WS Daniel, et al. Salmonellosis in white-tailed deer fawns. J
Wildl Dis 1970; 6: 389-96.
Rodriguez-Palacios, A, HR Staempfli, T Duffield, et al. Clostridium difficile in retail
ground meat, Canada. Emerg Infect Dis 2007; 13: 485-7.
Rodriguez-Palacios, A, HR Stampfli, T Duffield, et al. Clostridium difficile PCR
ribotypes in calves, Canada. Emerg Infect Dis 2006; 12: 1730-6.
Rodriguez-Palacios, A, RJ. Reid-Smith, HR Stampfli, et al. Possible seasonality of
Clostridium difficle in retail meat, Canada. Emerg Infect Dis 2009; 15: 802-5.
Rupnik, M, W. Andreas, O. Zimmermann et al. Clostridium difficile Toxinotype V,
Ribotype 078, in Animals and Humans. J. Clin Microbiol 2008: 46: 2146
159
Sato, Y, C Kobayash, K Ichikawa, et al. An occurrence of Salmonella typhimurium
infection in sika deer (Cervus nippon). J Vet Med Sci 2000; 62: 313-5.
Service, NAS. 2007 Census of Agriculture, Ohio State and County Data, Volume 1,
Geographic Area Series, Part 35. USDA.2009
Sheskin, D. Handbook of Parametric and Nonparametric Statistical Procedures. Boca
Raton, Chapman & Hall/CRC.2000
Songer, JG, HT Thrin, Killgore, GE, et al. Clostridium difficile in Retail Meat Products,
USA, 2007. Emerg Infect Dis 2009; 15: 819-21.
Sumner, JL, IR Perry and CA Reay. Microbiology of New Zealand farmed venison. J Sci
Food Agric 1977a; 28: 1105-8.
Sumner, JL, IR Perry and CA Reay. Microbiology of New Zealand feral venison. J Sci
Food Agric 1977b; 28: 829-32.
Tham, W, E Bannerman, J Bille, et al. Listeria monocytogenes subtypes associated with
mortality among fallow deer (Dama dama). J Zoo Wildl Med 1999; 30: 545-9.
vercauteren, K, M Lavelle, N Seward, et al. Fince-line contact between wild and farmed
white-tailed deer in Michigan: Potential for disease transmission. Journal of
Wildlife Management 2006; 71: 1603-1606.
Zhang, W and SJ Knabel. Multiplex PCR assay simplifies serotyping and sequence
typing of Listeria monocytogenes associated with human outbreaks. Journal of
Food Protection 2005; 68: 1907-10.
Zheng, H, Y Sun, Z Mao, et al. Investigation of virulence genes in clinical isolates of
Yersinia enterocolitica. FEMS Immunol Med Microbiol 2008; 53: 368-74.
160
Chapter 8 ― Clostridium difficile in anthropogenic wild bird ecosystems and
dissemination potential
(For submission)
Abstract
Here we examined at the effect of wild birds in the dissemination of this pathogen
and indentify ecological aspect that are relevant to the epidemiology of C difficile
infections at the interface between a rural and an urban area in a region with high density
of food animal productivity. A systematic review of literature and cross-sectional,
longitudinal and experimental infection studies were conducted along with radiotelemetry
analysis were conducted to determine the dissemination potential of C. difficile by wild
birds. Toxigenic strains of C. difficile were significantly more frequently identified in
wild birds inhabiting aquatic environments (Geese and ducks), than invasive terrestrial
birds inhabiting dairy farms (European starlings); no toxigenic C difficile was identified
from terrestrial small passerine birds inhabiting forest. The disemination potential of C.
difficile via wild birds was documented with molecular epidemiology methods and
telemetry.
161
Introduction
Concerns about emerging strains of enteric human pathogen have spread globally,
as the pathogen continues to reach distant and isolated regions (e.g., Australia) (Clements
et al. 2010; Gerding 2010). Clostridium difficile has been associated with severe enteric
disease since its first description in 1935 but little is known about the ecological factors
that might explain increased persistence and global dissemination (Clements et al. 2010;
Gerding 2010). Because, emerging strains have been found in food animals and foods of
regions where C. difficile is public health burden (Clements et al. 2010), some have
naturally looked at international trade as a reasonable route for disease expansion.
Likewise, the high rapid traffic of people from endemic regions in international flights
has been obviously considered (Clements et al. 2010). Here we examined at the effect of
wild birds in the dissemination of this pathogen and indentify ecological aspect that are
relevant to the epidemiology of C difficile infections at the interface between a rural and
an urban area in a region with high density of food animal productivity.
In the USA, C. difficile infections cause over half a million cases documented per
annum, with over 25,000 deaths (Campbell et al. 2009). In this context, C. difficile
infections are often more serious and frequent than the infamous methicillin-resistant
Staphylococcus aureus (MRSA) infections. Like other clostridia, most C. difficile strains
produce spores that survive for years in the environment (Baverud et al. 2003). Disease
occurs, in humans and some animals, when spores are ingested and activated in the gut.
Rapidly growing generations of daughter cells that produce toxins (A, B, and/or binary)
then destroy the gut lining. After active multiplication, induced by unfavorable conditions
162
within the gut, decaying generations turn back into spores to be excreted in the feces.
Fecal contamination thus contributes to environmental dissemination and spore
persistence (Baverud et al. 2003).
Due to the natural tendency of some birds to have regular contact with excrements
from food animals, mostly due to the presence of easy to get undigested grains that were
undigested and come with the feces, interspecies transmission of C. difficile strains is
plausible. Considered by some producers as ―farm pests‖ pigeons, European Starlings,
grackles and House sparrows are among the most frequent species seen in association
with cattle. In winter, the comingling between cattle and birds increases. Determining to
which extent birds contribute to spore dissemination is important because emerging C.
difficile strains (namely PCR ribotype 078) is increasingly common in food animals, and
foods. These strains now cause more human infections in the community. Recent studies
have documented that the farm level prevalence of strains PCR ribotype 078 can reach
100% (Debast et al. 2009) in North America and Europe. But it is uncertain if it is less
frequent in more distant regions. Here, we first describe the basis of our hypothesis that
wild birds are effective disseminators of C difficile; then we tested it using observational
studies and animal experimentation.
The study
Retrospective spatial analysis of emerging C. difficile in cattle
To test the basis of our hypothesis we first conducted a retrospective spatial
analysis of the first known emerging C. difficile strains reported in food animals in North
163
America to determine the distribution pattern of farming operations affected with
emerging epidemic strains (Rodriguez-Palacios et al. 2006). When looking at the three
most common genotypes in a study of calves sampled in 102 farms in 2004 in a 150.000
km2 region surrounded by the Great Lakes, two apparent patterns of spatial distribution
were observed. One was represented by a then newly variant C difficile strain (functional
mutation in gene for toxin A production) found first in severe outbreaks of human disease
in Central Canada at 4,000 km from the closest farm, was widely disseminated among the
farms, i.e., PCR ribotype 017 (Figure 14). The second pattern corresponded to the
remaining common ribotypes (including PCR ribotypes 078 and 027) were seemingly
present in clusters. The fact that the first isolate from PCR type 017 reported in humans
had been identified 5 years prior to the time of sampling and the location was
geographically divided by large bodies of water suggested that terrestrial compared to
aerial distribution was less likely. However earlier dissemination and paralell
independent evolution of the strains in the two regions is also possible.
Considerations about aerial dissemination
Recent reports indicate that aerial dissemination of C. difficile in indoor
environments in hospitals is possible. This might suggest that airborne spores could also
explain long range dissemination of spores from patients in Manitoba to calves farmed in
a vast region of Ontario. Is it that the calves acquired the PCR 017 variant strain from
humans, or was it the opposite? A connection between the calves and those patients
might be at attribute to global meat markets because meats can harbor cattle derived
emerging C. difficile strains (Rodriguez-Palacios et al. 2007; Rodriguez-Palacios et al.
164
2009). But the inverse connection between the patients and those calves would be more
difficult to explain. Airborne dissemination of hospital-associated spores (or from other
hypothetical distant source) through large-scale eastbound winds could be a possibility.
This is because the major winds blow almost all year round moving particles often in the
same direction, as revealed by wild forest fire smoke (NASA 2010). However, that region
have had the lowest average values of airborne fine particulate matter pollution (<2.5 µm
in diameter) compared to the rest of the world (van Donkelaar et al. 2010). Also, it is
important to note that central Canada is very cold (often frozen) during at least six of 12
months, which lowers the values of air pollution, and the carrying potential.
Supporting the difficulty to explain long-range aerial dissemination of spores, a
recent study in Europe (with inherent high values of air pollution) showed that indeed
spores can disseminate out of intense swine farming facilities with the wind; but the
detection decreased to undetectable levels with 30 meters of the facility (Keessen et al.
2010). Aerial dilution of spores as they radiate further apart from farm and spore
precipitation or atmospheric elevation could be alternate outcomes that would evade
detection. Although aerial dissemination might contribute to spore dissemination, other
more effective vehicles of transmission might play a role in disease dynamics. Since
invasive wild birds are common infamous inhabitants of farms in North America,
especially during winter when we previously documented a higher rate of C. difficile in
dairy calves, we further considered that birds could be involved in seasonal and continued
patters of spore dissemination.
Clostridium spp. in wild birds over the past 40 years
165
To assess the potential for birds to carry pathogenic C. difficile in their digestive
tract a systematic review of peer-reviewed literature since 1970 was conducted. In total,
over 720 hits were obtained. No reports describing any form of testing for C. difficile
fecal shedding in any type of wild birds were available. Only one study, published in the
1980 (Borriello et al. 1983), tested and identified C. difficile in domestic pet ducks (n=2)
and geese (n=2) and a parakeet (n=1) and found toxigenic C. difficile in two birds.
To explore the potential risk of wild bird exposure to any other pathogenic
clostridia we expanded the search to investigate the spatial distribution of reports
describing toxicoinfections associated with C. botulinum (largely lethal in birds and
thought associated with the presence of spores in aquatic sediments and carcasses) and C.
perfringens enteritis. Extrapolating the obtained data indicated that the risk of ingestion
of C difficile spores should be widely spread. A large body of spatial evidence exist
documenting C. botulinum associated with disease of almost exclusively aquatic wild
birds, in primarily in coastal lands associated with both sea and fresh water bodies.
Increasing description of this problem has emerged from various parts of the world but
particularly in Korea. In the USA, cases have been reported sporadically through the past
century. Reports of enteritis associated to C. perfringens are significantly less, and they
are often associated with sporadic cases. A few other species of Clostridium spp. have
also been reported in association with enteric disease. Collectively, these results indicated
that it is very likely to find C. difficile spores in the feces of wild birds, especially if there
are present in aquatic environments such as rivers and oceans (al Saif and Brazier 1996;
Zidaric et al. 2010). For most cases of avian botulism, spores are often isolated from
166
aquatic sediments. We thus tested this hypothesis with a cross sectional observational
study.
Prevalence of C. difficile in wild birds from anthropogenic and forest ecosystems
To quantify the relative risk of C. difficile carriage by wild birds we conducted a
cross-sectional study of C. difficile shedding in birds representing three distinct
ecosystems, one aquatic, and two terrestrials (cattle farms, and urban and rural forests).
Because the most common aquatic wild bird identified in North America are mallards
and Canada geese, which can be resident in public parks or water bodies used for
agricultural purposes, or truly large-range migrants, we decided to test this birds to
represent aquatic ecosystems. Among the most common terrestrial invasive bird in North
America are European starlings, which are short range migrants and at large residents of
farming operations, we decided to include these birds as a representative population that
have direct contact with food animals but minimal exposure to aquatic environments.
Lastly, to control for exposure to known sources of C. difficile spores (i.e., aquatic
environments, and food animals), we sampled small terrestrial birds in urban forests and
strategic migratory landmarks.
Results from this cross-sectional study, conducted with individual samples from
540 birds, in an area of squared kilometers in Ohio, USA, showed a significantly higher
rate of C. difficile shedding in waterfowl, followed by the European starlings sampled in
the cattle farms. The group of smallest birds was the less likely to yield a positive sample
(figure 15). PCR analysis showed that toxigenic strains were more likely to be present in
waterfowl. This study confirmed that waterfowl sampled from public places where a
167
small body of water was present are healthy acrriers of toxigenic C. difficile. This finding
is of major significance because public parks commonly inhabited by geese or ducks are
commonly heavily contaminated with bird excrements that are difficult to avoid. The
inadvertent exposure of elderly people (Rodriguez-palacios 2011 unpublished data) who
commonly visit or fish in those parks could be another source of C difficile for the
community; especially, because food and drink consumption is a common practice during
leisure times.
Verification of low shedding prevalence in small migratory birds
To further validate the results from the previous cross sectional study, and
additional historic collection of fecal specimens from small migratory birds were tested.
The purpose was to increase the study power to identify a low shedding prevalence. Thus
two types of study samples were screened; this is historic and prospective sampling. The
historic samples corresponded to fecal samples from birds trapped during migration in
forest reserves in the south coast of Lake Erie (a mandatory migration hallmark for many
species that reach the lake at this point) in 2005 that had been either frozen or stored on
70% ethanol. The prospectively collected samples corresponded to samples that were
subsequently collected during 2008 and 2009 in the preserved marsh area where ~50% of
the samples from 2007 had been collected. This time, composite and individual small
fresh droppings were collected from perch areas near the research station. Together,
testing of those samples (n=310) confirmed that C. difficile is not common in small
migratory birds, in agreement with date from 400 more recent samples from Eastern
Europe (Bandelj et al. 2011).
168
Experimental infection of wild birds with C difficile
Information derived from a cross sectional study only provides information
relevant to a single time point but it does not provide dynamical information. To assess
the dynamics of fecal shedding in wild birds and therefore of dissemination potential
across farms, we documented the dynamics of fecal shedding in European starlings
following experimental inoculation with C. difficile. Following infection of 28 wild
starlings (trapped at cattle farms) with three different strains of C. difficile it was possible
to document that starlings are also healthy carriers. Under experimental conditions, those
birds shed C. difficile for up to 5-8 days, but showed no signs of disease. At the end of the
study, the score of fat tissue deposits present in the mesentery and associated viscera was
similar for all infected groups and controls. Results indicate that infected birds were able
to maintain their appetite and body mass in captivity and after infection. Of interest, C.
difficile was isolated from the liver of some birds indicating translocation from the
intestine was possible.
Telemetry study of bird movements and C. difficile typing
To determine whether birds could move from farm to farm within those 5 days
and disseminate C. difficile to cattle in dairy farms we visited farms to trapped birds and
concurrently sample cattle. We thus determined the risk of dissemination and
transmission of C. difficile spores among wild birds and from birds to cattle by
combining PCR ribotyping data (a fingerprinting genetic technique) and telemetricassessed movement data. Starlings captured were tagged and released in strategically
selected farms over time, in 2008. Proving the dissemination principle, culture analysis
169
and PCR ribotyping showed that strains of C. difficile collected from birds accompanying
tracked birds had identical molecular to identify strains obtained from cattle (figure 16).
Further, we documented that during the period of the study that the C. difficile positive
farms had recent telemetric records indicating that at least one tagged bird (and likely her
companions) had moved repetitively between three farms during that time frame (figure
17).
Discussion
In summary, we confirmed that C. difficile can be carried and disseminated by
wild birds that commonly inhabit anthropogenic habitats such as cattle farms and public
parks. The most common wild bird associated with C. difficile shedding were waterfowls,
as we had predicted from the disease ecology from Clostridium botulinum in avian
species over tha past 40 years. In contrast, small migratory and resident forest birds are
not natural reservoirs of C. difficile. This disproves the somewhat general view that C.
difficile is a normal inhabitant of the intestinal tract of animals and that spores are widely
present in the environment. Lastly, our studies indicate that wild birds that shed C
difficile are healthy, even after experimental infection (and stressful housing/handling
conditions) with high doses of emerging hypervirulent strains of C. difficile relevant for
disease in humans.
Understanding the role of birds in the epidemiology and its relationship with
anthropogenic landscape ecology is relevant because global warming promotes more
residence, and less migration, in wild birds. It is expected that more of these prolific
(invasive) bird species be in close contact with an increasing ageing population, who is
170
already at higher risk for CDI and who frequently visit parks inhabited with wildlife, and
often in company of their grandchildren. The inadvertent exposure of humans to
unwanted microorganisms via wild bird droppings is not a novel concept, what is to
highlight from this study however is that this is the first clostridium organism identified
in wild birds that actually causes severe economic and social costs to several health care
systems in the world. Clostridium spp produce spores which have inherent different
disease dynamics, and which resist common disinfectants. Major infection control efforts
have not been able to control the ongoing CDI epidemic indicating that it necessary to
consider sources of C difficile exposure outside of hospitals. Although not all, some
public parks inhabited by C. difficile carrying wild birds could be a new unrecognized
source for spores for communities at risk. The role of transcontinental dissemination of
emerging strain between C. difficile regional hotspots also needs further consideration.
171
Acknowledgements
This study was supported by federal funds allocated to the Ohio Agricultural
Research and Development Center, The Ohio State University. A.R.P. is a Public Health
Preparedness for Infectious Diseases Program Research Fellow. Thanks to M. Kauffman,
J. Schrock and P. Schlegel, M. Morash for technical assistance. A.R.P conceived the
study, conducted the experiments, and wrote the first manuscript draft. In this thesis,
other collaborating individuals have not been acknowledged.
172
1999- Location of
first variant PCR
ribotype 017
outbreak isolate in
humans
2004-Location of
PCR ribotype 017
isolates from cattle
farms
1000 km
Landmark islands
for migratory birds
Figure 14. Retrospective spatial analysis of data from a study of C. difficile in calves,
Canada, 2004(Rodriguez-Palacios et al. 2006). Note the distribution of farms with PCR
ribotype 017 (variant without gene for toxin A; tcdA-/tcdB+cdtAB-). PCR ribotype 017
is responsible for severe disease in people in many countries.
173
% of samples yielding C. difficile
5 types
6
2x3 table
Fisher exact
(P=0.034)
a
5
3 types
4
5.8%
3
2
b
1 type
2.8%
1
11/189
6/215
Waterfowl
Starlings
0.7%
1/136
0
Songbirds
Figure 15. Cross sectional study of fecal prevalence of C. difficile in wild birds in Ohio,
USA, 2007.
174
Between birds and cattle
Between bird species
500bp
200bp
E. starling
E. starling
E. starling
Cattle
Cattle
1000bp
E. starling
E. starling
E. starling
Waterfowl
Waterfowl
Waterfowl
Waterfowl
1000bp
C
A
B
500bp
200bp
Figure 16. PCR ribotyping of isolates from wild waterfowl (geese or ducks), European
starlings and dairy cattle sampled within the same period and region. Strains subtypes AC.
175
0.6 miles
Starl.
Same Starling
Sept
8 miles
Aug
Starlings
Sept
Same day/farm
A
A
A
A
A
Figure 17. PCR ribotyping and telemetry data indicate that C. difficile (Subtype A in this
example) can be isolated from several birds (E. starlings) captured over time from farms
that had network connectivity through repetitive bird movements.
176
References
Bandelj, P., T. Trilar, J. Racnik, et al. (2011). "Zero prevalence of Clostridium difficile in
wild passerine birds in Europe." FEMS microbiology letters 321(2): 183-185.
Baverud, V., A. Gustafsson, A. Franklin, et al. (2003). "Clostridium difficile: prevalence
in horses and environment, and antimicrobial susceptibility." Equine Vet J 35(5):
465-471.
Borriello, S. P., P. Honour, T. Turner, et al. (1983b). "Household pets as a potential
reservoir for Clostridium difficile infection." J Clin Pathol 36(1): 84-87.
Campbell, Robert J. P., L. M. P. H. Giljahn, K. M. P. H. Machesky, et al. (2009).
"Clostridium difficile Infection in Ohio Hospitals and Nursing Homes During
2006." Infection Control and Hospital Epidemiology 30(6): 526-533.
Clements, A. C., R. J. Magalhaes, A. J. Tatem, et al. (2010). "Clostridium difficile PCR
ribotype 027: assessing the risks of further worldwide spread." Lancet Infectious
Diseases 10(6): 395-404.
Debast, S. B., L. A. van Leengoed, A. Goorhuis, et al. (2009). "Clostridium difficile PCR
ribotype 078 toxinotype V found in diarrhoeal pigs identical to isolates from
affected humans." Environ Microbiol 11(2): 505-511.
Gerding, D. N. (2010). "Global epidemiology of Clostridium difficile infection in 2010."
Infect Control Hosp Epidemiol 31 Suppl 1: S32-34.
Keessen, E. C., C. J. Donswijk, S. P. Hol, et al. (2010) "Aerial dissemination of
Clostridium difficile on a pig farm and its environment. Available from:
http://igitur-archive.library.uu.nl/student-theses/2011-0518-200326/artikel.doc.
Accessed July 21, 2011."
177
NASA (2010). "NASA Video Shows Global Reach of Pollution from Fires."
Rodriguez-Palacios, A., R. J. Reid-Smith, H. R. Staempfli, et al. (2009a). "Possible
seasonality of Clostridium difficile in retail meat, Canada." Emerg Infect Dis
15(5): 802-805.
Rodriguez-Palacios, A., H. R. Staempfli, T. Duffield, et al. (2007a). "Clostridium difficile
in retail ground meat, Canada." Emerg Infect Dis 13(3): 485-487.
Rodriguez-Palacios, A., H. R. Stampfli, T. Duffield, et al. (2006). "Clostridium difficile
PCR ribotypes in calves, Canada." Emerg Infect Dis 12(11): 1730-1736.
van Donkelaar, A., R. V. Martin, M. Brauer, et al. (2010). "Global estimates of ambient
fine particulate matter concentrations from satellite-based aerosol optical depth:
development and application." Environ Health Perspect 118(6): 847-855.
178
Chapter 9 ― Moist heat resistance, spore aging, and superdormancy in Clostridium
difficile
As published in:
Rodriguez-Palacios, A. and J. T. LeJeune (2011). "Moist-heat resistance, spore aging,
and superdormancy in Clostridium difficile." Appl. Environ. Microbiol. 77(9):
3085-3091.
Abstract
Clostridium difficile spores can survive extended heating at 71oC (160oF), a
minimum temperature commonly recommended for adequate cooking of meats. To
determine the extent to which higher temperatures would be more effective at killing C.
difficile, we quantified (D-values) the effect of moist heat at 85oC (145oF, for 0 to 30
minutes) on C. difficile spores, and compared it to 71 and 63oC. Fresh (1-week-old) and
aged (≥20-week-old) C. difficile spores from food and food animals were tested in
multiple experiments. Heating at 85oC markedly reduced spore recovery in all
experiments (5-6 log10 with 15 minutes, P < 0.001), regardless of spore age. In ground
beef, the inhibitory effect of 85oC was also reproducible (P < 0.001), but heating at 96oC
reduced 6 log10 within 1-2 minutes. Mechanistically, optical density and enumeration
experiments indicated that 85oC inhibits cell division but not germination, but the effect
179
was reversible in some spores. Heating at 63oC reduced counts for fresh spores (1 log10,
30 minutes, P< 0.04), but increased counts of 20-week-old spores by 30% (15 minutes, P
< 0.02) indicating sublethal heat treatment reactivates superdormant spores.
Superdormancy is an increasingly recognized characteristic in Bacillus sp., and it is likely
to occur in C. difficile as spores age. The potential for reactivation of (super)dormant
spores with sub-lethal temperatures may be a food safety concern, but it has potential
diagnostic value. Ensuring food is heated to >85oC would be a simple and important
intervention to reduce the risk of inadvertent ingestion of C. difficile spores.
Introduction
Clostridium difficile is a spore-forming enteric pathogen associated with
increasing outbreak frequency and disease severity in humans worldwide (Pepin et al.
2005; Warny et al. 2005; McFarland 2008). Since 2006, molecular studies have shown
that subtypes of C. difficile isolates recovered from food animals and retail foods are
genetically indistinguishable from human-derived strains (namely, PCR ribotypes 027,
077 and 078) indicating that foods could be a source of inadvertent infection for humans
(Rodriguez-Palacios et al. 2006; Rodriguez-Palacios et al. 2007; Rupnik 2007; Bakri et
al. 2009; Rodriguez-Palacios et al. 2009; Songer et al. 2009). Although ingestion of
spores may not necessarily result in disease induction, the presence of C. difficile spores
in a large proportion of raw and ready-to-eat meat products in North America (6-42%)
highlights the need to consider this potential risk factor in disease epidemiology
(Rodriguez-Palacios et al. 2009; Songer et al. 2009). Identification of effective measures
180
to reduce C. difficile exposure, especially among higher risk (e.g., elderly, or individuals
consuming antimicrobials or antacids) individuals, would further reduce the likelihood of
this organism becoming an important foodborne pathogen.
Although interventions to control food contamination with C. difficile may be
applied both during production and at harvest, little is known about this pathogen in food
animals and their environment. Thus, until additional information about the epizootiology
and ecology of this pathogen is better understood, post-harvest strategies, such as proper
cooking and handling, may be among the most effective measures to control foodborne
exposure to C. difficile in the short-term. Minimal internal meat cooking temperatures
ranging from 63 to 85oC (145 to 185oF) have been promoted by government and industry
organizations to control other foodborne pathogens (e.g., Salmonella spp., Escherichia
coli O157) (USDA ; FSA 2007; CFIA 2010). For ground meats, a frequent target is to
reach 71oC and maintain that temperature for a few seconds to reduce between 6 and 7
log10 units of most recognized non-spore forming pathogens (USDA 1999; FSA 2007).
But C. difficile produce spores which are more thermotolerant than vegetative cells.
Recent thermo-resistance studies on C. difficile where spores were heated at 71oC for two
hours (Rodriguez-Palacios et al. 2010), and studies with other clostridia (Sanchez-Plata et
al. 2005), indicate that C. difficile spores survive cooking temperatures recommended for
ground meats, and if heated spores survive, they could possibly proliferate during the
post-cooking chilling phase (Sanchez-Plata et al. 2005). Previous thermal studies with C.
difficile have been single-time point studies and conducted largely with human derived
strains (Meisel-Mikolajczyk et al. 1995; Wultanska et al. 2004). However, effective
181
comparisons often required multiple time point experiments and the determination of Dvalues (Juneja et al. 2001). The main objective of the present study was to quantify and
compare the inhibitory effect of moist-heat at 63, 71, 85 and 96oC on C. difficile spores,
from food animal and retail foods, in liquid media and in ground beef. In addition, we
tested similar thermal effect on spores aged under different conditions, and preliminarily
determined whether the mechanism of thermal inhibition was by impairing spore
germination or by impairing cell division after germination.
Materials and Methods
Clostridium difficile strains
Progressively, three experiment types: i) Thermal inhibition in liquid culture
media, ii) Mechanism of thermal inhibition, and, iii) Thermal inhibition on ground beef
and gravy were conducted with multiple C. difficile strains (n = 22; 13 genotypes).
Genotypes tested included toxigenic PCR ribotypes 014, 027, 033, 077, and 078, and
toxigenic strain ATCC 9569 (Fig. 23). All strains, except the ATCC strain, were isolated
from food animals and foods and represent genotypes of public health relevance in North
America (Rodriguez-Palacios et al. 2006; Rodriguez-Palacios et al. 2007).
Spore production and aging
Prior testing, spores were produced by using the nutrient exhaustion method
(Young and Mandelstam 1979). To enhance heterogeneity and thus external validity
(Richter et al. 2009), spores were produced in two nutritious broths. In brief, frozen C.
difficile isolates were revived and purified on tryptic soya agar (Acumedia, Lansing, MI)
supplemented with 5% defibrinated sheep blood (Hemostat Laboratories, Dixon, CA)
182
(blood agar). Single, 72-hour-old colonies were used to inoculate broth supplemented
with sodium taurocholate and antibiotics for C. difficile as in previous studies (Wilson et
al. 1982; Levett and Margaritis-Bassoulis 1985; Aspinall and Hutchinson 1992;
Rodriguez-Palacios et al. 2007). Alternatively for some experiments, spores were
produced in antibiotic-free brain heart infusion broth supplemented with 5% yeast extract
(Acumedia, Lansing, MI) (BHI broth). Inoculated broths were incubated anaerobically at
37oC for 10 days to allow C. difficile to grow and produce spores following nutrient
exhaustion of the media. In all cases spores were harvested (7,000 x g for 10 minutes)
after 10 days of incubation and re-suspended in plastic transparent tubes using phosphate
buffered saline (PBS), and left on a bench top to age at room temperature (23 ± 3oC). If
the spores were heat tested within 2 days of production, these spores were considered
fresh spores. But to simulate aging conditions likely to occur in food, feces, or the
environment, some spores were left at this temperature an additional 20 or 52 weeks:
these are referred as aged spores.
Thermal inhibition in liquid media
Thermal testing was conducted in PBS containing the spore suspensions using
conventional PCR microtubes (210 ± 10 µl) which were submerged in automatic water
baths synchronized to the desired temperatures using the average of two external
thermometers (EU620-0918, WVN International). Temperatures tested were 63 and
85oC, which encompass the thermal spectrum of widely publicized food safety guidelines
for household minimal internal cooking of meat meals (USDA ; FSA 2007; CFIA 2010).
Based on pilot trials, timing started 10-15 seconds after the PCR tubes were immersed in
183
the water bath, when the temperature inside tested tubes reached target temperatures.
Then, at specific time intervals (0, 15 and 30 minutes for 63oC; and 0, 3, 6, 9, 12, 15 and
30 minutes for 85oC) the heated aliquots were removed, cooled in icy water and
enumerated using the spread plating method on blood agar. Colonies were plated in
duplicate and counted after 72 h of anaerobic incubation at 37oC. To confirm that
survivor colonies corresponded to the strains tested, PCR ribotyping (Bidet et al. 1999;
Rodriguez-Palacios et al. 2006; Rodriguez-Palacios et al. 2007) was used to verify the
genotypes of the inoculum and heated recovered organisms were identical. The D-values
(decimal reduction times; i.e., time needed to inhibit 90% of spores) were determined
using at least two linear equation fits of the inhibition curves excluding the graph
shoulders, i.e., the initial and final time points (Juneja et al. 2001). Because precedent
enumeration studies of C. difficile in feces indicate that spore recoverability during
storage decreases over time (Weese et al. 2000), we also determined if refrigeration
(4±2oC) would reduce the recoverability of heated and nonheated spores over time. Thus,
replicate aliquots of all non heated strains, and replicate aliquots of heated PCR ribotype
078 strain (85oC) at multiple time points (3, 6, 9, 12 and 15 minutes) were enumerated
after 14 and 18 days of refrigeration. Paired analysis and D-values derived from PCR
ribotype 078 aliquots were used to assess variability.
Mechanism of thermal inhibition
To determine if heat inactivation of C. difficile spores was due to impaired
germination, or inhibition of cell division following germination as it has been reported
for Bacillus sp. (Ghosh and Setlow 2009; Ghosh et al. 2009), we first used germination
184
studies to determine if heated spores could germinate in a nutritious broth, and then,
followed it with outgrowth studies to determine if these (germinated) spores would then
be able to divide to produce visible colonies on blood agar. Normally, during early stages
of germination, often within 30 minutes of incubation, most spores release dipicolinic
acid and calcium from their core, changing its refractancy and reducing the optical
density of the liquid medium that contains them (Paredes-Sabja et al. 2008; ParedesSabja et al. 2008). If germination pathways are affected with heat, no germination-related
optical density (OD600) changes would occur in the incubation broth. If germination
occurred normally, and the cell had normal division, the formation of colonies in agar
surfaces would also be as expected. However, if heat impaired cell division mechanisms,
a reduced number of colonies would then be visible or not at all apparent on the agar.
In brief, the complementary germination-outgrowth experiments were conducted
as follows. First, aged spores from C. difficile PCR ribotypes 027, 078 and 077, and
toxigenic ATCC 9569 strain were resuspended and heat shocked individually in BHI
broth at 63, 71 and 85oC for 15 minutes. To prevent inadvertent spore germination,
spores suspensions were maintained in icy water in a cold room set at 4oC. Spore
suspensions were seeded in replicates (210 µl; OD600 0.78 ± 0.03) in microtitre plate
(Corning Incorporated, Corning, NY) wells to monitor the OD600 during subsequent
anaerobic incubation in the BHI at 37oC. Although germination of C. difficile spores is
not impaired with room air (Jump et al. 2007), the wells were layered with 40 µl of 0.2
µm-filtered pre-reduced heavy mineral oil (Rafii and Park 2005) to minimize potential
confounding effects due to air diffusion into the BHI during readings outside of the
185
anaerobic chamber (<4 minutes). An ELISA reader (Spectra max 340PC, Molecular
Devices, Sunnyvale, CA) was used to assess the OD600 changes of the spore suspensions
after heat shock (time 0) and thereafter at 0.5, 1, 1.5, 2, 3, 4, 5, 6 and 7 hours of anaerobic
incubation. Nonheated spores (controls) were concurrently tested to document normal
germination OD600 changes. The mean OD600 time point values were derived from at least
4 replicates. To quantify the ability of heated spores to divide and produce countable
colonies (outgrowth study), spores were plated and enumerated on blood agar before and
after incubation in BHI. However, because the production of daughter vegetative cells
from spores could have been possible during the 7 h of incubation, the BHI broths were
treated with 96% ethanol (1:1, v/v, 30 min) (Clabots et al. 1989), washed and
resuspended in PBS prior to enumeration to eliminate potential vegetative cells and infer
the number of dormant spores capable of normal division.
Thermal inhibition in ground beef and gravy
Thermal resistance data from PBS or culture broth studies should not be directly
extrapolated to food matrices due to the physicochemical complexity of foods (Juneja et
al. 2001). Therefore, we also quantified the extent of spore inhibition using retail ground
beef (2-kg packages purchased at a local retailer) and a beef gravy model (5% beef
extract, 1.7% soluble starch, 1.5% proteose peptone and 0.5% yeast extract) (Juneja et al.
1998). In these experiments, the strains studied in the abovementioned germinationoutgrowth study were tested following a protocol for E. coli O157 reported by Juneja et
al (Juneja et al. 1998) with minor modifications.
186
In brief, Juneja et al (Juneja et al. 1998) used a mixture of four strains of E. coli (a
―cocktail‖), experimentally inoculated single random aliquots of ground beef (3 grams),
and individually heated those aliquots using Whirlpack™ bags and a water bath. Here,
we used the same Whirlpack bag technique, but we assessed the effect of heating for
1:1:1:1 (107 spores · ml-1) mixture of four C. difficile genotypes (PCR ribotype 027, 078
and 077, and an ATCC toxigenic strain). We determined the D-values in a complete
block design using three food matrices (i.e., a nutritious fat-free gravy model, retail ‗lean‘
7% fat ground beef, and retail 30% fat ground beef) (Juneja et al. 2001) and four heating
temperatures (63, 71, 85, and purportedly more effective 96oC). The 96oC was chosen as
a sub-boiling water temperature for the city of Wooster where the altitude adjusted
boiling temperature is about 98.8oC. These experiments were conducted in duplicate with
spore enumeration conducted in triplicate; the spores used here had been aged for 52
weeks at the time of testing. We used Juneja‘s gravy model to control and quantify the
effect of spore entrapment within the beef, which shrinks with heating. Following
experimental inoculation, the 3-gram beef aliquots were heated for 0, 2, 5, 7.5, 10, 15, 20,
and 30 minutes, and then mixed with 3 ml of PBS (1:1, v/w), stomached for one minute
and enumerated as abovementioned for D-value determination. To control for the natural
level of contamination of retail meats with C. difficile, we also quantified the spore
concentration of noninoculated nonheated aliquots of both ground beef types purchased.
187
Statistics
Univariate and multivariable analysis of repeated measures of transformed (log10)
and normalized enumeration data was conducted as indicated by Davis (Davis 2002),
using STATA (College Station, TX, v.10) and Minitab 15 (Upper Saddle River, NJ).
Comparisons among D-values were analyzed with generalized linear regression; relevant
descriptive statistics and 95% confidence intervals (95% CI) with pooled errors are
reported. Percentages of effect were computed as 10 exponentiated to the log10 unit of
interest.
Results
Thermal inhibition in liquid media
Controlling for strain variability, heating C. difficile spores in PBS at 63oC for 30
minutes was not as effective as heating at 85oC in reducing the spore counts (Fig. 18).
However, the effect of heating at 63oC depended on whether the spores were fresh or
aged. While heating fresh spores at 63oC resulted in a steady significant inhibitory effect
overtime (median reduction of 0.37 log10, i.e., 57.3% of spores after 30 minutes, 95% CI
= 0.13, 0.67 log10; GLM, P = 0.002), heat treatment of aged spores resulted in a 30%
increased recovery compared to baseline counts within 15 minutes of heating (GLM, P =
0.049). Heating at 85oC resulted in much faster inhibition for all strains (99.9998% of
spores after 15 minutes) with estimated D-value at 85oC ranging from 6.0 to 8.5 minutes,
irrespective of spore aging (median reduction of 5.6 log10, 95% CI = 4.48, 5.94 log10;
GLM, P < 0.0001; Fig. 18).
188
When the effect of refrigeration on the recovery of nonheated aged spores was
assessed, storage significantly reduced the spore counts by day 18 (median reduction of
0.28 log10, i.e., 42.5% of spores, 95% CI = 0.045, 0.915 log10; Wilcoxon signed, P =
0.014; Fig. 19A). At the strain level, two isolates had a 2-log10 reduction by days 14 and
18. Complementarily, repeated quantification of heated aliquots of PCR-ribotype 078 at
85oC showed that refrigeration did not affect spore enumeration and the D-values therein
derived (Fig. 19B).
Heat treatment at 85oC affects cell division but not spore germination
Assessment of heat-shocked spores (15 minutes) in BHI during 7 hours of
incubation demonstrated that the OD600 germination curves of all heated strains were
normal as they were similar to the germination curves of nonheated controls (GLM, P >
0.1; Fig. 20A). However, when we assessed the effect of heat shock on cell division,
determined as the number of countable colonies on agar after normal germination
(curves), multivariable analysis indicated that 15 minutes of heat shock at 85oC was the
only temperature that significantly reduced the number of countable colonies. On
average, 85oC induced a 4.2 log10 reduction (99.994% of spores) compared to nonheated
controls (95% CI= 3.9, 4.5; GLM, P < 0.0001; Fig. 20B).
Thermal inhibition at 85oC can revert during incubation in broth
Paired analysis (before vs. after 7 hours of incubation in BHI) of ethanol-resistant
spores showed that 85oC resulted in complete inhibition of most isolates immediately
after 15 minutes of heat shock (Fig 21A, panel 85oC, 0h). However, when those heat
shocked spores (85oC) were left incubating in BHI, the number of countable colonies
189
after 7 hours of incubation increased significantly (Fig. 21A, panel 85oC, 7h) indicating
that the thermal inhibitory effect observed immediately after heat shock can be reverted
during subsequent incubation. In contrast, heat shock with 63 and 71oC increased the rate
of spore recovery compared to nonheated controls immediately after heat shock. These
two temperatures also enhanced the rate of spore germination (assessed as the number of
ethanol susceptible cells) during the subsequent 7 h of incubation in BHI. The highest
rate of germination was observed with 63oC (0.4 log10 increase, 95% CI = 0.1, 0.9; GLM,
P < 0.01; Fig. 21B).
Thermal inhibition in ground beef and gravy
Quantification of C. difficile contamination in the ground beef packages indicated
that natural contamination was present in both products (different processing plants),
varied within the packages, and ranged from 0 to 0.5-3.3 log10 · gr-1. Mathematically, the
highest level of natural contamination corresponded to less than 0.02% of the spore
concentration achieved with experimental contamination and was deemed minimal to
bias the estimation of the D-values. Multivariable analysis of D-value data indicated that
85 and 96oC have the highest inhibitory effect against C. difficile experimentally
inoculated in beef matrices (GLM, P< 0.001; Fig 22). In contrast, 63 and 71oC had
comparable and a minimal effect in inhibiting the C. difficile mixture (see D-values in
supplementary table S2). Although significant differences were observed among the
matrices tested (which varied in up to 30% in fat content), the direction of such effects
varied with temperature indicating matrix-temperature interactions (Fig 22).
190
Discussion
This study quantified the inhibitory effect of moist heat at 63, 71, 85 and 96oC on
C. difficile spores in liquid media and in ground beef. Our quantitative report also
evaluated for the first time fresh and aged spores, the latter which are likely to occur
naturally in the environment and within contaminated foods. To contribute to food safety
validation protocols, temperatures tested included widely publicized food safety
guidelines of minimal internal cooking temperatures for meat meals (USDA ; FSA 2007;
CFIA 2010). Briefly, publicized temperatures range from 63oC (for 30-minute batch milk
pasteurization, and for preparation of fish, and steaks/roasts of beef, veal and lamb as
―medium-rare‖) to 85oC for whole poultry and turkey meals. However, cooking at 71oC
is often cited in the ―safe handling instructions‖ of most retail meat labels, and is
indicated for ―well-done‖ purposes of beef, veal and lamb steaks, for ground beef, and for
pork products (USDA ; CIPHI 2003; Silvestre et al. 2008; CFIA 2010; FDA 2010). Of
concern, recent data indicate that 71oC results only in a 2 log10 reduction after heating C.
difficile spores for 2 hours (Rodriguez-Palacios et al. 2010).
Under the conditions of our experiments, moist heat treatment at 85oC,
irrespective of spore aging and test media, markedly inhibited (up to 6 log10) spores from
most C. difficile strains within 15 minutes of treatment. Considering that the
concentration of C. difficile spores in naturally contaminated retail raw meat products
may be of less than 3.3 log10 · g-1 of product (Weese et al. 2009), cooking at least 85oC
could markedly reduce human exposure to C. difficile. Here we found that heating
191
aliquots containing less than 4 log10 at 85oC in liquid media always yielded no cultivable
spores after 15 minutes of heating. For other foodborne pathogens such as Salmonella
spp., it is known that there is a greater risk of pathogen persistence in the final ready-toeat meat products after heat treatment if the initial pathogen load in the raw products is
high (USDA 1999). However, it is unknown whether C. difficile proliferates in retail
meals and if bacterial proliferation reaches more than 4 log10 · g-1. Likewise, it is
uncertain to what extent spore aging within food matrices parallels the aging process
induced in culture broths in our study and the subsequent magnitude of thermal tolerance.
To date most studies assessing C. difficile germination and sporulation have used
freshly prepared spores leaving aged spores largely uncharacterized. As aging of C.
difficile spores could occur in the environment or in foods, studying the effect of heat is
critical for understanding spore survivability patterns. Here, we showed that 85oC was
markedly inhibitory for aged and fresh spores; however, of food safety relevance was the
finding that 63 and 71oC enhanced germination in aged spores. The food safety concern
resides in that heat treatment is a well known germination factor that triggers the
synchronic activation of most spore-forming organisms associated with food, and that
spore germination during the subsequent chilling period is possible in other clostridia
(Sanchez-Plata et al. 2005). Furthermore, sublethal heat shock of spores in other
clostridia is known to significantly enhance the thermal tolerance of surviving spores
(Juneja et al. 2003).
Storage under refrigeration resulted in lower spore counts over time. Although
there are no comparable studies assessing aging of C. difficile in pure suspensions under
192
refrigeration, our finding is comparable to that of a study assessing the C. difficile
recoverability in inoculated feces (Weese et al. 2000). In that study, the rate of isolation
of C. difficile decreased over time (Weese et al. 2000). Although, another similar study
stated there was little effect (Freeman and Wilcox 2003), our findings indicate that
factors associated with decreased recovery (e.g., super dormancy) might be enhanced
during storage. Superdormancy is a recently described characteristic in Bacillus species
(Ghosh and Setlow 2009), and a rising concern for the food industry as superdormant
Bacillus spores have higher moist-heat resistance, lower water core content than regular
spores, and can recover from heat-stress injuries during refrigeration (Ghosh et al. 2009;
Ghosh and Setlow 2010; Wei et al. 2010). Understanding the properties of
superdormancy in Bacillus spp. could also indicate ways to reduce the potential impact of
C. difficile on food safety.
To date, some epidemiological studies assessing C. difficile contamination of
foods have used thermal treatment (80oC for 10 minutes) to reduce contaminants and
facilitate spore recovery (Songer et al. 2009). However, the observed inhibition with
85oC in our study indicates that culture-based studies based on thermal methods for spore
selection that use between 71 and 85oC might introduce selection bias, favoring more
thermo-resistant spores. Hence, it is advisable to use temperatures below 71oC, being
63oC a pasteurization temperature potentially beneficial for spore selection in diagnostic
settings.
Thermal inhibition at 85oC is a promising preventive measure since cooking
guidelines at this temperature are readily available for preparation of whole turkey meals,
193
and could be easily adopted. However, this study indicates that more than a few seconds
are required to lower the number of C difficile spores to the desired 5-6 log10 reduction. It
is expected that extended cooking at 85oC could maintain the nutritional quality of the
cooked foods, but it is uncertain how extended heating would impact the nutritional and
palatability characteristics of non-turkey meals. A balance between food safety risks,
nutritional value and food quality needs to be taken into consideration. The marked
inhibitory effect of 96oC here documented, further highlight the potential benefits of
revising cooking time-temperature recommendations, at least for individuals at risk. This
study generated the first D-values for C. difficile in a food matrix. These data can be used
for policy and management decision making relevant to food safety and infection control
strategies to disinfect hospital reusable materials (Alfa et al. 2008).
Comparing our estimated D-values at 85oC (<8 minutes) to those reported for
other clostridia, it becomes evident that C. difficile appears to be more susceptible to heat.
For instance, the D-values at 100oC for C. perfringens, the most common spore-forming
foodborne pathogen in the USA (Scallan et al. 2011), in beef gravy varies from 16 to 21
minutes (Juneja et al. 2003), and in distilled water the D-value at 85oC was 55 minutes
(Heredia et al. 1997). Thus, cooking guidelines for industrial purposes against C.
perfringens (Crouch and Golden 2005) may also be a good basis for the refinement of
household guidelines against C. difficile.
In summary, we quantified the inhibitory effect of heat on C. difficile spores,
estimated D-values (63-96oC) with confidence intervals for ground beef and gravy, and
mechanistically determined that the inhibitory effect of heat shock at 85oC, the highest
194
cooking recommended internal minimal temperature, is due to inhibition of cell division
but not germination, and that it can be reversible. Spore germination and recoverability
was enhanced with 63oC, especially in aged spores. However, with concern, but as
expected, we documented the sub-lethal effect of 71oC and its moderate effect to enhance
germination. Thermoresistance experimentation also indicated that spore superdormancy,
an increasingly recognized characteristic in Bacillus sp., is likely to occur in C. difficile
as spores age. Ensuring food is heated to more than 85oC would be a simple and
important intervention to reduce the risk of inadvertent ingestion of C. difficile spores.
Acknowledgements
This study was partly supported by state and US federal funds allocated to the
Ohio Agricultural Research and Development Center, The Ohio State University,
Wooster, Ohio. ARP is a research fellow supported by the Public Health Preparedness for
Infectious Diseases (Targeted Investments in Excellence) Program, The Ohio State
University. Special thanks to Dr. Michele Williams for comments during the final
revision of the manuscript. Authors declare no conflicts of interest.
195
Normalized values of mean LOg CFU
vs time values of mean LOg CFU vs time
Normalized
1.2
6
Individual Value Plot of VALUESLogNorm
vs min
Value Plot of VALUESLogNorm
Individual
Individual
VALUESLogNormvs
vs min
min
Fresh
spores
Aged
spores
9 12 15
18 21 24 27 30 95%
0 CI
3 for6the9Mean
12 15 .
.
.
. 30
1.2
0.8
oC
6363
oC
1.0
0.8
0.6
0.6
0.4
0.4
0.2
0.2
95%
95% CI
the Mean
Mean
CI for
for the
85
63
Pairedtt,,PP==0.009
Paired
0.009
1.4
1.0
Normalized Log CFU - spores
Normalized CFU
[ log10 at time
i *1 ] / log10 at time 0
norm_value
63
3
oCo
8585
C
0.0
0
3
6
9
ariable: temp_1
0.0
12 15 18 21 24 27 30 0
time
9
12
3 0 6 3 9 612 15
. 15
.
. . . . 30 .
minutes
time
minutes of heating
85
Pairedt,t,PP==0.005
0.005
Paired
1.4
1.4
1.4
NormalizedLog
Log CFU
CFU -- spores
spores
Normalized
0
.
Pairedt,t,PP==0.027
0.027
Paired
1.2
1.2
1.2
oC
oC
6363
1
1.0
1.0
1.0
0.8
0.8
0.8
0.6
0.6
0.6
0.4
0.4
0.4
0.2
0.2
0.2
0.0
0.0
0.0
30
85
85ooCC
0
00
3
33
6
66
9
99
12
12
12
15
15
15
.
..
.
..
minutes
minutes
of heating
minutes
minutes
.
..
.
..
0
30
30
30
0
00
Panel variable: temp_1
Figure 18. Effect of moist heat in phosphate buffered saline at 63 and 85oC on viability of
fresh (1-week-old) and aged (20-week-old) spores of C. difficile.
Open symbols represent isolate means; grey squares with interval bars represent 95% CI
of the means.
196
3
33
6
66
9
99
12
12
12
15
15
15
B
1
0
Reduced counts
-1
-2
(day 18)-(day 0)
(day 14)-(day 0)
4
a y = -0.1373x + 3.514
3.5
R² = 0.886
3
b y = -0.1017x + 3.353
R² = 0.907
2.5
90%
reduction
One-Sample Wilcoxon signed rank,
P = 0.014
log10 CFU · g-1
2
Increased counts
Spore cultivability relative to baseline
log10 CFU · g-1
A
2
c y = -0.1529x + 3.531
R² = 0.999
1.5
d y = -0.1527x + 3.648
1
R² = 0.826
D-value
~6 min
0.5
0
(day 18)-(day 14)
0
3
6
9
12
15
18
21
24
27
30
minutes of heating at 85o C
days compared
Figure 19. Effect of storage on cultivability of aged C. difficile spores during 18 days of
refrigeration at 4ºC.
A) Nonheated spores illustrate reduced cultivability over time (12 strains; triangles and
interval bars represent means and 95% CI). B) Heated spores at 85oC indicate
reproducible estimation and cultivability (PCR ribotype 078 strain; Regression fits shown
for 0, 3, 14 and 18 days).
197
oC
62
63oC
oC
71
71oC
oC
85
85oC
B
1
0.95
0.9
PCR type 078
0.85
0.8
0.75
0.75
0.7
0.7
0
100
100
200
200
300
300
400
400
minutes of incubation at 37oC after 15 min of heat shock
Number of ethanol resistant cells after
7 h of incubation – log10 CFU · g-1
Not heated
1.05
ATCC
Normalized OD600 changes of BHI
spore suspensions
A
6
GLM, P < 0.001
5
4
3
2
1
0
oC:
temp
4 62 71 85
strain_ C PCR type
R1 078
4 62 71 85
R3 077
PCR type
4 62 71 85
R5
ATCC
4 62 71 85
R6 027
PCR type
Figure 20. Germination and outgrowth of C. difficile
Germination and outgrowth studies on C. difficile to determine if 15 minutes of heat
shock impairs spore germination or the subsequent cell division. A) Germination study.
Optical density (OD600) reduction of heavily inoculated BHI spore suspensions, incubated
at 37oC for 7h, is similar for heated and nonheated spores indicating normal germination
of heated spores. Only PCR ribotype 078 and the ATCC strain are shown; means ± SD.
B) Outgrowth study. Enumeration of ethanol resistant cells from aliquots in Figure 3A
after 7h of incubation. Countable colonies on blood agar indicate adequate cell division.
Reduced counts after heat shock with 85oC (and normal germination in Fig 3A) indicate
impaired cell division. Dots represent replica averages; means±95% CI (solid grey bars).
198
Heat shock temperature (15 minutes, prior to incubation in BHI)
7
4oC
63oC
71oC
n=26
n=18
n=18
7
7
85oC
6
6
6
5
5
5
5
4
4
4
4
3
3
3
3
2
2
2
0
y = 0.067x + 4.407
Slope 95%CI: -0.09, 0.22
0h
y = -1.072x + 6.248
Slope 95%CI: -1.26,-0.88
1
0
7h
0
2
y = -0.444x + 5.227
Slope 95%CI: -0.78,-0.11
1
0
0h
7h
One-ANOVA F test, P < 0.0001
2
enhanced
1
0
reduced
-1
-2
y = 0.201x - 0.0728
Slope 95%CI: -0.15, 0.55
Not heated
4C (4oC)
63oC
63C
71oC
71C
Germination in BHI after 7h of incubation
B
1
0h
7h
0h
7h
Hours of incubation in BHI after heat shock
ethanol susceptible cells - log10 CFU · g-1
1
n=18
7
6
Spore germination relative to baseline
Ethanol-resistant spores
log10 CFU · g-1
A
85oC
85C
Heat shock temperature
(15 min prior incubation)
Figure 21. Effect of heat shock on immediate cultivability and subsequent germination of
C. difficile spores determined by enumeration of ethanol-resistant cells before and after
7h of incubation in BHI.
A) Line plots represent paired enumeration of heat-shocked and nonheated individual
spore replicas; four genotypes. Linear equations and extended dashed lines represent
average effects (95% CI of means). The highest and steepest dashed line (63oC) indicates
immediate enhanced cultivability and subsequently increased spore germination. 85oC
shows immediate inhibition and subsequent increased cultivability.
B) Spore germination relative to baseline. Normalized paired differences (before and
after BHI incubation). Heat shock with 63 and 71oC enhanced germination in spores that
otherwise would have remained dormant after adding 7h of incubation in BHI and 72 h in
blood agar at 37oC (here referred to as superdormant spores). Dots represent average
difference per replica; means±95% CI (See Table 10).
199
200
e
minutes
Minutes of heat to inhibit 90% of spores
D-value (minutes to inhibit 90% of spores)
(D- values for Clostridium difficile)
30% fat - Ground beef
0% fat - Gravy
3% fat - Lean ground beef
150
e
5
4
d
d
e
3
2
100
d
e
1
d
temp
50
temp
d
d,e
d,e
0
85
e
95
0
63
a
71
a
85 b
95 c
Temperature ( oC)
Figure 22. Minutes of heat needed to inhibit a mixture of aged C. difficile spores heated
in three food matrices.
Dots indicate individual D-value estimates; means±95% CI. Identical superscripts
connect variables with nonsignificant differences (i.e., GLM, P > 0.05), whereas different
superscripts indicate significance. See also supplementary table S2 for coefficient
estimates.
200
Pearson correlation (Opt:7.00%) [0.0%-100.0%]
100
PCR-Bidet
80
60
PCR-Bidet
A
D
PCR 078
PCR 033
PCR 017
K
I
M
C
H
F
K
G
E
ATCC
PCR M26
PCR M31
PCR 027
PCR 077
PCR 014
Supplementary material
Figure 23. Genetic diversity of representative isolates tested.
All strains were toxigenic except PCR ribotype M26 which lacks pathogenicity locus
(Rodriguez-Palacios et al. 2006; Rodriguez-Palacios et al. 2007).
201
Table 9. Coefficient estimates of heat-shock enhancement of spore germination after 7h
of incubation in nutritious BHI broth
Median
95%CI of median
Heat-shock treatment
difference
log10 CFU · ml-1
log10 CFU · ml-1
P*
Univariate
4oC, Not heated
-0.04
-0.21, 0.12
referent
63oC
1.04
0.84, 1.28
<0.0001*
71oC
0.52
0.13, 0.80
0.021*
85oC
-0.12
-0.62, 0.00
0.272
63 minus 4oC
1.12
0.88, 1.34
<0.0001*
71 minus 4oC
0.49
0.17, 0.81
0.027*
85 minus 4oC
-0.17
-0.51, 0.18
1.000
71 minus 63oC
-0.63
-0.99, -0.26
0.005*
85 minus 63oC
-1.29
-1.68, -0.89
<0.0001*
85 minus 71oC
-0.65
-1.14, -0.18
0.004*
Pair-wise differences
* Null hypothesis: median effect difference equals zero; significance held at P < 0.05
derived from one-sample Signed Wilcoxon Rank and pair-wise Bonferroni significance
tests. Univariate analysis compares value to referent control, i.e., 4oC.
202
Table 10. Minutes of heat needed to inhibit 90% of spores (D-values) estimated for a
1:1:1:1 mixture of C. difficile of PCR ribotypes 027, 078, 077 and ATCC 9689 strains in
three food matrices
D-value means from two experiments ± SD
(95% confidence intervals)
o
63 C
71oC
85oC
96oC
(145oF)
(160oF)
(185oF)
(205oF)
D-value means
Gravy, 0% fat
Lean ground beef, 3% fat
Ground beef, 30% fat
55.9±21.8
56.8±30.5
2.51±0.43
0.59±0.32
(21.3 to 90.6)
(18.9 to 94.7)
(1.83 to 3.2)
(0.07 to 1.1)
99.4±62.3
71.2±38.7
3.0±1.11
0.62±0.29
(22 to 176)
(9.6 to 132)
(1.83 to 4.17)
(0.15 to 1.08)
45.8±10.8
46.8±10.9
3.27±1.02
1.19±0.67
(28.7 to 63)
(33.1 to 60.4)
(2.2 to 4.33)
(0.49 to 1.89)
69.6±45.4
57.3±27.8
2.97±0.94
0.85±0.55
(42.1 to 96.9)
(41.3 to 73.4)
(2.48 to 3.48)
(0.53 to 1.17)
-0.018515
-0.025223
-0.36619
-1.6678
Cumulative D-value
mean
95% CI
Linear function, statistics
Slope
mean
S.D.
0.008362
0.018412
0.10826
1.0015
95% CI, lower limit
-0.023568
-0.036349
-0.42388
-2.2460
95% CI, upper limit
-0.013462
-0.014097
-0.30850
-1.0896
0.65±0.23
0.67±0.28
0.86±0.14
0.89±0.15
13
13
16
14
r2, mean ± SD*
n= estimated D-values
*r2 is a measure of goodness-of-fit of the linear regressions.
203
References
Alfa, M. J., N. Olson and L. Buelow-Smith (2008). "Simulated-use testing of bedpan and
urinal washer disinfectors: evaluation of Clostridium difficile spore survival and
cleaning efficacy." American Journal of Infection Control 36(1): 5-11.
Aspinall, S. T. and D. N. Hutchinson (1992). "New selective medium for isolating
Clostridium difficile from faeces." Journal of Clinical Pathology 45(9): 812-814.
Bakri, M. M., D. J. Brown, J. P. Butcher, et al. (2009b). "Clostridium difficile in ready-toeat salads, Scotland." Emerging Infectious Diseases 15(5): 817-818.
Bidet, P., F. Barbut, V. Lalande, et al. (1999). "Development of a new PCR-ribotyping
method for Clostridium difficile based on ribosomal RNA gene sequencing."
FEMS Microbiology Letters 175(2): 261-266.
CFIA. (2010). "Canadian Food Inspection Agency. Food Thermometer Food Safety Tips:
Preventing foodborne illness." Retrieved July 14, from
http://www.inspection.gc.ca/english/fssa/concen/tipcon/thermoe.shtml.
CIPHI. (2003). "Canadian Institute of Public Health Inspectors. Food safety factors for
cooking Christmas Turkey " Retrieved December 19, from
http://www.ciphi.ca/pdf/turkey.pdf.
Clabots, C. R., S. J. Gerding, M. M. Olson, et al. (1989). "Detection of asymptomatic
Clostridium difficile carriage by an alcohol shock procedure." Journal of Clinical
Microbiology 27(10): 2386-2387.
Crouch, E. and N. Golden. (2005). "Risk Assessment for Clostridium perfringens in
Ready-to-Eat and partly cooked meat and poultry products. USDA - Food Safety
Inspection Service." Retrieved September, 1, from
http://www.fsis.usda.gov/PDF/CPerfringens_Risk_Assess_Sep2005.pdf.
204
Davis, C. S. (2002). Statistical methods for the analysis of repeated measurements. New
York, Springer: 415.
FDA. (2010). "United States Food and Drug Administration. Grade "A" Pasteurized Milk
Ordinance - Public Health Service/Food and Drug Administration (2009
Revision)." Retrieved April 26, from
http://www.fda.gov/downloads/Food/FoodSafety/ProductSpecificInformation/MilkSafety/NationalConferenceonInterstateMilkShipmentsN
CIMSModelDocuments/UCM209789.pdf.
Freeman, J. and M. H. Wilcox (2003). "The effects of storage conditions on viability of
Clostridium difficile vegetative cells and spores and toxin activity in human
faeces." Journal of Clinical Pathology 56(2): 126-128.
FSA. (2007). "Food Standards Agency, United Kingdom. Advisory Committee on the
Microbiological Safety of Food: Report on the Safe Cooking of Burgers."
Retrieved July 31, from
http://www.food.gov.uk/news/newsarchive/2007/jul/burgers and
http://www.food.gov.uk/multimedia/pdfs/acmsfburgers0807.pdf.
Ghosh, S. and P. Setlow (2009). "Isolation and characterization of superdormant spores
of Bacillus species." Journal of Bacteriology 191(6): 1787-1797.
Ghosh, S. and P. Setlow (2010). "The preparation, germination properties and stability of
superdormant spores of Bacillus cereus." J Appl Microbiol 108(2): 582-590.
Ghosh, S., P. Zhang, Y. Q. Li, et al. (2009). "Superdormant spores of Bacillus species
have elevated wet-heat resistance and temperature requirements for heat
activation." J Bacteriol 191(18): 5584-5591.
Heredia, N. L., G. A. Garcia, R. Luevanos, et al. (1997). "Elevation of the heat resistance
of vegetative cells and spores of Clostridium perfringens type A by sublethal heat
shock." Journal of food protection 60(8): 998-1000.
Jump, R. L. P., M. J. Pultz and C. J. Donskey (2007). "Vegetative Clostridium difficile
Survives in Room Air on Moist Surfaces and in Gastric Contents with Reduced
Acidity: a Potential Mechanism To Explain the Association between Proton Pump
205
Inhibitors and C. difficile-Associated Diarrhea?" Antimicrob. Agents Chemother.
51(8): 2883-2887.
Juneja, V. K., B. S. Eblen and H. M. Marks (2001a). "Modeling non-linear survival
curves to calculate thermal inactivation of salmonella in poultry of different fat
levels." International journal of food microbiology 70(1-2): 37-51.
Juneja, V. K., B. S. Eblen and G. M. Ransom (2001b). "Thermal Inactivation of
Salmonella spp. in Chicken Broth, Beef, Pork, Turkey, and Chicken:
Determination of D- and Z-values." Journal of Food Science 66(1): 146-152.
Juneja, V. K., P. G. Klein and B. S. Marmer (1998). "Heat shock and thermotolerance of
Escherichia coli O157:H7 in a model beef gravy system and ground beef." J Appl
Microbiol 84(4): 677-684.
Juneja, V. K., J. S. Novak, L. Huang, et al. (2003). "Increased thermotolerance of
Clostridium perfringens spores following sublethal heat shock." Food Control
14(3): 163-168.
Levett, P. N. and G. Margaritis-Bassoulis (1985). "Enrichment media for isolation of
Clostridium difficile from faeces." Eur J Clin Microbiol 4(2): 135-136.
McFarland, L. V. (2008). "Update on the changing epidemiology of Clostridium difficileassociated disease." Nat Clin Pract Gastroenterol Hepatol 5(1): 40-48.
Meisel-Mikolajczyk, F., E. Kaliszuk-Kaminska and G. Martirosian (1995). "[Study of the
thermoresistance of Clostridium difficile spores]." Med Dosw Mikrobiol 47(3-4):
177-181.
Paredes-Sabja, D., C. Bond, R. J. Carman, et al. (2008a). "Germination of spores of
Clostridium difficile strains, including isolates from a hospital outbreak of
Clostridium difficile-associated disease (CDAD)." Microbiology 154(8): 22412250.
Paredes-Sabja, D., J. A. Torres, P. Setlow, et al. (2008b). "Clostridium perfringens spore
germination: characterization of germinants and their receptors." J Bacteriol
190(4): 1190-1201.
206
Pepin, J., M. E. Alary, L. Valiquette, et al. (2005). "Increasing risk of relapse after
treatment of Clostridium difficile colitis in Quebec, Canada." Clin Infect Dis
40(11): 1591-1597.
Rafii, F. and M. Park (2005). "Effects of gyrase mutation on the growth kinetics of
ciprofloxacin-resistant strains of Clostridium perfringens." Anaerobe 11(4): 201205.
Richter, S. H., J. P. Garner and H. Wurbel (2009). "Environmental standardization: cure
or cause of poor reproducibility in animal experiments?" Nat Methods 6(4): 257261.
Rodriguez-Palacios, A., R. J. Reid-Smith, H. R. Staempfli, et al. (2009a). "Possible
seasonality of Clostridium difficile in retail meat, Canada." Emerg Infect Dis
15(5): 802-805.
Rodriguez-Palacios, A., R. J. Reid-Smith, H. R. Staempfli, et al. (2010). "Clostridium
difficile survives minimal temperature recommended for cooking ground meats."
Anaerobe 16: 540-542.
Rodriguez-Palacios, A., H. R. Staempfli, T. Duffield, et al. (2007a). "Clostridium difficile
in retail ground meat, Canada." Emerg Infect Dis 13(3): 485-487.
Rodriguez-Palacios, A., H. R. Stampfli, T. Duffield, et al. (2006). "Clostridium difficile
PCR ribotypes in calves, Canada." Emerg Infect Dis 12(11): 1730-1736.
Rupnik, M. (2007). "Is Clostridium difficile-associated infection a potentially zoonotic
and foodborne disease?" Clin Microbiol Infect 13(5): 457-459.
Sanchez-Plata, M. X., A. Amezquita, E. Blankenship, et al. (2005). "Predictive model for
Clostridium perfringens growth in roast beef during cooling and inhibition of
spore germination and outgrowth by organic acid salts." J Food Prot 68(12):
2594-2605.
Scallan, E., R. M. Hoekstra, F. J. Angulo, et al. (2011). "Foodborne illness acquired in
the United States--major pathogens." Emerg Infect Dis 17(1): 7-15.
207
Silvestre, D., P. Ruiz, C. Martinez-Costa, et al. (2008). "Effect of Pasteurization on the
Bactericidal Capacity of Human Milk." J Hum Lact 24(4): 371-376.
Songer, J. G., H. T. Trinh, G. E. Killgore, et al. (2009). "Clostridium difficile in retail
meat products, USA, 2007." Emerg Infect Dis 15(5): 819-821.
USDA "United States Department of Agriculture - Food Safety and Inspection Service.
Food Safety Education - Is it done yet? From:
http://www.fsis.usda.gov/is_it_done_yet/brochure_text/index.asp#SMIT
(Accessed January 12, 2011)."
USDA. (1999). "Food Safety and Inspection Agency - United States Department of
Agriculture. Time-temperature tables for cooking ready-to-eat poultry products."
Retrieved February 7, from
http://www.fsis.usda.gov/oppde/rdad/fsisnotices/rte_poultry_tables.pdf.
Warny, M., J. Pepin, A. Fang, et al. (2005). "Toxin production by an emerging strain of
Clostridium difficile associated with outbreaks of severe disease in North America
and Europe." Lancet 366(9491): 1079-1084.
Weese, J. S., B. P. Avery, J. Rousseau, et al. (2009a). "Detection and enumeration of
Clostridium difficile in retail beef and pork." Appl Environ Microbiol 75(15):
5009-5011.
Weese, J. S., H. R. Staempfli and J. F. Prescott (2000). "Survival of Clostridium difficile
and its toxins in equine feces: implications for diagnostic test selection and
interpretation." J Vet Diagn Invest 12(4): 332-336.
Wei, J., I. M. Shah, S. Ghosh, et al. (2010). "Superdormant spores of Bacillus species
germinate normally with high pressure, peptidoglycan fragments, and bryostatin."
J Bacteriol 192(5): 1455-1458.
Wilson, K. H., M. J. Kennedy and F. R. Fekety (1982). "Use of sodium taurocholate to
enhance spore recovery on a medium selective for Clostridium difficile." J Clin
Microbiol 15(3): 443-446.
208
Wultanska, D., H. Pituch, P. Obuch-Woszczatynski, et al. (2004). "[Comparative study of
thermoresistance spores of Clostridium difficile strains belonging to different
toxigenicity groups]." Med Dosw Mikrobiol 56(2): 155-159.
Young, M. and J. Mandelstam (1979). Early events during bacterial endospore formation.
Advances in Microbial Physiology. A. H. Rose, Academic Press, Inc. London. 20
103-162.
209
Chapter 10 ― Cooking and the Selection of Emerging Pathogens: Clostridium
difficile
(Submitted for Publication)
Abstract
Despite our growing understanding of the role of cooking on human evolution,
little is known about the role of that practice on the selection (and evolution) of emerging
pathogens. Using prototype strains of internationally known enteric pathogen
Clostridium difficile, we experimentally demonstrated that cooking at recommended
minimum temperatures do not modify the population structure of the pathogen in foods.
However, higher temperatures, likely used in routine cooking, strikingly favored the
selection of newly emerging hypervirulent C. difficile strains (namely PCR ribotype 078).
Cooking, a unique trait of humans, may certainly contribute to the selection (and possibly
evolution) of emerging enteric pathogens.
Introduction
Cooking favored the evolution of hominids after the invention of fire by
promoting social transformations, maximizing food digestibility, and possibly by
reducing energetic expenditures associated with foodborne infections (Carmody and
210
Wrangham 2009). Despite our growing understanding of the role of cooking on human
evolution, little is known about the role of that practice on the selection (and evolution) of
emerging pathogens. Here, we experimentally demonstrate that cooking favors the
selection of spores of a variant genotype of spore-forming enteric pathogen Clostridium
difficile (PCR ribotype 078) which emerged in the mid 2000s as a hypervirulent pathogen
of humans and animals, and a predominant food contaminant, internationally (Table 11).
Since the 1970s, C. difficile-associated diseases have been linked to the ingestion
and germination of pathogenic spores with toxin production in the gut. Because spores of
emerging hypervirulent C. difficile (PCR ribotypes 027 and 078) have been isolated from
foods since 2006 (Rodriguez-Palacios et al. 2007; Songer et al. 2009), there are concerns
about foodborne transmission. Although cooking at minimum recommended
temperatures controls most foodborne pathogens (60-74oC; Table 12), it is unknown if
(and how) cooking modifies the ecology of C. difficile in fully and partially heated foods.
The resistance of C. difficile to recommended cooking temperatures, the presence of
emerging C. difficile strains in ready-to-eat meats, the ability of spores to recover from
heat-injuries and increase germination when heated, and constant outbreaks of foodborne
diseases linked to partially cooked foods (Songer et al. 2009; Rodriguez-Palacios et al.
2010; Alfano-Sobsey et al. 2011; Rodriguez-Palacios and LeJeune 2011), led us to
hypothesize that cooking could systematically favor the selection of emerging enteric
pathogens.
211
Materials and Methods
We tested our hypothesis by heating and comparing the survival probabilities of
four relevant prototypes of C. difficile: Emerging PCR ribotypes 027 and 078
(hypervirulent toxinotypes III and V), and PCR ribotype 077 and strain ATCC-9689
(regular toxinotypes 0). Using one-year-old spores, strains were inoculated into retail
ground beef, heat tested individually, and then collectively, at minimum and exceeding
recommended household cooking temperatures.
To quantify the resistance of C. difficile to heat, spores of each strain were first
heated individually at 85oC, for 0-30 minutes. Surviving spores were then quantified on
blood agar. Multivariable generalized linear analysis indicated that 85oC strongly
inhibited C. difficile within 10 minutes (~5 log reduction, P < 0.001), but the effect
significantly varied between strains. Strain ATCC-9689 was the most susceptible
(reduction of 10 million to 1-10 spores/gram, within 10 minutes), and PCR ribotype 027
the most resistant, with lower but comparable inhibition for the other ribotypes (Fig. 25).
To determine how cooking influences the ecology of C. difficile in fully and
partially heated foods, we then used four temperatures (63-96oC) and two types of retail
ground beef to quantify the relative individual survival probability for a mixture of three
C. difficile strains (resistant:intermediate:susceptible, 1:1:1 ratio). In duplicated
randomized complete block design, beef aliquots inoculated with the spores were heated
at various time-temperatures. Surviving spores, which grew as colonies on agar were
systematic and randomly selected, and PCR ribotyped to quantify surviving subtypes. We
212
also quantified naturally occurring C. difficile PCR ribotypes in the retail beef used for
experiments, with and without treatment at 63oC (Fig. 26).
Results
Microbial analysis confirmed natural contamination of the beef used with various
C. difficile genotypes (including ribotype-078), which survived 63oC for 30 minutes
(pasteurization). Subtyping analysis of beef experimentally contaminated with the spore
mixture confirmed that neither cooking at 63 nor 71oC affected the population structure
of C. difficile. However, heating at 85 and 96oC had a major selective effect, depending
on the length of cooking; favoring the selection of PCR ribotype 078 as cooking times
increased (Fig. 24A). Multivariable multinomial logistic regression showed the survival
probability for PCR ribotype 078 increased as the lethality of cooking increased (Fig.
24B).
Discussion
In the past, cooking with boiling water (90-100oC) or direct fire (>100oC) was the
norm. But in modern societies, there is increasing dependence on quick-to-cook foods.
Protecting public health, food safety guidelines recommend minimum cooking
temperatures between 63-74oC (Rodriguez-Palacios and LeJeune 2011). Here we showed
that those recommendations did not affect the population structure of C. difficile, but
importantly, that cooking at higher temperatures (likely in routine cooking) favored the
selection of emerging strains. Is this why hypervirulent strains predominate in ready-to213
eat meats? (Songer et al. 2009). The extent to which our findings apply to other
strains/pathogens is uncertain, but the ability of enteric pathogens to survive cooking is
noteworthy (6).
We might be selecting emerging enteric pathogens without noticing. As we lower
the cooking temperatures used by our ancestors, to attain desirable food texture and
palatability, we may ingest more heat-injured microorganisms. Does heat-injury promote
microbial mutagenesis or transformations within the gut, or foods? Our results, and the
remarkable association between heat-shock proteins and the potentiation of rapid
evolution of pathogenic traits in fungi (Cowen and Lindquist 2005), indicate that heatdriven evolution might occur with cooking (and possibly disinfection), and be clinically
relevant for humans and animals consuming heated foods.
Acknowledgements
Supported by federal funds allocated to the Ohio Agricultural Research and
Development Center, The Ohio State University. A.R.P. is a Public Health Preparedness
for Infectious Diseases Program Research Fellow. Thanks to M. Kauffman, J. Schrock
and P. Schlegel for technical assistance. A.R.P conceived the study; both authors
analyzed the data and wrote the manuscript. No conflict of interests.
214
A
Frequency of recovery of C. difficile during cooking of beef
Naturally
contaminated
log10 CFU - spores
88
66
89
inoculated 0 0 0-inoculum 0PCR078
7.9
PCR-078
8
8.8
PCR-027
8.6
6
22
4
22
7.8
20
20
30
30
5
96oC
0
00
10
10
10
20
20
20
0
30
30
30
minutes of heat — (63-96o C)
Hottest
96otreatment
C, 30min
Thermal
63oC, 0-30min
a
0.9 pp
078
PCR-078
a,b
pp
123
PCR-027
0.8 pp 314
PCR-ATCC b
0.9
0.7
0.7
0.7
0.6
0.6
0.6
0.5
0.5
0.5
0.4
0.4
0.4
0.3
0.3
0.3
0.2
0.2
0.2
0.1
0.1
0.1
0.8
0
5
85oC
Hot
predicted survival probability
pp 314 pp 314
10
GLM, P < 0.02
8
non-inoculated 63 2 2-leanPCR078
30 0.50
7.6
0
0
pp 123 pp 123
10
Flora (f at beef )
PCR123
4
8.2
non-inoculated
63 3PCR078
3-fat
15
0.50
PCR078
non-inoculated
63 2 2-lean
15
1.48
10
10
0.9
15
PCR123
non-inoculated 0 2 2-lean 0PCR078
2.30
pp 078 pp 078
15
Flora (lean beef )
63oC
71oC
PCR078
inoculated 71 2 2-lean 30
5.83
8.4 Flora (f at beef )
Other
non-inoculated 0 3 3-fat 0 3.00
PCR078
B
20
inoculated 63 3 3-fat 30 6.15
PCR078
PCR078
inoculated 71 3 3-fat 15 6.02
PCR078
6
minutes of heat — (63oC)
20
inoculated 63 3 3-fat 15
6.20
inoculated
71 1 1-gravy 20 6.45
PCR078
inoculated 85 2 2-lean 2PCR078
4.38
non-inoculated 63 3 3-fat 2PCR078
3.29
44
00
25
inoculated 63 1 1-gravy 2 6.60
PCR078
inoculated
71 3 3-fat
Flora (lean
beef
) 5 6.00PCR078
PCR- ATCC
00
25
Experimentally
contaminated
0
0 25
0.8
0
20
15
10
5
0
number of surviving colonies PCR ribotyped
(fewer colonies indicate stronger thermal inhibition)
Figure 24. Cooking favored the selection of C. difficile PCR ribotype 078 (and 027 to a
lesser extent) as time-temperatures increased.
(A) C. difficile single-colony PCR ribotyping summary data (PCR-legend and associated
pie charts; N=152 independent-random beef aliquots); and reduction of background beef
flora with pasteurization temperature 63oC (dashed lines used as referent in both plots,
mean ± SD, 3-5 replicas/point). (B) Predicted survival probability derived from
multivariable multinomial logistic regression. Distinct superscripts on PCR types indicate
differences (adjusted P < 0.05).
215
*
Clostridium difficile
(1+log10 CFU/ 0.1mL of beef extract)
*
8
PCR type 027
PCR type 078
PCR type 077
6
ATCC
4
2
*
0
0
10
20
30
minutes of heat at 85oC
Figure 25. Experiment 1 – Individual thermal testing of prototype C. difficile PCR
ribotypes in lean ground beef at the highest reported minimal cooking temperature
recommended for poultry meals in Canada.
Asterisks indicate significantly different effect estimates (Generalized linear model, P <
0.01)
216
New meat type
ATCC
New meat type
Other
ref .
bp
1500
PCR- 077
100
80
60
40
PCR 077
PCR-077
PCR-027
ref . PCR077
PCR-078
Other
PCR-027
PCR-Bidet
PCR-Bidet
PCR-078
B
Pearson correlation (Opt:7.00%) [0.0%-100.0%]
ref .
A
PCR 077
PCR-077
PCR 077
PCR-077
PCR 077
PCR-077
ATCC ATCC
500
PCR 027
PCR-027
PCR 027
PCR-027
PCR 078
PCR-078
200
Figure 26. PCR ribotyping (Bidet et al. 1999) was used to differentiate between the
inoculated, the naturally present, and the heat-survivor C. difficile strains recovered on
blood agar.
(A) The prototype strains tested in this study were easily distinguishable by PCR
ribotyping. (B). Two distinguishable C. difficile PCR ribotypes, not previously identified
in our laboratory, were naturally present in the retail beef used in this study. They were
detected in two single beef aliquots after 5 and 30 minutes of heating at 63oC. ‗ref.‘
indicates prototype inoculated strain.
217
Table 11. Evidence that C. difficile PCR ribotype 078 is increasing internationally.
Country
Comment or actual quotes
Year (ref.)
Relative occurrence in humans with CDI
Germany
Scotland
Netherlands
USA
Third most frequent PCR ribotype (4.7%) in humans with CDI after type
001 (70%) and 027 (4.8%).
November 2007-December 2009. Three PCR ribotypes predominated
amongst the Scottish isolates of C. difficile; ribotype 106 (29.4%),
ribotype 001 (22%) and ribotype 027 (12.6%) followed by the less
prevalent ribotypes including 002, 015, 014, 078, 005, 023 and 020.
(Reil et al. 2011)
34-month prospective case-control study: hospitalized CDI patients,
patients without diarrhea, and patients with non-CDI diarrhea. The
incidence of CDI was 17.5 per 10,000 hospital admissions. PCR ribotype
078 was the second most frequent type (12.7%).
PCR ribotype 078 (Toxinotype V) was the third most prevalent type
(10%) in community-associated C. difficile disease.
(Hensgens et al. 2011)
(Wiuff et al. 2011)
(Limbago et al. 2009)
Increased incidence in people
Europe
Netherlands
Canada
France
The incidence of C. difficile infection in hospitals was 4·1 per 10,000
patient-days per hospital (range 0·0-36·3). Isolates available from 389
cases corresponded to 65 different PCR ribotypes. Ribotype 078 was the
third most frequent (8%) after ribotypes 014/020 (16%) and 001 (9%).
Between 2005 and 2009 (2,788 samples): PCR ribotype 027 decreased
by 2006. An increase of CDI detected in patients with diarrhea in the
community coincided with emergent new ribotypes. Type 078 increased
(9.1%) since the end of 2006, and became the third most frequent type in
the Netherlands, after ribotypes 001 (27.5%) and 014 (9.3%).
Data Canadian Nosocomial Infections Surveillance Program CDI
surveillance 2004-2008. The incidence of CDI associated with ribotype
078 increased from 0.5% (2004) to 1.6% (2008).
(Bauer et al. 2011)
Data National Reference Laboratory. PCR ribotype 078 significantly
increased in France from 3.25% (8/246) in July to December 2006 to
11.1% (16/144) in July to December 2007.
(Rupnik et al. 2008)
(Hensgens et al. 2009)
(Mulvey et al. 2010)
Recent ‘first reports’ of PCR ribotype 078
New Zealand
Ireland
Ireland
China
1/97 (1%) patients with CDI. First PCR ribotype 078 identified in the
country.
1/133 (0.75%) patients with CDI. First clindamycin-resistant PCR
ribotype 078.
First report of PCR ribotype 078 (clindamycin susceptible).
First report. 1/110 (0.91%) isolates from humans with CDI were PCR
ribotype 078.
(Roberts et al. 2011)
(Solomon et al. 2011)
(Burns et al. 2010)
(Huang et al. 2010)
Relative occurrence in food animals
Canada
Most common ribotype in commercial young dairy calves in farms (25%),
and in veal calves (53%) in a commercial feedlot cattle operation.
France
CDI rates have increased in the community and new genotypes (e.g.
PCR ribotype 078) are emerging in both humans and animals- Update.
Predominant type in calves, piglets, and present in finishing cattle and
farmed deer.
USA
(Costa et al. 2011)
(Rodriguez-Palacios et
al. 2006)
(Barbut et al. 2011)
(Rodriguez-Palacios et
al. 2011)
(Norman et al. 2011)
(French et al. 2010)
(Keel et al. 2007)
Relative occurrence in foods
Canada
All isolates from chicken meats contaminated with C. difficile were PCR
ribotype 078.
(Weese et al. 2010)
Canada
USA
(Metcalf et al. 2010)
(Songer et al. 2009)
3/5 isolates from retail vegetables were PCR ribotype 078.
Ribotype 078 was predominant (75%) in ready-to-eat meats
contaminated
218
Table 12. Recommended safe minimum cooking temperatures and holding times to
reduce exposure to foodborne pathogens.
Food category
Leftovers &
Casseroles
Poultry
Ground Meats
Fresh Beef, Veal,
Lamb
Seafood
Pork and Ham
Example Items
Temperature
-
74 C (165 F)
Chicken, Turkey, Duck, Goose; whole or
parts
Turkey, Chicken
Beef, Pork, Veal, Lamb
Steaks, roasts, chops
74 C (165 F)
Fin Fish, Shrimp, lobster, and crabs,
Clams, oysters, and mussels; Scallops
Fresh pork /raw ham
Precooked ham (to reheat)
o
o
o
o
None
o
o
None
None
3 minutes
74 C (165 F)
o
o
71 C (160 F)
o
o
63 C (145 F)
o
Holding
time
None
o
63 C (145 F) or cook until
None
flesh is opaque and
separates easily with a fork.
o
o
63 C (145 F)
3 minutes
o
o
60 C (140 F)
None
Adapted from chart publicly available at www.foodsafety.gov/keep/charts/mintemp.html.
In Canada, recommendations state that whole (and stuffed) poultry should be cooked to at
least 85oC (185oF), available at http://www.hc-sc.gc.ca/fn-an/securit/illintox/info/poultry-volaille-questions-eng.php. Holding time refers to the number of
minutes needed at the recommended temperature to inhibit at least 6 log10 units of major
foodborne pathogens (e.g., Salmonella spp.).Current reports indicate that some seafood
will need to be heated to 90oC to inactivate foodborne human viruses (6).
219
Supplementary Materials and Methods
Rationale for temperatures tested and hypothesis
In order to achieve 5-6 Log10 reductions in C. difficile spore counts (to meet food
safety standards), we recently reported (Rodriguez-Palacios and LeJeune 2011) the need
to heat foods at temperatures higher than 71oC/160oF―the temperature typically
recommended as safe for cooking most meals (Table S2) (USDA). Heat treatment of
meat experimentally inoculated with C. difficile indicated that heating at 96oC/205oF
(sub-boiling temperature) was effective against 99.9 % of spores within 1 minute (95%
CI of minutes to destroy 90% of spores [D-value96] = 0.8, 1.4) (Rodriguez-Palacios and
LeJeune 2011). However, many spores survived several minutes of heating at 96oC
(Rodriguez-Palacios and LeJeune 2011). The presence of thermoresistant spores after
cooking indicated that heat applied to foods could destroy some C. difficile strains while
favoring the selection of others. Here we tested that hypothesis.
Clostridium difficile strains
We experimentally quantified the survival probability to moist-heat cooking of
toxigenic C. difficile PCR ribotypes 078/toxinotype-V, 027/toxinotype-III, PCR ribotype
077/toxinotype-0 of bovine origin isolated from calves in dairy farms of Canada in 2004
(Rodriguez-Palacios et al. 2006) and prototypic human-origin strain ATCC9569/toxinotype-0 isolated from a human affected with C. difficile infection in the 1990s.
Since spores likely persist in the environment before coming into contact with foods,
220
spores from all strains tested were aged on phosphate buffered saline for 52 weeks as
previously described (Rodriguez-Palacios and LeJeune 2011).
Experiment 1 – Individual thermal testing
For individual testing, duplicate samples of retail beef (6.6 g, prepared from a
package of 97% lean ground beef) were individually inoculated with PBS suspensions
containing single C. difficile strains (6.9 ± 0.6 Log10 CFU/0.1 g) and manually mixed.
Inoculated samples were then split into three pairs of smaller aliquots (1.1 g) for
duplicate testing and 3 time points. These aliquots, placed inside 4-ounce Whirlpack
bags, were pressed to thin layers (~64 cm2, without air) and heated in a water bath at
85oC/185oF (highest minimal cooking temperature recommended for difficult-to-cook
meats) following a protocol used for Escherichia coli O157:H7 reported by Juneja et al.
(USDA) with slight modifications for C. difficile (Rodriguez-Palacios and LeJeune
2011). After 10, 20 and 30 minutes of heating, survivor spores were enumerated from
two 100-µL aliquots of beef extract (leakage from heated beef) per aliquot, using the
spread plate method on tryptic soy agar supplemented with 5% defibrinated sheep blood.
Experiment 2 – Collective thermal testing of C. difficile mixture
With observed thermo-resistance differences between the strains, another
experiment was conducted in duplicate to determine the concurrent relative probability of
survival for the three C. difficile strains in a spore mixture (1:1:1 ratio). This time, we
assessed the effect of four cooking temperatures (63, 71, 85, and 96oC) in two types of
ground beef (3 and 30% fat content). In brief, random 3-g aliquots (n=128) of ground
beef were inoculated with 7-8 Log10 of C. difficile. Inoculated and non-inoculated control
221
(n=24) aliquots were heated in water baths and survivor spores were enumerated on
blood agar after 0, 2.5, 5, 7.5, 10, 15, 20 and 30 minutes of heating. To determine the
probability of survival for each individual strain after exposure to various timetemperatures, we randomly selected survivor C. difficile colonies from representative
blood agar plates and generated a genetic fingerprint, PCR ribotyping (Bidet et al. 1999),
for each colony selected.
Systematic random sampling of survivor C. difficile
Sampling of survivor C. difficile colonies was conducted using a systematic
random scheme to prevent selection bias and to represent the effectiveness of the thermal
treatment. Starting with the most effective temperature (e.g., the ones with the lowest
colony counts), and the last two-fold dilutions yielding C. difficile, all countable colonies
were subculture for PCR ribotyping if survivor colonies were less than 10 per dilution.
When countable colonies were between 10 and 40, all colonies within a random quadrant
of the plate were subcultured and typed. No plates were sampled if the colonies were not
separated from each other to ensure proper single colony testing.
Natural contamination in beef with C difficile and background microbial flora as
controls
As a control, the natural level of C. difficile contamination was quantified in
random aliquots of the ground beef and determined if ‗naturally occurring‘ spores,
sampled as above, could be differentially selected with pasteurization temperature at
63oC. Further the effect of thermal treatments on the total number of non-C. difficile
colonies (background beef flora) was determined on 10-fold serial dilutions inoculated
222
onto blood agar after 48 hours of anaerobic incubation to monitor the effectiveness of
thermal treatments.
Statistical Analysis
Experiment 1 – The effect of heat on the log10 CFU of C. difficile data was
analyzed using a generalized linear model with minutes of heat, replicate, and strain type
as predictors.
Experiment 2 – The relative frequency of isolation data for each PCR ribotype
identified in beef was analyzed as continuous variables using generalized linear
regression. Predictors tested included minutes of heating, temperature, type of beef, log10
CFU of C. difficile, number of isolates PCR ribotyped per aliquot, and replica. To
determine the survival probability of each PCR ribotype within a given beef aliquot, the
frequencies of isolation data were converted into a categorical variable (yes/no) for each
ribotype. Multivariable multinomial logistic regression was then used to compare and
estimate predicted survival probabilities as a function of the variables described above
(UCLA). Intercooled Stata software (v10.1, College Station, TX) was used. Adjusted pvalues are reported.
223
References
Alfano-Sobsey, E., D. Sweat, A. Hall, et al. (2011). "Norovirus outbreak associated with
undercooked oysters and secondary household transmission." Epidemiol Infect: 17.
Barbut, F., G. Jones and C. Eckert (2011). "Epidemiology and control of Clostridium
difficile infections in healthcare settings: an update." Current Opinion in
Infectious Diseases 24(4): 370-376.
Bauer, M. P., D. W. Notermans, B. H. van Benthem, et al. (2011). "Clostridium difficile
infection in Europe: a hospital-based survey." Lancet 377(9759): 63-73.
Burns, K., M. Morris-Downes, W. N. Fawley, et al. (2010). "Infection due to C. difficile
ribotype 078: first report of cases in the Republic of Ireland." Journal of Hospital
Infection 75(4): 287-291.
Carmody, R. N. and R. W. Wrangham (2009). "The energetic significance of cooking." J
Hum Evol 57(4): 379-391.
Costa, M. C., H. R. Stampfli, L. G. Arroyo, et al. (2011). "Epidemiology of Clostridium
difficile on a veal farm: Prevalence, molecular characterization and tetracycline
resistance." Vet Microbiol.
Cowen, L. E. and S. Lindquist (2005). "Hsp90 potentiates the rapid evolution of new
traits: drug resistance in diverse fungi." Science 309(5744): 2185-2189.
French, E., A. Rodriguez-Palacios and J. T. LeJeune (2010). "Enteric bacterial pathogens
with zoonotic potential isolated from farm-raised deer." Foodborne Pathog Dis
7(9): 1031-1037.
224
Hensgens, M. P., A. Goorhuis, D. W. Notermans, et al. (2009). "Decrease of
hypervirulent Clostridium difficile PCR ribotype 027 in the Netherlands." Euro
Surveill 14(45).
Hensgens, M. P., A. Goorhuis, C. M. van Kinschot, et al. (2011). "Clostridium difficile
infection in an endemic setting in the Netherlands." Eur J Clin Microbiol Infect
Dis 30(4): 587-593.
Huang, H., A. Weintraub, H. Fang, et al. (2010). "Antimicrobial susceptibility and
heteroresistance in Chinese Clostridium difficile strains." Anaerobe 16(6): 633635.
Keel, K., J. S. Brazier, K. W. Post, et al. (2007). "Prevalence of PCR ribotypes among
Clostridium difficile isolates from pigs, calves, and other species." J Clin
Microbiol 45(6): 1963-1964.
Limbago, B. M., C. M. Long, A. D. Thompson, et al. (2009b). "Clostridium difficile
strains from community-associated infections." J Clin Microbiol 47(9): 30043007.
Metcalf, D. S., M. C. Costa, W. M. Dew, et al. (2010). "Clostridium difficile in
vegetables, Canada." Lett Appl Microbiol 51(5): 600-602.
Mulvey, M. R., D. A. Boyd, D. Gravel, et al. (2010). "Hypervirulent Clostridium difficile
strains in hospitalized patients, Canada." Emerg Infect Dis 16(4): 678-681.
Norman, K. N., H. M. Scott, R. B. Harvey, et al. (2011). "Prevalence and Genotypic
Characteristics of Clostridium difficile in a Closed and Integrated Human and
Swine Population." Appl Environ Microbiol.
Reil, M., M. Erhard, E. J. Kuijper, et al. (2011). "Recognition of Clostridium difficile
PCR-ribotypes 001, 027 and 126/078 using an extended MALDI-TOF MS
system." Eur J Clin Microbiol Infect Dis.
Roberts, S., H. Heffernan, N. Al Anbuky, et al. (2011). "Molecular epidemiology and
susceptibility profiles of Clostridium difficile in New Zealand, 2009." N Z Med J
124(1332): 45-51.
225
Rodriguez-Palacios, A. and J. T. LeJeune (2011). "Moist-heat resistance, spore aging,
and superdormancy in Clostridium difficile." Appl. Environ. Microbiol. 77(9):
3085-3091.
Rodriguez-Palacios, A., C. Pickworth, S. Loerch, et al. (2011b). "Transient Fecal
Shedding and Limited Animal-to-Animal Transmission of Clostridium difficile by
Naturally Infected Finishing Feedlot Cattle." Appl Environ Microbiol 77(10):
3391-3397.
Rodriguez-Palacios, A., R. J. Reid-Smith, H. R. Staempfli, et al. (2010). "Clostridium
difficile survives minimal temperature recommended for cooking ground meats."
Anaerobe 16: 540-542.
Rodriguez-Palacios, A., H. R. Staempfli, T. Duffield, et al. (2007a). "Clostridium difficile
in retail ground meat, Canada." Emerg Infect Dis 13(3): 485-487.
Rodriguez-Palacios, A., H. R. Stampfli, T. Duffield, et al. (2006). "Clostridium difficile
PCR ribotypes in calves, Canada." Emerg Infect Dis 12(11): 1730-1736.
Rupnik, M., A. Widmer, O. Zimmermann, et al. (2008). "Clostridium difficile toxinotype
V, ribotype 078, in animals and humans." J Clin Microbiol 46(6): 2146.
Solomon, K., S. Fanning, S. McDermott, et al. (2011). "PCR ribotype prevalence and
molecular basis of macrolide-lincosamide-streptogramin B (MLSB) and
fluoroquinolone resistance in Irish clinical Clostridium difficile isolates." J
Antimicrob Chemother.
Songer, J. G., H. T. Trinh, G. E. Killgore, et al. (2009). "Clostridium difficile in retail
meat products, USA, 2007." Emerg Infect Dis 15(5): 819-821.
Weese, J. S., R. J. Reid-Smith, B. P. Avery, et al. (2010). "Detection and characterization
of Clostridium difficile in retail chicken." Lett Appl Microbiol 50(4): 362-365.
Wiuff, C., D. J. Brown, H. Mather, et al. (2011). "The epidemiology of Clostridium
difficile in Scotland." J Infect 62(4): 271-279.
226
Chapter 11 ― Clostridium difficile and higher cooking temperatures for
communities at risk
(Submitted for Publication)
Clostridium difficile infection (CDI) in hospitals and the community continues to
be a growing international public health concern and a major financial burden despite
strengthened control actions. Evolving antimicrobial prescription and hygiene policies
have been widely publicized to protect hospitalized individuals from developing severe
CDI. However, communities that were previously unaffected are becoming increasingly
susceptible to severe CDI, and these groups have not been targeted for prevention. The
elderly, children, pregnant women, and young adults are at high risk for CDI within the
community, especially after receiving antimicrobials or antacids, undergoing surgery, or
suffering chronic medical conditions. Community-associated CDI have also been
associated with emergent antimicrobial-resistant and hypervirulent C. difficile strains, but
the source for infection is often undetermined.
The fecal-oral transmission of C. difficile has been linked to environments
contaminated with spores (e.g., health-care centers), a problem potentiated by improper
hand washing. In the past, CDI were exclusively attributed to exposure to health-care
227
settings, and control actions have primarily targeted health professionals to reduce spore
dissemination between patients. However, since 2006, the frequent isolation of emergent
C. difficile strains from raw and ready-to-eat foods (6-42%) in North America (Gould and
Limbago 2010) indicate that exposure to foodborne spores is possibly a risk factor,
although causality ties with CDI have not been confirmed. Although it is uncertain how
many ingested spores are needed to cause disease in people, the presence of <9,000
contaminating spores of C difficile per gram of food (Weese et al. 2009; RodriguezPalacios and LeJeune 2011), indicate that reducing food contamination may be
beneficial.
Infection control recommendations against C. difficile are not readily available for
patients or household settings. Here we propose a simple intervention to reduce the
qualitative risk of ingesting foodborne spores based on results from recent thermal studies
that assessed the sensitivity of epidemic C. difficile strains to various cooking timetemperature combinations (Rodriguez-Palacios et al. 2010; Rodriguez-Palacios and
LeJeune 2011). We propose accompanying medical prescriptions, treatments, or
diagnoses that predispose to CDI, with recommendations to: i) improve kitchen hygiene,
ii) increase cooking temperatures, and iii) minimize exposure to documented sources of C
difficile spores (Box 1). Because spores are necessary for disease, these measures might
mitigate the incidence of CDI among susceptible individuals.
Based on PCR ribotyping (a molecular classification method) there are over 200
C. difficile subtypes isolated from humans, animals and the environment. However, the
frequent isolation of only a few specific highly epidemic (Goorhuis et al. 2008) C.
228
difficile subtypes (PCR ribotypes 001, 078, 027) from foods (Gould and Limbago 2010)
is concerning. With the presence of C. difficile spores in raw and ready-to-eat foods and
people‘s preference for rare/medium cooked meats, an obvious question is whether
cooking reduces the foodborne exposure risk.
At present, C. difficile is not considered a foodborne pathogen (Gould and
Limbago 2010). Thus, recommendations for its control are not widely publicized. There
are however cooking recommendations for less heat-resistant pathogens. Most food
safety guidelines recommend achieving a minimum internal temperature of 71ºC (160oF)
for a few seconds to make meals safe.(Rodriguez-Palacios et al. 2010) For more
difficult-to-cook whole-poultry meals, some recommendations state that at least 85oC
(185oF) is needed.(Rodriguez-Palacios and LeJeune 2011) At recommended cooking
temperatures, most vegetative pathogens (e.g., E. coli) can be eliminated within seconds,
but clostridia produce resistant spores. Although there is abundant information on thermal
inhibition of Clostridium perfringens and Clostridium botulinum (worldwide foodborne
pathogens), species differences prevent extrapolating data to C. difficile.
Recently, complementary thermal studies on C. difficile have expanded our
understanding on this pathogen.(Rodriguez-Palacios et al. 2010; Rodriguez-Palacios and
LeJeune 2011) The first report documented that heating C. difficile spores of epidemic
PCR ribotypes from food and food animals at minimum recommendations for ground
meats (71oC, <2 minutes) causes no spore inhibition.(Rodriguez-Palacios et al. 2010) Not
surprisingly, extending the cooking time to two hours only reduced the number of spores
by 2-3 Log10 units (from 10 million to 100 thousand spores/gram), which is far less than
229
the 6 Log10 units (from 10 million to 1-10 spores/gram) reduction desired by processing
technologies. Hence, given the level and frequency of meat contamination, the risk of
ingesting C. difficile in foods is real even if current cooking recommendations are
followed.
To further assess the effects of cooking C. difficile at higher temperatures, we
assessed their spore survival in ground beef and gravy at different fat contents (030%).(Rodriguez-Palacios and LeJeune 2011) Assessed temperatures included 63
(145oF), 71, 85 and 96oC (205oF). Cooking at 85oC markedly reduced C. difficile spores
within 10-15 minutes, but the inhibition was more pronounced at sub-boiling temperature
(96oC). It approached the desired 6 Log10 spore reduction within 1-2 minutes of heating
(Table). Since foods can be contaminated with up to 4 Log10 (<9,000) of spores per
gram,(Weese et al. 2009; Rodriguez-Palacios and LeJeune 2011) cooking at >96oC could
markedly reduce the risk of ingesting C. difficile.
Of concern, the same experiments also showed that cooking at 63 and 71oC
(widely recommended) enhanced the rate of spore germination and total C. difficile
counts in most strains.(Rodriguez-Palacios and LeJeune 2011) Furthermore, experiments
illustrated that a fraction of the spores that were inhibited with 85oC, recovered within 7
hours if left incubating in nutrient-rich media. Inadequate cooking followed with
improper cooling and storage could therefore favor the proliferation of C difficile within
foods, which is precisely the factor responsible for C. perfringens food poisoning in
humans. The importance of thorough cooking needs to be emphasized, especially,
because recent studies indicated that sublethal heating favors the selection of
230
hypervirulent C. difficile strains (Rodriguez-Palacios, A & LeJeune, submitted).
In addition to ‗better-cooking-practices‘, other preventive measures could be
publicized for susceptible communities (Box 2). Hand washing should be reinforced
when preparing foods, after visiting bathrooms, and especially after having contact with
animals (or their environment) because hypervirulent C. difficile strains are often found
in excrements (Gould and Limbago 2010). Susceptible communities should also avoid
direct exposure to farm/companion animals. Particular attention should be placed on dogs
fed raw meats or that are used as ‗therapy animals‘ in health-care settings as they may
carry C. difficile (Gould and Limbago 2010). Public health authorities, physicians and
those persons cooking for susceptible communities, should consider these factors when
making recommendations and preparing foods.
Accompanying relevant prescriptions (or consultations) with recommendations to
reinforce better-cooking practices and hygiene will be beneficial until validated methods
to reduce food contamination are available. Revised food safety guidelines are needed to
provide at-risk communities with adequate cooking temperatures, and times (currently
not indicated in most guidelines), to prevent inadvertent ingestion of clostridial spores. If
C. difficile from foodborne sources indeed results in illness, then minimizing crosscontamination during food production and processing could also reduce the incidence of
CDI.
Adoption of higher-temperature cooking and enhanced hygiene practices may be
limited: Some individuals may be reluctant to adopt recommendations or may not believe
food/environmental contamination is an issue, while others may prefer to continue eating
231
undercooked foods due to personal or cultural preferences. Nevertheless, providing the
information to patients and the community ensures that the highest standards of practice
are applied. According on the Oxford Centre for Evidence-Based Medicine guidelines
(http://www.cebm.net), it is safe to say there is already enough epidemiological and
bench research data from diverse sources (Weese et al. 2009; Gould and Limbago 2010;
Rodriguez-Palacios et al. 2010; Rodriguez-Palacios and LeJeune 2011), theoretical
rationale and mechanistic-based reasoning to harmlessly and systematically advise
susceptible communities to cook their foods better (Table 13) and lower their exposure to
potential sources of C. difficile (Fig 27 and 28), at least during periods considered of high
susceptibility for CDI. With coordination, monitoring the impact of such measures could
further assess whether CDI is linked to foods, especially among patients prone to
recurrence.
Acknowledgements
Supported by federal funds allocated to the Ohio Agricultural Research and
Development Center, The Ohio State University. A.R.P. is a Public Health Preparedness
for Infectious Diseases Program Research Fellow. Special thanks to Qiuhong Wang (MD,
PhD) and Sayaka Takanashi (MD, PhD) for thoughtful medical insights during final
stages of manuscript preparation. No conflict of interests. A.R.P wrote first manuscript
draft.
232
Table 13. Minutes of heat needed to inhibit 90% of C. difficile spores at achievable
household cooking temperatures.
Currently recommended minimum internal temperatures for cookinga
Steaks, fish, and
Most
Whole poultry and
batch pasteurization
ground meats
turkey meals
o
o
63 C
71 C
85oC
o
o
(145 F)
(160 F)
(185oF)
Consider
Boiling water
temperature
96oC
(205oF)
Gravy, 0% fat;
mean ± SD
(95% C.I.)
55.9 ± 21.8
(21.3 to 90.6)
56.8 ± 30.5
(18.9 to 94.7)
2.51 ± 0.43
(1.83 to 3.2)
0.59 ± 0.32
(0.07 to 1.1)
99.4 ± 62.3
(22 to 176)
71.2 ± 38.7
(9.6 to 132)
3.0 ± 1.11
(1.83 to 4.17)
0.62 ± 0.29
(0.15 to 1.08)
45.8 ± 10.8
(28.7 to 63)
46.8 ± 10.9
(33.1 to 60.4)
3.27 ± 1.02
(2.2 to 4.33)
1.19 ± 0.67
(0.49 to 1.89b)
69.6 ± 45.4
(42.1 to 96.9)
57.3 ± 27.8
(41.3 to 73.4)
2.97 ± 0.94
(2.48 to 3.48)
0.85 ± 0.55
(0.53b to 1.17)
Lean ground beef, 3% fat
mean ± SD
(95% C.I.)
Ground beef, 30% fat
mean ± SD
(95% C.I.)
Collectively, mean ± SD
95% C.I. of the mean
1
a
Food safety agencies recommend the use of commercially available ‗Food
Thermometers‘ to ensure proper cooking
(http://www.fsis.usda.gov/is_it_done_yet/brochure_text/index.asp#SMIT).
b
To achieve food safety standards of 6 log10 units of microbial reduction, cooking
temperatures between 3.2 and 11.4 minutes at 96oC (6 x [95% C.I.]; superscripted ‗b‘)
would be a safe starting point. Adapted from Rodriguez-Palacios & LeJeune
(2011)(Rodriguez-Palacios and LeJeune 2011) with permission from the American
Society of Microbiology (http://aem.asm.org/cgi/content/full/77/9/3085/DC1).
233
Supplementary material Boxes 1 and 2.
Box 1. Risk and exposure factors for acquisition of C. difficile spores and infections
Well-known Risk Factors (these can be host- or environment-dependent):
 Antimicrobials/antacids increases with ―sophisticated‖ drugs, combinations or long treatments.
 Elderly (over 65 years old) are more likely to get sick.
 Debilitating illness such as cancer, or immune-suppressive conditions/medications.
 Colon diseases such as inflammatory bowel disease or colorectal cancer.
 Abdominal surgery or gastrointestinal procedure.
 Having had C. difficile infection already.
 Current or past hospitalization or contact with CDI patients.
 Living in a nursing home/long-term care facility.
Possible exposure risk factors (environment-dependent; investigation is underway):
 Through contaminated foods (difficult to predict/notice; variable; C. difficile has been found in
food with good quality – smell and taste).
 Healthy animals shedding C. difficile (difficult to predict/notice; animals look normal).
 Diseased animals suffering CDI that shed C. difficile in feces (predictable, possible but unclear role
in human disease; animal-to-animal has been partly documented).
 Animal waste (occupational risk is possible but unknown; occupational risk has been documented
for laboratory personnel working with human fecal specimens).
 Contaminated environment (difficult to predict/notice unless obvious influence of potential source
of C. difficile; for instance, sewage from human/animal facilities; C. difficile has been found in
recreational waters).
1
Figure 27. Box 1 - Risk and exposure factors for acquisition of C. difficile spores and
infection
234
Box 2. Expanded recommendations to reduce the risk of acquiring C. difficile spores
 Hand washing. Hospital guidelines recommend an alcohol-based hand sanitizer to eliminate most
microorganisms but it is largely ineffective in destroying C. difficile spores; thorough hand washing
with soap and water is effective in removing spores from hands—this is the best. Health care
workers in human/animal hospitals, affected patients, and everyone who suspects exposure would
benefit from hand washing—this would also benefit others. Very relevant for household kitchens
and for personnel working in farms, food processing plants or restaurants.
 Contact precautions. In [human/animal] hospitals, affected subjects can be isolated [in private
rooms/stables/cages], and workers wear disposable gloves and gowns to prevent spread. Apart from
voluntary isolation in the community, isolating animals discharged from hospitals with a diagnosis
of CID should be encouraged and communicated to related household relatives or farm workers,
whichever applies. Contact with hospital visitation pets should be minimized outside hospitals (in
hospitals their emotional benefit to patients could be balanced with reinforcing other precautions;
controversial). Avoid exposure to animal or animal manure regardless of their apparent health status.
Beware that C. difficile spores that go down the drain or hill will remain dormant and alive for
months or years.
 Thorough cleaning. C. difficile spores can survive routine household disinfectants. In any setting,
all surfaces and equipment should be carefully cleaned with a detergent and a hospital-grade
disinfectant approved for C. difficile or chlorine bleach diluted with water at a 1:10 ratio (contact
exposure should be of at least 5-10 minutes; note that the effect of bleach disappears with
evaporation or in presence of dirt or abundant organic matter). Especially important on food
preparation areas, avoid cross-contamination because C. difficile has been isolated from raw meats
and vegetables. Keep pets outside of food preparation areas as they may be carriers of C. difficile.
 Use antimicrobials and antacids only if necessary. Antibiotics are often prescribed for viral
illnesses that aren't helped by these drugs. Consult your doctor if you think an antibiotic is needed.
Ask your doctor to prescribe one that has a narrow range and that you can take for the shortest time
possible. How antacids predispose to C. difficile infections is not fully understood, but it may be in
part because they block the killing effect of gastric acidity and thus allow the ingestion of otherwise
acid-susceptible microorganisms. Since antibiotics and antacids may disrupt the intestinal flora for
several (2-6 weeks) try to follow this precautions for as long as possible. Ideally make them a
healthy habit for your benefit and the benefit of your loved ones.
 Increase cooking time and temperatures. Studies have shown that C. difficile spores survive 2
hours of extended cooking if using internationally recommended minimum internal cooking
temperature (71oC, as cited on retail food labels). The currently recommended minimum
temperatures for difficult-to-cook meals (whole turkey) is 85oC; although at this temperature C.
difficile spores can be markedly inhibited, it may require at least 15 minutes, leaving the possibility
that surviving spores may re-grow within foods after heating. Cooking at boiling temperatures for at
least 3 minutes seems to be a reasonable strategy to inhibit most spores from foods. However, proper
chilling and storage of foods soon after heating is absolutely needed to prevent bacterial growth.
 Adopt complementary food safety measures available for at-risk populations. Although
publicized minimum temperatures are inadequate to destroy C. difficile, currently available
guidelines have practical measures that may be suitable to enhance food safety against clostridial
spores, including chilling and storage practices. (http://www.fsis.usda.gov/is_it_done_yet)
 Consider risk assessment features against C. perfringens. Food poisoning associated with C.
perfringens is one of the major foodborne illnesses in the USA. A freely available comprehensive
report exists addressing the risk of such pathogen in partly cooked and ready-to-eat meat and poultry
products. These data could be useful as C. difficile appears to be more susceptible to heat than C.
perfringens. (http://www.fsis.usda.gov/PDF/CPerfringens_Risk_Assessment_Tagged.pdf)
Figure 28. Box 2 - Expended recommendations to reduce the risk of acquiring C. difficile
spores
235
References
Goorhuis, A., D. Bakker, J. Corver, et al. (2008). "Emergence of Clostridium difficile
infection due to a new hypervirulent strain, polymerase chain reaction ribotype
078." Clin Infect Dis 47(9): 1162-1170.
Gould, L. H. and B. Limbago (2010). "Clostridium difficile in Food and Domestic
Animals: A New Foodborne Pathogen?" Clin Infect Dis 51(5): 577-582.
Rodriguez-Palacios, A. and J. T. LeJeune (2011). "Moist-heat resistance, spore aging,
and superdormancy in Clostridium difficile." Appl. Environ. Microbiol. 77(9):
3085-3091.
Rodriguez-Palacios, A., R. J. Reid-Smith, H. R. Staempfli, et al. (2010). "Clostridium
difficile survives minimal temperature recommended for cooking ground meats."
Anaerobe 16: 540-542.
Weese, J. S., B. P. Avery, J. Rousseau, et al. (2009a). "Detection and enumeration of
Clostridium difficile in retail beef and pork." Appl Environ Microbiol 75(15):
5009-5011.
236
Chapter 12 ― Conclusions
•
C. difficile can be shed by food animals at the time of harvest at concentrations
equivalent to super-shedding thresholds reported for E. coli O157. Despite the
isolation of emerging PCR ribotype 078 from finishing cattle 24 hours prior to
harvest, fecal shedding of C. difficile was transient (no animal tested twice in
consecutive relevant samplings which corresponded to 24h and 1 week apart)
•
The shedding in feedlot cattle did not increase with increased use of
antimicrobials in over a half of all animals in the lot. Antimicrobials were used
during the first month for treatment of typical, but serious feedlot cattle
respiratory diseases.
•
The prevalence of C. difficile in food animals during the summer is very low (00.6%) compared to horses (7.2%). This indicates that the risk of food
contamination at harvest is low.
•
High shedding levels from few food animals might be sufficient to contaminate
processing facilities. Studies sampling more farms and processing facilities in
winter would help to determine factor to explain why the prevalence of food
contamination is the highest in winter.
237
•
The potential for carcass and food contamination at the time of animal harvest
exist and could be amplified if derived foods are improperly cooked and stored.
Thermal studies showed that spores can recover, germinate, and multiply if
heating is not lethal.
•
High sub-lethal cooking temperatures favor the selection of emerging C. difficile
strains, especially emerging PCR ribotype 078. Therefore cooking
recommendations need to be accompanied with preharvest and postharvest
intervention measures to lower the load of contaminating spores.
•
Wild birds and possibly wild ruminants may play an important role on
dissemination of C. difficile in agricultural ecosystems. Starlings may allow for
efficient dissemination of C. difficile in an agricultural region, while long-range
migratory waterfowl could do it a larger distances. Waterfowl can also
contaminate environments visited by elderly, like public parks, and therefore
serve as a novel factor for C. difficile exposure.
•
Food safety risk associated with ingestion of toxigenic C. difficile from foods is
therefore real, especially in times of growing trends for cooking and eating
minimally processed foods.
•
Further studies are needed to identify critical control points to reduce the risk of
food and environemntal contamination with C. difficile. If foods and animals are
sources for zoonotic transmission, interventions to minimize exposure should
result in a reduction of C difficile infections in the community.
238
All References
Ackermann, G., D. Adler, et al. (2003). "In vitro activity of linezolid against Clostridium
difficile." J Antimicrob Chemother 51(3): 743-745.
al-Barrak, A., J. Embil, et al. (1999). "An outbreak of toxin A negative, toxin B positive
Clostridium difficile-associated diarrhea in a Canadian tertiary-care hospital." Can
Commun Dis Rep 25(7): 65-9.
al Saif, N. and J. S. Brazier (1996). "The distribution of Clostridium difficile in the
environment of South Wales." J Med Microbiol 45(2): 133-7.
Alfa, M. J., A. Kabani, et al. (2000). "Characterization of a Toxin A-Negative, Toxin BPositive Strain of Clostridium difficile Responsible for a Nosocomial Outbreak of
Clostridium difficile-Associated Diarrhea." J. Clin. Microbiol. 38(7): 2706-2714.
Alfa, M. J., N. Olson, et al. (2008). "Simulated-use testing of bedpan and urinal washer
disinfectors: evaluation of Clostridium difficile spore survival and cleaning
efficacy." American Journal of Infection Control 36(1): 5-11.
Alfano-Sobsey, E., D. Sweat, et al. (2011). "Norovirus outbreak associated with
undercooked oysters and secondary household transmission." Epidemiol Infect: 17.
Archibald, L. K., S. N. Banerjee, et al. (2004). "Secular trends in hospital-acquired
Clostridium difficile disease in the United States, 1987-2001." J Infect Dis 189(9):
1585-9.
Arroyo, L. G., S. A. Kruth, et al. (2005). "PCR ribotyping of Clostridium difficile isolates
originating from human and animal sources." J Med Microbiol 54(Pt 2): 163-6.
239
Arroyo, L. G., J. Rousseau, et al. (2005). "Use of a selective enrichment broth to recover
Clostridium difficile from stool swabs stored under different conditions." J Clin
Microbiol 43(10): 5341-5343.
Arroyo, L. G., J. S. Weese, et al. (2004). "Experimental Clostridium difficile enterocolitis
in foals." J Vet Intern Med 18(5): 734-8.
Aspinall, S. T. and D. N. Hutchinson (1992). "New selective medium for isolating
Clostridium difficile from faeces." Journal of Clinical Pathology 45(9): 812-814.
Bachman, J. (2008, May 30th, 2008). "Visualizing One Billion Dollars
" Retrieved January 15, 2010, from <http://www.wallstats.com/blog/visualizing-onebillion-dollars/>.
Bakri, M. M., D. J. Brown, et al. (2009). "Clostridium difficile in ready-to-eat salads,
Scotland." Emerg Infect Dis 15(5): 817-8.
Bakri, M. M., D. J. Brown, et al. (2009). "Clostridium difficile in ready-to-eat salads,
Scotland." Emerging Infectious Diseases 15(5): 817-818.
Balassiano, I. T., K. R. Miranda, et al. (2009). "Characterization of Clostridium difficile
strains isolated from immunosuppressed inpatients in a hospital in Rio de Janeiro,
Brazil." Anaerobe 15(3): 61-4.
Bandelj, P., T. Trilar, et al. (2011). "Zero prevalence of Clostridium difficile in wild
passerine birds in Europe." FEMS microbiology letters 321(2): 183-5.
Barbut, F., G. Jones, et al. (2011). "Epidemiology and control of Clostridium difficile
infections in healthcare settings: an update." Current Opinion in Infectious
Diseases 24(4): 370-376.
Bartlett, J. G. (2008). "Historical perspectives on studies of Clostridium difficile and C.
difficile infection." Clin Infect Dis 46 Suppl 1: S4-11.
Bauer, M. P., D. W. Notermans, et al. (2011). "Clostridium difficile infection in Europe: a
hospital-based survey." Lancet 377(9759): 63-73.
240
Baverud, V. (2004). "Clostridium difficile diarrhea: infection control in horses." Vet Clin
North Am Equine Pract 20(3): 615-30.
Baverud, V., A. Gustafsson, et al. (2003). "Clostridium difficile: prevalence in horses and
environment, and antimicrobial susceptibility." Equine Vet J 35(5): 465-71.
Bell, N., N. Schuurman, et al. (2008). "Are injuries spatially related? Joinâ€―count spatial
autocorrelation for small-area injury analysis." Injury Prevention 14(6): 346-353.
Bidet, P., F. Barbut, et al. (1999). "Development of a new PCR-ribotyping method for
Clostridium difficile based on ribosomal RNA gene sequencing." FEMS
Microbiology Letters 175(2): 261-266.
Biller, P., B. Shank, et al. (2007). "Moxifloxacin therapy as a risk factor for Clostridium
difficile-associated disease during an outbreak: attempts to control a new epidemic
strain." Infect Control Hosp Epidemiol 28(2): 198-201.
Boes, J., L. Nersting, et al. (2005). "Prevalence and diversity of Campylobacter jejuni in
pig herds on farms with and without cattle or poultry." Journal of Food Protection
68(4): 722-7.
Borgmann, S., M. Kist, et al. (2008). "Increased number of Clostridium difficile
infections and prevalence of Clostridium difficile PCR ribotype 001 in southern
Germany." Euro Surveill 2008;13(49):pii=19057.
(http://www.eurosurveillance.org/ViewArticle.aspx?ArticleId=19057).
Borriello, S. P., P. Honour, et al. (1983). "Foodborne transmission of Clostridium
difficile." Gastroenterology 84(1): 201.
Borriello, S. P., P. Honour, et al. (1983). "Household pets as a potential reservoir for
Clostridium difficile infection." J Clin Pathol 36(1): 84-7.
Brown, L. D., T. T. Cai, et al. (2001). "Interval Estimation for a Binomial Proportion."
Statistical Science 16(2): 101-117.
241
Burns, K., M. Morris-Downes, et al. (2010). "Infection due to C. difficile ribotype 078:
first report of cases in the Republic of Ireland." Journal of Hospital Infection
75(4): 287-291.
Campbell, Robert J. P., L. M. P. H. Giljahn, et al. (2009). "Clostridium difficile Infection
in Ohio Hospitals and Nursing Homes During 2006." Infection Control and
Hospital Epidemiology 30(6): 526-533.
Carmody, R. N. and R. W. Wrangham (2009). "The energetic significance of cooking." J
Hum Evol 57(4): 379-91.
Centers for Disease Control and Prevention (CDC), U. S. (2008). "Surveillance for
community-associated Clostridium difficile--Connecticut, 2006." MMWR Morb
Mortal Wkly Rep 57(13): 340-3.
CFIA. (2010). "Canadian Food Inspection Agency. Food Thermometer Food Safety Tips:
Preventing foodborne illness." Retrieved July 14, from
http://www.inspection.gc.ca/english/fssa/concen/tipcon/thermoe.shtml.
Chase-Topping, M., D. Gally, et al. (2008). "Super-shedding and the link between human
infection and livestock carriage of Escherichia coli O157." Nat Rev Micro 6(12):
904-912.
Chowell, G., J. M. Hyman, et al. (2009). The Effective Reproduction Number as a
Prelude to Statistical Estimation of Time-Dependent Epidemic Trends.
Mathematical and Statistical Estimation Approaches in Epidemiology, Springer
Netherlands: 103-121.
Cid, A., A. R. Juncal, et al. (1998). "Clostridium difficile Bacteremia in an
Immunocompetent Child." J. Clin. Microbiol. 36(4): 1167-1168.
CIPHI. (2003). "Canadian Institute of Public Health Inspectors. Food safety factors for
cooking Christmas Turkey " Retrieved December 19, from
http://www.ciphi.ca/pdf/turkey.pdf.
242
Clabots, C. R., S. J. Gerding, et al. (1989). "Detection of asymptomatic Clostridium
difficile carriage by an alcohol shock procedure." Journal of Clinical
Microbiology 27(10): 2386-2387.
Clements, A. C., R. J. Magalhaes, et al. (2010). "Clostridium difficile PCR ribotype 027:
assessing the risks of further worldwide spread." Lancet Infectious Diseases
10(6): 395-404.
Clooten, J., S. Kruth, et al. (2008). "Prevalence and risk factors for Clostridium difficile
colonization in dogs and cats hospitalized in an intensive care unit." Vet
Microbiol 129(1-2): 209-14.
Clopper, C. J. and E. S. Pearson (1934). "The use of confidence or fiducial limits
illustrated in the case of the binomial " Biometrika 26: 404-413
Cohen, S. H., D. N. Gerding, et al. (2010). "Clinical practice guidelines for Clostridium
difficile infection in adults: 2010 update by the society for healthcare
epidemiology of America (SHEA) and the infectious diseases society of America
(IDSA)." Infect Control Hosp Epidemiol 31(5): 431-55.
Costa, M. C., L. Arroyo, et al. (2009). "Incidence of Clostridium difficile in feces of male
Holstein veal calves. American College of Veterinary Internal Medicine Forum.
http://www.acvim.org/websites/acvim/File/Costa.pdf." Retrieved 15 March
2011.
Costa, M. C., H. R. Stampfli, et al. (2011). "Epidemiology of Clostridium difficile on a
veal farm: Prevalence, molecular characterization and tetracycline resistance."
Vet Microbiol.
Council, N. R., Ed. (2000). Nutrient requirements of beef cattle. Washington, DC., Natl.
Acad. Press.
Cowen, L. E. and S. Lindquist (2005). "Hsp90 potentiates the rapid evolution of new
traits: drug resistance in diverse fungi." Science 309(5744): 2185-9.
Crouch, E. and N. Golden. (2005). "Risk Assessment for Clostridium perfringens in
Ready-to-Eat and partly cooked meat and poultry products. USDA - Food Safety
243
Inspection Service." Retrieved September, 1, from
http://www.fsis.usda.gov/PDF/CPerfringens_Risk_Assess_Sep2005.pdf.
Dallal, R. M., B. G. Harbrecht, et al. (2002). "Fulminant Clostridium difficile: an
underappreciated and increasing cause of death and complications." Ann Surg
235(3): 363-72.
Davis, C. S. (2002). Statistical methods for the analysis of repeated measurements. New
York, Springer: 415.
Debast, S. B., L. A. van Leengoed, et al. (2009). "Clostridium difficile PCR ribotype 078
toxinotype V found in diarrhoeal pigs identical to isolates from affected humans."
Environ Microbiol 11(2): 505-511.
Dodson, K. and J. LeJeune (2005). "Escherichia coli O157:H7, Campylobacter jejuni,
and Salmonella Prevalence in cull dairy cows marketed in northeastern Ohio."
Journal of Food Protection 68(5): 927-31.
Dohoo, I., W. Martin, et al. (2003). Cohort studies. Veterinary Epidemiologic Research.
S. M. McPike. Charlottetown, Prince Edward Island, Canada., AVC, Inc.: 69,
151-162
Dubberke, Erik R., Anne M. Butler, et al. (2009). "Multicenter Study of the Impact of
Community-Onset Clostridium difficile Infection on Surveillance for C. difficile
Infection." Infection Control and Hospital Epidemiology 30(6): 518-525.
Dubberke, E. R. and A. I. Wertheimer (2009). "Review of current literature on the
economic burden of Clostridium difficile infection." Infect Control Hosp
Epidemiol 30(1): 57-66.
Ehrich, M., B. D. Perry, et al. (1984). "Acute diarrhea in horses of the Potomac River
area: examination for clostridial toxins." J Am Vet Med Assoc 185(4): 433-5.
Elder, R. O., J. E. Keen, et al. (2000). "Correlation of enterohemorrhagic Escherichia coli
O157 prevalence in feces, hides, and carcasses of beef cattle during processing."
Proc Natl Acad Sci U S A 97(7): 2999-3003.
244
FARAD. (2011). "Food Animal Residue Avoidance and Depletion Program. Prohibited
Drug List " Retrieved 15 March 2011, from
http://www.farad.org/eldu/prohibit.html.
FDA. (2002). "September 19. Reminder - Extra-label use of fluoroquinolones prohibited.
United States Food and Drug Administration." Retrieved 31 August 2010, from
http://www.fda.gov/AnimalVeterinary/NewsEvents/CVMUpdates/ucm129354.ht
m.
FDA. (2007). "March 19. Clindamycin hydrochloride oral drops (for use in dogs and
cats). United States Food and Drug Administration. ANADA 200-398 original
approval." May 20 2011, from
http://www.fda.gov/downloads/AnimalVeterinary/Products/ApprovedAnimalDru
gProducts/FOIADrugSummaries/ucm061348.pdf.
FDA. (2009). "November 18. United States Food and Drug Administration takes action
against dairy farm and owner - Farm selling animals with illegal drug residues in
violation of federal law." FDA NEWS RELEASE, May 20 2011, from
http://www.fda.gov/NewsEvents/Newsroom/PressAnnouncements/2009/ucm1912
32.htm.
FDA. (2009). "November 19. United States Food and Drug Administration takes action
against Maryland veal calf dealer - Man allegedly sold animals for human
consumption that contained illegal drug residues." May 20 2011, from
http://www.fda.gov/NewsEvents/Newsroom/PressAnnouncements/2009/ucm1914
46.htm.
FDA. (2010). "United States Food and Drug Administration. Grade "A" Pasteurized Milk
Ordinance - Public Health Service/Food and Drug Administration (2009
Revision)." Retrieved April 26, from
http://www.fda.gov/downloads/Food/FoodSafety/ProductSpecificInformation/MilkSafety/NationalConferenceonInterstateMilkShipmentsN
CIMSModelDocuments/UCM209789.pdf.
Fedorko, D. P. and E. C. Williams (1997). "Use of cycloserine-cefoxitin-fructose agar
and L-proline-aminopeptidase (PRO Discs) in the rapid identification of
Clostridium difficile." J Clin Microbiol 35(5): 1258-1259.
245
Fluckey, W. M., G. H. Loneragan, et al. (2009). "Diversity and Susceptibility of
Enterococcus Isolated from Cattle before and after Harvest." Journal of Food
Protection 72: 766-774.
Freeman, J. and M. H. Wilcox (2003). "The effects of storage conditions on viability of
Clostridium difficile vegetative cells and spores and toxin activity in human
faeces." Journal of Clinical Pathology 56(2): 126-128.
French, E., A. Rodriguez-Palacios, et al. (2010). "Enteric bacterial pathogens with
zoonotic potential isolated from farm-raised deer." Foodborne Pathog Dis 7(9):
1031-7.
FSA. (2007). "Food Standards Agency, United Kingdom. Advisory Committee on the
Microbiological Safety of Food: Report on the Safe Cooking of Burgers."
Retrieved July 31, from
http://www.food.gov.uk/news/newsarchive/2007/jul/burgers and
http://www.food.gov.uk/multimedia/pdfs/acmsfburgers0807.pdf.
Garcia, A., J. Perez, et al. (2000). "Evaluation of four rapid methods for the investigation
of the toxigenic capacity of Clostridium difficile strains isolated in a selective
medium." Enferm Infecc Microbiol Clin 18(3): 109-112.
Gastmeier, P., D. Weitzel-Kage, et al. (2009). "Surveillance of Clostridium difficileassociated diarrhoea with the German nosocomial infection surveillance system
KISS (CDAD-KISS)." International journal of antimicrobial agents
33(Supplement 1): S19-S23.
Gérard, M., N. Defresne, et al. (1989 ). "Polymicrobial septicemia with Clostridium
difficile in acute diverticulitis." Eur J Clin Microbiol Infect Dis 8(4): 300-302.
Gerding, D. N. (2010). "Global epidemiology of Clostridium difficile infection in 2010."
Infect Control Hosp Epidemiol 31 Suppl 1: S32-4.
Geric, B., R. J. Carman, et al. (2006). "Binary toxin-producing, large clostridial toxinnegative Clostridium difficile strains are enterotoxic but do not cause disease in
hamsters." J Infect Dis 193(8): 1143-50.
246
Geric, B., M. Rupnik, et al. (2004). "Distribution of Clostridium difficile variant
toxinotypes and strains with binary toxin genes among clinical isolates in an
American hospital." J Med Microbiol 53(Pt 9): 887-94.
Ghosh, S. and P. Setlow (2009). "Isolation and characterization of superdormant spores
of Bacillus species." Journal of Bacteriology 191(6): 1787-1797.
Ghosh, S. and P. Setlow (2010). "The preparation, germination properties and stability of
superdormant spores of Bacillus cereus." J Appl Microbiol 108(2): 582-590.
Ghosh, S., P. Zhang, et al. (2009). "Superdormant spores of Bacillus species have
elevated wet-heat resistance and temperature requirements for heat activation." J
Bacteriol 191(18): 5584-5591.
Gill, C. O. (2007). "Microbiological conditions of meats from large game animals and
birds." Meat Science 77(2): 149-160.
Goorhuis, A., D. Bakker, et al. (2008). "Emergence of Clostridium difficile infection due
to a new hypervirulent strain, polymerase chain reaction ribotype 078." Clin
Infect Dis 47(9): 1162-1170.
Gould, L. H. and B. Limbago (2010). "Clostridium difficile in Food and Domestic
Animals: A New Foodborne Pathogen?" Clin Infect Dis 51(5): 577-82.
Goulding, D., H. Thompson, et al. (2009). "Distinctive profiles of infection and
pathology in hamsters infected with Clostridium difficile strains 630 and B1."
Infect Immun 77(12): 5478-85.
Gurian, L., T. T. Ward, et al. (1982). "Possible foodborne transmission in a case of
pseudomembranous colitis due to Clostridium difficile: influence of
gastrointestinal secretions on Clostridium difficile infection." Gastroenterology
83(2): 465-9.
Hammitt, M. C., D. M. Bueschel, et al. (2008). "A possible role for Clostridium difficile
in the etiology of calf enteritis." Vet Microbiol 127(3-4): 343-52.
247
He, M., M. Sebaihia, et al. "Evolutionary dynamics of Clostridium difficile over short and
long time scales." Proceedings of the National Academy of Sciences 107(16):
7527-7532.
Hecht, D. W. and D. Citron (2007). "Clinical Laboratory Standards Institute. M11-A7Methods for antimicrobial susceptibility testing of anaerobic bacteria - Seven
Edition." 27(2): 64p.
Hensgens, M. P., A. Goorhuis, et al. (2009). "Decrease of hypervirulent Clostridium
difficile PCR ribotype 027 in the Netherlands." Euro Surveill 14(45).
Hensgens, M. P., A. Goorhuis, et al. (2011). "Clostridium difficile infection in an
endemic setting in the Netherlands." Eur J Clin Microbiol Infect Dis 30(4): 58793.
Heredia, N. L., G. A. Garcia, et al. (1997). "Elevation of the heat resistance of vegetative
cells and spores of Clostridium perfringens type A by sublethal heat shock."
Journal of food protection 60(8): 998-1000.
Holzel, C. S., K. S. Harms, et al. (2010). "Resistance to Linezolid in a Porcine
Clostridium perfringens Strain Carrying a Mutation in the rplD Gene Encoding
the Ribosomal Protein L4." Antimicrobial Agents and Chemotherapy 54(3):
1351-1353.
Huang, H., A. Weintraub, et al. (2010). "Antimicrobial susceptibility and heteroresistance
in Chinese Clostridium difficile strains." Anaerobe 16(6): 633-635.
Iizuka, M., H. Itou, et al. (2004). "Elemental diet modulates the growth of Clostridium
difficile in the gut flora." Aliment Pharmacol Ther 20 Suppl 1: 151-7.
Indra, A., H. Lassnig, et al. (2009). "Clostridium difficile: a new zoonotic agent?" Wien
Klin Wochenschr 121(3-4): 91-5.
Janvilisri, T., J. Scaria, et al. (2009). "Microarray identification of Clostridium difficile
core components and divergent regions associated with host origin." J Bacteriol
191(12): 3881-91.
248
Jhung, M. A., A. D. Thompson, et al. (2008). "Toxinotype V Clostridium difficile in
humans and food animals." Emerg Infect Dis 14(7): 1039-1045.
Jones, M. A. and D. Hunter (1983). "Isolation of Clostridium difficile from pigs." Vet
Rec 112(11): 253.
Jones, R. N., T. R. Fritsche, et al. (2007). "LEADER surveillance program results for
2006: an activity and spectrum analysis of linezolid using clinical isolates from
the United States (50 medical centers)." Diagn Microbiol Infect Dis 59(3): 30917.
Jones, R. N., S. Kohno, et al. (2009). "ZAAPS International Surveillance Program (2007)
for linezolid resistance: results from 5591 Gram-positive clinical isolates in 23
countries." Diagn Microbiol Infect Dis 64(2): 191-201.
Jones, R. N. and R. Masterton (2001). "Determining the value of antimicrobial
surveillance programs." Diagn Microbiol Infect Dis 41(4): 171-5.
Jump, R. L. P., M. J. Pultz, et al. (2007). "Vegetative Clostridium difficile Survives in
Room Air on Moist Surfaces and in Gastric Contents with Reduced Acidity: a
Potential Mechanism To Explain the Association between Proton Pump Inhibitors
and C. difficile-Associated Diarrhea?" Antimicrob. Agents Chemother. 51(8):
2883-2887.
Juneja, V. K., B. S. Eblen, et al. (2001). "Modeling non-linear survival curves to
calculate thermal inactivation of salmonella in poultry of different fat levels."
International journal of food microbiology 70(1-2): 37-51.
Juneja, V. K., B. S. Eblen, et al. (2001). "Thermal Inactivation of Salmonella spp. in
Chicken Broth, Beef, Pork, Turkey, and Chicken: Determination of D- and Zvalues." Journal of Food Science 66(1): 146-152.
Juneja, V. K., P. G. Klein, et al. (1998). "Heat shock and thermotolerance of Escherichia
coli O157:H7 in a model beef gravy system and ground beef." J Appl Microbiol
84(4): 677-684.
249
Juneja, V. K., J. S. Novak, et al. (2003). "Increased thermotolerance of Clostridium
perfringens spores following sublethal heat shock." Food Control 14(3): 163-168.
Kaatz, G. W., S. D. Gitlin, et al. (1988). "Acquisition of Clostridium difficile from the
hospital environment." Am J Epidemiol 127(6): 1289-94.
Keel, K., J. S. Brazier, et al. (2007). "Prevalence of PCR ribotypes among Clostridium
difficile isolates from pigs, calves, and other species." J Clin Microbiol 45(6):
1963-1964.
Keessen, E. C., C. J. Donswijk, et al. (2010) "Aerial dissemination of Clostridium
difficile on a pig farm and its environment. Available from: http://igiturarchive.library.uu.nl/student-theses/2011-0518-200326/artikel.doc. Accessed July
21, 2011."
Keessen, E. C., A. J. van den Berkt, et al. (2011). "The relation between farm specific
factors and prevalence of Clostridium difficile in slaughter pigs." Vet Microbiol.
Kim, H., T. V. Riley, et al. (2008). "Increasing prevalence of toxin A-negative, toxin Bpositive isolates of Clostridium difficile in Korea: impact on laboratory
diagnosis." J Clin Microbiol 46(3): 1116-7.
Kiss, D. and G. Bilkei (2005). "A new periparturient disease in Eastern Europe,
Clostridium difficile causes postparturient sow losses." Theriogenology 63(1): 1723.
Kuijper, E. J., F. Barbut, et al. (2008). "Update of Clostridium difficile infection due to
PCR ribotype 027 in Europe, 2008." Euro Surveill 13(31).
Kuijper, E. J. and J. T. van Dissel (2008). "Spectrum of Clostridium difficile infections
outside health care facilities." CMAJ 179(8): 747-8.
Kyne, L. (2010). "Clostridium difficile--beyond antibiotics." N Engl J Med 362(3): 264-5.
Lawley, T. D., S. Clare, et al. (2009). "Antibiotic Treatment of Clostridium difficile
Carrier Mice Triggers a Supershedder State, Spore-Mediated Transmission, and
250
Severe Disease in Immunocompromised Hosts." Infect. Immun. 77(9): 36613669.
Lefebvre, S. L., L. G. Arroyo, et al. (2006). "Epidemic Clostridium difficile strain in
hospital visitation dog." Emerg Infect Dis 12(6): 1036-7.
Lemee, L., A. Dhalluin, et al. (2004). "Multiplex PCR targeting tpi (triose phosphate
isomerase), tcdA (Toxin A), and tcdB (Toxin B) genes for toxigenic culture of
Clostridium difficile." J Clin Microbiol 42(12): 5710-5714.
Levett, P. N. and G. Margaritis-Bassoulis (1985). "Enrichment media for isolation of
Clostridium difficile from faeces." Eur J Clin Microbiol 4(2): 135-136.
Libby, D. B. and G. Bearman (2009). "Bacteremia due to Clostridium difficile--review of
the literature." International Journal of Infectious Diseases 13(5): e305-e309.
Limbago, B., C. Long, et al. (2009). "Clostridium difficile Strains from Community-1 Associated
Infections." J of Clinical Micro Published online ahead of print July 1, 2009.
Limbago, B. M., C. M. Long, et al. (2009). "Clostridium difficile strains from
community-associated infections." J Clin Microbiol 47(9): 3004-3007.
Loo, V. G., L. Poirier, et al. (2005). "A predominantly clonal multi-institutional outbreak
of Clostridium difficile-associated diarrhea with high morbidity and mortality." N
Engl J Med 353(23): 2442-9.
Lyras, D., J. R. O'Connor, et al. (2009). "Toxin B is essential for virulence of Clostridium
difficile." Nature 458(7242): 1176-9.
McFarland, L. V. (2008). "Update on the changing epidemiology of Clostridium difficileassociated disease." Nat Clin Pract Gastroenterol Hepatol 5(1): 40-48.
Medina-Torres, C. E., J. S. Weese, et al. (2011). "Prevalence of Clostridium difficile in
horses." Veterinary Microbiology.
251
Meisel-Mikolajczyk, F., E. Kaliszuk-Kaminska, et al. (1995). "[Study of the
thermoresistance of Clostridium difficile spores]." Med Dosw Mikrobiol 47(3-4):
177-181.
Metcalf, D. S., M. C. Costa, et al. (2010). "Clostridium difficile in vegetables, Canada."
Lett Appl Microbiol 51(5): 600-602.
Morgan, O. W., B. Rodrigues, et al. (2008). "Clinical severity of Clostridium difficile
PCR ribotype 027: a case-case study." PLoS One 3(3): e1812.
Mulvey, M. R., D. A. Boyd, et al. (2010). "Hypervirulent Clostridium difficile strains in
hospitalized patients, Canada." Emerg Infect Dis 16(4): 678-681.
Muto, C. A., M. Pokrywka, et al. (2005). "A large outbreak of Clostridium difficileassociated disease with an unexpected proportion of deaths and colectomies at a
teaching hospital following increased fluoroquinolone use." Infect Control Hosp
Epidemiol 26(3): 273-80.
Nagy, E. and M. J. Dowzicky (2010). "In vitro activity of tigecycline and comparators
against a European compilation of anaerobes collected as part of the Tigecycline
Evaluation and Surveillance Trial (TEST)." Scand J Infect Dis 42(1): 33-8.
NASA (2010). "NASA Video Shows Global Reach of Pollution from Fires."
NCBA (2002). "April 18. National Cattlemen's Beef Association. News archive:
Increased Export of U.S. Trimmings Could Benefit McDonald‘s, Help Improve
Value for U.S. Producers.
http://www.beefusa.org/NEWSIncreasedExportofUSTrimmingsCouldBenefitMc
DonaldsHelpImproveValueforUSProducers4927.aspx." September
14(September 14).
Norman, K. N., R. B. Harvey, et al. (2009). "Varied prevalence of Clostridium difficile in
an integrated swine operation." Anaerobe 15(6): 256-260.
Norman, K. N., H. M. Scott, et al. (2011). "Prevalence and Genotypic Characteristics of
Clostridium difficile in a Closed and Integrated Human and Swine Population."
Appl Environ Microbiol.
252
O'Brien, J. A., B. J. Lahue, et al. (2007). "The emerging infectious challenge of
Clostridium difficile-associated disease in Massachusetts hospitals: clinical and
economic consequences." Infect Control Hosp Epidemiol 28(11): 1219-27.
O'Neill, G. L., F. T. Ogunsola, et al. (1996). "Modification of a PCR Ribotyping Method
for Application as a Routine Typing Scheme forClostridium difficile." Anaerobe
2(4): 205-209.
Oliver, S. P., D. A. Patel, et al. (2009). "ASAS Centennial Paper: Developments and
future outlook for preharvest food safety." J. Anim Sci. 87(1): 419-437.
Organization, F. a. A. O. o. t. U. N. a. W. H. (2008). Microbiological hazards in fresh
leafy vegetables and herbs: Meeting Report. Microbiological Risk Assessment
Series. , FAO/WHO.
Paredes-Sabja, D., C. Bond, et al. (2008). "Germination of spores of Clostridium difficile
strains, including isolates from a hospital outbreak of Clostridium difficileassociated disease (CDAD)." Microbiology 154(8): 2241-2250.
Paredes-Sabja, D., J. A. Torres, et al. (2008). "Clostridium perfringens spore
germination: characterization of germinants and their receptors." J Bacteriol
190(4): 1190-1201.
Pearce, M. C., D. Fenlon, et al. (2004). "Distribution of Escherichia coli O157 in Bovine
Fecal Pats and Its Impact on Estimates of the Prevalence of Fecal Shedding."
Appl. Environ. Microbiol. 70(10): 5737-5743.
Pell, A. N. (1997). "Manure and microbes: public and animal health problem?" J Dairy
Sci 80(10): 2673-81.
Pepin, J., M. E. Alary, et al. (2005). "Increasing risk of relapse after treatment of
Clostridium difficile colitis in Quebec, Canada." Clin Infect Dis 40(11): 15911597.
Pepin, J., L. Valiquette, et al. (2004). "Clostridium difficile-associated diarrhea in a
region of Quebec from 1991 to 2003: a changing pattern of disease severity."
CMAJ 171(5): 466-72.
253
Persson, S., M. Torpdahl, et al. (2008). "New multiplex PCR method for the detection of
Clostridium difficile toxin A (tcdA) and toxin B (tcdB) and the binary toxin
(cdtA/cdtB) genes applied to a Danish strain collection." Clin Microbiol Infect
14(11): 1057-1064.
Petrie, A. and P. Watson (1999). Statistics for Veterinary and Animal Sciences. Oxford,
UK., Blackwell Science.
Pirs, T., M. Ocepek, et al. (2008). "Isolation of Clostridium difficile from food animals in
Slovenia." J Med Microbiol 57(Pt 6): 790-792.
Porter MC, Reggiardo C, et al. (2002). Association of Clostridium difficile with bovine
neonatal diarrhea [abstract]. In: Abstracts of the Proceedings of the American
Association of Veterinary Laboratory Diagosticians. 45th annual meeting. Saint
Louis; 2002 Oct 19-21. Davis (CA): American Association of Veterinary
Laboratory Diagosticians.
Post, K. W., B. H. Jost, et al. (2002). "Evaluation of a test for Clostridium difficile toxins
A and B for the diagnosis of neonatal swine enteritis." J Vet Diagn Invest 14(3):
258-9.
Rafii, F. and M. Park (2005). "Effects of gyrase mutation on the growth kinetics of
ciprofloxacin-resistant strains of Clostridium perfringens." Anaerobe 11(4): 201205.
Randall, E. J. (2003) "Horse Manure Management: The Nitrogen Enhancement System."
Redelings MD, Sorvillo F, et al. (2007). "Increase in Clostridium difficile-related
mortality rates, United States, 1999-2004." Emerg Infect Dis 13(9): 1417-1419.
Reil, M., M. Erhard, et al. (2011). "Recognition of Clostridium difficile PCR-ribotypes
001, 027 and 126/078 using an extended MALDI-TOF MS system." Eur J Clin
Microbiol Infect Dis.
Rello, J. (2005). "Pharmacokinetics, pharmacodynamics, safety and tolerability of
tigecycline." J Chemother 17 Suppl 1: 12-22.
254
Ricciardi, R., D. A. Rothenberger, et al. (2007). "Increasing prevalence and severity of
Clostridium difficile colitis in hospitalized patients in the United States." Arch
Surg 142(7): 624-31; discussion 631.
Richter, S. H., J. P. Garner, et al. (2009). "Environmental standardization: cure or cause
of poor reproducibility in animal experiments?" Nat Methods 6(4): 257-261.
Riley, T. V. (1998). "Clostridium difficile: a pathogen of the nineties." Eur J Clin
Microbiol Infect Dis 17(3): 137-41.
Roberts, K., C. F. Smith, et al. (2008). "Aerial dissemination of Clostridium difficile
spores." BMC Infect Dis 8: 7.
Roberts, S., H. Heffernan, et al. (2011). "Molecular epidemiology and susceptibility
profiles of Clostridium difficile in New Zealand, 2009." N Z Med J 124(1332):
45-51.
Rodriguez-Palacios, A., M. Koohmaraie, et al. (2011). "Prevalence, Enumeration, and
Antimicrobial Agent Resistance in Clostridium difficile in Cattle at Harvest in the
United States." J Food Prot October, In press.
Rodriguez-Palacios, A. and J. T. LeJeune (2011). "Moist-heat resistance, spore aging,
and superdormancy in Clostridium difficile." Appl. Environ. Microbiol. 77(9):
3085-3091.
Rodriguez-Palacios, A., C. Pickworth, et al. (2011). "Transient Fecal Shedding and
Limited Animal-to-Animal Transmission of Clostridium difficile by Naturally
Infected Finishing Feedlot Cattle." Appl Environ Microbiol 77(10): 3391-3397.
Rodriguez-Palacios, A., R. J. Reid-Smith, et al. (2009). "Possible seasonality of
Clostridium difficile in retail meat, Canada." Emerg Infect Dis 15(5): 802-805.
Rodriguez-Palacios, A., R. J. Reid-Smith, et al. (2010). "Clostridium difficile survives
minimal temperature recommended for cooking ground meats." Anaerobe 16:
540-542.
255
Rodriguez-Palacios, A., H. R. Staempfli, et al. (2007). "Clostridium difficile in retail
ground meat, Canada." Emerg Infect Dis 13(3): 485-7.
Rodriguez-Palacios, A., H. R. Stampfli, et al. (2006). "Clostridium difficile PCR
ribotypes in calves, Canada." Emerg Infect Dis 12(11): 1730-6.
Rodriguez-Palacios, A., H. R. Stampfli, et al. (2007). "Natural and experimental infection
of neonatal calves with Clostridium difficile." Vet Microbiol 124(1-2): 166-172.
Rodriguez-Palacios, A., H. R. Stampfli, et al. (2007). "Natural and experimental infection
of neonatal calves with Clostridium difficile." Vet Microbiol 124(1-2): 166-72.
Rodriguez-Palacios, A., H. R. Stampfli, et al. (2009). "Possible seasonality of
Clostridium difficile in retail meat, Canada." Emerg Infect Dis: 831-834.
Rupnik, M. (2007). "Is Clostridium difficile-associated infection a potentially zoonotic
and foodborne disease?" Clin Microbiol Infect 13(5): 457-459.
Rupnik, M., B. Dupuy, et al. (2005). "Revised nomenclature of Clostridium difficile
toxins and associated genes." J Med Microbiol 54(Pt 2): 113-117.
Rupnik, M., A. Widmer, et al. (2008). "Clostridium difficile toxinotype V, ribotype 078,
in animals and humans." J Clin Microbiol 46(6): 2146.
Rupnik, M., M. H. Wilcox, et al. (2009). "Clostridium difficile infection: new
developments in epidemiology and pathogenesis." Nat Rev Microbiol 7(7): 52636.
Russmann, H., K. Panthel, et al. (2007). "Evaluation of three rapid assays for detection of
Clostridium difficile toxin A and toxin B in stool specimens." Eur J Clin
Microbiol Infect Dis 26(2): 115-9.
Sailhamer, E. A., K. Carson, et al. (2009). "Fulminant Clostridium difficile colitis:
patterns of care and predictors of mortality." Arch Surg 144(5): 433-9; discussion
439-40.
256
Sanchez-Plata, M. X., A. Amezquita, et al. (2005). "Predictive model for Clostridium
perfringens growth in roast beef during cooling and inhibition of spore
germination and outgrowth by organic acid salts." J Food Prot 68(12): 2594-2605.
Sargeant, J. M., A. M. O‘Connor, et al. (2010). "The REFLECT Statement: Reporting
Guidelines for Randomized Controlled Trials in Livestock and Food Safety:
Explanation and Elaboration†." Zoonoses and Public Health 57(2): 105-136.
Sathish, S. and K. Swaminathan (2008). "Molecular characterization of the diversity of
Clostridium chauvoei isolates collected from two bovine slaughterhouses:
Analysis of cross-contamination." Anaerobe 14(3): 190-199.
Scallan, E., R. M. Hoekstra, et al. (2011). "Foodborne illness acquired in the United
States--major pathogens." Emerg Infect Dis 17(1): 7-15.
Schwan, C., B. r. Stecher, et al. (2009). "<italic>Clostridium difficile</italic> Toxin CDT
Induces Formation of Microtubule-Based Protrusions and Increases Adherence of
Bacteria." PLoS Pathog 5(10): e1000626.
Shim, J. K., S. Johnson, et al. (1998). "Primary symptomless colonisation by Clostridium
difficile and decreased risk of subsequent diarrhoea." Lancet 351(9103): 633-636.
Shin, B. M., E. Y. Kuak, et al. (2008). "Multicentre study of the prevalence of toxigenic
Clostridium difficile in Korea: results of a retrospective study 2000-2005." J Med
Microbiol 57(Pt 6): 697-701.
Shin, B. M., E. Y. Kuak, et al. (2008). "Emerging toxin A-B+ variant strain of
Clostridium difficile responsible for pseudomembranous colitis at a tertiary care
hospital in Korea." Diagn Microbiol Infect Dis 60(4): 333-7.
Silvestre, D., P. Ruiz, et al. (2008). "Effect of Pasteurization on the Bactericidal Capacity
of Human Milk." J Hum Lact 24(4): 371-376.
Simango, C. (2006). "Prevalence of Clostridium difficile in the environment in a rural
community in Zimbabwe." Transactions of the Royal Society of Tropical
Medicine and Hygiene 100(12): 1146-1150.
257
Simango, C. (2006). "Prevalence of Clostridium difficile in the environment in a rural
community in Zimbabwe." Trans R Soc Trop Med Hyg 100(12): 1146-50.
Simango, C. and S. Mwakurudza (2008). "Clostridium difficile in broiler chickens sold at
market places in Zimbabwe and their antimicrobial susceptibility." Int J Food
Microbiol 124(3): 268-70.
Smith, L. D. and E. O. King (1962). "Occurrence of Clostridium difficile in infections of
man." J Bacteriol 84: 65-7.
Solomon, K., S. Fanning, et al. (2011). "PCR ribotype prevalence and molecular basis of
macrolide-lincosamide-streptogramin B (MLSB) and fluoroquinolone resistance
in Irish clinical Clostridium difficile isolates." J Antimicrob Chemother.
Songer, J. G. (2004). "The emergence of Clostridium difficile as a pathogen of food
animals." Anim Health Res Rev 5(2): 321-6.
Songer, J. G. (2009). "Clostridia as agents of zoonotic disease." Vet Microbiol.
Songer, J. G., R. Jones, et al. (2007). "Prevention of porcine Clostridium difficileassociated disease by competitive exclusion with nontoxigenic organisms." Vet
Microbiol 124(3-4): 358-361.
Songer, J. G., H. T. Trinh, et al. (2009). "Clostridium difficile in retail meat products,
USA, 2007." Emerg Infect Dis 15(5): 819-821.
Spigaglia, P., F. Barbanti, et al. (2008). "Fluoroquinolone resistance in Clostridium
difficile isolates from a prospective study of C. difficile infections in Europe." J
Med Microbiol 57(Pt 6): 784-9.
Spigaglia, P. and P. Mastrantonio (2002). "Molecular analysis of the pathogenicity locus
and polymorphism in the putative negative regulator of toxin production (TcdC)
among Clostridium difficile clinical isolates." J Clin Microbiol 40(9): 3470-3475.
Stabler, R. A., D. N. Gerding, et al. (2006). "Comparative phylogenomics of Clostridium
difficile reveals clade specificity and microevolution of hypervirulent strains." J
Bacteriol 188(20): 7297-305.
258
Stare, B. G., M. Delmee, et al. (2007). "Variant forms of the binary toxin CDT locus and
tcdC gene in Clostridium difficile strains." J Med Microbiol 56(Pt 3): 329-35.
Sullivan, N. M., S. Pellett, et al. (1982). "Purification and characterization of toxins A
and B of Clostridium difficile." Infect. Immun. 35(3): 1032-1040.
Tanner, H. E., K. J. Hardy, et al. (2010). "Co-existence of multiple MLVA sub-types of
Clostridium difficile PCR-Ribotype 027 strains within faecal specimens." J Clin
Microbiol.
Terhes, G., E. Urban, et al. (2004). "Community-Acquired Clostridium difficile Diarrhea
Caused by Binary Toxin, Toxin A, and Toxin B Gene-Positive Isolates in
Hungary." J. Clin. Microbiol. 42(9): 4316-4318.
Thakur, S., M. Putnam, et al. (2010). "Prevalence of antimicrobial resistance and
association with toxin genes in Clostridium difficile in commercial swine."
American Journal of Veterinary Research 71(10): 1189-1194.
Thitaram, S. N., J. F. Frank, et al. (2011). "Clostridium difficile from Healthy Food
Animals: Optimized Isolation and Prevalence." J Food Prot 74(1): 130-133.
Thomas, J. A., K. C. Newman, et al. (2011). "Bacteraemia from an unrecognized source
(occult bacteraemia) occurring during Clostridium difficile infection."
Scandinavian Journal of Infectious Diseases 43(4): 269-274.
Thonnard, J., F. Carreer, et al. (1996). "Toxin A detection on Clostridium difficile
colonies from 24-h cultures." Clin Microbiol Infect 2(1): 50-54.
Turcs, aacute, et al. (2001). "Occurrence of Anaerobic Bacterial, Clostridial, and
Clostridium perfringens Spores in Raw Goose Livers from a Poultry Processing
Plant in Hungary." Journal of food protection 64: 1252-1254.
UCLA "Multinomial Logistic Regression. UCLA: Academic Technology Services,
Statistical Consulting Group. From
http://www.ats.ucla.edu/stat/stata/dae/mlogit.htm (Accessed July 10, 2011)."
259
USDA "United States Department of Agriculture - Food Safety and Inspection Service.
Food Safety Education - Is it done yet? From:
http://www.fsis.usda.gov/is_it_done_yet/brochure_text/index.asp#SMIT
(Accessed January 12, 2011)."
USDA. (1999). "Food Safety and Inspection Agency - United States Department of
Agriculture. Time-temperature tables for cooking ready-to-eat poultry products."
Retrieved February 7, from
http://www.fsis.usda.gov/oppde/rdad/fsisnotices/rte_poultry_tables.pdf.
USDA. (2010, September 14). "Estimated red meat production. CattleFax and United
States Department of Agriculture reports: Livestock slaughter and meat
production." Retrieved May 14 2010, from http://www.cattlefax.com/quicklinksdata/usda-reports/.
USDA, F. S. a. I. S. (1997) "U.S.D.A. Promotes Horse & Goat Meat."
van Donkelaar, A., R. V. Martin, et al. (2010). "Global estimates of ambient fine
particulate matter concentrations from satellite-based aerosol optical depth:
development and application." Environ Health Perspect 118(6): 847-55.
Vengust, M., L. G. Arroyo, et al. (2003). "Preliminary evidence for dormant clostridial
spores in equine skeletal muscle." Equine Vet J 35(5): 514-516.
Von Abercron, S. M., F. Karlsson, et al. (2009). "Low occurrence of Clostridium difficile
in retail ground meat in Sweden." J Food Prot 72(8): 1732-4.
Warny, M., J. Pepin, et al. (2005). "Toxin production by an emerging strain of
Clostridium difficile associated with outbreaks of severe disease in North America
and Europe." Lancet 366(9491): 1079-1084.
Weese, J. S. (2010). "Clostridium difficile in food--innocent bystander or serious threat?"
Clin Microbiol Infect 16(1): 3-10.
Weese, J. S. and J. Armstrong (2003). "Outbreak of Clostridium difficile-associated
disease in a small animal veterinary teaching hospital." J Vet Intern Med 17(6):
813-6.
260
Weese, J. S., B. P. Avery, et al. (2009). "Detection and enumeration of Clostridium
difficile in retail beef and pork." Appl Environ Microbiol 75(15): 5009-5011.
Weese, J. S., R. Finley, et al. (2010). "Evaluation of Clostridium difficile in dogs and the
household environment." Epidemiol Infect: 1-5.
Weese, J. S., R. J. Reid-Smith, et al. (2010). "Detection and characterization of
Clostridium difficile in retail chicken." Lett Appl Microbiol 50(4): 362-5.
Weese, J. S., H. R. Staempfli, et al. (2000). "Survival of Clostridium difficile and its
toxins in equine feces: implications for diagnostic test selection and
interpretation." J Vet Diagn Invest 12(4): 332-336.
Weese, J. S., L. Toxopeus, et al. (2006). "Clostridium difficile associated diarrhoea in
horses within the community: predictors, clinical presentation and outcome."
Equine Vet J 38(2): 185-8.
Weese, J. S., H. E. Weese, et al. (2001). "Suspected Clostridium difficile-associated
diarrhea in two cats." J Am Vet Med Assoc 218(9): 1436-9, 1421.
Wei, J., I. M. Shah, et al. (2010). "Superdormant spores of Bacillus species germinate
normally with high pressure, peptidoglycan fragments, and bryostatin." J
Bacteriol 192(5): 1455-1458.
Wilcox, M. H. (2004). "Descriptive study of intravenous immunoglobulin for the
treatment of recurrent Clostridium difficile diarrhoea." J Antimicrob Chemother
53(5): 882-4.
Wilcox, M. H., W. Fawley, et al. (2000). "In vitro activity of new generation
fluoroquinolones against genotypically distinct and indistinguishable Clostridium
difficile isolates." J Antimicrob Chemother 46(4): 551-556.
Wilcox, M. H. and W. N. Fawley (2000). "Hospital disinfectants and spore formation by
Clostridium difficile." Lancet 356(9238): 1324.
Wilcox, M. H., L. Mooney, et al. (2008). "A case-control study of community-associated
Clostridium difficile infection." J Antimicrob Chemother 62(2): 388-96.
261
Wilson, K. H., M. J. Kennedy, et al. (1982). "Use of sodium taurocholate to enhance
spore recovery on a medium selective for Clostridium difficile." J Clin Microbiol
15(3): 443-446.
Wise, R., J. M. Andrews, et al. (1998). "The in-vitro activity of linezolid (U-100766) and
tentative breakpoints." J Antimicrob Chemother 42(6): 721-728.
Wiuff, C., D. J. Brown, et al. (2011). "The epidemiology of Clostridium difficile in
Scotland." J Infect 62(4): 271-279.
Wultanska, D., H. Pituch, et al. (2004). "[Comparative study of thermoresistance spores
of Clostridium difficile strains belonging to different toxigenicity groups]." Med
Dosw Mikrobiol 56(2): 155-159.
Yates, W. D. (1982). "A review of infectious bovine rhinotracheitis, shipping fever
pneumonia and viral-bacterial synergism in respiratory disease of cattle." Can J
Comp Med 46(3): 225-63.
Young, M. and J. Mandelstam (1979). Early events during bacterial endospore formation.
Advances in Microbial Physiology. A. H. Rose, Academic Press, Inc. London. 20
103-162.
Zhang, J., S. K. Wall, et al. (2010). "Contamination Rates and Antimicrobial Resistance
in Bacteria Isolated from Grass-Fed Labeled Beef Products." Foodborne
Pathogens and Disease 7(11): 1331-1336.
Zidaric, V., S. Beigot, et al. (2010). "The occurrence and high diversity of Clostridium
difficile genotypes in rivers." Anaerobe.
Zidaric, V., M. Zemljic, et al. (2008). "High diversity of Clostridium difficile genotypes
isolated from a single poultry farm producing replacement laying hens." Anaerobe
14(6): 325-7.
Zidaric, V., M. Zemljic, et al. (2008). "High diversity of Clostridium difficile genotypes
isolated from a single poultry farm producing replacement laying hens." Anaerobe
14(6): 325-327.
262