ABSTRACT

A robust high-throughput multilocus sequence typing (MLST) scheme for Clostridium difficile was developed and validated using a diverse collection of 50 reference isolates representing 45 different PCR ribotypes and 102 isolates from recent clinical samples. A total of 49 PCR ribotypes were represented overall. All isolates were typed by MLST and yielded 40 sequence types (STs). A web-accessible database was set up (http://pubmlst.org/cdifficile/ ) to facilitate the dissemination and comparison of C. difficile MLST genotyping data among laboratories. MLST and PCR ribotyping were similar in discriminatory abilities, having indices of discrimination of 0.90 and 0.92, respectively. Some STs corresponded to a single PCR ribotype (32/40), other STs corresponded to multiple PCR ribotypes (8/40), and, conversely, the PCR ribotype was not always predictive of the ST. The total number of variable nucleotide sites in the concatenated MLST sequences was 103/3,501 (2.9%). Concatenated MLST sequences were used to construct a neighbor-joining tree which identified four phylogenetic groups of STs and one outlier (ST-11; PCR ribotype 078). These groups apparently correlate with clades identified previously by comparative genomics. The MLST scheme was sufficiently robust to allow direct genotyping of C. difficile in total stool DNA extracts without isolate culture. The direct (nonculture) MLST approach may prove useful as a rapid genotyping method, potentially benefiting individual patients and informing hospital infection control.
Clostridium difficile is an obligate anaerobic Gram-positive bacillus carried asymptomatically in the gut of approximately 2 to 7% of healthy human adults (28, 35). Nosocomial acquisition of C. difficile in humans is common, and symptoms ranging from mild diarrhea to severe pseudomembranous colitis can develop during antibiotic treatment or shortly afterwards (1, 28, 40). Symptoms are caused by toxins A and B encoded by the tcdA and tcdB genes located within the pathogenicity locus (PaLoc) and potentially an additional binary toxin (reviewed in reference 7).
The rate and severity of nosocomial infections increased between the years 2000 and 2008 (27, 46, 47), coincident with the emergence of a hypervirulent fluoroquinolone-resistant clone, designated PCR ribotype 027 (31, 34). Outbreaks of C. difficile infection (CDI) caused by the 027 clone have been reported in North America and throughout Europe (15, 18, 23, 31, 32). This strain is now endemic, causing 36% of cases in England, United Kingdom, from April 2008 to March 2009 (10). Mortality can range from 6 to 15% (2), and the economic burden of CDI is substantial, with an estimated direct cost of over $6,000 per case in the United States (37).
Studies of global epidemiology require easily comparable genotyping data for large numbers of bacterial isolates. Genotyping methods in common use for C. difficile include PCR ribotyping, pulsed-field gel electrophoresis (PFGE), and restriction-endonuclease analysis (REA) (4, 10, 26, 41). These techniques are generally labor- and resource-intensive, not easily adapted to very high-throughput, and often restricted to reference laboratories. Furthermore, interlaboratory comparison of data can be difficult when they are based on gel banding patterns. Multilocus sequence typing (MLST) is a microbial genotyping method facilitating isolate discrimination using nucleotide sequences of housekeeping gene fragments (24). Each unique combination of alleles is assigned a sequence type (ST) number. The MLST technique is scalable, according to the question to be addressed or the resources available, and amenable to automation using very-high-throughput robotic systems (12). Searchable Internet-accessible MLST databases (as at http://pubmlst.org/ ) allow laboratories performing MLST to maintain ownership of their data. However, having a single laboratory as curator of the database avoids the potential confusion that arises when new allele and ST numbers are assigned (14). MLST data are also a powerful tool for studying the population biology of bacterial species (25, 42). An MLST scheme for C. difficile has been described (20) but has not been widely adopted. This may be partly because the ddl locus was a null allele and failed to amplify in certain strains (L. Lemee, personal communication). Furthermore, a curated Internet-accessible database was not available.
Our aim was to develop a robust MLST scheme for C. difficile and to set up an Internet-accessible database to allow the simple depositing, retrieval, and comparison of data. The scheme was validated using two collections of isolates, one representing a wide range of PCR ribotypes and another consisting of isolates cultured from recent clinical samples. The scheme was also sufficiently robust to allow direct MLST typing of C. difficile in total DNA extracts from stool samples, avoiding the need for isolate culture and enabling rapid genotyping (3.5 days for 24 samples) to be performed.

MATERIALS AND METHODS

C. difficile stools and culture.

A total of 215 human stool specimens submitted to the Clinical Microbiology Laboratory, John Radcliffe Hospital, Oxford, United Kingdom, between 12 July and 17 October 2008 were included in this study. Stools were from both hospital and community patients. Stools were chosen so that half were sequential, enzyme-linked immunosorbent assay (ELISA) positive (n = 107) with sufficient stool remaining, and half were ELISA negative (n = 108) submitted during the same time period (Premier Toxins A&B Enzyme Immunoassay; Meridian Bioscience Europe, Villa Cortese, Italy). All stool samples underwent culture for C. difficile. Industrial methylated spirits ([IMS] 0.5 ml) was added to a 0.5-ml fecal sample (pea-sized portion if the stool was formed), and the sample was vortex mixed and incubated at room temperature for 1 h. A loopful was then cultured onto modified Brazier's cycloserine-cefoxitin-egg yolk (CCEY) agar (CCEY agar base containing cycloserine-cefoxitin supplement and 5% defibrinated horse blood), and the plates were incubated anaerobically at 37°C for up to 7 days. A single colony was subcultured onto a Columbia blood agar (CBA) plate and incubated for 48 h, after which colonies giving the characteristic odor and fluorescence under UV illumination were obtained. For long-term storage, isolates were emulsified in nutrient broth containing 10% glycerol and stored at −80°C.
An additional 50 isolates were obtained from a collection held at Leeds General Infirmary (reference laboratory for the C. difficile Ribotyping Network for England and Northern Ireland). They represented 45 different PCR ribotypes (plus five duplicates) chosen to represent the overall genetic diversity of C. difficile (determined by PCR ribotyping) and were used to validate the MLST scheme.

Extraction of total stool DNA.

Total DNA was extracted from stool samples using a FastPrep homogenizer (MP Biomedicals Europe, Illkirch, France) to lyse cells and spores, followed by DNA purification using a FastDNA Spin kit for soil (MP Biomedicals). The manufacturer's protocol was followed with the following refinements. Stool samples (100 μl or equivalent volume if the stool was formed) were added to 978 μl of sodium phosphate buffer in impact-resistant 2.0-ml tubes containing matrix E, which comprises 1.4-mm ceramic spheres, 0.1-mm silica spheres, and one 4-mm glass bead in buffer. MT buffer (MP Biomedicals) (122 μl) was added, and stools were homogenized for 40 s at a speed of 6.0 m per s. The lysate was clarified by centrifugation at 13,000 rpm for 15 min. Proteins were precipitated using 250 μl of protein precipitation solution (PPS) and removed by centrifugation for 10 min. The supernatant was mixed with 1 ml of silica binding matrix for 2 min to take up DNA, and then the matrix was allowed to settle for 5 min. The binding matrix was transferred to a SPIN filter (MP Biomedicals) and washed using 500 μl of SEWS-M (salt-ethanol wash solution). After the sample was air dried at room temperature, DNA was eluted from the matrix in 100 μl of DNA elution solution ([DES] DNase and pyrogen-free water).

Preparation of chromosomal DNA from cultured C. difficile isolates.

Isolates were cultured onto CBA and incubated anaerobically for 48 h. A few colonies were emulsified in TE (Tris-EDTA) buffer (Sigma-Aldrich Co., Ltd., Gillingham, United Kingdom) and heated at 100°C for 10 min. Debris was removed by centrifugation at 13,500 rpm for 2 min, and the supernatant was removed for use in MLST. DNA was stored at −20°C.

C. difficile nucleotide sequence alignment and choice of candidate loci for MLST.

Ten publicly available C. difficile genome sequences, including six of PCR ribotype 027, were aligned using Mauve alignment software (5). The annotated C. difficile 630 genome (36) was included as a reference. This alignment contained several large gaps and was refined using BLAST to position fragments of each genome that were left unaligned by Mauve. Candidate regions for MLST were determined from the refined alignment as follows. Using windows of 500 bp (10 to 60 variable sites), the numbers of variable sites and the numbers of gaps were calculated. MLST loci were chosen such that there was a significant degree of divergence across the 500 bp, and no gaps were present. Fragments were annotated according to their orthologues in C. difficile 630 to ensure that they spanned housekeeping genes. Candidate fragments were tested in silico for suitability for primer design. The standard BLASTn search for “short nearly exact matches” was used, and this search is equivalent to BLASTn with the following parameters: word size, 7; low-complexity filter (DUST) off; expect value, 1,000. The database was the entire GenBank nonredundant database (nr). Details are found at: http://www.ncbi.nlm.nih.gov/blast/producttable.shtml#shortn . PCR was carried out on a subset of C. difficile isolates to verify amplification efficiency of the new MLST fragment. The seven loci and primers chosen for MLST are shown in Table 1 .

C. difficile high-throughput multilocus sequence typing.

MLST was performed as described below by setting up PCR and sequencing reactions in 24-, 48-, or 96-well plates (Fig. 3 shows the procedure for 24 samples, but it can easily be scaled up to 48 samples). Seven PCR amplicons were obtained for each isolate using the primers shown in Table 1. Each 50-μl PCR mixture contained 39.75 μl of molecular biology-grade water (Sigma-Aldrich Co., Ltd.), 5 μl of 10× PCR buffer (Qiagen Ltd., Crawley, United Kingdom), 1 μl of a 10 μM concentration of each forward and reverse primer, 1 μl of 10 mM deoxynucleoside triphosphate (dNTP) mix (Invitrogen Corp., Paisley, United Kingdom), 0.25 μl of HotStart Taq DNA polymerase (Qiagen Ltd.), and 2 μl of C. difficile chromosomal DNA (approximately 10 ng) or extracted total stool DNA. The amplification conditions were 95°C for 15 min, followed by 35 cycles of 94°C for 30 s, 50°C for 40 s, and 72°C for 70 s, with a final extension at 72°C for 5 min and storage at 15°C. The amplification products were purified by precipitation with 20% polyethylene glycol (molecular weight, 8,000) and 2.5 M NaCl, and their nucleotide sequences were determined on each DNA strand using the amplification primers and BigDye Ready Reaction Mix (Applied Biosystems, Warrington, United Kingdom) as follows. Each 10-μl sequencing reaction mixture comprised 2 μl of PCR amplicon, 4 μl of a 1:15 dilution of either forward or reverse PCR primer (0.66 μM), 0.25 μl of BigDye Ready Reaction Mix, 1.875 μl of 5× sequencing buffer (20 ml of stock solution comprised 200 μl of 1 M MgCl2, 8 ml of 1 M Tris-HCl, pH 9, and 11.8 ml of molecular biology-grade water [all from Sigma-Aldrich Co., Ltd.]), and 1.875 μl of molecular biology-grade water. Dilution of the BigDye Ready Reaction Mix using 5× sequencing buffer reduces the cost of high-throughput sequencing without any compromise in sequence quality. The reaction conditions were 30 cycles of denaturation at 96°C for 10 s, annealing at 50°C for 5 s, and extension at 60°C for 2 min. Unincorporated dye terminators were removed by precipitation of the termination products with 2 volumes of ethanol and 0.1 volume of 3 M sodium acetate (pH 5.2), followed by centrifugation, and the resulting pellet was then washed with 70% ethanol. The reaction products were separated and detected using a 3730 XL DNA analyzer (Applied Biosystems). For each sample, the program STARS (Sequence Typing Analysis Retrieval System [http://pubmlst.org/software/assembly/ ]) was used to rapidly collate paired reads, determine sequences, and identify alleles. The data for C. difficile alleles and STs were deposited in a newly developed C. difficile MLST database, which is accessible at http://pubmlst.org/cdifficile . Phylogenetic analysis was performed using the program MEGA, version 4 (Molecular Evolutionary Genetics Analysis [http://www.megasoftware.net/ ]).

Detection of PaLoc genes by PCR.

The oligonucleotide primers used to detect the tcdA (encoding toxin A), tcdB (encoding toxin B), and tcdC (encoding a negative regulator of toxins A and B) sequences found within the pathogenicity locus operon (PaLoc) are summarized in Table 1. The tcdA assay was published by Lemee et al. (21) and amplifies a 369-bp amplicon for toxin A-positive B-positive (A+ B+) strains and a 110-bp amplicon for A-negative (A) B+ strains, which contain a deletion in the tcdA gene. The reaction conditions were 95°C for 15 min, followed by 35 cycles of 94°C for 30 s, 52°C for 30 s, and 72°C for 40 s, with a final extension at 72°C for 5 min and storage at 15°C. The tcdB primers amplify a 688-bp amplicon under the same reaction conditions used for tcdA, except an annealing temperature of 50°C for 40 s and extension of 72°C for 70 s were used. The tcdC primers (33) amplify the 5′ region of the tcdC gene, giving a 475-bp amplicon under the same conditions used for tcdA. The absence of the PaLoc was demonstrated using primers lok1 and lok3 (3), which amplify a 769-bp amplicon in strains without the PaLoc. The reaction conditions were the same as those for tcdB above.

PCR ribotyping.

All PCR ribotyping of reference isolates and cultured isolates described in the present study was performed at the reference laboratory for the C. difficile Ribotyping Network for England and Northern Ireland, Leeds General Infirmary. PCR ribotyping was performed as described previously, with modifications (30). Briefly, bacterial growth was harvested from cultures raised on modified Brazier's CCEY agar with the omission of egg yolk and addition of 5 mg/liter lysozyme (CCEYL) (BioConnections, Wetherby, United Kingdom) for 48 h at 37°C (44). Template DNA was prepared using a QIAxtractor automated nucleic acid extraction system (Qiagen Ltd). Amplification reactions were performed in 50-μl volumes containing 50 pmol of both forward and reverse primers, 25 μl of HotStart Taq Plus PCR Master Mix (Qiagen Ltd.), 19 μl of water, and 5 μl of DNA template. The reaction mixtures were activated by heating to 95°C for 5 min and then subjected to 30 cycles of 92°C for 1 min, 55°C for 1 min, and 72°C for 1.5 min. A final cycle of 95°C for 1 min, 55°C for 45 s, and 72°C for 5 min was added. The resultant amplimer was concentrated to a final volume of approximately 20 μl by heating the opened reaction tubes at 75°C for 30 min. Amplification products were subjected to electrophoresis using 3% Metasieve agarose (Flowgen Bioscience, Nottingham, United Kingdom) at a field strength of 7.5 V/cm for approximately 2.5 h. Agarose gels were imaged using a GeneGenius camera system (Syngene, Cambridge, United Kingdom) after ethidium bromide staining. DNA profiles were analyzed and identified against a library of known PCR ribotypes using BioNumerics, version 4.6, software (Applied Maths, Belgium).

ID.

The index of discrimination (ID) for MLST and PCR ribotyping was calculated according to Hunter and Gaston (11). The ID expresses the average probability that two individuals in the collection will have the same MLST type.

RESULTS

C. difficile MLST.

The MLST scheme was designed to be technically robust, generating high amplicon yields for all C. difficile genotypes, under the same PCR conditions for all seven loci. Loci were chosen that were widely distributed around the chromosome (Fig. 1), had demonstrable discriminatory ability, and lacked insertions or deletions that caused changes in length. The longest possible sequence was used at each locus to maximize resolution, trim sites were located so that primer sequences were removed, and null alleles were avoided (as far as possible, given current information on the species) to ensure 100% typeability.

Validation of the C. difficile MLST scheme. (i) Typeability of all genotypes.

The ability of the MLST primers to amplify and sequence all C. difficile genotypes was assessed using DNA from 152 cultured isolates. These included 50 isolates of 45 different PCR ribotypes (from a collection held at Leeds General Infirmary, the reference laboratory for the C. difficile Ribotyping Network for England and Northern Ireland) representing the known diversity of the species (Fig. 2) and an additional 102 isolates from recent clinical samples (91 from 107 ELISA-positive stools and 11 from 108 ELISA-negative stools) (Tables 2 and 3). The clinical isolates were also PCR ribotyped, and consequently four additional PCR ribotypes were included, giving a total of 49 in the study as a whole (Fig. 2 and Tables 2 and 3). All isolates were successfully typed by MLST, and a total of 40 STs were identified (Fig. 2). These data have been made available at http://pubmlst.org/cdifficile/ .
The MLST primers amplified DNA from both toxigenic and nontoxigenic strains. The 91 isolates from ELISA-positive stools were presumed to be toxigenic, and all were typed by MLST. The 11 isolates from ELISA-negative stools were assessed by PCRs designed to amplify fragments of three genes located within the PaLoc (Table 3). Nine of 11 isolates were toxigenic. The other two were nontoxigenic and were assigned a unique genotype, ST-15. They both yielded a 769-bp amplicon with the lok1/lok3 primer pair (Tables 1 and 3), confirming the absence of the PaLoc. The MLST primers have been further validated using a collection of 37 isolates (including 26 nontoxigenic strains) from healthy infants aged under 2 years and a large collection of ∼1,600 clinical isolates collected at the John Radcliffe Hospital, Oxford, United Kingdom, between September 2006 and April 2009 (molecular epidemiology data are beyond the scope of this paper).

(ii) Discrimination among isolates.

The level of discrimination achieved with MLST was compared to PCR ribotyping. The total number of isolates with both ST and PCR ribotype data available was 152, comprising 50 in the Leeds PCR ribotype collection, 91 from ELISA-positive stools (Table 2), and 11 from ELISA-negative stools (Table 3). Among the 152 isolates, 32/40 STs occurred with a single PCR ribotype, but 8 STs were associated with two or more PCR ribotypes (Fig. 2). Conversely, the PCR ribotype was not always predictive of the ST, with three PCR ribotypes being associated with more than one ST (Fig. 2). The multiple PCR ribotypes associated with single STs had a high degree of similarity in DNA profile. For example, when PCR ribotypes share an ST (n = 8), 6/8 STs were PCR ribotypes that differed by a single band. The other two were PCR ribotypes 023 and 063 (both ST-5) (Fig. 2), which differ by four bands, and PCR ribotype 009 (ST-3), which differs from the other PCR ribotypes associated with ST-3 by four bands.
The set of 102 clinical isolates was used to compare the discriminatory ability of the two methods. The IDs (11) for MLST and PCR ribotyping were 0.90 and 0.92, respectively, a difference that is unlikely to be statistically significant.

(iii) MLST performed directly on total stool DNA.

The robustness of the MLST scheme was validated by typing strains contained in total DNA extracted from stool samples. C. difficile DNA isolated from cultured cells is a straightforward template for PCR amplification. However, when the same target is mixed in human feces, successful amplification requires efficient lysis of C. difficile cells and spores, PCR inhibitor removal (19, 29), and specific amplification from a complex mixture of DNAs (8, 22). To test the approach, MLST data were generated from both cultured isolates and total stool DNA extracts for 95 clinical samples (89/107 ELISA-positive and 6/108 ELISA-negative samples) (Tables 2 and 3). The direct and culture MLST data were generated by two different people, each without prior knowledge of the other's results. Direct MLST data were examined carefully for evidence of primer cross-reaction with other clostridial species in stools and coinfections with multiple C. difficile genotypes. Among 93 of the 95 stools, there was 100% agreement of the direct and culture MLST data. The two exceptions yielded occasional mixed peaks in the sequence chromatograms (2.1%), indicating a coinfection of more than one C. difficile genotype. This compares to 3/23 (13%) stools reported by Wroblewski et al. (45) using PCR detection of binary toxin gene(s) to detect heterogeneous isolate populations, and 2/23 (9%) patients with a first episode of CDI detected by van den Berg et al. (43) using PCR ribotyping. One additional stool DNA extract (Table 2, sample 104+) yielded very low levels of PCR amplicons, insufficient for sequencing at all seven loci.

Genetic variation and phylogenetic relationships among isolates.

The relationships among the 40 STs were examined using the concatenated sequences of the seven MLST loci to construct a neighbor-joining tree (Fig. 2). The total number of variable nucleotide sites was 103/3,501 (2.9%), falling to 59/3,501 (1.7%) if the outlier ST-11 (078) was excluded. The comparable data for amino acid sites were 22/1,167 (1.9%) and 17/1,167 (1.5%). Despite the relatively low level of genetic diversity, the STs clustered into four groups with one outlier (ST-11) (Fig. 2). The majority of sequence types clustered in group 1 with very low internal bootstrap values. Group 2 contained ST-1 (027), the hypervirulent clone; group 3 contained two STs, both of which were associated with PCR ribotype 023; and group 4 contained the toxin A B+ ST-37 (PCR ribotype 017). The outlier ST-11 is associated with PCR ribotype 078, which has been reported as an emergent hypervirulent clone.

DISCUSSION

MLST is a proven technology for understanding the molecular epidemiology and population biology of bacterial species (25). Although it has been applied to a diverse collection of C. difficile isolates (20), MLST has not been widely adopted for this organism, in contrast to the majority of clinically important bacterial species (25). Our aim was to further develop MLST for C. difficile, setting up a more robust method by the following steps: (i) replacing the null allele employed at one of the loci included in the previously published scheme (20) with an allele present in all strains, (ii) improving discrimination by using longer sequences for MLST, and (iii) establishing an Internet-accessible MLST database to allow straightforward accumulation of data over time and to simplify the comparison of data among laboratories. This MLST scheme for C. difficile was also sufficiently robust to allow typing to be performed directly on DNA extracted from stool, without culture. This could potentially be used to generate actionable genotyping data close to real time since the entire process can be completed for a batch of 24 isolates in 3.5 days (Fig. 3), at a consumables cost of £15 per stool (or $24.65 as of 29 October 2009) and the cost of one graduate-level member of staff.
The MLST scheme was sufficiently discriminatory to give typing data which can be interpreted with confidence; according to Hunter and Gaston (11) an ID greater than 0.90 is desirable to meet this requirement. For our 102 clinical isolates, MLST and PCR ribotyping had comparable discriminatory abilities (ID of 0.90 for MLST and of 0.92 for PCR ribotyping). The differences between the methods were generally consistent with a simple genetic explanation; multiple ribotypes for the same ST usually had very similar profiles, and multiple STs for the same ribotype generally had very closely related STs. Capillary gel electrophoresis-based PCR ribotyping is a promising tool to study subtypes within ribotypes, and it may assist the explanation of such observations (13). They may also be consistent with limited recombination that may be characteristic of C. difficile.
We calculated the ID as 0.958 for the previously published MLST scheme (34 STs, 62 PCR ribotypes, and 72 isolates) (20) and as 0.983 for PCR ribotyping for the same collection. However, this is not an entirely robust comparison since all the isolates were specifically chosen for their genetic diversity, and a true ID should reflect the capacity of a typing method to discriminate epidemiologically unrelated isolates within a population.
Pulsed-field gel electrophoresis is another genotyping technique widely used to characterize C. difficile. The IDs for PFGE, the previously published MLST scheme of Lemee et al. (20), and PCR ribotyping were found by Killgore et al. (17) to be 0.843 (PFGE), 0.699 (MLST), and 0.700 (PCR ribotyping) for a collection of 42 isolates from four countries representing epidemic strains and the next most commonly isolated strain types.
Despite the relatively low overall genetic diversity detected within these housekeeping loci, it was still possible to identify four different phylogenetic groups of C. difficile STs (Fig. 2). The majority of STs clustered in group 1, group 2 contained ST-1 (PCR ribotype 027), group 3 contained two STs associated with PCR ribotype 023, and group 4 contained toxin A B+ ST-37 (PCR ribotype 017). A single outlier, ST-11, was associated with PCR ribotype 078. The previously described MLST scheme for C. difficile identified three divergent lineages, one containing the A B+ isolates, which corresponds to our group 4 (20). Stabler et al. (39) used comparative genomics to identify four clades, and these appear to correlate with the four groups we have identified by MLST in the present study (Fig. 2). In that previous study HA1 (human and animal 1) (39) contained mainly human isolates with just a few animals, and this clade probably corresponds to our group 1, which contained the majority of our human isolate STs. HA2 probably corresponds to our group 3 as this contained mainly animal isolates (pig and bovine), with few isolates from humans. These data suggest that genotypes clustered by MLST may correlate with groups derived from whole-genome comparisons using DNA microarrays (39), implying that MLST may be an accurate proxy for whole-genome analysis. The newly emergent ST-11 (PCR ribotype 078) hypervirulent clone was a genetically distinct outlier. This genotype causes infection in humans, pigs, and calves (6, 16) and has been found in cooked and raw meat products (38). Multilocus variable-number tandem-repeat analysis (MLVA) data confirmed a strong degree of genetic relatedness between human and animal isolates belonging to this genotype in The Netherlands (9). ST data presented here suggest that ST-11 (078) has emerged from a single, genetically distinct clade. The other four ST groups may represent different C. difficile clonal complexes, with the level of nucleotide sequence divergence between STs representing each group ranging from 11/3,501 (0.3%) to 60/3,501 (1.7%).
A robust MLST scheme can now be applied to studies of C. difficile epidemiology and population structure. Direct MLST of C. difficile in stool provides a rapid genotyping method which generates data that are easily compared among laboratories using an Internet-accessible database. It will now be possible to test in a clinical setting the utility of MLST for outbreak identification, detection of transmission events among patients, and the identification of emergent hypervirulent clones, thereby assessing the potential benefits of MLST to individual patients and hospital infection control.
FIG. 1.
FIG. 1. The relative positions (to scale) of the seven housekeeping loci chosen for MLST on the C. difficile strain 630 chromosome, together with three genes of the pathogenicity locus (PaLoc) detected by PCRs referred to in this study and the genes encoding the binary toxin cdtA and cdtB.
FIG. 2.
FIG. 2. Neighbor-joining tree constructed using the concatenated sequences (3,501 nucleotides) of the seven loci used in MLST. Bootstraps were generated using 1,000 replicates, and low values were removed for clarity. STs are shown in bold. The PCR ribotype(s) found in association with each ST are indicted. The STs cluster into four groups, designated 1 to 4, with bootstraps greater than 80 and one outlier (ST-11). These correspond to groups defined previously by microarray analysis of whole genomes (39) which were designated HA1 (human and animal 1), hypervirulent (containing the 027 strain), HA2, and the A B+ clade (also containing other toxigenic types). The total number of variable sites was 103/3,501 (2.9%), and if the outlier ST-11 was excluded, the number was 59/3,501 (1.7%). ST-1 (indicated by *1) was associated with PCR ribotype 027 and also with a single PCR ribotype 036 isolate. DNA profiles for PCR ribotypes 036 and 027 are very similar and differ by only a single band.
FIG. 3.
FIG. 3. Flow diagram summarizing the time required to perform the laboratory work and sequence data analysis in high-throughput MLST.
TABLE 1.
TABLE 1. Nucleotide sequences of the oligonucleotide primers used in the studya
LocusPrimer namebPrimer sequence (5′-3′)Amplicon size (bp)Locus size (bp)cReference or source
adkadk1FTTACTTGGACCTCCAGGTGC635501This study
 adk1RTTTCCACTTCCTAAGGCTGC   
atpAatpA1FTGATGATTTAAGTAAACAAGCTG674555This study
 atpA1RAATCATGAGTGAAGTCTTCTCC   
dxrdxr3FGCTACTTTCCATTCTATCTG525411This study
 dxr4RCCAACTCTTTGTGCTATAAA   
glyAglyA1FATAGCTGATGAGGTTGGAGC625516This study
 glyA1RTTCTAGCCTTAGATTCTTCATC   
recArecA2FCAGTAATGAAATTGGGAGAAGC705564This study
 recA2RATTCAGCTTGCTTAAATGGTG   
sodAsodA5FCCAGTTGTCAATGTATTCATTTC585450This study
 sodA6RATAACTTCATTTGCTTTTACACC   
tpitpi2FATGAGAAAACCTATAATTGCAG640504This study
 tpi2RTTGAAGGTTTAACACTTCCACC   
tcdAtcdA-FAGATTCCTATATTTACATGACAATAT369 (A+ B+)dNA21
 tcdA-RGTATCAGGCATAAAGTAATATACTTT110 (A B+)d  
tcdBtcdB1TGATGAAGATACAGCAGAAGC688NAThis study
 tcdB2TGATTCTCCCTCAAAATTCTC   
tcdCtcdC-F(−17)AAAAGGGAGATTGTATTATGTTTTC479NA33
 tcdC-R(+462)CAATAACTTGAATAACCTTACCTTCA   
cdd1lok1(cdd1)AAAATATACTGCACATCTGTATAC769NA3
cdu1lok3 (cdu1)TTTACCAGAAAAAGTAGCTTTAA   
a
Primers were used (i) to perform C. difficile MLST, (ii) to detect the presence of three loci within the pathogenicity locus (tcdA, tcdB, and tcdC), and (iii) to confirm the absence of the pathogenicity locus in nontoxigenic strains (lok1/lok3).
b
In the primer names F indicates forward, and R indicates reverse.
c
NA, not applicable.
d
The size of the amplicon varies with the strain genotype.
TABLE 2.
TABLE 2. MLST of C. difficile from ELISA-positive stool samples determined both directly using total stool DNA and using isolates cultured from the same stools
Sample no.aELISA ODbMLST PCR result (direct)Direct STcCulture resultiCulture STdPCR ribotype
23+++++++1+++1027
17+0.353+1++1027
2+2.251+1+1027
8+++++++1+++1027
37+1.216+1+1027
43+0.663+1+1027
72+2.5+1+++1027
75+2.5+1++1027
85+0.388+1+1027
88+5.584+1+++1027
89+5.392+1++1027
94+7.098+1+++1027
98+2.5+1+++1027
101+1.208+1+++1027
107+2.5+1+++1027
60+++++++1++1027
53+1.8+2+2014
58+0.261+2++2020
27+++++++2+++2020
11+1.614+2+2020
12+0.968+2+++2020
13+7.718+2+++2020
14+3.648+2++2020
31+5.648+2+2014
41+0.843+2+2020
44+0.29+2+++2020
48+1.003+2++2076
63+2.5+2+++2014
66+2.5+2+++2020
67+2.5+2++2014
84+0.563+2+++2020
1+++++++2++2076
106+1.049+2+++2020
52+++++++3++3072
55+++++++3++3001
57+0.285+3+3072
16+++++++3++3001
64+0.434+3+++3072
65+2.5+4++4137
68+1.322+4++4137
69+2.294+5+++5023
76+2.5+5++5023
78+2.5+5++5023
95+2.5+5+++5023
59+4.083+6+6005
51+++++++6+++6005
26+1.155+6+6005
70+2.5+6++6005
92+1.896+6+++6005
93+0.7+6+++6005
3+0.679+7+7026
32+0.499+7+7026
74+2.5+7+++7026
39+0.27+8+++8002
77+2.427+8+++8002
96+2.282+8+++8002
97+2.379+8+++8002
105+2.136+8+++8002
73+0.499+8+++8002
79+1.163+9+9081
7+0.35+9+9081
80+2.5+10+++10015
99+0.796+10+++10015
15+0.202+10+10015
82+2.5+11++11078
4+2.13+11+++11078
61+0.567+11+11078
20+++++++12+12003
103+0.657+12+++12225
108+2.5+13+++13129
22+++++++14+++14014
5+1.658+14+14014
33+1.658+14++14014
42+0.317+14++14014
46+0.385+14+14014
62+2.5+14+++14014
19+2.257+14+14014
36+1.09+14++14014
28+0.447+19++19110
6+++++++22+22023
49+0.41+37++37017
9+++++++42+++42106
24+0.182+42+42174
40+++++++42++42106
56+3.296+54++54012
104+0.387(+)hNDe+++1027
100+0.67+Mixedg+5023
71+2.322+Mixedg+++11078
45+2.127+1  
47+6.186+1  
50+0.278 +2220
86+0.369 +7026
91+3.578 +++8002
90+1.155 +NDfNDf
102+0.885 +NDfNDf
54+0.26   
21+7.792   
25+0.67   
29+3.475   
30+0.875   
18+0.158   
34+1.487   
35+0.435   
38+0.270   
81+0.537   
83+0.475   
87+0.391   
a
The plus sign indicates an ELISA-positive result.
b
OD, optical density. +++++, value above the scale.
c
MLST was performed directly from stool samples.
d
MLST was performed from cultured isolates.
e
ND, not done due to very low PCR amplicon yields or no amplicon.
f
ND, not done because sample was negative on reculturing from frozen isolate stock.
g
Evidence of more than one genotype present in stool due to mixed peaks in the nucleotide sequence chromatogram.
h
Poor amplicon yield.
i
Crude estimate of viable C. difficile load. +, <100 colonies; ++, 100 to 1,000 colonies; +++, >1,000 colonies.
TABLE 3.
TABLE 3. MLST of C. difficile from ELISA-negative stool samples determined both directly using total stool DNA and using isolates cultured from the same stools
Sample no.aMLST PCR result (direct)Direct STdCulture resultgCulture STePCR ribotypePaLoc PCR resultlok1/lok3 PCR
22−+6+6005+
23−+15+15010+f
36−+6+6005+
43−+8+8002+
69−+2+2020+
79−+2+2020+
15− +69072+
55−b +1027+
85−c +1027+
99− ++15010+f
108− +8002+
a
The minus sign indicates an ELISA-negative result.
b
Patient provided a subsequent sample not included in the study (+7 days) that was ELISA positive and ST-1 by culture (sample not typed by direct MLST).
c
Patient provided a subsequent sample not included in the study (+39 days) that was ELISA positive and ST-1 by culture (sample not typed by direct MLST).
d
MLST was performed directly from stool samples.
e
MLST was performed from cultured isolates.
f
PCR yielded a769-bp amplicon, confirming the absence of the PaLoc.
g
Crude estimate of viable C. difficile load. +, <100 colonies; ++, 100 to 1,000 colonies; +++, >1,000 colonies.

Acknowledgments

This study was supported by the NIHR Biomedical Research Centre, Oxford, United Kingdom. Melina Kachrimanidou was supported by the Society of Biopathology of Northern Greece.
We thank the staff of the Clinical Microbiology Laboratory and Infection Control, John Radcliffe Hospital, Oxford, and Infection Control Laboratory staff, Leeds General Infirmary, for their assistance throughout this work.

REFERENCES

1.
Bartlett, J. G. 2002. Clinical practice. Antibiotic-associated diarrhea. N. Engl. J. Med. 31 : 334-339.
2.
Bartlett, J. G., and D. N. Gerding. 2008. Clinical recognition and diagnosis of Clostridium difficile infection. Clin. Infect. Dis. 46 : S12-S18.
3.
Braun, V., T. Hundsberger, P. Leukel, M. Sauerborn, and C. von Eichel-Streiber. 1996. Definition of the single integration site of the pathogenicity locus in Clostridium difficile. Gene 181 : 29-38.
4.
Clabots, C. R., S. Johnson, K. M. Bettin, P. A. Mathie, M. E. Mulligan, D. R. Schaberg, L. R. Peterson, and D. N. Gerding. 1993. Development of a rapid and efficient restriction endonuclease analysis typing system for Clostridium difficile and correlation with other typing systems. J. Clin. Microbiol. 31 : 1870-1875.
5.
Darling, A. C., B. Mau, F. R. Blattner, and N. T. Perna. 2004. Mauve: multiple alignment of conserved genomic sequence with rearrangements. Genome Res. 14 : 1394-1403.
6.
Debast, S. B., L. A. van Leengoed, A. Goorhuis, C. Harmanus, E. J. Kuijper, and A. A. Bergwerff. 2009. Clostridium difficile PCR ribotype 078 toxinotype V found in diarrhoeal pigs identical to isolates from affected humans. Environ. Microbiol. 11 : 505-511.
7.
Denève, C., C. Janoir, I. Poilane, C. Fantinato, and A. Collignon. 2009. New trends in Clostridium difficile virulence and pathogenesis. Int. J. Antimicrob. Agents 33(Suppl. 1): S24-S28.
8.
Fujisawa, T., K. Namba, K. Hirayama, W. K. Lee, and T. Mitsuoka. 1995. New selective media for isolation of clostridia from faecal specimens. J. Appl. Bacteriol. 78 : 481-486.
9.
Goorhuis, A., D. Bakker, J. Corver, S. B. Debast, C. Harmanus, D. W. Notermans, A. A. Bergwerff, F. W. Dekker, and E. J. Kuijper. 2008. Emergence of Clostridium difficile infection due to a new hypervirulent strain, polymerase chain reaction ribotype 078. Clin. Infect. Dis. 47 : 1162-1170.
10.
Health Protection Agency. 2009. Clostridium difficile Ribotyping Network for England and Northern Ireland report 2008/09. Health Protection Agency, London, United Kingdom. http://www.hpa.org.uk/web/HPAwebFile/HPAweb_C/1258560554236 .
11.
Hunter, P. R., and M. A. Gaston. 1988. Numerical index of discriminatory ability of typing systems: an application of Simpson's index of diversity. J. Clin. Microbiol. 26 : 2465-2466.
12.
Ibarz Pavón, A. B., and M. C. Maiden. 2009. Multilocus sequence typing. Methods Mol. Biol. 551 : 129-140.
13.
Indra, A., S. Huhulescu, M. Schneeweis, P. Hasenberger, S. Kernbichler, A. Fiedler, F. Wewalka, G. Allerberger, and E. J. Kuijper. 2008. Characterization of Clostridium difficile isolates using capillary gel electrophoresis-based PCR ribotyping. J. Med. Microbiol. 57 : 1377-1382.
14.
Jolley, K. A., M. S. Chan, and M. C. Maiden. 2004. MlstdbNet—distributed multi-locus sequence typing (MLST) databases. BMC Bioinformatics 5 : 86.
15.
Kazakova, S. V., K. Ware, B. Baughman, O. Bilukha, A. Paradis, S. Sears, A. Thompson, B. Jensen, L. Wiggs, J. Bessette, J. Martin, J. Clukey, K. Gensheimer, G. Killgore, and L. C. McDonald. 2006. A hospital outbreak of diarrhea due to an emerging epidemic strain of Clostridium difficile. Arch. Intern. Med. 166 : 2518-2524.
16.
Keel, K., J. S. Brazier, K. W. Post, S. Weese, and J. G. Songer. 2007. Prevalence of PCR ribotypes among Clostridium difficile isolates from pigs, calves, and other species. J. Clin. Microbiol. 45 : 1963-1964.
17.
Killgore, G., A. Thompson, S. Johnson, J. Brazier, E. Kuijper, J. Pepin, E. H. Frost, P. Savelkoul, B. Nicholson, R. J. van den Berg, H. Kato, S. P. Sambol, W. Zukowski, C. Woods, B. Limbago, D. N. Gerding, and L. C. McDonald. 2008. Comparison of seven techniques for typing international epidemic strains of Clostridium difficile: restriction endonuclease analysis, pulsed-field gel electrophoresis, PCR-ribotyping, multilocus sequence typing, multilocus variable-number tandem-repeat analysis, amplified fragment length polymorphism, and surface layer protein A gene sequence typing. J. Clin. Microbiol. 46 : 431-437.
18.
Kuijper, E. J., F. Barbut, J. S. Brazier, N. Kleinkauf, T. Eckmanns, M. L. Lambert, D. Drudy, F. Fitzpatrick, C. Wiuff, D. J. Brown, J. E. Coia, H. Pituch, P. Reichert, J. Even, J. Mossong, A. F. Widmer, K. E. Olsen, F. Allerberger, D. W. Notermans, M. Delmée, B. Coignard, M. Wilcox, B. Patel, R. Frei, E. Nagy, E. Bouza, M. Marin, T. Akerlund, A. Virolainen-Julkunen, O. Lyytikäinen, S. Kotila, A. Ingebretsen, B. Smyth, P. Rooney, I. R. Poxton, and D. L. Monnet. 2008. Update of Clostridium difficile infection due to PCR ribotype 027 in Europe, 2008. Euro Surveill. 31 : 18942.
19.
Lantz, P. G., M. Matsson, T. Wadström, and P. Rådström. 1997. Removal of PCR inhibitors from human faecal samples through the use of an aqueous two-phase system for sample preparation prior to PCR. J. Microbiol. Methods 28 : 159-167.
20.
Lemee, L., A. Dhalluin, M. Pestel-Caron, J. F. Lemeland, and J. L. Pons. 2004. Multilocus sequence typing analysis of human and animal Clostridium difficile isolates of various toxigenic types. J. Clin. Microbiol. 42 : 2609-2617.
21.
Lemee, L., A. Dhalluin, S. Testelin, M. A. Mattrat, K. Maillard, J. F. Lemeland, and J. L. Pons. 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 : 5710-5714.
22.
Ley, R. E., M. Hamady, C. Lozupone, P. J. Turnbaugh, R. R. Ramey, J. S. Bircher, M. L. Schlegel, T. A. Tucker, M. D. Schrenzel, R. Knight, and J. I. Gordon. 2008. Evolution of mammals and their gut microbes. Science 320 : 1647-1651.
23.
Loo, V. G., L. Poirier, M. A. Miller, M. Oughton, M. D. Libman, S. Michaud, A. M. Bourgault, T. Nguyen, C. Frenette, M. Kelly, A. Vibien, P. Brassard, S. Fenn, K. Dewar, T. J. Hudson, R. Horn, P. René, Y. Monczak, and A. Dascal. 2005. A predominantly clonal multi-institutional outbreak of Clostridium difficile-associated diarrhea with high morbidity and mortality. N. Engl. J. Med. 353 : 2442-2449.
24.
Maiden, M. C., J. A. Bygraves, E. Feil, G. Morelli, J. E. Russell, R. Urwin, Q. Zhang, J. Zhou, K. Zurth, D. A. Caugant, I. M. Feavers, M. Achtman, and B. G. Spratt. 1998. Multilocus sequence typing: a portable approach to the identification of clones within populations of pathogenic microorganisms. Proc. Natl. Acad. Sci. U. S. A. 95 : 3140-3145.
25.
Maiden, M. C. 2006. Multilocus sequence typing of bacteria. Annu. Rev. Microbiol. 60 : 561-588.
26.
McDonald, L. C., G. E. Killgore, A. Thompson, R. C. Owens, Jr., S. V. Kazakova, S. P. Sambol, S. Johnson, and D. N. Gerding. 2005. An epidemic, toxin gene-variant strain of Clostridium difficile. N. Engl. J. Med. 353 : 2433-2441.
27.
McFarland, L. V. 2009. Renewed interest in a difficult disease: Clostridium difficile infections—epidemiology and current treatment strategies. Curr. Opin. Gastroenterol. 25 : 24-35.
28.
McFarland, L. V., M. E. Mulligan, R. Y. Kwok, and W. E. Stamm. 1989. Nosocomial acquisition of Clostridium difficile infection. N. Engl. J. Med. 320 : 204-210.
29.
Monteiro, L., D. Bonnemaison, A. Vekris, K. G. Petry, J. Bonnet, R. Vidal, J. Cabrita, and F. Mégraud. 1997. Complex polysaccharides as PCR inhibitors in feces: Helicobacter pylori model. J. Clin. Microbiol. 35 : 995-998.
30.
O'Neill, G. L., F. T. Ogunsola, J. S. Brazier, and B. I. Duerden. 1996. Modification of a PCR ribotyping method for application as a routine typing scheme for C. difficile. Anaerobe 2 : 205-209.
31.
Pépin, J., L. Valiquette, M. E. Alary, P. Villemure, A. Pelletier, K. Forget, K. Pépin, and D. Chouinard. 2004. Clostridium difficile-associated diarrhea in a region of Quebec from 1991 to 2003: a changing pattern of disease severity. CMAJ 171 : 466-472.
32.
Pépin, J., L. Valiquette, and B. Cossette. 2005. Mortality attributable to nosocomial Clostridium difficile-associated disease during an epidemic caused by a hypervirulent strain in Quebec. CMAJ 173 : 1037-1042.
33.
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 : 1057-1064.
34.
Razavi, B., A. Apisarnthanarak, and L. M. Mundy. 2007. Clostridium difficile: emergence of hypervirulence and fluoroquinolone resistance. Infection 35 : 300-307.
35.
Rolfe, R. D. 1988. Asymptomatic intestinal colonization by Clostridium difficile, p. 201-225. In R. D. Rolfe and S. M. Finegold (ed.), Clostridium difficile: its role in intestinal disease. Academic Press, Inc., New York, NY.
36.
Sebaihia, M., B. W. Wren, P. Mullany, N. F. Fairweather, N. Minton, R. Stabler, N. R. Thomson, A. P. Roberts, A. M. Cerdeño-Tárraga, H. Wang, M. T. Holden, A. Wright, C. Churcher, M. A. Quail, S. Baker, N. Bason, K. Brooks, T. Chillingworth, A. Cronin, P. Davis, L. Dowd, A. Fraser, T. Feltwell, Z. Hance, S. Holroyd, K. Jagels, S. Moule, K. Mungall, C. Price, E. Rabbinowitsch, S. Sharp, M. Simmonds, K. Stevens, L. Unwin, S. Whithead, B. Dupuy, G. Dougan, B. Barrell, and J. Parkhill. 2006. The multidrug-resistant human pathogen Clostridium difficile has a highly mobile, mosaic genome. Nat. Genet. 38 : 779-786.
37.
Song, X., J. G. Bartlett, K. Speck, A. Naegeli, K. Carroll, and T. M. Perl. 2008. Rising economic impact of Clostridium difficile-associated disease in adult hospitalized patient population. Infect. Control Hosp. Epidemiol. 29 : 823-828.
38.
Songer, J. G., H. T. Trinh, G. E. Killgore, A. D. Thompson, L. C. McDonald, and B. M. Limbago. 2009. Clostridium difficile in retail meat products, USA, 2007. Emerg. Infect. Dis. 15 : 819-821.
39.
Stabler, R. A., D. N. Gerding, J. G. Songer, D. Drudy, J. S. Brazier, H. T. Trinh, A. A. Witney, J. Hinds, and B. W. Wren. 2006. Comparative phylogenomics of Clostridium difficile reveals clade specificity and microevolution of hypervirulent strains. J. Bacteriol. 188 : 7297-7305.
40.
Starr, J. 2005. Clostridium difficile associated diarrhoea: diagnosis and treatment. BMJ 331 : 498-501.
41.
Stubbs, S. L., J. S. Brazier, G. L. O'Neill, and B. I. Duerden. 1999. PCR targeted to the 16S-23S rRNA gene intergenic spacer region of Clostridium difficile and construction of a library consisting of 116 different PCR ribotypes. J. Clin. Microbiol. 37 : 461-463.
42.
Urwin, R., and M. C. Maiden. 2003. Multi-locus sequence typing: a tool for global epidemiology. Trends Microbiol. 11 : 479-487.
43.
van den Berg, R. J., H. A. Ameen, T. Furusawa, E. C. Claas, E. R. van der Vorm, and E. J. Kuijper. 2005. Coexistence of multiple PCR-ribotype strains of Clostridium difficile in faecal samples limits epidemiological studies. J. Med. Microbiol. 54 : 173-179.
44.
Wilcox, M. H., W. N. Fawley, and P. Parnell. 2000. Value of lysozyme agar incorporation and alkaline thioglycollate exposure for the environmental recovery of Clostridium difficile. J. Hosp. Infect. 44 : 65-69.
45.
Wroblewski, D., G. E. Hannett, D. J. Bopp, G. K. Dumyati, T. A. Halse, N. B. Dumas, and K. A. Musser. 2009. Rapid molecular characterization of Clostridium difficile, and assessment of the population of C. difficile in stool specimens. J. Clin. Microbiol. 47 : 2142-2148.
46.
Zilberberg, M. D. 2009. Clostridium difficile-related hospitalizations among US adults, 2006. Emerg. Infect. Dis. 15 : 122-124.
47.
Zilberberg, M. D., A. F. Shorr, and M. H. Kollef. 2008. Increase in adult Clostridium difficile-related hospitalizations and case-fatality rate, United States, 2000-2005. Emerg. Infect. Dis. 14 : 929-931.

Information & Contributors

Information

Published In

cover image Journal of Clinical Microbiology
Journal of Clinical Microbiology
Volume 48Number 3March 2010
Pages: 770 - 778
PubMed: 20042623

History

Received: 11 September 2009
Revision received: 24 October 2009
Accepted: 19 December 2009
Published online: 1 March 2010

Permissions

Request permissions for this article.

Contributors

Authors

David Griffiths
Nuffield Department of Clinical Medicine, Oxford University, John Radcliffe Hospital, Oxford OX3 9DU, United Kindgom
National Institute for Health Research, Oxford Biomedical Research Centre Programme, John Radcliffe Hospital, Oxford OX3 9DU, United Kingdom
Warren Fawley
Department of Microbiology, The General Infirmary, Old Medical School, Leeds LS1 3EX, United Kingdom
Melina Kachrimanidou
Nuffield Department of Clinical Medicine, Oxford University, John Radcliffe Hospital, Oxford OX3 9DU, United Kindgom
National Institute for Health Research, Oxford Biomedical Research Centre Programme, John Radcliffe Hospital, Oxford OX3 9DU, United Kingdom
Rory Bowden
Department of Statistics, University of Oxford, 1 South Parks Road, Oxford OX1 3TG, United Kingdom
Derrick W. Crook
Nuffield Department of Clinical Medicine, Oxford University, John Radcliffe Hospital, Oxford OX3 9DU, United Kindgom
National Institute for Health Research, Oxford Biomedical Research Centre Programme, John Radcliffe Hospital, Oxford OX3 9DU, United Kingdom
Rowena Fung
Nuffield Department of Clinical Medicine, Oxford University, John Radcliffe Hospital, Oxford OX3 9DU, United Kindgom
National Institute for Health Research, Oxford Biomedical Research Centre Programme, John Radcliffe Hospital, Oxford OX3 9DU, United Kingdom
Tanya Golubchik
Department of Statistics, University of Oxford, 1 South Parks Road, Oxford OX1 3TG, United Kingdom
Rosalind M. Harding
Department of Zoology, Oxford University, South Parks Road, Oxford OX1 3PS, United Kingdom
Katie J. M. Jeffery
Department of Microbiology, Oxford Radcliffe NHS Trust, John Radcliffe Hospital, Oxford OX3 9DU, United Kingdom
Keith A. Jolley
Department of Zoology, Oxford University, South Parks Road, Oxford OX1 3PS, United Kingdom
Richard Kirton
Department of Microbiology, Oxford Radcliffe NHS Trust, John Radcliffe Hospital, Oxford OX3 9DU, United Kingdom
Tim E. Peto
Nuffield Department of Clinical Medicine, Oxford University, John Radcliffe Hospital, Oxford OX3 9DU, United Kindgom
National Institute for Health Research, Oxford Biomedical Research Centre Programme, John Radcliffe Hospital, Oxford OX3 9DU, United Kingdom
Gareth Rees
Department of Microbiology, Oxford Radcliffe NHS Trust, John Radcliffe Hospital, Oxford OX3 9DU, United Kingdom
Nicole Stoesser
Nuffield Department of Clinical Medicine, Oxford University, John Radcliffe Hospital, Oxford OX3 9DU, United Kindgom
National Institute for Health Research, Oxford Biomedical Research Centre Programme, John Radcliffe Hospital, Oxford OX3 9DU, United Kingdom
Alison Vaughan
Nuffield Department of Clinical Medicine, Oxford University, John Radcliffe Hospital, Oxford OX3 9DU, United Kindgom
National Institute for Health Research, Oxford Biomedical Research Centre Programme, John Radcliffe Hospital, Oxford OX3 9DU, United Kingdom
A. Sarah Walker
Nuffield Department of Clinical Medicine, Oxford University, John Radcliffe Hospital, Oxford OX3 9DU, United Kindgom
National Institute for Health Research, Oxford Biomedical Research Centre Programme, John Radcliffe Hospital, Oxford OX3 9DU, United Kingdom
Bernadette C. Young
Department of Microbiology, Oxford Radcliffe NHS Trust, John Radcliffe Hospital, Oxford OX3 9DU, United Kingdom
Mark Wilcox
Department of Microbiology, Institute of Molecular and Cellular Biology, University of Leeds, Leeds LS2 9JT, United Kingdom
Department of Microbiology, The General Infirmary, Old Medical School, Leeds LS1 3EX, United Kingdom
Kate E. Dingle [email protected]
Nuffield Department of Clinical Laboratory Sciences, Oxford University, John Radcliffe Hospital Oxford OX3 9DU, United Kingdom
National Institute for Health Research, Oxford Biomedical Research Centre Programme, John Radcliffe Hospital, Oxford OX3 9DU, United Kingdom

Metrics & Citations

Metrics

Note:

  • For recently published articles, the TOTAL download count will appear as zero until a new month starts.
  • There is a 3- to 4-day delay in article usage, so article usage will not appear immediately after publication.
  • Citation counts come from the Crossref Cited by service.

Citations

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. For an editable text file, please select Medlars format which will download as a .txt file. Simply select your manager software from the list below and click Download.

View Options

Figures and Media

Figures

Media

Tables

Share

Share

Share the article link

Share with email

Email a colleague

Share on social media

American Society for Microbiology ("ASM") is committed to maintaining your confidence and trust with respect to the information we collect from you on websites owned and operated by ASM ("ASM Web Sites") and other sources. This Privacy Policy sets forth the information we collect about you, how we use this information and the choices you have about how we use such information.
FIND OUT MORE about the privacy policy