emi4374-sup-0001-si

advertisement
1
Supplementary online material
2
3
TEXT
4
Experimental procedures
5
6
Sampling procedures. In October 2006, and in June 2008, we conducted two sampling
7
campaigns in Trois-Sauts, an isolated French Guiana village. Characteristics of the village,
8
the humans and the human sampling are described elsewhere (Ruimy et al., 2010; Woerther et
9
al., 2010). Briefly, the village is located in the Amazonian pristine forest, in the south of
10
French Guiana, nearby the Oyapock river source and in a protected area, where access is
11
administratively controlled (Fig. S1). The village is divided in 4 hamlets, Zidock, Roger, Pina
12
and Yawapa from where human samples came from (Fig. S1). The villagers are native
13
Amerindian Wayampis. Their way of life is comparable to traditional societies as they share
14
large huts with no access to hygienic facilities, and eat local food (traditional agriculture,
15
fishing and hunting). However they receive education and medical care from metropolitan
16
French teachers and nurses living in the village. Human associated animals were sampled in
17
Zidock, Roger and Pina (Fig. S1), they mostly belong to dog and chicken populations, living
18
free in the hamlets, hence not as closed to villagers as could be pets from industrialised
19
country inhabitants. Wild animals were sampled by trapping along two transects (one
20
connecting Zidock and Roger and the other issued from Zidock and ending 20 km away in
21
North-West direction in the Amazonian forest) and then released alive in the forest or killed
22
and kept as representative specimens of rare species. Wild animals were also sampled thanks
23
to hunting by the villagers for food purpose. In this case, they were brought immediately to
24
the village and sampled as fast as possible. Main characteristics of individuals including
25
Linnean denomination, order, class, body mass, diet and sampling localisation [obtained by
1
1
the Global Positionning System (GPS)] were collected (Table S1). The study was approved
2
by ad hoc Guadeloupe Ethics Committee (Comité de protection des personnes de
3
Guadeloupe, France; no 06-05).
4
We sampled 393 individuals comprising 162 adult Wayampi Amerindians, 33 human
5
associated animals living in the village and 198 wild animals. Among the 162 human
6
samples, 94 (58 %), 37 (23%), 15 (9%) and 16 (10%) were sampled in the hamlets Zidock,
7
Roger, Pina and Yawapa, respectively (Fig. S1). Among the 33 human associated animal
8
samples, 22 (67%) came from Zidock, 2 (6%) from Roger and 9 (27%) from Pina (Fig. S1).
9
Among the 198 wild animal sampled during the two collects, 122 in 2006 and 76 in 2008, 117
10
(59%) were collected in or close to the hamlets Zidock, Roger, Pina and Yawapa [69 (57%) in
11
2006 and 48 (63%) in 2008] and 81 (41%) were collected outside the hamlets along two
12
transects or captured by the villagers around the village [53 (43%) in 2006 and 28 (37%) in
13
2008] (Fig. S1).
14
15
Strain isolation. Fresh faecal samples and rectal swabs (from humans and animals,
16
respectively) were inoculated extemporaneously onto Drigalski agar slants in screw-cup
17
tubes, stored at room temperature and sent to metropolitan France two weeks later. There, the
18
whole culture from each tube was suspended in 1.5 mL of brain-heart infusion broth with
19
10% glycerol and stored at −80°C. 100µL aliquots of each stored broth were cultured on
20
chromogenic plates (Uriselect®; BioRad). Pink colonies were tested for their indole
21
production and identified as E. coli if positive, according to the manufacturer’s
22
recommendations. The first strain identified as E. coli was chosen to be representative of each
23
sample and considered as randomly selected.
24
2
1
Antibiotic resistance pattern. The antimicrobial susceptibility of the strains to 32
2
antibiotics (see list in Table S3) was determined using the disk-diffusion method, as described
3
elsewhere (http://www.sfm.asso.fr). For each strain, phenotype of resistance was classified
4
into 2 categories: S (sensitivity to all antibiotics tested) or R (resistance to at least one
5
antibiotic). Antibiotic resistance scores were also determined as the relative proportion of the
6
number of identified resistances on the total number of tested resistances *100.
7
8
The presence of penicillinase TEM and CMY-type ß-lactamases were detected by
PCR as in (Branger et al., 2005; Courpon-Claudinon et al., 2011).
9
10
Strain genotyping. Phylogenetic groups were firstly assigned to one of the seven groups A0,
11
A1, B1, B22, B23, D1 and D2 using the triplex PCR method (Clermont et al., 2000; Escobar-
12
Paramo et al., 2004). MLST was then performed using the Pasteur Institute scheme
13
(http://www.pasteur.fr/recherche/genopole/PF8/mlst/EColi.html) (Jaureguy et al., 2008) on 96
14
B22, B23, D1 and D2 strains allowing the assignation of the strains to the B2, D, E and F
15
phylogroups (Tenaillon et al., 2010). C group strains (Moissenet et al., 2010) were determined
16
among A1 strains using an allele-specific C group PCR on the trpA gene developed in this
17
study. The PCR steps were as follows: denaturation for 4 min at 94°C, 30 cycles of 5 s at
18
94°C and 10 s at 59°C, and a final extension step of 5 min at 72°C using the primers
19
trpAgpC1
20
TCTGCGCCGGTCACGCCC-3’) (product size: 219 bp). An allele-specific E group PCR was
21
also developed to verify that no strain typed as A0, A1, and B1 by the triplex PCR method was
22
belonging to this group. The chosen target was the arpA gene using the ArpAgpE.f 5’-
23
GATTCCATCTTGTCAAAATATGCC-3’
24
GAAAAGAAAAAGAATTCCCAAGAG-3’ primers amplifying a 301 bp fragment. The
25
PCR conditions were as above except that the annealing temperature was of 57°C. Among
(5’-AGTTTTATGCCCAGTGCGAG-3’)
and
and
trpAgpC2
ArpAgpE.r
(5’-
5’-
3
1
the 176 A0, A1, and B1 strains, no one belongs to the E group. At the opposite, all strains
2
classified as E by the MLST were positive with our E group PCR assay. A maximum-
3
likelihood phylogenetic tree was reconstructed with the PHYML program (Guindon et al.,
4
2005) using the concatenated MLST sequences from the 96 E. coli strains included in this
5
study, 35 strains of the ECOR collection (including 15, 6, 3, 6, 2, 2, and 1 strains belonging to
6
the B2, D, E/UG, F, A, B1 and C phylogenetic groups, respectively) (Ochman and Selander,
7
1984; Lescat et al., 2009), 13 representative B2 phylogenetic subgroup strains previously
8
analysed by Le Gall et al. (Le Gall et al., 2007), 5 representative D and F group strains
9
(Touchon et al., 2011) and the reference strains ED1a, E2348/69, 536, TN03, 042 and
10
EDL933. The tree was rooted on E. fergusonii ATCC 35469T. Detection of clade strains was
11
performed by PCR as described by Clermont et al. (Clermont et al., 2011a).
12
Extraintestinal virulence genes (hly, cnf, aer, papC, iroN, traT, fyuA, sfa, and kpsE)
13
(Bingen-Bidois et al., 2002; Johnson et al., 2006) and intraintestinal virulence genes (afaD,
14
stx1, stx2, ipaH, eae, bfpA, ST and LT coding genes, aaiC) (Clermont et al., 2011b) were
15
detected by PCR. Bacteriocins, including colicins and microcins, were detected as follow.
16
Colicins have firstly been detected by a slightly modified phenotypic method (Schamberger
17
and Diez-Gonzalez, 2005). At first a suspension in phosphate buffered saline of E. coli K-12
18
(at 0.5 Mc Farland) as a sensitive strain has been plated on a Luria Bertani (LB) agar medium
19
containing mitomycin (concentration at 0.25 mg/L). Then, 10 μl of an over night (O/N)
20
culture in LB medium of each strain were spotted on the mitomycin-LB agar plate. After an
21
O/N culture at 37°C, the presence of colicin (or phage) has been detected for the strains
22
surrounded by a halo traducing an inhibition of the culture of the E. coli K-12 strain. Strains
23
positive for the phenotypic test have then been tested by PCR for the presence of the most
24
frequent colicin genes (colIa/Ib, colE1, colB) (Gordon and O'Brien, 2006). The detection of
25
the main microcins genes (micH47 and micV), which production is related to the iron
4
1
concentration in the medium and known to be associated to the iroN gene (Waters and Crosa,
2
1991), have been detected among the iroN positive strains (Gordon and O'Brien, 2006). For
3
each strain, genotypes of extraintestinal virulence, intraintestinal virulence and bacteriocin
4
were separately classified in 2 categories: absence or presence of at least one character.
5
Relevant scores for each type of characters were then determined by the relative proportion of
6
the number of identified extraintestinal, intraintestinal virulence factors and bacteriocins on
7
the total number of tested characters of each type *100.
8
9
Septicemia mouse model. A mouse model of systemic infection was used to assess the
10
intrinsic virulence of 8 B2 strains chosen to be representative of the B2 phylogenetic
11
subgroups, according to the tree reconstructed from the MLST data obtained in this study
12
(Picard et al., 1999). Ten outbreed female OF1 mice (6-week-old, 14-16 g) were challenged
13
subcutaneously in the neck with a standardized bacterial inoculum for each strain (0.2 ml of a
14
Ringer solution containing 109 cfu/ml of log-phase bacteria). Mortality was assessed over 7
15
days post-challenge. Each experimental series included a positive control (urosepsis strain
16
CFT073) and a negative control (commensal derived strain K-12 MG1655). In this model,
17
lethality is a rather clear-cut parameter and strains are usually classified either as non-killer
18
(strains killing none or one mouse out of 10) or killer (strains killing 9 or 10 mice out of 10)
19
(Johnson et al., 2006). Strains that did not fall in these two categories were considered as
20
being intermediate killer. Animal experimentations were done according to the authorization
21
n° 6665 given by the Ministère de l’Agriculture, France.
22
23
Factorial analysis of correspondence (FAC). A FAC was used to describe associations
24
among the data (Greenacre, 1992). FAC was conducted with SPAD.N software (Cisia, Saint
25
Mandé, France) from a two-way table. This table had 272 rows, one for each studied E. coli
5
1
strains and 6 columns corresponding to the 6 variables: the origins of the strains [human
2
strains (HS), human associated animal strains (HAAS) and wild animal strains (WAS)], the
3
phylogenetic group (A, B1, B2, C, D, E and F) according to the phylogrouping method and
4
the MLST data, the presence of extra-intestinal virulence determinants, of intra-intestinal
5
virulence determinants and of bacteriocins, the presence of resistance to antibiotics. For each
6
column, the character of each strain was coded as a binary code: present =1, absent = 0.
7
8
Comparison of nucleotide diversity of the MLST data. Nucleotide diversity per site (Pi) of
9
the strains obtained from Trois Sauts village and belonging to the B2, D, E and F
10
phylogenetic groups (96 strains) was compared to the nucleotide diversity per site of strains
11
belonging to the same phylogroups but obtained from (i) the ECOR collection representative
12
of the E. coli species genetic diversity (Ochman and Selander, 1984) (31 strains), (ii) the
13
Broad
14
(http://www.broadinstitute.org/annotation/genome/escherichia_antibiotic_resistance/MultiHo
15
me.html) from which strains have been chosen for their origin from various species and
16
completely sequenced (40 strains) and (iii) a personal collection of 137 strains (ROAR
17
collection) composed of 38, 36 and 63 strains from human, domestic and wild animals,
18
respectively, sampled in metropolitan France in the 2000’s (Skurnik, Clermont, Brisse and
19
Denamur, personal data). Nucleotide diversities per site of the 4 collections were estimated
20
using DnaSP (Rozas, 2009) and then compared with a t-test.
collection
21
6
1
2
References
3
Bingen-Bidois, M., Clermont, O., Bonacorsi, S., Terki, M., Brahimi, N., Loukil, C. et al.
4
(2002) Phylogenetic analysis and prevalence of urosepsis strains of Escherichia coli bearing
5
pathogenicity island-like domains. Infect Immun 70: 3216-3226.
6
Branger, C., Zamfir, O., Geoffroy, S., Laurans, G., Arlet, G., Thien, H.V. et al. (2005)
7
Genetic background of Escherichia coli and extended-spectrum beta-lactamase type. Emerg
8
Infect Dis 11: 54-61.
9
Clermont, O., Bonacorsi, S., and Bingen, E. (2000) Rapid and simple determination of the
10
Escherichia coli phylogenetic group. Appl Environ Microbiol 66: 4555-4558.
11
Clermont, O., Gordon, D.M., Brisse, S., Walk, S.T., and Denamur, E. (2011a)
12
Characterization of the cryptic Escherichia lineages: rapid identification and prevalence.
13
Environ Microbiol 13: 2468-2477.
14
Clermont, O., Olier, M., Hoede, C., Diancourt, L., Brisse, S., Keroudean, M. et al. (2011b)
15
Animal and human pathogenic Escherichia coli strains share common genetic backgrounds.
16
Infect Genet Evol 11: 654-662.
17
Courpon-Claudinon, A., Lefort, A., Panhard, X., Clermont, O., Dornic, Q., Fantin, B. et al.
18
(2011) Bacteraemia caused by third-generation cephalosporin-resistant Escherichia coli in
19
France: prevalence, molecular epidemiology and clinical features. Clin Microbiol Infect.
20
Escobar-Paramo, P., Grenet, K., Le Menac'h, A., Rode, L., Salgado, E., Amorin, C. et al.
21
(2004) Large-scale population structure of human commensal Escherichia coli isolates. Appl
22
Environ Microbiol 70: 5698-5700.
23
Gordon, D.M., and O'Brien, C.L. (2006) Bacteriocin diversity and the frequency of multiple
24
bacteriocin production in Escherichia coli. Microbiology 152: 3239-3244.
7
1
Greenacre, M. (1992) Correspondence analysis in medical research. Stat Methods Med Res 1:
2
97-117.
3
Guindon, S., Lethiec, F., Duroux, P., and Gascuel, O. (2005) PHYML Online--a web server
4
for fast maximum likelihood-based phylogenetic inference. Nucleic Acids Res 33: W557-559.
5
Jaureguy, F., Landreau, L., Passet, V., Diancourt, L., Frapy, E., Guigon, G. et al. (2008)
6
Phylogenetic and genomic diversity of human bacteremic Escherichia coli strains. BMC
7
Genomics 9: 560.
8
Johnson, J.R., Clermont, O., Menard, M., Kuskowski, M.A., Picard, B., and Denamur, E.
9
(2006) Experimental mouse lethality of Escherichia coli isolates, in relation to accessory
10
traits, phylogenetic group, and ecological source. J Infect Dis 194: 1141-1150.
11
Le Gall, T., Clermont, O., Gouriou, S., Picard, B., Nassif, X., Denamur, E., and Tenaillon, O.
12
(2007) Extraintestinal virulence is a coincidental by-product of commensalism in B2
13
phylogenetic group Escherichia coli strains. Mol Biol Evol 24: 2373-2384.
14
Lescat, M., Hoede, C., Clermont, O., Garry, L., Darlu, P., Tuffery, P. et al. (2009) aes, the
15
gene encoding the esterase B in Escherichia coli, is a powerful phylogenetic marker of the
16
species. BMC Microbiology 9: 723.
17
Moissenet, D., Salauze, B., Clermont, O., Bingen, E., Arlet, G., Denamur, E. et al. (2010)
18
Meningitis caused by Escherichia coli producing TEM-52 extended-spectrum beta-lactamase
19
within an extensive outbreak in a neonatal ward: epidemiological investigation and
20
characterization of the strain. J Clin Microbiol 48: 2459-2463.
21
Ochman, H., and Selander, R.K. (1984) Standard reference strains of Escherichia coli from
22
natural populations. J Bacteriol 157: 690-693.
23
Picard, B., Garcia, J.S., Gouriou, S., Duriez, P., Brahimi, N., Bingen, E. et al. (1999) The link
24
between phylogeny and virulence in Escherichia coli extraintestinal infection. Infect Immun
25
67: 546-553.
8
1
Rozas, J. (2009) DNA sequence polymorphism analysis using DnaSP. Methods Mol Biol 537:
2
337-350.
3
Ruimy, R., Angebault, C., Djossou, F., Dupont, C., Epelboin, L., Jarraud, S. et al. (2010) Are
4
host genetics the predominant determinant of persistent nasal Staphylococcus aureus carriage
5
in humans? J Infect Dis 202: 924-934.
6
Schamberger, G.P., and Diez-Gonzalez, F. (2005) Assessment of resistance to colicinogenic
7
Escherichia coli by E. coli O157:H7 strains. J Appl Microbiol 98: 245-252.
8
Tenaillon, O., Skurnik, D., Picard, B., and Denamur, E. (2010) The population genetics of
9
commensal Escherichia coli. Nat Rev Microbiol 8: 207-217.
10
Touchon, M., Charpentier, S., Clermont, O., Rocha, E.P., Denamur, E., and Branger, C.
11
(2011) CRISPR distribution within the Escherichia coli species is not suggestive of
12
immunity-associated diversifying selection. J Bacteriol 193: 2460-2467.
13
Waters, V.L., and Crosa, J.H. (1991) Colicin V virulence plasmids. Microbiol Rev 55: 437-
14
450.
15
Woerther, P.L., Angebault, C., Lescat, M., Ruppe, E., Skurnik, D., Mniai, A.E. et al. (2010)
16
Emergence and dissemination of extended-spectrum beta-lactamase-producing Escherichia
17
coli in the community: lessons from the study of a remote and controlled population. J Infect
18
Dis 202: 515-523.
19
20
9
1
FIGURE LEGENDS
2
3
Figure S1: Location of study site and sample collection points in Trois-Sauts, French Guiana.
4
The study was located in a protected area where human access is restricted. Locations of
5
human, human associated animal and wild animal samples are indicated in white, grey and
6
black circles, according to the Global Positioning System (GPS) collected for each sample.
7
Circles are proportional to the numbers of sampled individuals.
8
9
TABLE TITLES
10
11
Table S1. Main characteristics of individuals sampled during the 1st collect in October 2006
12
[humans (H), human associated animals (HAA) and wild animals (S1-WA)] and the 2nd
13
collect in June 2008 [wild animals (S2-WA)] as well as the E. coli presence in the faeces.
14
Table S2. Presence of extraintestinal virulence factors, intraintestinal virulence factors,
15
bacteriocins and corresponding scores of strains of E. coli issued from humans, human
16
associated animals and wild animals from both collects (2006-2008).
17
Table S3. Presence of resistance to antibiotics and resistance score of E. coli issued from
18
humans, human associated animals and wild animals from both collects (2006-2008).
19
Table S4. Character mapping of the 96 E. coli strains belonging to the B2, D, E and F
20
phylogroups and ordered as in the Fig. 1.
21
22
23
24
25
10
Download