Skip to main content
BioMed Research International logoLink to BioMed Research International
. 2014 Apr 15;2014:465054. doi: 10.1155/2014/465054

Molecular Screening of Virulence Genes in Extraintestinal Pathogenic Escherichia coli Isolated from Human Blood Culture in Brazil

Vanessa L Koga 1, Geizecler Tomazetto 1, Paula S Cyoia 1, Meiriele S Neves 1, Marilda C Vidotto 1, Gerson Nakazato 1, Renata K T Kobayashi 1,*
PMCID: PMC4009324  PMID: 24822211

Abstract

Extraintestinal pathogenic Escherichia coli (ExPEC) is one of the main etiological agents of bloodstream infections caused by Gram-negative bacilli. In the present study, 20 E. coli isolates from human hemocultures were characterized to identify genetic features associated with virulence (pathogenicity islands markers, phylogenetic group, virulence genes, plasmid profiles, and conjugative plasmids) and these results were compared with commensal isolates. The most prevalent pathogenicity island, in strains from hemoculture, were PAI IV536, described by many researchers as a stable island in enterobacteria. Among virulence genes, iutA gene was found more frequently and this gene enconding the aerobactin siderophore receptor. According to the phylogenetic classification, group B2 was the most commonly found. Additionally, through plasmid analysis, 14 isolates showed plasmids and 3 of these were shown to be conjugative. Although in stool samples of healthy people the presence of commensal strains is common, human intestinal tract may serve as a reservoir for ExPEC.

1. Introduction

Escherichia coli is one of the most common microorganisms of the human intestinal microbiota. However, a small percentage of E. coli is capable of causing extraintestinal infections (extraintestinal pathogenic Escherichia coli—ExPEC), and these ExPECs are considered some of the main etiological agents of bacteremia caused by Gram-negative bacilli [1, 2]. According to the phylogenetic classification, ExPECs typically belong to group B2 and less commonly to group D, whereas commensal intestinal strains belong to group A or B1 [3]. ExPEC pathogenicity is due to the presence of genes, located on plasmids or chromosomes that encode virulence factors. When present on the chromosome, these genes are typically found in specific regions called pathogenicity islands (PAI). Given that the severity of bacterial infections is often due to the genetic features of the pathogenic agent and that few studies have investigated genetic aspects of ExPEC isolates from bacteremia in Brazil [4, 5], the aim of the present study was to characterize 20 E. coli isolates from human hemocultures for genetic features associated with ExPEC. The investigation was based on the screening for PAI associated sequences, determination of phylogenetic group, genotypic identification of the major virulence factors of ExPEC, and plasmid analyses, and these results compared with commensal strains. This information may help us better understand the pathogenesis of these bacteria.

2. Materials and Methods

2.1. Bacterial Isolates

To perform the study, 20 E. coli strains isolated from human hemocultures were kindly provided by Professor Marilda C. Vidotto (Brazil) [6] and 51 E. coli strains were obtained from stools of 19 healthy Brazilians. The strains were stored in brain heart infusion with 20% glycerol at −20°C. The polymerase chain reaction (PCR) was used for the genetic characterization of PAI associated sequences, presence of virulence genes, and for phylogenetic classification. PCR amplicons were visualized on 1.0% agarose gels stained with GelRed (Biotium). After gel electrophoresis the images were captured using Image Capture Systems (LPixImageHE).

2.2. Detection of PAI Markers

The presence of sequences associated with seven different PAIs, previously characterized in uropathogenic E. coli (UPEC), was determined (PAI I536, II536, IV536, ICFT073, IICFT073, IJ96 and IIJ96) [7] (Table 1).

Table 1.

Primers for detection of PAIs markers, phylogenetic analysis, virulence genes, and their respective virulence factors.

Genes Sequence (5′ to 3′) Size of product (bp) Virulence factors Reference
PAI I 536 TAA TGC CGG AGA TTC ATT GTC
AGG ATT TGT CTC AGG GCT TT
1.800 α-Haemolysin, CS12 fimbriae, and F17-like fimbrial adhesin Sabaté et al., 2006 [7]
PAI II 536 CAT GTC CAA AGC TCG AGC C
CTA CGT CAG GCT GGC TTT G
1.000 α-Haemolysin and P-related fimbriae Sabaté et al., 2006 [7]
PAI IV 536 AAG GAT TCG CTG TTA CCG GAC
TCG TCG GGC AGC GTT TCT TCT
300 Yersiniabactin siderophore system Sabaté et al., 2006 [7]
PAI I CFT073 GGA CAT CCT GTT ACA GCG CGC A
TCG CCA CCA ATC ACA GC GAA C
930 α-Haemolysin, P-fimbriae, and aerobactin Sabaté et al., 2006 [7]
PAI II CFT073 ATG GAT GTT GTA TCG CGC
ACG AGC ATG TGG ATC TGC
400 P-fimbriae and iron-regulated genes Sabaté et al., 2006 [7]
PAI I J96 TCG TGC TCA GGT CCG GAA TTT
TGG CAT CCC ACA TTA TCG
400 α-Haemolysin and P-fimbriae Sabaté et al., 2006 [7]
PAI II J96 GGA TCC ATG AAA ACA TGG TTA ATG GG
GAT ATT TTT GTT GCC ATT GGT TAC C
2.300 α-Haemolysin, Prs-fimbriae, and cytotoxic necrotizing factor 1 Sabaté et al., 2006 [7]
chuA GAC GAA CCA ACG GTC AGG AT
TGC CGC CAG TAC CAA AGA CA
279 Hemetransport in enterohemorrhagic O157:H7 E. coli Clermont et al., 2000 [3]
yjaA TGA AGT GTC AGG AGA CGC TG
ATG GAG AAT GCG TTC CTC AAC
211 Protein of function unknown Clermont et al., 2000 [3]
 TSPE4.C2 GAG TAA TGT CGG GGC ATT CA
CGC GCC AAC AAA GTA TTA CG
152 Putative DNA fragment (TSPE4.C2) in E. coli Clermont et al., 2000 [3]
kpsII GCG CAT TTG CTG ATA CTG TTG
CAT CCA GAC GAT AAG CAT GAG CA
272 Capsule synthesis K1 e K5 Johnson and Stell, 2000 [8]
 K1 TAG CAA ACG TTC TAT TGG TGC
CAT CCA GAC GAT AAG CAT GAG CA
156 Capsule K1 Johnson and Stell, 2000 [8]
cvaC CAC ACA CAA ACG GGA GCT GTT
CTT CCC GCA GCA TAG TTC CAT
680 Colicin V Johnson and Stell, 2000 [8]
iutA GGC TGG ACA TCA TGG GAA CTG G
CGT CGG GAA CGG GTA GAA TCG
302 Aerobactin siderophore receptor Johnson et al., 2008 [9]
fyuA TGA TTA ACC CCG CGA CGG AA
CGC AGT AGG CAC GAT CTT GTA
880 Yersiniabactin Johnson and Stell, 2000 [8]
papC GAC GGC TGT ACT GCA GGG TGT GGC G
ATA TCC TTT CTG CAG GCA GGG TGT GGC
328 P fimbriae Le Bouguénec et al., 1992 [10]
papG CTG TAA TTA CGG AAG TGA TTT CTG
ACT ATC CGG CTC CGG ATA AAC CAT
1070 P fimbriae Johnson and Stell, 2000 [8]
sfaA CTC CGG AGA ACT GGG TGC ATC TTA C
CGG AGG AGT AAT TAC AAA CCT GGC A
410 Sfa fimbriae Le Bouguénec et al., 1992 [10]
sfaS GTG GAT ACG ACG ATT ACT GTG
CCG CCA GCA TTC CCT GTA TTC
240 Sfa fimbriae Johnson and Stell, 2000 [8]
Afa GGC AGA GGG CCG GCA ACA GGC
CCC GTA ACG CGC CAG CAT CTC
559 M fimbriae Johnson and Stell, 2000 [8]
ibe10 AGG CAG GTG TGC GCC GCG TAC
TGG TGC TCC GGC AAA CCA TGC
170 Invasion of brain endothelium Johnson and Stell, 2000 [8]
hlyA AAC AAG GAT AAG CAC TGT TCT GGC
ACC ATA TAA GCG GTC ATT CCC GTC
1.177 Hemolysin Johnson and Stell, 2000 [8]
cnf1 AGG ATG GAG TTT CCT ATG CAG GAG
CAT TCA GAG TCC TGC CCT CAT TAT T
498 Cytotoxic necrotizing factor 1 Johnson and Stell, 2000 [8]
cnf2 AAT CTA ATT AAA GAG AAC
CAT GCT TTG TAT ATC TA
543 Cytotoxic necrotizing factor 2 Blanco et al., 1996 [34]
traT GGT GTG GTG CGA TGA GCA CAG
CAC GGT TCA GCC ATC CCT GAG
290 Serum resistance Johnson and Stell, 2000 [8]
iroN AAT CCG GCA AAG AGA CGA ACC GCC T
GTT CGG GCA ACC CCT GCT TTG ACT TT
553 Salmochelin siderophore receptor Johnson et al., 2008 [9]
ompT TCA TCC CGG AAG CCT CCC TCA CTA CTA T
TAG CGT TTG CTG CAC TGG CTT CTG ATA C
496 Episomal outer membrane protease Johnson et al., 2008 [9]
hlyF GGC CAC AGT CGT TTA GGG TGC TTA CC
GGC GGT TTA GGC ATT CCG ATA CTC AG
450 Putative avian hemolysin Johnson et al., 2008 [9]
Iss CAG CAA CCC GAA CCA CTT GAT G
AGC ATT GCC AGA GCG GCA GAA
323 Episomal increased serum survival Johnson et al., 2008 [9]

This PCR contained 1 U Taq DNA polymerase (Invitrogen) in 2x PCR buffer (Invitrogen), 0.2 mM of each dNTP, 2.5 mM MgCl2, and 20 pmol/μL of each primer (Table 1). The program consisted of 94°C for 5 min, followed by 30 cycles of 94°C for 1 min, 55°C for 1 min, and 72°C for 1 min, with a final extension step at 72°C for 10 min [7]. The positive control used in the PCR was J96.

2.3. Phylogenetic Classification

Phylogenetic classification showed that the E. coli strains belonged to four groups (A, B1, B2, or D) based on the presence of the chuA and yjaA genes and the DNA fragment (TSPE4.C2). This PCR contained 1.25 U Taq DNA polymerase (Invitrogen) in 1x PCR buffer (Invitrogen), 20 pmol of each dNTP, 2.5 mM MgCl2, and 1 μM of each primer (Table 1). The program of PCR consisted of 94°C for 4 min, followed by 30 cycles of 94°C for 5 seg and 54°C for 10 seg, with a final extension step at 72°C for 5 min [3].

2.4. Virulence Factors Genes

The pathogenicity of E. coli is associated with the presence of virulence factors that can be encoded by chromosomal and plasmid genes, and thus 19 genes encoding virulence factors were investigated. The genes selected were specific for hemolysins (hlyA and hlyF), cytotoxic necrotizing factors (cnf1 and cnf2), colicin V (cvaC), aerobactin (iutA), yersiniabactin (fyuA), salmochelin (iroN), P-fimbriae (papC and papG), S-fimbrial adhesin (sfaA and sfaS), afimbrial adhesin (afa), serum resistance (iss and traT), brain microvascular endothelium invasion (ibe10), K1 capsule (kpsII and K1), and ompT outer membrane protein (ompT) [810]. This PCR contained 1.25 U Taq DNA polymerase (Invitrogen) in 1x PCR buffer (Invitrogen), 0.2 mM of each dNTP, 2.5 mM MgCl2, and 1 μM of each primer (Table 1). The program of PCR consisted of 94°C for 5 min, followed by 30 cycles of 94°C for 1 min, 55°C for 1 min, and 72°C for 1 min, with a final extension step at 72°C for 10 min.

2.5. Plasmids Profile

To analyze the plasmid profile, the plasmids of wild strains and the plasmids R27 (110 MDa), JPN11 (66 MDa), PSA (23 MD), and pRK (13, 2 MDa) used as markers of molecular mass were extracted by alkaline lysis [11]; the molecular weight of the plasmids was measured (LabImage 1D software) and the ability to transfer was determined.

The strains that harbored plasmids were chosen for mating experiments. The strains were grown in LB (Luria Bertani Broth) until the exponential phase. 1.2 mL of this culture was transferred to a flask containing 0.4 mL of the recipient culture in the stationary phase, E. coli K12-711 [12]. The mixture was incubated at 37°C for 3 hours. Transconjugants resistant to drugs were selected on MacConkey agar containing inhibitory concentrations of nalidixic acid (resistance present in E. coli K12-711), tetracycline, ampicillin, or kanamycin. The colonies grown on each selective plate were tested for the presence of virulence genes and pathogenicity islands, according to the virulence pattern of the donor strain. The resistance profile, necessary for this test, is shown in Table 2.

Table 2.

Plasmid and resistance profile of strains from hemocultures.

Strains Plasmids Molecular size of the plasmids (MDa) Resistance profile
1 p1a 78 ApCbCfCmKnTr
2 p2a, b, c, d 113, 112, 86, 82 ApCbSmFoKnTr
3 p3a, b, c, d, e 112, 99, 78, 58, 29 ApCbCfCmSmFoGnKnSiSuTbTcTr
4 p4a 59 ApCfFoTcTr
5 p5a, b 99, 71 ApCbCfCmSmFoKnSuTcTr
6 p6a 112 ApCbCfCmSmFoGnKnSiTcTr
7 p7a 99 ApCbSmFoSu
8 NP ApFo
9 NP ApFo
10 p10a 82 ApCmFoTc
11 p11a 68 ApCbCmSmFo
12 NP Ap
13 p13a, b, c, d, e 87, 68, 59, 50, 29 ApCbCfSmFoGnKnSuTbTcTr
14 p14a, b 62, 44 ApCbSm
15 NP Ap
16 NP Ap
17 p17a 87 ApCfCmSmKnTcTr
18 p18a 62 Ap
19 p19a 92 ApCbCmSmKnSuTcTr
20 NP Ap

NP: no plasmid; Ap: ampicillin; Cb: carbenicillin; Cf: cephalothin; Cm: chloramphenicol; Sm: streptomycin; Fo: fosfomycin; Gn: gentamicin; Kn: kanamycin; Si: sisomicin; Su: sulfonamide; Tb: tobramycin; Tc: tetracycline; Tr: trimethoprim.

3. Results and Discussion

Extraintestinal pathogenic Escherichia coli (ExPEC) is one of the leading causes of bloodstream infections (BSI) worldwide. In Brazil, it was observed a high mortality associated with BSI [13]. Despite the importance of E. coli bloodstream infections due to their high morbidity and mortality, the pathogenesis is not well known [14, 15] and not studied enough in South America [4, 5].

In this study we screened 20 strains from bacteremia from newborns (10%), children aged between 6 months and 14 years (20%), and adults (70%) and from two Brazilian teaching hospitals for the determination of phylogenetic group, pathogenicity islands (PAI) associated sequences, and important virulence factors responsible for extraintestinal pathogenesis. The results were compared with those obtained with isolates from the stools of healthy humans. The relationship between the presence of PAIs, virulence genes, and the phylogenetic group was analyzed.

Among the 20 isolates from the hemocultures tested, 70% of the isolates displayed PAI associated sequences (total of 22 islands), while in the commensal strains 52.94% of them displayed PAI (total of 45 islands). In agreement with previously published data [7, 16, 17], PAI IV536 was the most prevalent in both groups, following the PAI ICFT073 and PAI IICFT073, as shown in Tables 3 and 4. PAI IV536 has been described by many researchers as a stable island, and it is one of the most commonly found PAIs in enterobacteria [7, 18]. Our results show that the islands present in UPEC, although poorly researched in septicemic strains, are also found in E. coli isolated from hemocultures. This similarity can be associated with the fact that the urinary tract infection is one of the most common infections and bacteraemia is often a complication of this infection. But there are other ways for the presence of bacteria in the blood, such as meningitis and polymicrobial intra-abdominal infections. The presence of bacteria in the bloodstream suggests the ability of these pathogens to survive in an environment with scarce free iron and to resist the bactericidal activity in the blood [4].

Table 3.

Study of pathogenicity islands, phylogenetic classification, and virulence genes in 20 ExPEC strains from human hemocultures.

Strains PC Pathogenicity islands Virulence genes
IV536 ICFT073 IICFT073 IJ96 kpsII K1 fyuA papC papG sfaA Ibe10 hly cnf1 cnf2 traT iutA iroN ompT hlyF iss
1 B2 + + + + + + + + + +
2 B2 + + + + + + +
3 B2 + + + + +
4 B2 + + + + +
5 A + + + + +
6 A + + + + +
7 B2 + + + + + + + + + +
8 B2 + +
9 B2 + + + + + + + + + +
10 B1 + + + + + + + + +
11 A + + + + + +
12 B1 + + +
13 A + + + + + + + +
14 A + + + + + + +
15 A +
16 B2 +
17 D + + + +
18 B1 + + +
19 D + + + + + +
20 B1 + + + +

PC: phylogenetic classification; pathogenicity islands: PAI IV536, PAI ICFT073, PAI IICFT073, and PAI IJ96; adhesions: papC, papG, sfaA, sfaS, and afa; invasion: ibe10; toxins: cnf2, cnf1, hlyA, hlyF, and cvaC; iron uptake systems: fyuA, iutA, and iroN; serum resistance: kps II, K1, traT, and iss; proteases: ompT.

Table 4.

Distribution of pathogenicity islands (PAI) according to phylogenetic classification (PC), among ExPEC and commensal strains.

PC PAI I536 PAI II536 PAI IV536 PAI IJ96 PAI ICFT073 PAI IICFT073 Total of PAIs Total of strains
B2 (ExPEC) 0 0 6 1 6 2 15 8
B2 (commensal) 0 0 2 0 2 1 5 6
D (ExPEC) 0 0 2 0 0 0 2 2
D (commensal) 0 0 0 0 0 0 0 3
B1 (ExPEC) 0 0 3 0 0 1 3 4
B1 (commensal) 0 0 0 0 2 0 2 5
A (ExPEC) 0 0 1 0 0 0 1 6
A (commensal) 3 3 21 0 7 4 38 37

It is not clear yet if all E. coli from the intestinal tract of healthy people can be considered commensal, as some isolates showed up to five pathogenicity islands. Already it has been reported that ExPEC can asymptomatically colonize the intestinal tract [7].

In this study, genes related to toxin and hemolysin production were researched and included cnf2, cnf1, hlyA, hlyF, and cvaC. The hlyA gene is frequently detected in ExPEC and it was the most prevalent gene in our strains of hemoculture (30%) [19, 20]. Escherichia coli hemolysin (hlyA) is a pore-forming bacterial exotoxin that may contribute to the virulence of bacteria during bloodstream infection and sepsis [20, 21]. The PAIs I536, II536, ICFT073, IJ96, and IIJ96 harbor a copy of hlyABCD system encoding α-hemolysin, and in our study, six strains harbor hlyA gene and three also had these PAIs (Table 3). Commensal strains showed a large prevalence of the hlyA gene too, with 52.94% of the strains, and did not show a good correlation with the PAIs, since only 3 of 27 isolates that had the hlyA had the respective PAI. Of the virulence genes encoding adhesins (papC, papG, sfaA, sfaS, and afa), papC and papG were the most prevalent in the strains of hemoculture. These two genes were present in 30% of our strains and were always found together in the isolates, including in commensal strains. These genes are part of the mannose-resistant P-fimbriae operon and have been associated with E. coli isolated from bacteremia [22, 23]. The PAIs IJ96, ICFT073, and IICFT073 harbor genes encoding P-fimbriae. Of the six strains of hemocultures containing the genes papC and papG, three contained corresponding PAIs (Table 3). Meanwhile, all the commensal strains containing the papC and papG genes also have PAI ICFT073, showing a good correlation between them.

Siderophore production is important for bacterial survival in the bloodstream. The aerobactin siderophore system is an important virulence factor that contributes to bacterial growth in host tissues and fluids where iron availability is limited [24, 25]. The aerobactin receptor (IutA) is commonly associated with extraintestinal E. coli and those isolated from bacteremia [22, 26]. In the present study, iutA was the most commonly found virulence gene, present in 65% of the isolates tested (Table 3), different from commensal which did not show this gene. PAI ICFT073 contains aerobactin genes, and 5 of the 13 strains containing the iutA gene also contained sequences similar to PAI ICFT073. Genes corresponding to other iron uptake systems were also identified: iroN, encoding the salmochelin receptor, was present in 55% of isolates; and fyuA, encoding the yersiniabactin receptor [24, 25], was present in 45% of isolates. PAI IV536 contains yersiniabactin encoding genes, and 8 of the 9 isolates exhibiting the fyuA gene contained PAI IV536 related sequences. Thus, there is a good correlation between the presence of this island and the genes encoding yersiniabactin. There was a good correlation between the commensal strains too and of the 24 isolates that contained the fyuA gene, 22 isolates had also PAI IV536.

Of the genes that confer serum resistance (kpsII, K1, traT, and iss), traT, which encodes an outer membrane lipoprotein that contributes to serum resistance [27] was detected in 50% of the isolates.

The invasion determinant encoded by the ibe10 gene was present in 10% of the isolates, whereas the ompT gene, encoding an outer membrane protease, was present in 15% of the isolates.

These results demonstrate that the virulence of septicemic E. coli is not summarized by the presence of a single virulence factor, since each step in the infection process can be mediated by different virulence factors [14, 28], but the expression of the combination of virulence factors together with the imbalance between immune defenses and characteristics of the environment determines a multifactorial outcome [29].

Several studies have demonstrated that isolates belonging to phylogenetic group B2 are more commonly extraintestinal pathogenic strains [3, 15, 22]. Our results demonstrated that group B2 (40%) was the most common group among the E. coli strains from hemoculture (Table 4). Strains belonging to group B2 also had the most PAI associated sequences, and 15 of the total 22 PAIs identified were present in this group. In contrast, in the commensal strain the most prevalent group was A (72.54%) and this group had a greater number of PAIs (86.95%) (Table 4). Moreover, seven of eight strains of group B2 had PAIs, against only one of six strains of group A. However, despite reports in the literature that isolates belonging to groups A and B1 are more often strictly commensal strains from the intestinal microbiota [3, 7, 15, 22], ten of the isolates sampled belonged to groups A and B1 (Table 3), demonstrating that these groups are also capable of causing systemic infection. The results of the current study indicate that isolates that are phylogenetically characterized as mainly commensal can in some cases be isolated from bloodstream infections, reinforcing the concept that virulence is associated with the presence of multiple virulence factors and is dependent on the host's immune system [29]. The results also showed that group B2 E. coli, despite being uncommon among commensal strains, can be present in intestinal flora (11.76% of our commensal strains) (Table 4), suggesting that they may act as a reservoir for bacteria that can cause extraintestinal infection [7].

In previous reports, ExPEC strains from group B2 have been shown to contain more virulence factors than those from groups A and B1 [15, 22, 30]. However, our strains showed on average 4 to 5 virulence factors genes, regardless of the phylogenetic group, and despite the correlation between the presence of virulence genes and strains belonging to phylogenetic group B2, some isolates from group B2 were found to have few virulence genes, whereas some isolates from groups A and B1 had up to 8 virulence genes, while some strains of group B2 with a greater number of virulence factors had 8 virulence genes too (Table 3). Thus, although some isolates from groups A and B1 are limited in their virulence gene content and are not likely to be highly virulent, others in these groups contained multiple virulence factors genes that could contribute to extraintestinal virulence.

Similarly, the virulence genes are not only associated with the PAIs, because some ExPEC also harbor virulence genes in plasmids [31]. Furthermore, PAI are located in regions of high genetic mobility, which show elements that allow recombination and, consequently, PAI rearrangement, deletion, and/or acquisition of foreign DNA [7].

Given the importance of plasmids as mobile elements in the horizontal gene transfer, the plasmid profile was investigated. In this study, 14 strains showed these plasmids, which ranged from 1 to 5 plasmids per strain, and three plasmids transferred to another E. coli by conjugation, appearing to be conjugative plasmids. The transconjugants (sample 5, 17, and 19) received genes iutA, ompT, hlyF, iroN, traT, and iss, showing the possible presence of these virulence genes in plasmids (Table 5). Thus, the presence of transferable plasmids in E. coli isolated from hemocultures can contribute to the horizontal transfer of virulence genes to nonpathogenic isolates and interestingly the conjugative plasmids 5.2, from strain 5, had iss, iroN, ompT, and hlyF genes, whose genes are generally present in typical APEC plasmids in a conserved virulence plasmidic (CVP) region [32]. These findings support the idea that APEC or E. coli from commercial chicken carcasses has a potential zoonotic risk, as well as serves as a reservoir for virulence genes for ExPEC strains [9, 28, 33].

Table 5.

Genetic characterization of ExPEC transconjugants.

Strains iss traT iutA iroN ompT hlyF
E. coli K12-711
E. coli 5 + + + + +
5.2 + + + + +
E. coli 17 + +
17.1 + +
E. coli 19 + +
19.1 +

Serum resistance: iss and traT; iron uptake systems: iutA and iroN; proteases: ompT; toxin: hlyF.

Strains (donator of plasmids) = 5, 17, and 19; Tranconjugants = 5.2, 17.1, and 19.1.

As ExPEC pathogenicity is due to genetic features such as virulence genes, pathogenicity islands, and plasmids associated with the virulence, the study of genetic factors is important to better understand these important pathogens. Thus, for the screening and even prevention of blood-borne diseases caused by ExPECs, further research should be conducted on the genetic features associated with the virulence of these pathogens. Although in stool samples of healthy people the presence of commensal strains is common, our results showed that the intestinal microbiota may harbor E. coli of phylogenetic group B2. Also E. coli with PAIs and virulence genes suggest that the intestinal microbiota may act as a reservoir of ExPEC with virulence genetic factors present at the E. coli from blood stream infection.

Acknowledgment

This study was supported by Decit/SCTIE/MS/CNPq, FundaçãoAraucária (Edital PPSUS: Gestão Compartilhada em Saúde—2009).

Conflict of Interests

The authors declare that they have no conflict of interests.

References

  • 1.Sader HS, Jones RN, Andrade-Baiocchi S, Biedenbach DJ. Four-year evaluation of frequency of occurrence and antimicrobial susceptibility patterns of bacteria from bloodstream infections in Latin American medical centers. Diagnostic Microbiology and Infectious Disease. 2002;44(3):273–280. doi: 10.1016/s0732-8893(02)00469-8. [DOI] [PubMed] [Google Scholar]
  • 2.Laupland KB, Gregson DB, Church DL, Ross T, Pitout JDD. Incidence, risk factors and outcomes of Escherichia coli bloodstream infections in a large Canadian region. Clinical Microbiology and Infection. 2008;14(11):1041–1047. doi: 10.1111/j.1469-0691.2008.02089.x. [DOI] [PubMed] [Google Scholar]
  • 3.Clermont O, Bonacorsi S, Bingen E. Rapid and simple determination of the Escherichia coli phylogenetic group. Applied and Environmental Microbiology. 2000;66(10):4555–4558. doi: 10.1128/aem.66.10.4555-4558.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Ananias M, Yano T. Serogroups and virulence genotypes of Escherichia coli isolated from patients with sepsis. Brazilian Journal of Medical and Biological Research. 2008;41(10):877–883. doi: 10.1590/s0100-879x2008001000008. [DOI] [PubMed] [Google Scholar]
  • 5.Conceição RA, Ludovico MS, Andrade CGTJ, Yano T. Human sepsis-associated Escherichia coli (SEPEC) is able to adhere to and invade kidney epithelial cells in culture. Brazilian Journal of Medical and Biological Research. 2012;45:376–472. doi: 10.1590/S0100-879X2012007500057. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Tomazetto G, Silva ACM, Vidotto MC. Fatores de virulência de amostras de Escherichia coli septicêmicas. Biosaúde. 2004;6:3–16. [Google Scholar]
  • 7.Sabaté M, Moreno E, Pérez T, Andreu A, Prats G. Pathogenicity island markers in commensal and uropathogenic Escherichia coli isolates. Clinical Microbiology and Infection. 2006;12(9):880–886. doi: 10.1111/j.1469-0691.2006.01461.x. [DOI] [PubMed] [Google Scholar]
  • 8.Johnson JR, Stell AL. Extended virulence genotypes of Escherichia coli strains from patients with urosepsis in relation to phylogeny and host compromise. The Journal of Infectious Diseases. 2000;181(1):261–272. doi: 10.1086/315217. [DOI] [PubMed] [Google Scholar]
  • 9.Johnson TJ, Wannemuehler Y, Doetkott C, Johnson SJ, Rosenberger SC, Nolan LK. Identification of minimal predictors of avian pathogenic Escherichia coli virulence for use as a rapid diagnostic tool. Journal of Clinical Microbiology. 2008;46(12):3987–3996. doi: 10.1128/JCM.00816-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Le Bouguénec CL, Archambaud M, Labigne A. Rapid and specific detection of the pap, afa, and sfa adhesin-encoding operons in uropathogenic Escherichia coli strains by polymerase chain reaction. Journal of Clinical Microbiology. 1992;30(5):1189–1193. doi: 10.1128/jcm.30.5.1189-1193.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Kado CI, Liu S-T. Rapid procedure for detection and isolation of large and small plasmids. Journal of Bacteriology. 1981;145(3):1365–1373. doi: 10.1128/jb.145.3.1365-1373.1981. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Laporta MZ, Silva MLM, Scaletsky ICA, Trabulsi LR. Plasmids coding for drug resistance and localized adherence to HeLa cells in Enteropathogenic Escherichia coli O55:H- and O55:H6. Infection and Immunity. 1986;51(2):715–717. doi: 10.1128/iai.51.2.715-717.1986. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Sales JAL, Jr., David CM, Hatum R, et al. Sepse Brasil: estudo epidemiológico da sepse em unidades de terapia intensiva brasileiras. Revista Brasileira de Terapia Intensiva. 2006;18:9–17. [PubMed] [Google Scholar]
  • 14.Köhler C-D, Dobrindt U. What defines extraintestinal pathogenic Escherichia coli? International Journal of Medical Microbiology. 2011;301(8):642–647. doi: 10.1016/j.ijmm.2011.09.006. [DOI] [PubMed] [Google Scholar]
  • 15.Skjot-Rasmussen L, Ejrnaes K, Lundgren B, Hammerum AM, Frimodt-Moller N. Virulence factors and phylogenetic grouping of Escherichia coli isolates from patients with bacteraemia of urinary tract origin relate to sex and hospital- vs. Community-acquired origin. International Journal of Medical Microbiology. 2012;302(3):129–134. doi: 10.1016/j.ijmm.2012.03.002. [DOI] [PubMed] [Google Scholar]
  • 16.Ostblom A, Adlerberth I, Wold AE, Nowrouzian FL. Pathogenicity island markers, virulence determinants malX and usp, and the capacity of Escherichia coli to persist in infants’ commensal microbiotas. Applied and Environmental Microbiology. 2011;77(7):2303–2308. doi: 10.1128/AEM.02405-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Schubert S, Rakin A, Heesemann J. The Yersinia high-pathogenicity island (HPI): evolutionary and functional aspects. International Journal of Medical Microbiology. 2004;294(2-3):83–94. doi: 10.1016/j.ijmm.2004.06.026. [DOI] [PubMed] [Google Scholar]
  • 18.Middendorf B, Hochhut B, Leipold K, Dobrindt U, Blum-Oehler G, Hacker J. Instability of Pathogenicity Islands in Uropathogenic Escherichia coli 536. Journal of Bacteriology. 2004;186(10):3086–3096. doi: 10.1128/JB.186.10.3086-3096.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Bert F, Huynh B, Dondero F, et al. Molecular epidemiology of Escherichia coli bacteremia in liver transplant recipients. Transplant Infectious Disease. 2011;13(4):359–365. doi: 10.1111/j.1399-3062.2011.00618.x. [DOI] [PubMed] [Google Scholar]
  • 20.Sonnem AFP, Henneke P. Role of pore-forming toxins in neonatal sepsis. Clinical and Developmental Immunology. 2013;2013:13 pages. doi: 10.1155/2013/608456.608456 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Bien J, Sokolova O, Bozko P. Role of uropathogenic Escherichia coli virulence factors in development of urinary tract infection and kidney damage. International Journal of Nephrology. 2012;2012:15 pages. doi: 10.1155/2012/681473.681473 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Moreno E, Planells I, Prats G, Planes AM, Moreno G, Andreu A. Comparative study of Escherichia coli virulence determinants in strains causing urinary tract bacteremia versus strains causing pyelonephritis and other sources of bacteremia. Diagnostic Microbiology and Infectious Disease. 2005;53(2):93–99. doi: 10.1016/j.diagmicrobio.2005.05.015. [DOI] [PubMed] [Google Scholar]
  • 23.Otto G, Magnusson M, Svensson M, Braconier JH, Svanborg C. pap Genotype and P fimbrial expression in Escherichia coli causing bacteremic and nonbacteremic febrile urinary tract infection. Clinical Infectious Diseases. 2001;32(11):1523–1531. doi: 10.1086/320511. [DOI] [PubMed] [Google Scholar]
  • 24.Garénaux A, Caza M, Dozois CM. The Ins and Outs of siderophore mediated iron uptake by extra-intestinal pathogenic Escherichia coli . Veterinary Microbiology. 2011;153(1-2):89–98. doi: 10.1016/j.vetmic.2011.05.023. [DOI] [PubMed] [Google Scholar]
  • 25.Gao Q, Wang X, Xu H, et al. Roles of iron acquisition systems in virulence of extraintestinal pathogenic Escherichia coli: salmochelin and aerobactin contribute more to virulence than heme in a chicken infection model. BMC Microbiology. 2012;143:1–12. doi: 10.1186/1471-2180-12-143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Hilali F, Ruimy R, Saulnier P, et al. Prevalence of virulence genes and clonality in Escherichia coli strains that cause bacteremia in cancer patients. Infection and Immunity. 2000;68(7):3983–3989. doi: 10.1128/iai.68.7.3983-3989.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Johnson JR, Russo TA. Uropathogenic Escherichia coli as agents of diverse non-urinary tract extraintestinal infections. The Journal of Infectious Diseases. 2002;186(6):859–864. doi: 10.1086/342490. [DOI] [PubMed] [Google Scholar]
  • 28.Mokady D, Gophna U, Ron EZ. Extensive gene diversity in septicemic Escherichia coli strains. Journal of Clinical Microbiology. 2005;43(1):66–73. doi: 10.1128/JCM.43.1.66-73.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Bidet P, Bingen BE. Facteurs de pathogénicité et physiopathologie des Escherichia coli extra-intestinaux. Archives de Pédiatrie. 2012;19:S80–S92. doi: 10.1016/S0929-693X(12)71279-4. [DOI] [PubMed] [Google Scholar]
  • 30.Johnson JR, Clermont O, Menard M, Kuskowski MA, Picard B, Denamur E. Experimental mouse lethality of Escherichia coli isolates, in relation to accessory traits, phylogenetic group, and ecological source. The Journal of Infectious Diseases. 2006;194(8):1141–1150. doi: 10.1086/507305. [DOI] [PubMed] [Google Scholar]
  • 31.Pérez-Mendoza D, de la Cruz F. Escherichia coli genes affecting recipient ability in plasmid conjugation: are there any? BMC Genomics. 2009;10, article 71 doi: 10.1186/1471-2164-10-71. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Lemaître C, Mahjoub-Messai F, Dupont D, et al. A conserved virulence plasmidic region contributes to the virulence of the multiresistant Escherichia coli meningitis strain S286 belonging to phylogenetic group C. PLOS ONE. 2013;8(9) doi: 10.1371/journal.pone.0074423.e74423 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Kobayashi RKT, Aquino I, Ferreira ALDS, Vidotto MC. EcoR phylogenetic analysis and virulence genotyping of avian pathogenic Escherichia coli strains and Escherichia coli isolates from commercial chicken carcasses in southern Brazil. Foodborne Pathogens and Disease. 2011;8(5):631–634. doi: 10.1089/fpd.2010.0726. [DOI] [PubMed] [Google Scholar]
  • 34.Blanco M, Blanco JE, Alonso MP, Blanco J. Virulence factors and O 277 groups of Escherichia coliisolates from patients with acute pyelonephritis, cystitis, 278 and asymptomatic bacteriuria. European Journal of Epidemiology. 1996;12:191–198. doi: 10.1007/BF00145506. [DOI] [PubMed] [Google Scholar]

Articles from BioMed Research International are provided here courtesy of Wiley

RESOURCES