首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The study of phylogenetic groups and pathogenicity island (PAI) markers in commensal Escherichia coli strains from asymptomatic Chinese people showed that group A strains are the most common and that nearly half of all fecal strains which were randomly selected harbor PAIs.Escherichia coli is a well-diversified commensal species in the intestine of healthy humans but also includes intestinal or extraintestinal pathogens. It has been reported that pathogenic E. coli may be derived from fecal strains by acquisition of virulence determinants (11). The relationship between the E. coli genetic background and the acquisition of virulence factors is now better understood (1, 5). Extraintestinal E. coli strains may harbor several virulence factors, such as adhesins, fimbriae, and hemolysin, which can contribute to bacterial pathogenesis. These traits are usually encoded on pathogenicity islands (PAIs), which have been studied in pathogenic E. coli previously (15). The E. coli population includes 4 major phylogroups (A, B1, B2, and D) (2). Pathogenic strains belong mainly to groups B2 and D, while most fecal isolates belong to groups A and B1. Strains of groups B2 and D often carry virulence factors that are lacking in group A and B1 strains (3, 9, 13).In this study, we examined the distribution of phylogroups and the prevalence of PAIs in commensal E. coli strains isolated from asymptomatic persons in one region of China.  相似文献   

2.
Multiplex PCR analyses of DNAs from genotypically unique Escherichia coli strains isolated from the feces of 138 humans and 376 domesticated animals from Jeonnam Province, South Korea, performed using primers specific for the chuA and yjaA genes and an unknown DNA fragment, TSPE4.C2, indicated that none of the strains belonged to E. coli phylogenetic group B2. In contrast, phylogenetic group B2 strains were detected in about 17% (8 of 48) of isolates from feces of 24 wild geese and in 3% (3 of 96) of isolates obtained from the Yeongsan River in Jeonnam Province, South Korea. The distribution of E. coli strains in phylogenetic groups A, B1, and D varied depending on the host examined, and there was no apparent seasonal variation in the distribution of strains in phylogenetic groups among the Yeongsan River isolates. The distribution of four virulence genes (eaeA, hlyA, stx1, and stx2) in isolates was also examined by using multiplex PCR. Virulence genes were detected in about 5% (38 of 707) of the total group of unique strains examined, with 24, 13, 13, and 9 strains containing hlyA, eaeA, stx2, and stx1, respectively. The virulence genes were most frequently present in phylogenetic group B1 strains isolated from beef cattle. Taken together, results of these studies indicate that E. coli strains in phylogenetic group B2 were rarely found in humans and domesticated animals in Jeonnam Province, South Korea, and that the majority of strains containing virulence genes belonged to phylogenetic group B1 and were isolated from beef cattle. Results of this study also suggest that the relationship between the presence and types of virulence genes and phylogenetic groupings may differ among geographically distinct E. coli populations.Escherichia coli is a normal inhabitant of the lower intestinal tract of warm-blooded animals and humans. While the majority of E. coli strains are commensals, some are known to be pathogenic, causing intestinal and extraintestinal diseases, such as diarrhea and urinary tract infections (42). Phylogenetic studies done using multilocus enzyme electrophoresis and 72 E. coli strains in the E. coli reference collection showed that E. coli strains can be divided into four phylogenetic groups (A, B1, B2, and D) (20, 41, 48). Recently, a potential fifth group (E) has also been proposed (11). Since multiplex PCR was developed for analysis of phylogenetic groups (6), a number of studies have analyzed a variety of E. coli strains for their phylogenetic group association (10, 12, 17, 18, 23, 54). Duriez et al. (10) reported the possible influence of geographic conditions, dietary factors, use of antibiotics, and/or host genetic factors on the distribution of phylogenetic groups among 168 commensal E. coli strains isolated from human stools from three geographically distinct populations in France, Croatia, and Mali. Random-amplified polymorphic DNA analysis of the intraspecies distribution of E. coli in pregnant women and neonates indicated that there was a correlation between the distribution of phylogenetic groups, random-amplified polymorphic DNA groups, and virulence factors (54). Moreover, based on comparisons of the distribution of E. coli phylogenetic groups among humans of different sexes and ages, it has been suggested that E. coli genotypes are likely influenced by morphological, physiological, and dietary differences (18). In addition, climate has also been proposed to influence the distribution of strains within E. coli phylogenetic groups (12). There are now several reports indicating that there is a potential relationship between E. coli phylogenetic groups, age, and disease. For example, E. coli isolates belonging to phylogenetic group B2 have been shown to predominate in infants with neonatal bacterial meningitis (27) and among urinary tract and rectal isolates (55). Also, Nowrouzian et al. (39) and Moreno et al. (37) reported that strains belonging to phylogenetic group B2 persisted among the intestinal microflora of infants and were more likely to cause clinical symptoms.Boyd and Hartl (2) reported that among the E. coli strains in the E. coli reference and the diarrheagenic E. coli collections, strains in phylogenetic group B2 carry the greatest number of virulence factors, followed by those in group D. Virulence factors carried by group B2 strains are thought to contribute to their strong colonizing capacity; a greater number of virulence genes have been detected in resident strains than in transient ones (38). Moreover, a mouse model of extraintestinal virulence showed that phylogenetic group B2 strains killed mice at greater frequency and possessed more virulence determinants than strains in other phylogenetic groups, suggesting a link between phylogeny and virulence genes in E. coli extraintestinal infection (45). In contrast, Johnson and Kuskowski (25) suggested that a group B2 ancestral strain might have simply acquired virulence genes by chance and that these genes were vertically inherited by group members during clonal expansion. However, numerous studies published to date suggest that there is a relationship between the genomic background of phylogenetic group B2 and its association with virulence factors (12, 28, 35, 39, 45).Both enteropathogenic and enterohemorrhagic E. coli (EPEC and EHEC, respectively) strains are among the most important food-borne pathogens worldwide, often causing severe gastrointestinal disease and fatal infections (13). While EPEC strains cause diarrhea and generally do not produce enterotoxin, they possess an adherence factor which is controlled by the chromosomal gene eaeA, encoding intimin (8). Unlike the EPEC strains, however, the EHEC strains typically contain the hlyA, stx1, and stx2 virulence genes, encoding hemolysins and Shiga-like type 1 and 2 toxins, respectively, and eaeA. The ability to detect EHEC has been greatly facilitated by the use of multiplex PCR (13, 44, 53). Several studies have shown that strains producing Shiga-like toxin 2 are more frequently found in cases of hemolytic-uremic syndrome than are those containing Shiga-like toxin 1 (30, 43, 46, 49).In the study reported here, we examined the distribution of phylogenetic groups and the prevalence of virulence genes in 659 genotypically unique E. coli strains isolated from humans and domestic animals in South Korea. In addition, we also tested 48 and 96 nonunique E. coli isolates from wild geese and the Yeongsan River, respectively, for phylogenetic distribution and virulence gene profiles. Here, we report that contrary to what has been previously reported in other parts of the world, no E. coli strains belonging to phylogenetic group B2 were found in domesticated animals and in humans from Jeonnam Province, South Korea. We also report that among the strains we examined, virulence genes were mainly found in phylogenetic group B1 strains isolated from beef cattle. Results of these studies may prove to be useful for the development of risk management strategies to maintain public health.  相似文献   

3.
4.
To discern the possible spread of the Escherichia coli O25b:H4-ST131 clonal group in poultry and the zoonotic potential of avian strains, we made a retrospective search of our strain collection and compared the findings for those strains with the findings for current strains. Thus, we have characterized a collection of 19 avian O25b:H4-ST131 E. coli strains isolated from 1995 to 2010 which, interestingly, harbored the ibeA gene. Using this virulence gene as a criterion for selection, we compared those 19 avian strains with 33 human O25b:H4-ST131 ibeA-positive E. coli strains obtained from patients with extraintestinal infections (1993 to 2009). All 52 O25b:H4-ST131 ibeA-positive E. coli strains shared the fimH, kpsMII, malX, and usp genes but showed statistically significant differences in nine virulence factors, namely, papGIII, cdtB, sat, and kpsMII K5, which were associated with human strains, and iroN, kpsMII K1, cvaC, iss, and tsh, which were associated with strains of avian origin. The XbaI macrorestriction profiles of the 52 E. coli O25b:H4-ST131 ibeA-positive strains revealed 11 clusters (clusters I to XI) of >85% similarity, with four clusters including strains of human and avian origin. Cluster VII (90.9% similarity) grouped 10 strains (7 avian and 3 human strains) that mostly produced CTX-M-9 and that also shared the same virulence profile. Finally, we compared the macrorestriction profiles of the 12 CTX-M-9-producing O25b:H4-ST131 ibeA strains (7 avian and 5 human strains) identified among the 52 strains with those of 15 human O25b:H4-ST131 CTX-M-14-, CTX-M-15-, and CTX-M-32-producing strains that proved to be negative for ibeA and showed that they clearly differed in the level of similarity from the CTX-M-9-producing strains. In conclusion, E. coli clonal group O25b:H4-ST131 ibeA has recently emerged among avian isolates with the new acquisition of the K1 capsule antigen and includes CTX-M-9-producing strains. This clonal group represents a real zoonotic risk that has crossed the barrier between human and avian hosts.Strains of the extensively antimicrobial-resistant Escherichia coli clonal group of sequence type (ST) 131 (ST131) belonging to serotype O25b:H4 have recently been recognized to be important human pathogens worldwide (9, 33). Although it is commonly associated with the dissemination of CTX-M-15 extended-spectrum cephalosporin resistance, E. coli O25b:H4-ST131 also occurs as a fluoroquinolone (FQ)-resistant but cephalosporin-susceptible pathogen (5, 22, 26, 27). Currently, it is assumed that O25b:H4-ST131 strains circulate not only among humans but also among animal hosts (13, 21, 37), which would contribute to the ongoing global emergence of O25b:H4-ST131, in the case of regular transmission between animals and humans. Even though CTX-M-15 is the most widely distributed extended-spectrum beta-lactamase (ESBL) linked to this clonal group, other, different variants of CTX-M have recently been reported, such as CTX-M-9, CTX-M-14, and CTX-M-32 (4, 34, 36, 39). Noteworthy was the detection, for the first time on poultry farms, of this clonal group producing CTX-M-9 that had macrorestriction profiles and virulence genes very similar to those observed in clinical human isolates (10).Extraintestinal pathogenic E. coli (ExPEC) strains, which include avian pathogenic E. coli (APEC) and human uropathogenic E. coli (UPEC), septicemic E. coli, and newborn meningitis-causing E. coli (NMEC) strains, exhibit considerable genome diversity and have a wide range of virulence-associated factors (12, 18). While infections caused by APEC strains initially start as a respiratory tract disease which evolves to a systemic infection of the internal organs and, finally, to sepsis, the most frequent origin of human sepsis is urinary tract infection (UTI), especially pyelonephritis (2, 3, 11). However, APEC strains have been recognized to share common traits with human isolates (29, 30, 31), including the K1 capsule antigen (23, 24, 29) and the ibeA gene (14). In addition, retail chicken products have been found to carry nalidixic-resistant ExPEC strains (17, 19), and although it is drug susceptible, an E. coli strain belonging to the O25b:H4-ST131 clonal group has even recently been detected in retail chicken (41), supporting the urgent necessity for the implementation of food control measures.The aim of the present study was to discern the possible spread of the O25b:H4-ST131 clonal group, especially CTX-M-9-producing strains, in poultry and the zoonotic potential of avian isolates. For this purpose, we made a retrospective search of our human and avian strain collections and compared the findings for those strains with the findings for current strains. Identification of this emerging clone among avian sources and comparison of the clone with clinical human isolates will shed new light on the epidemiology of the O25b:H4-ST131 clonal group.  相似文献   

5.
The impact of feed supplementation with bambermycin, monensin, narasin, virginiamycin, chlortetracycline, penicillin, salinomycin, and bacitracin on the distribution of Escherichia coli pathotypes in broiler chickens was investigated using an E. coli virulence DNA microarray. Among 256 E. coli isolates examined, 59 (23%) were classified as potentially extraintestinal pathogenic E. coli (ExPEC), while 197 (77%) were considered commensal. Except for chlortetracycline treatment, the pathotype distribution was not significantly different among treatments (P > 0.05). Within the 59 ExPEC isolates, 44 (75%) were determined to be potentially avian pathogenic E. coli (APEC), with the remaining 15 (25%) considered potentially “other” ExPEC isolates. The distribution within phylogenetic groups showed that 52 (88%) of the ExPEC isolates belonged to groups B2 and D, with the majority of APEC isolates classified as group D and most commensal isolates (170, 86%) as group A or B1. Indirect assessment of the presence of the virulence plasmid pAPEC-O2-ColV showed a strong association of the plasmid with APEC isolates. Among the 256 isolates, 224 (88%) possessed at least one antimicrobial resistance gene, with nearly half (107, 42%) showing multiple resistance genes. The majority of resistance genes were distributed among commensal isolates. Considering that the simultaneous detection of antimicrobial resistance tet(A), sulI, and blaTEM genes and the integron class I indicated a potential presence of the resistance pAPEC-O2-R plasmid, the results revealed that 35 (14%) of the isolates, all commensals, possessed this multigene resistance plasmid. The virulence plasmid was never found in combination with the antimicrobial resistance plasmid. The presence of the ColV plasmid or the combination of iss and tsh genes in the majority of APEC isolates supports the notion that when found together, the plasmid, iss, and tsh serve as good markers for APEC. These data indicate that different resistant E. coli pathotypes can be found in broiler chickens and that the distribution of such pathotypes and certain virulence determinants could be modulated by antimicrobial agent feed supplementation.Several classes of antimicrobial agents, such as glycolipids (bambermycin), cyclic peptides (bacitracin), ionophores (monensin and salinomycin), streptogramins (virginiamycin), and β-lactams (penicillin), are widely used as food additives in modern animal husbandry to prevent infections and promote growth (6). Increasing antimicrobial resistance in animals and its potential threat to human health led to the ban of bacitracin, spiramycin, tylosin, and virginiamycin as feeding additives by the European Union in 1999 (7, 46). Although this precautionary measure is still controversial because of being seen as having a negligible impact on human health, negative consequences for animal health and welfare, including economic losses for farmers, were subsequently observed in Europe (7). In stark contrast, however, the ban has been beneficial in reducing the total quantity of antibiotics administered to food animals (7, 47). Under good production conditions and correct use of antibiotics, poultry production is reported to be competitive (14, 47, 48) and even beneficial in reducing antimicrobial resistance in important food animal reservoirs and thus the potential threat to public health (48).Escherichia coli is generally considered a commensal member of the normal gastrointestinal microflora in humans and animals, yet some strains are known to cause serious morbidity and mortality. The expression of various virulence factors, which affect cellular processes, can result in different clinical diseases, such as cystitis, pyelonephritis, sepsis/meningitis, and gastroenteritis. The possession of different virulence gene subsets can further define the E. coli pathotype (31). The extraintestinal pathogenic E. coli (ExPEC) strains are epidemiologically and phylogenetically distinct from both intestinal pathogenic and commensal strains (43). In North America, annually, several million cases of urinary tract infections, abdominal infections, pelvic infections, pneumonia, meningitis, and sepsis are caused by ExPEC (42). In poultry production, avian pathogenic E. coli (APEC) is responsible for significant economic losses. APEC strains induce extraintestinal diseases such as air sacculitis, colibacillosis, polysorositis, and septicemia in birds (9, 21, 22, 31, 35, 45). Although no specific set of virulence factors has been clearly linked to APEC strains, most identified virulence factors are similar to those frequently associated with ExPEC (36).Bearing in mind that the avian intestinal environment has been considered a reservoir of E. coli having zoonotic potential (15) and the possible contamination of poultry products with such bacteria during slaughter, the impact of antimicrobial feeding additives on the distribution and dissemination of bacterial pathotypes and antibiotic resistance needs to be explored to address human, animal, and environmental health concerns. To this end, an E. coli DNA virulence microarray previously employed to assess the genotypes (virulence and antibiotic resistance genes) of E. coli strains isolated from different environmental ecosystems and from the chicken intestinal tract (1, 10, 19, 20, 33) was used. The aim of the present trial was to investigate the distributions of pathotypes and of virulence and antibiotic resistance genes in E. coli isolates from broilers fed with antimicrobial supplementation diets including bambermycin, penicillin, salinomycin, bacitracin, chlortetracycline, virginiamycin, monensin, and narasin.  相似文献   

6.
DNA sequence-based molecular subtyping methods such as multilocus sequence typing (MLST) are commonly used to generate phylogenetic inferences for monomorphic pathogens. The development of an effective MLST scheme for subtyping Escherichia coli O157:H7 has been hindered in the past due to the lack of sequence variation found within analyzed housekeeping and virulence genes. A recent study suggested that rhs genes are under strong positive selection pressure, and therefore in this study we analyzed these genes within a diverse collection of E. coli O157:H7 strains for sequence variability. Eighteen O157:H7 strains from lineages I and II and 15 O157:H7 strains from eight clades were included. Examination of these rhs genes revealed 44 polymorphic loci (PL) and 10 sequence types (STs) among the 18 lineage strains and 280 PL and 12 STs among the 15 clade strains. Phylogenetic analysis using rhs genes generally grouped strains according to their known lineage and clade classifications. These findings also suggested that O157:H7 strains from clades 6 and 8 fall into lineage I/II and that strains of clades 1, 2, 3, and 4 fall into lineage I. Additionally, unique markers were found in rhsA and rhsJ that might be used to define clade 8 and clade 6. Therefore, rhs genes may be useful markers for phylogenetic analysis of E. coli O157:H7.Escherichia coli O157:H7 was first described in 1983 as the causative agent of a food-borne outbreak attributed to contaminated ground beef patties (35), and it has subsequently emerged as a very important food-borne pathogen. Diseases caused by E. coli O157:H7, such as hemorrhagic colitis and hemolytic uremic syndrome, can be very severe or even life-threatening. Cattle are believed to be the main reservoir for E. coli O157:H7 (5, 15, 41), although other animals may also carry this organism (6, 21). Outbreaks are commonly associated with the consumption of beef and fresh produce that come into contact with bovine feces or feces-contaminated environments, such as food contact surfaces, animal hides, or irrigation water (12, 21, 30, 38).It is well-established that strains of E. coli O157:H7 vary in terms of virulence and transmissibility to humans and that strains differing in these characteristics can be distinguished using DNA-based methods (22, 29, 42). For example, octamer-based genome scanning, which is a PCR approach using 8-bp primers, provided the first evidence that there are at least two lineages of O157:H7, termed lineage I and lineage II (22). Strains classified as lineage I are more frequently isolated from humans than are lineage II strains (42). A later refinement of this classification system was coined the lineage-specific polymorphism assay (LSPA), which classified strains based upon the amplicon size obtained using PCRs targeting six chromosomal regions of E. coli O157:H7 and assigned a six-digit code based upon the pattern obtained (42). Most strains of lineage I grouped into LSPA type 111111, while the majority of lineage II strains fell into LSPA types 211111, 212111, and 222222. More recently, it was suggested that LSPA type 211111 strains comprise a separate group called lineage I/II (45).To gain greater insight into the recent evolution of E. coli O157:H7, a method that is more discriminatory than the LSPA method is desirable. Multilocus sequence typing (MLST) is a method that discriminates between strains of a bacterial species by identifying DNA sequence differences in six to eight targeted genes. Satisfactory MLST schemes exist for other bacterial pathogens (28, 43); however, due to the lack of sequence variations in previously targeted gene markers in E. coli O157:H7 (13, 33), MLST approaches for subtyping this pathogen have been more difficult to develop. More recently, high-throughput microarray and sequencing platforms have been used to identify hundreds of single nucleotide polymorphisms (SNPs) that are useful for discriminating between strains of E. coli O157:H7 during epidemiologic investigations and for drawing phylogenetic inferences (11, 20, 29, 44). Particularly noteworthy, Manning et al. (29) developed a subtyping scheme based upon the interrogation of 32 putative SNP loci. This method separated 528 strains into 39 distinct SNP genotypes, which were grouped into nine statistically supported phylogenetic groups called clade 1 through clade 9. By analyzing the rates of hemolytic uremic syndrome observed in patients infected with strains of clades 2, 7, and 8, it was also concluded that clade 8 strains are more virulent to humans than other strains (29).One drawback of current DNA sequence-based subtyping schemes for E. coli O157:H7 is that they require screening of at least 32 SNP loci. We were interested in asking whether a simpler approach that targets a few informative gene markers could be developed for rapid strain discrimination and phylogenetic determination. A recent analysis of E. coli genomes predicted that rearrangement hot spot (rhs) genes are under the strongest positive selection of all coding sequences analyzed (34). Therefore, we hypothesized that these genes would display significant sequence variations for subtyping O157:H7 strains. The rhs genes were first discovered as elements mediating tandem duplication of the glyS locus in E. coli K-12 (26); however, their function remains unknown. There are nine rhs genes within the genome of the prototypical E. coli O157:H7 strain Sakai, and these genes are designated rhsA, -C, -D, -E, -F, -G, -I, -J, and -K (see Table S1 in the supplemental material) (16). Three of these nine rhs genes, rhsF, -J, and -K, were previously studied by Zhang et al. (44), and a number of SNPs were identified among these genes. However, no studies have been conducted to comprehensively investigate rhs genes as markers in an MLST scheme for subtyping E. coli O157:H7.The primary purpose of the present study was to investigate whether there are sufficient DNA sequence variations among rhs genes to develop an MLST approach for subtyping E. coli O157:H7. In this study, a greater level of DNA sequence variation was observed among rhs genes than in gene markers targeted in previous studies (13, 33). Furthermore, phylogenetic analysis using these rhs genes generally agreed with the established lineage and clade classifications of O157:H7 strains defined previously. We also wanted to determine whether there is a correlation between the lineage classification of O157:H7 strains (42) and the recently proposed clade classification (29). The present study reports evidence that O157:H7 strains from clade 8 are classified as lineage I/II, which is a different lineage from well-studied E. coli O157:H7 outbreak strains, such as EDL933 and Sakai. Therefore, we suggest that outbreaks of O157:H7 are caused by two lineages of this pathogen, lineage I and lineage I/II.  相似文献   

7.
Staphylococcal enterotoxins (SE) can cause toxin-mediated disease, and those that function as superantigens are implicated in the pathogenesis of allergic diseases. The prevalence of 19 enterotoxin genes was determined by PCR in clinical S. aureus strains derived from wounds (108) and blood (99). We performed spa typing and multilocus sequence typing (MLST) to determine clonal origin, and for selected strains staphylococcal enterotoxin B (SEB) production was measured by enzyme-linked immunosorbent assay. Strains carried a median of five SE genes. For most SE genes, the prevalence rates among methicillin-resistant and methicillin-sensitive S. aureus isolates, as well as wound- and blood-derived isolates, did not differ. At least one SE gene was detected in all except two S. aureus isolates (>99%). Complete egc clusters were found in only 11% of S. aureus isolates, whereas the combination of sed, sej, and ser was detected in 24% of clinical strains. S. aureus strains exhibited distinct combinations of SE genes, even if their pulsed-field gel electrophoresis and MLST patterns demonstrated clonality. USA300 strains also showed considerable variability in SE content, although they contained a lower number of SE genes (mean, 3). By contrast, SE content was unchanged in five pairs of serial isolates. SEB production by individual strains varied up to 200-fold, and even up to 15-fold in a pair of serial isolates. In conclusion, our results illustrate the genetic diversity of S. aureus strains with respect to enterotoxin genes and suggest that horizontal transfer of mobile genetic elements encoding virulence genes occurs frequently.As a commensal, Staphylococcus aureus colonizes the nasal mucosa of 20 to 40% of humans (54), and as a pathogen it causes pyogenic diseases and toxin-mediated diseases (38). S. aureus produces many different virulence factors, including enterotoxins (SEs), which can cause defined toxic shock syndromes (4). The characterization of some of these toxins led to the discovery of superantigens (41), which bind to major histocompatibility complex class II molecules and Vβ chains of T-cell receptors, resulting in the activation of large numbers of T cells (20 to 30%) and massive cytokine production (10, 18). These superantigen-induced “cytokine storms” are responsible for the toxic effects seen in staphylococcal entertoxin B (SEB)- and toxic shock syndrome toxin (TSST)-associated shock syndromes in S. aureus infections (13, 40, 47). To date, 19 SEs have been identified based on sequence homologies, and studies have reported enterotoxin genes in up to 80% of all S. aureus strains (4, 21). Although many new enterotoxins have been identified, i.e., seg ser and seu (33, 37, 44, 49), their precise functions have not been characterized yet. The majority of experimental work with SEs is still done with SEB, toxic shock syndrome toxin 1, and SEA (27, 31), because these toxins are commercially available. Most SEs are located on mobile elements in bacterial genomes such as plasmids or pathogenicity islands and can thus be easily transferred horizontally between strains (5, 34, 35). Certain SE genes are grouped together. For instance seg, sei, sem, sen, and seo are commonly found in a gene cluster (egc) on genomic island νSAβ (34), and sel and sek are often found together with seb or sec on S. aureus pathogenicity islands. Other staphylococcal superantigen genes are encoded on plasmids (sed, sej, and ser) or are linked to the antibiotic resistance cassette SCCmec (seh) (44, 55). Phage φ3 carries either sea (strain Mu50), sep (N315), or sea sek seq (MW2) (1, 29).Although a few clinical studies have attempted to correlate shock and outcome with the presence of certain SEs in patients with S. aureus infections (17, 28), the contribution of these toxins to outcome is still unclear. Recent papers have proposed the SEs are immunomodulators and that colonization with S. aureus strains that produce SEB may contribute to the pathogenesis of asthma, chronic rhinitis, and dermatitis (2, 36, 46, 48, 56). The superantigen function of SEs in supernatants of S. aureus cultures can be neutralized by serum of colonized patients (21, 23). With new data emerging implicating SEs in the pathogenesis of chronic allergic syndromes, production of monoclonal antibodies and or vaccine strategies targeting SEs may be considered (6, 24, 26, 30) in the future. It is therefore important to characterize the prevalence of SE genes in clinical S. aureus strains.In this study, we analyzed SE content in both methicillin-resistant S. aureus (MRSA) and methicillin-sensitive S. aureus (MSSA) strains that were cultured from wounds (including USA300) and bloodstream infections of patients from a defined geographical area. In addition, SEB production was quantified by enzyme-linked immunosorbent assay (ELISA) in S. aureus strains carrying the seb gene, and spa typing confirmed clonal diversity among S. aureus isolates from different patients, as well as clonal stability in serial isolates, and multilocus sequence typing (MLST) done on a subset of less common spa types. We conclude that SE genes are abundant in S. aureus strains, albeit less abundant in USA300. SE content and combination are highly diverse and therefore more discriminatory than pulsed-field gel electrophoresis (PFGE) and MLST typing, albeit stable in serial isolates. Quantification of SEB production demonstrates that enterotoxin secretion can vary greatly among strains, even if they belong to the same S. aureus lineage. Given the complexities of SE prevalence, regulation, and possible function, we propose that the association of these toxins with chronic allergic diseases or outcome may be oversimplified at present. Precise characterizations of SE function and secretion patterns in individual S. aureus clones are warranted.  相似文献   

8.
Third-generation cephalosporin resistance of Salmonella and commensal Escherichia coli isolates from cattle in the United States is predominantly conferred by the cephamycinase CMY-2, which inactivates β-lactam antimicrobial drugs used to treat a wide variety of infections, including pediatric salmonellosis. The emergence and dissemination of blaCMY-2--bearing plasmids followed and may in part be the result of selection pressure imposed by the widespread utilization of ceftiofur, a third-generation veterinary cephalosporin. This study assessed the potential effects of ceftiofur on blaCMY-2 transfer and dissemination by (i) an in vivo experimental study in which calves were inoculated with competent blaCMY-2-bearing plasmid donors and susceptible recipients and then subjected to ceftiofur selection and (ii) an observational study to determine whether ceftiofur use in dairy herds is associated with the occurrence and frequency of cephalosporin resistance in Salmonella and commensal E. coli. The first study revealed blaCMY-2 plasmid transfer in both ceftiofur-treated and untreated calves but detected no enhancement of plasmid transfer associated with ceftiofur treatment. The second study detected no association (P = 0.22) between ceftiofur use and either the occurrence of ceftiofur-resistant salmonellosis or the frequency of cephalosporin resistance in commensal E. coli. However, herds with a history of salmonellosis (including both ceftiofur-resistant and ceftiofur-susceptible Salmonella isolates) used more ceftiofur than herds with no history of salmonellosis (P = 0.03) These findings fail to support a major role for ceftiofur use in the maintenance and dissemination of blaCMY-2-bearing plasmid mediated cephalosporin resistance in commensal E. coli and in pathogenic Salmonella in these dairy cattle populations.The major mechanism of third-generation cephalosporin resistance among U.S. human and veterinary clinical isolates of Salmonella enterica subsp. enterica is the beta-lactamase CMY-2 (12, 17, 43, 44, 46). blaCMY-2, which likely originated from the chromosomal AmpC locus of Citrobacter freundii, is disseminated among a group of similar plasmids harbored by diverse Enterobacteriaceae species (1, 2, 20, 26, 30, 31, 42, 45). In Salmonella, blaCMY-2-bearing plasmids have been observed in more than 30 serovars, notably including serovar Newport, which has gained specific attention from public health officials as a rapidly emerging threat (2, 6, 31).Commensal Escherichia coli frequently harbors blaCMY-2-bearing plasmids (15, 33, 44), and these plasmids may be transferable to pathogens, since blaCMY-2 plasmids isolated from E. coli and S. enterica share extensive sequence similarity in addition to the blaCMY-2 open reading frame (5, 12, 42, 44). This transfer may occur in the gastrointestinal tracts of cattle, where these bacterial species periodically coexist and where transconjugants may be subjected to specific antimicrobial selection pressure. In fact, in vivo transfer of blaCMY-2 in the gastrointestinal tract has been reported between a Klebsiella pneumoniae blaCMY-2 plasmid donor and a Salmonella enterica serovar Typhimurium isolate in cattle and goats (29).Ceftiofur is the only third-generation cephalosporin antimicrobial drug that is used in cattle production systems and is labeled for the treatment of pneumonia, postpartum metritis, necrotizing pododermatitis, and mastitis. Two ceftiofur preparations, ceftiofur sodium (Naxcel) and ceftiofur hydrochloride (Excenel) (Pfizer Animal Health, New York, NY), are unique in the veterinary pharmacopeia because they require no withholding and discard of milk collected from treated cows, making them frequent therapeutic choices in lactating animals (19, 35). Ceftiofur was licensed in 1988 (41) and its resistance in Salmonella spp. isolated from U.S. cattle, presumably conferred by blaCMY-2, was first documented in 1998 (6).The effects of ceftiofur use on selection of blaCMY-2-bearing commensal E. coli has been examined for cattle both epidemiologically and experimentally. Tragesser et al. studied 18 Ohio dairy herds and determined that the 11 herds that used ceftiofur in any capacity (labeled indications and/or extralabel use) were 25 times more likely to have E. coli with reduced susceptibility to ceftriaxone (an expected blaCMY-2 phenotype) than the seven herds that reported no ceftiofur use (40). Interestingly, however, within eight herds that had detailed treatment records, no association was detected between the prevalence of E. coli with reduced susceptibility to ceftriaxone and use of ceftiofur on an individual-animal basis (40). In an experimental study by Jiang et al., ceftiofur administered to dairy calves was correlated with a 14% increase in ceftriaxone-resistant fecal E. coli compared to untreated controls (21). Together, these studies show a correlation between selection pressure within the gastrointestinal tracts at the individual-animal level and show that ceftiofur use may promote the dissemination of resistance in commensal E. coli at the whole-herd level.Whether or not ceftiofur treatment directly affects in vivo horizontal transfer of blaCMY-2-bearing elements among E. coli and Salmonella has yet to be addressed. The diversity of blaCMY-2 plasmid-bearing bacterial hosts is consistent with wide dissemination of this genetic element. One hypothesis that could explain this wide dissemination is that ceftiofur may itself promote the in vivo horizontal transfer of blaCMY-2-bearing plasmids. Specifically, due to the relatively slow bactericidal activity of aminothiazolyl cephalosporins such as ceftiofur, it has been suggested that exposure to these compounds promotes filament formation in gram-negative bacteria prior to cell death that may increase the surface area and increase receptiveness of the cells for resistance plasmids (11).Because blaCMY-2 may be disseminated by horizontal transfer of R plasmids and/or clonal expansion of individual strains, we examined the effect of ceftiofur use on these processes with two approaches; the first approach specifically considered the issue of horizontal transfer in an experimental in vivo calf model, while the second approach, a field study, assessed the overall relationship between ceftiofur use and blaCMY-2 prevalence in the primary agricultural animal niche where it is used.  相似文献   

9.
FimH is an adhesive subunit of type 1 fimbriae expressed by different enterobacterial species. The enteric bacterium Klebsiella pneumoniae is an environmental organism that is also a frequent cause of sepsis, urinary tract infection (UTI), and liver abscess. Type 1 fimbriae have been shown to be critical for the ability of K. pneumoniae to cause UTI in a murine model. We show here that the K. pneumoniae fimH gene is found in 90% of strains from various environmental and clinical sources. The fimH alleles exhibit relatively low nucleotide and structural diversity but are prone to frequent horizontal-transfer events between different bacterial clones. Addition of the fimH locus to multiple-locus sequence typing significantly improved the resolution of the clonal structure of pathogenic strains, including the K1 encapsulated liver isolates. In addition, the K. pneumoniae FimH protein is targeted by adaptive point mutations, though not to the same extent as FimH from uropathogenic Escherichia coli or TonB from the same K. pneumoniae strains. Such adaptive mutations include a single amino acid deletion from the signal peptide that might affect the length of the fimbrial rod by affecting FimH translocation into the periplasm. Another FimH mutation (S62A) occurred in the course of endemic circulation of a nosocomial uropathogenic clone of K. pneumoniae. This mutation is identical to one found in a highly virulent uropathogenic strain of E. coli, suggesting that the FimH mutations are pathoadaptive in nature. Considering the abundance of type 1 fimbriae in Enterobacteriaceae, our present finding that fimH genes are subject to adaptive microevolution substantiates the importance of type 1 fimbria-mediated adhesion in K. pneumoniae.Klebsiella pneumoniae is recognized as an important opportunistic pathogen that frequently causes urinary tract infections (UTI), septicemia, or pneumonia, particularly in immunocompromised individuals (25). K. pneumoniae is responsible for up to 10% of all nosocomial bacterial infections (12, 35). In recent years, a high incidence of community-acquired K. pneumoniae pyogenic liver abscess with a high mortality rate has been reported, especially from Taiwan, but also from other Asian countries, Europe, and North America (6, 8, 19, 27, 44). Furthermore, 15% to 30% of K. pneumoniae isolates are resistant to broad-spectrum cephalosporins via plasmid-encoded extended-spectrum β-lactamases (5).In contrast to many other bacterial pathogens, K. pneumoniae is ubiquitous in nature. Its nonclinical habitats include environmental locations, such as vegetation, soil, and surface waters, as well as transient commensal colonization of mucosal surfaces in humans and other animals (1). Several studies have reported K. pneumoniae isolates of environmental origin to be nearly identical to clinical isolates with respect to several phenotypic properties (16, 22, 23, 25, 30). It has been suggested that environmental isolates of K. pneumoniae may be as virulent as clinical isolates (24, 39).Several virulence factors have been identified in K. pneumoniae (25, 38). The prominent polysaccharide capsule expressed by most isolates, together with the lipopolysaccharide layer, protects the bacteria against phagocytosis and the bactericidal activity of serum. Fimbrial adhesins expressed by the bacteria are protein structures able to recognize molecular receptors and to facilitate adherence to specific tissue surfaces in the host. K. pneumoniae produces two major fimbrial adhesion organelles, type 1 and type 3 fimbriae (9). Type 1 fimbriae have mannose-sensitive hemagglutinins, while type 3 fimbriae have mannose-resistant hemagglutinins (21).Type 1 fimbriae are the most common adhesive organelle in Enterobacteriaceae and have been most extensively studied in Escherichia coli. The type 1 fimbrial structures of K. pneumoniae are homologous to those of E. coli with regard to genetic composition and regulation (37). Type 1 fimbriae and the adhesive subunit FimH, in particular, play an important role in UTI caused by both K. pneumoniae and E. coli (3, 15, 17, 30, 37). Analysis of E. coli fimH variation at the population level has revealed that the FimH adhesin in urinary E. coli isolates accumulates amino acid replacements that increase its tropism toward the uroepithelium and various components of basement membranes (14, 26, 31, 33, 46). Most of the replacements increase the monomannose binding capability of FimH under low shear by altering allosteric catch bond properties of the protein (40). The natural FimH mutants were shown to provide an advantage in colonization of the urinary tract in a mouse model (32) and correlate with the overall extraintestinal virulence of E. coli (11). Thus, FimH mutations are pathoadaptive in nature. No such population-wide analysis has been performed for K. pneumoniae fimH.Population genetic analysis involves comparison of the nucleotide and structural variability of the locus of interest across multiple bacterial strains of different clonalities and geographic origins. The clonal structure of the strains can be determined by multiple-locus sequence typing (MLST), in which 400- to 500-bp sequences of multiple genetically unlinked loci are determined in order to define the phylogenetic relationship of the strains and the extent of interclonal gene recombination (horizontal gene transfer). MLST has been used to reveal the epidemiological relationship of ceftazidime- and ciprofloxacin-resistant K. pneumoniae isolates of nosocomial origin (4). In addition, the analysis of gene variability enables the determination of the type of selection processes acting on loci of interest, with possible identification of mutational changes of functional significance that could enhance the organism''s ability to cause disease, i.e., that could be of a pathoadaptive nature.In this study, the population dynamics of the K. pneumoniae FimH adhesin were determined by analysis of fimH allelic diversity in strains of environmental and various clinical origins in the context of K. pneumoniae clonal structure based on the allelic diversity of three loci—tonB, mdh and fumC—commonly used for MLST.  相似文献   

10.
The diversity of the Escherichia coli species is in part due to the large number of mobile genetic elements that are exchanged between strains. We report here the identification of a new integrative and conjugative element (ICE) of the pKLC102/PAGI-2 family located downstream of the tRNA gene pheU in the E. coli strain BEN374. Indeed, this new region, which we called ICEEc2, can be transferred by conjugation from strain BEN374 to the E. coli strain C600. We were also able to transfer this region into a Salmonella enterica serovar Typhimurium strain and into a Yersinia pseudotuberculosis strain. This transfer was then followed by the integration of ICEEc2 into the host chromosome downstream of a phe tRNA gene. Our data indicated that this transfer involved a set of three genes encoding DNA mobility enzymes and a type IV pilus encoded by genes present on ICEEc2. Given the wide distribution of members of this family, these mobile genetic elements are likely to play an important role in the diversification of bacteria.The fantastic diversity of the Escherichia coli species has been known for a long time. With modern sequencing strategies, the molecular bases of this diversity are now being unraveled (49). Analyzing the genome of 20 E. coli strains, Touchon et al. recently showed that only a minority of genes, approximately 1,900 genes, were shared by all E. coli strains and constituted the core genome of the E. coli species (50). Additionally, the total number of genes found in all E. coli strains, the pan-genome, is an order of magnitude larger than this core genome (50). The non-core genome of a strain, also called flexible gene pool, is therefore made of a wide diversity of genes. This genetic diversity of the E. coli species translates into a diversity of phenotypic properties. While most E. coli strains are commensal of the gastrointestinal tract of humans and warm-blooded animals, a significant number are responsible for different diseases in humans and animals (22), including extraintestinal infections in chickens; strains isolated from such cases are designated by the term APEC for avian pathogenic E. coli (10).This diversity arises from frequent horizontal gene transfers of mobile genetic elements such as transposons, plasmids, phages, genomic islands, or integrative and conjugative elements (ICEs) (11, 21, 34). Among these mobile genetic elements, ICEs have a particular place as they share properties with both plasmids, genomic islands, and transposons; they can be defined as elements that encode all the necessary machineries that allow their excision from the chromosome, their transfer to a recipient strain, and their integration into the recipient strain''s genome (5, 6, 46, 54). Well-known representatives of this class of genetic elements include Tn916 discovered in Enterococcus faecalis, the conjugative transposon CTnDOT in Bacteroides thetaiotaomicron, ICEKp1 in Klebsiella pneumoniae, SXT/R391-related elements, PFGI-1 in Pseudomonas fluorescens, and the clc element in Pseudomonas sp. strain B13 as well as ICEBs1 in Bacillus subtilis and ICEEc1 in the E. coli strain ECOR31 (1, 39, 44, 46, 54). Typically, ICEs contain at least three modules that are required for key steps in the ICE''s life cycle: an excision/integration module, a transfer module, and a regulation module (54). Besides these, ICEs often contain cargo regions that confer on their host a diverse array of properties, such as virulence properties (ICEEc1), antibiotic resistance (SXT), or degradation of chemical compounds (clc). Because of their self-transfer abilities and their diverse accessory gene repertoires, ICEs are very likely to play a major role in bacteria evolution (46).A new family of ICEs has recently gained interest and was named the pKLC102/PAGI-2 family. The first element of this family, the clc element, was discovered in Pseudomonas sp. strain B13 and confers on the bacteria the possibility to degrade aromatic compounds (42). The transfer of this element was discovered long before its complete sequence was characterized (16). Other members of this family include several elements present in Pseudomonas strains such as PAGI-1 and PAGI-2 as well as the pKLC102 element first considered to be a plasmid but later on shown to be an ICE because of its ability to integrate into the chromosome of its host (23, 52). pKLC102/PAGI-2 elements share a set of core genes (33) and, like most ICEs and genomic islands, are all integrated downstream of tRNA genes (26, 52). The transfer between strains has been demonstrated, albeit with different frequencies, for only a few members, such as the clc element, Pseudomonas aeruginosa pathogenicity island 1 (PAPI-1), and ICEHin1056 from Haemophilus influenzae (20, 37, 41); this transfer involves the type IV pilus (20), the integrase (40), and in some cases the formation of a circular intermediate of the excised ICE (24).In order to identify new accessory genes of APEC strains, we previously described tRNA loci in the E. coli genome that could represent potential insertion sites for new genomic islands (18). We had already used this strategy to characterize the AGI-3 region that is involved in the virulence of an avian pathogenic E. coli strain and that confers the ability to grow on fructooligosaccharides (7, 43). During this tRNA screening, we showed that genomic islands might potentially be present downstream of the tRNA genes argW, leuX, pheU, pheV, selC, serU, and thrW in several APEC strains.In this report, we describe the identification of a new genomic island located downstream of pheU in the APEC strain BEN374. This region, which we named ICEEc2, was fully sequenced, and its properties were analyzed in detail; ICEEc2 is a new ICE found in E. coli and belongs to the pKLC102/PAGI-2 family described above.  相似文献   

11.
Disulfide bond (DSB) formation is catalyzed by disulfide bond proteins and is critical for the proper folding and functioning of secreted and membrane-associated bacterial proteins. Uropathogenic Escherichia coli (UPEC) strains possess two paralogous disulfide bond systems: the well-characterized DsbAB system and the recently described DsbLI system. In the DsbAB system, the highly oxidizing DsbA protein introduces disulfide bonds into unfolded polypeptides by donating its redox-active disulfide and is in turn reoxidized by DsbB. DsbA has broad substrate specificity and reacts readily with reduced unfolded proteins entering the periplasm. The DsbLI system also comprises a functional redox pair; however, DsbL catalyzes the specific oxidative folding of the large periplasmic enzyme arylsulfate sulfotransferase (ASST). In this study, we characterized the DsbLI system of the prototypic UPEC strain CFT073 and examined the contributions of the DsbAB and DsbLI systems to the production of functional flagella as well as type 1 and P fimbriae. The DsbLI system was able to catalyze disulfide bond formation in several well-defined DsbA targets when provided in trans on a multicopy plasmid. In a mouse urinary tract infection model, the isogenic dsbAB deletion mutant of CFT073 was severely attenuated, while deletion of dsbLI or assT did not affect colonization.Disulfide bonds bridging cysteine pairs impart structural stability and protease resistance to secreted and membrane-associated proteins. Most organisms contain specific mechanisms for the formation of disulfide bonds in proteins, a process called oxidative protein folding. In bacteria, this folding process is catalyzed by the disulfide bond family of proteins (18, 22). The best-characterized bacterial disulfide bond machinery is the Escherichia coli K-12 oxidative system, which consists of two enzymes, the periplasmic DsbA and the inner-membrane DsbB (25, 35). DsbA is a monomeric protein comprising a thioredoxin (TRX) domain with an embedded helical insertion and a redox-active CPHC motif (34). This highly oxidizing protein introduces disulfide bonds into unfolded polypeptides by donating its redox-active disulfide (2, 4, 5), and as a result, the two cysteines contained in the CPHC catalytic motif become reduced. DsbB reoxidizes this cysteine pair and restores the oxidizing activity of DsbA, enabling it to assist the folding of a new substrate protein (21).The DsbAB oxidative protein folding system plays a well-documented part in bacterial virulence. Several studies have demonstrated a direct role for both enzymes, particularly DsbA, in the biogenesis of virulence factors utilized by bacterial pathogens in various stages of the infection process (19). The protein forming the P-ring of E. coli flagella, FlgI, was one of the first DsbA substrates identified (10) and flagellum-mediated motility was subsequently demonstrated to require the presence of functional DsbA in several gram-negative pathogens, including Salmonella enterica (1), Proteus mirabilis (8), Erwinia carotovora subsp. atroseptica (9), Burkholderia cepacia (17), and Campylobacter jejuni (42). In Yersinia pestis, S. enterica, Shigella flexneri, and enteropathogenic E. coli, deletion of dsbA results in defective type III secretion, a major virulence mechanism employed by these enteric pathogens to manipulate the host during infection. The defect was shown in each case to involve the outer membrane secretin (YscC, SpiA, Spa32, and EscC, respectively), which requires a single intramolecular disulfide bond to adopt a functional conformation (23, 36, 37, 49). Fimbria-mediated adhesion is a crucial first step of the infection process as it allows host colonization by mucosal pathogens. DsbA is required for functional assembly of several types of fimbriae, including P fimbriae of uropathogenic E. coli (UPEC) (24), bundle-forming pili (Bfp) of enteropathogenic E. coli (55), mannose-resistant Proteus-like (MR/P) fimbriae of Proteus mirabilis (8), plasmid-encoded fimbriae (Pef) of Salmonella enterica (6), type IV pili of Neisseria meningitidis (47), and toxin-coregulated pili (Tcp) of Vibrio cholerae (41). A number of studies have reported that dsbA and/or dsbB mutants are attenuated in infection models (9, 16, 41, 48, 52).The recent exponential increase in sequenced genomes has offered a first glimpse at the diversity of disulfide bond systems present in bacteria (13). In addition, it is now evident that several bacterial species encode multiple DsbA paralogues, often with demonstrated differences in substrate specificity. Neisseria meningitidis, for example, encodes three DsbA oxidoreductases: two inner membrane-associated lipoproteins (DsbA1 and DsbA2) and one periplasmic enzyme (DsbA3). While redundancy was observed in the oxidative folding of virulence-associated proteins by DsbA1 and DsbA2, DsbA3 alone was unable to restore important meningococcal virulence traits, such as type IV pilus-mediated adhesion to human endothelial cells (47). Recently, a second E. coli disulfide bond system (DsbLI) was identified in the genome-sequenced UPEC strain CFT073 and was demonstrated to be a functional paralogue of the prototypic DsbAB system (14). The oxidoreductase DsbL has the strongest oxidizing potential of all DsbA homologues characterized to date. Although the crystal structure of DsbL revealed a similar overall fold and domain architecture to DsbA, DsbL contains a longer helical insertion and deletions in the TRX domain that result in a truncated peptide binding groove. Moreover, DsbL shows different surface properties, including a distinct basic patch around the active site, which was suggested to allow stricter substrate specificity than the highly hydrophobic surface surrounding the active site of DsbA. Grimshaw and colleagues (14) demonstrated the specificity of the DsbLI system for the periplasmic enzyme arylsulfate sulfotransferase (ASST) encoded by assT, a gene found immediately upstream of dsbL and dsbI on the CFT073 chromosome. ASST belongs to a group of poorly characterized large bacterial ASSTs that are proposed to mediate detoxification of phenolic substances by catalyzing the transfer of sulfuryl groups from phenolic sulfates to phenol (26-28, 30). A reason for the specificity of DsbLI for ASST folding could be the presence of an allosteric disulfide bond, recently revealed by the enzyme''s crystal structure (33). This class of disulfide bond forms between Cα atoms of cysteines in unusually close proximity (3.8 Å in the case of ASST) and has higher steric strain energy than catalytic or structural disulfide bonds, thus explaining the requirement for the stronger DsbL oxidase for its formation (33). The activity of DsbL and DsbI was studied using plasmids introduced into E. coli K-12 strains with the native DsbAB system deleted. As yet, the role of the DsbLI system in UPEC virulence has not been investigated.E. coli CFT073 is a prototypic UPEC strain isolated from a female patient with acute pyelonephritis (38). UPEC strains are the causative agent of >80% of community-acquired urinary tract infections (UTIs) and >30% of nosocomial infections (7). The uropathogenic lifestyle of UPEC CFT073 is reflected in its genome, which contains several factors with an established role in urovirulence, including the well-studied type 1 and P fimbriae (50). Genomic comparison of CFT073—and other recently sequenced UPEC strains—with E. coli strains with distinct lifestyles (gut commensals, enteric pathogens, and avian pathogens) allows the discovery of genes unique to genomes of uropathogenic bacteria that are potentially novel urovirulence factors. One such UPEC-specific gene is assT, the gene located upstream of dsbL and dsbI in the chromosome of CFT073 (32).Here we characterize the DsbLI system in its native genetic background of UPEC CFT073 and compare and contrast the contribution of each of the two paralogous disulfide bond systems of CFT073 in the production of UPEC-associated virulence factors and in vivo uropathogenesis. Using isogenic dsbAB and dsbLI deletion mutants of CFT073, we demonstrate that the recently identified DsbLI oxidative protein folding machinery of UPEC CFT073 plays a secondary role in the production of urovirulence factors and does not appear to contribute to virulence in the mouse infection model used in this study. We also show that in the same infection model, an isogenic assT deletion mutant of CFT073 is not attenuated.  相似文献   

12.
Rapid and specific detection of Shiga toxin-producing Escherichia coli (STEC) strains with a high level of virulence for humans has become a priority for public health authorities. This study reports on the development of a low-density macroarray for simultaneously testing the genes stx1, stx2, eae, and ehxA and six different nle genes issued from genomic islands OI-122 (ent, nleB, and nleE) and OI-71 (nleF, nleH1-2, and nleA). Various strains of E. coli isolated from the environment, food, animals, and healthy children have been compared with clinical isolates of various seropathotypes. The eae gene was detected in all enteropathogenic E. coli (EPEC) strains as well as in enterohemorrhagic E. coli (EHEC) strains, except in EHEC O91:H21 and EHEC O113:H21. The gene ehxA was more prevalent in EHEC (90%) than in STEC (42.66%) strains, in which it was unequally distributed. The nle genes were detected only in some EPEC and EHEC strains but with various distributions, showing that nle genes are strain and/or serotype specific, probably reflecting adaptation of the strains to different hosts or environmental niches. One characteristic nle gene distribution in EHEC O157:[H7], O111:[H8], O26:[H11], O103:H25, O118:[H16], O121:[H19], O5:H−, O55:H7, O123:H11, O172:H25, and O165:H25 was ent/espL2, nleB, nleE, nleF, nleH1-2, nleA. (Brackets indicate genotyping of the flic or rfb genes.) A second nle pattern (ent/espL2, nleB, nleE, nleH1-2) was characteristic of EHEC O103:H2, O145:[H28], O45:H2, and O15:H2. The presence of eae, ent/espL2, nleB, nleE, and nleH1-2 genes is a clear signature of STEC strains with high virulence for humans.Since the early 1980s, Shiga toxin-producing Escherichia coli (STEC) has emerged as a major cause of food-borne infections (17, 30). STEC can cause diarrhea in humans, and some STEC strains may cause life-threatening diseases, such as hemorrhagic colitis (HC) and hemolytic uremic syndrome (HUS). On the basis of its human pathogenicity, this subset of STEC strains was also designated enterohemorrhagic E. coli (EHEC) (22, 25). Numerous cases of HC and HUS have been attributed to EHEC serotype O157:H7 strains, but it has now been recognized that other serotypes of STEC belong to the EHEC group. The STEC seropathotype classification is based upon the serotype association with human epidemics, HUS, and diarrhea and has been developed as a tool to assess the clinical and public health risks associated with non-O157 EHEC and STEC strains (18). Only a few serotypes of STEC have been reported as most frequently associated with severe disease in humans. Besides E. coli O157:[H7], five other serotypes, namely O26:[H11], O103:H2, O111:[H8], O121:[H19], and O145:[H28], account for the group of typical EHEC (25). (Brackets indicate genotyping of the flic or rfb genes; the absence of brackets indicates data obtained with the conventional serotyping approach using specific antisera, as described in Materials and Methods.) Atypical EHEC group strains of serotypes O91:[H21], O113:H21, and O104:H21 are less frequently involved in hemorrhagic diseases than typical EHEC but are a frequent cause of diarrhea (8, 12, 25). Recent data from Enter-Net, a global surveillance consortium of 35 countries that tracks enteric infectious diseases, showed that the number of human cases of illness caused by non-O157 EHEC increased globally by 60.5% between 2000 and 2005, while at the same time the number of cases linked to EHEC O157 increased by only 13% (1). In the past few years, new serotypes of EHEC that differ from those previously known as typical and atypical EHEC have emerged (6, 8, 23, 24, 31). These EHEC strains were identified as important causes of food-borne infections in humans and were described as “new emerging EHEC.”The production of Shiga toxin (Stx) by EHEC is the primary virulence trait responsible for HUS, but many E. coli non-O157:H7 strains that produce Stx do not cause HUS. Identification of human-virulent STEC by detection of unique stx genes may be misleading, since not all STEC strains are clinically significant for humans (11). Besides the ability to produce one or more types of Shiga toxins, typical EHEC strains harbor a genomic island called the “locus of enterocyte effacement” (LEE). Atypical EHEC strains are negative for the LEE but may carry other factors for colonization of the human intestine (6, 25). The LEE carries genes encoding functions for bacterial colonization of the gut and for destruction of the intestinal mucosa, thus contributing to the disease process (25). The LEE eae gene product intimin is directly involved in the attaching and effacing (A/E) process (37). The LEE includes regulatory elements, a type III secretion system (TTSS), secreted effector proteins, and their cognate chaperon (13, 29). In addition to the intimin, most of the typical EHEC strains harbor the plasmid-borne enterohemolysin (ehxA), which is considered an associated virulence factor (6, 25).A number of other pathogenicity island (PAI) candidates, including O island 122 (OI-122) and O island 71 (OI-71), have been found in EHEC and EPEC strains, but their role in disease is not fully clear. Within the EHEC group, both O157:H7 strains (19, 34) and non-O157 strains (18, 35) present a variable repertoire of virulence determinants, including a collection of non-LEE-encoded effector (nle) genes that encode translocated substrates of the type III secretion system (9, 20). Our objective was to identify type III secreted virulence factors that distinguish EHEC O157 and non-O157 strains constituting a severe risk for human health from STEC strains that are not associated with severe and epidemic disease, a concept called “molecular risk assessment” (MRA) by Coombes et al. (9). Supporting the MRA approach requires the development of diagnostic tests based on multiplex nucleic acid amplification and microfluidics-based detection using standardized platforms applicable in hospital service or public health laboratories. It is now feasible to develop low-density DNA arrays that can be used to examine the gene inventory from isolated strains, offering a genetic bar coding strategy. A recent innovation in this field is the introduction of the GeneSystems PCR technology (5, 36). In this study, we have developed a GeneDisc array designed for simultaneous detection of genes encoding Shiga toxins 1 and 2 (stx1 and stx2), intimins (eae), enterohemolysin (ehxA), and six different nle genes derived from genomic islands OI-71 and OI-122. We focused our efforts on the detection of the OI-122 genes, ent/espL2 (Z4326), nleB (Z4328), and nleE (Z4329), and the OI-71 genes, nleF (Z6020), nleH1-2 (Z6021), and nleA (Z6024). The macroarray presented here was evaluated for its specificity and ability to discriminate between STEC causing serious illness in humans and other E. coli strains.  相似文献   

13.
14.
15.
The purpose of this work was to evaluate the evolutionary history of Campylobacter coli isolates derived from multiple host sources and to use microarray comparative genomic hybridization to assess whether there are particular genes comprising the dispensable portion of the genome that are more commonly associated with certain host species. Genotyping and ClonalFrame analyses of an expanded 16-gene multilocus sequence typing (MLST) data set involving 85 isolates from 4 different hosts species tentatively supported the development of C. coli host-preferred groups and suggested that recombination has played various roles in their diversification; however, geography could not be excluded as a contributing factor underlying the history of some of the groups. Population genetic analyses of the C. coli pubMLST database by use of STRUCTURE suggested that isolates from swine form a relatively homogeneous genetic group, that chicken and human isolates show considerable genetic overlap, that isolates from ducks and wild birds have similarity with environmental water samples and that turkey isolates have a connection with human infection similar to that observed for chickens. Analysis of molecular variance (AMOVA) was performed on these same data and suggested that host species was a significant factor in explaining genetic variation and that macrogeography (North America, Europe, and the United Kingdom) was not. The microarray comparative genomic hybridization data suggested that there were combinations of genes more commonly associated with isolates derived from particular hosts and, combined with the results on evolutionary history, suggest that this is due to a combination of common ancestry in some cases and lateral gene transfer in others.Campylobacter species are a leading bacterial cause of gastroenteritis within the United States and throughout much of the rest of the developed world. According to the CDC, there are an estimated 2 million to 4 million cases of Campylobacter illness each year in the United States (37). Campylobacter jejuni is generally recognized as the predominant cause of campylobacteriosis, responsible for approximately 90% of reported cases, while the majority of the remainder are caused by the closely related sister species Campylobacter coli (27). Not surprisingly, therefore, the majority of research on Campylobacter has centered on C. jejuni, and C. coli is a less studied organism.A multilocus sequence typing (MLST) scheme of C. jejuni was first developed by Dingle et al. (13) on the basis of the genome sequence of C. jejuni NCTC 11168. There have also been a number of studies using the genome sequence data to develop microarrays for gene presence/absence determination across strains of C. jejuni and to identify the core genome components for the species (6, 15, 32, 33, 42, 43, 53, 57). Although C. coli is responsible for fewer food-borne illnesses than C. jejuni, the impact of C. coli is still substantial, and there is also evidence that C. coli may carry higher levels of resistance to some antibiotics (1). C. coli and C. jejuni also tend to differ in their relative prevalences in animal host species and various environmental sources (4, 48, 58), and there is some evidence that both taxa may include groups of host-specific putative ecotype strains (7, 36, 38, 39, 52, 56). At present, there is only a single draft genome sequence available for C. coli, and there are no microarray comparative genomic hybridization data for C. coli strains. Thus, there is no information on intraspecies variability in gene presence/absence in C. coli and how such variability might correlate with host species.The purpose of this work was to develop and apply an expanded 16-locus MLST genotyping scheme to evaluate the evolutionary history of Campylobacter coli isolates derived from multiple host sources and to use microarray comparative genomic hybridization to assess whether there are particular genes comprising the dispensable portion of the genome that are more commonly associated with isolates derived from different host species.  相似文献   

16.
The nanATEK-yhcH, yjhATS, and yjhBC operons in Escherichia coli are coregulated by environmental N-acetylneuraminic acid, the most prevalent sialic acid in nature. Here we show that YjhS (NanS) is a probable 9-O-acetyl N-acetylneuraminic acid esterase required for E. coli to grow on this alternative sialic acid, which is commonly found in mammalian host mucosal sites.The coregulated nanATEK-yhcH, yjhATS, and yjhBC operons involved in sialic acid catabolism in Escherichia coli are thought to be induced by the most common sialic acid, N-acetylneuraminic acid (Neu5Ac), through reversible inactivation of the NanR repressor encoded by nanR mapping immediately upstream of nanA (15, 27, 28; http://vetmed.illinois.edu/path/sialobiology/). Sialic acids are a family of over 40 naturally occurring 9-carbon keto sugar acids found mainly in metazoans of the deuterostome (starfish to human) developmental lineage and in some, mostly pathogenic, bacteria, where sialic acids expressed at the microbial cell surface inhibit host innate immunity (27). By contrast, most bacterial commensals and pathogens catabolize sialic acids as sole carbon and nitrogen sources, indicating exploitation of the sialic acid-rich host mucosal environment by a wide range of species (2, 27, 28). Interestingly, in vivo experimental evidence further indicates that sialic acid catabolism functions directly (nutrition) or indirectly (surface decoration and cell signaling) in host-microbe commensal and pathogenic interactions in organisms such as E. coli, Haemophilus influenzae, Pasteurella multocida, Salmonella enterica serovar Typhi, Streptococcus pneumoniae, Vibrio vulnificus, and Vibrio cholerae (1, 3, 5, 6, 10, 14, 23, 24, 26, 29). The animal species used for these studies include rodent models and natural hosts such as cattle and turkeys. The structural diversity of sialic acids at the terminal positions on glycoconjugates (glycoproteins and glycolipids) of mucosal surfaces of these hosts requires sialidases, acetyl esterases, and probably other enzymes that convert alternative or at least minor sialic acids to the more digestible Neu5Ac form (8, 9). We have previously demonstrated that E. coli has an epicurean propensity for metabolizing alternative sialic acids (30, 31). In the current communication, we show that YjhS is required for growth of E. coli on 9-O-acetyl-N-acetylneuraminic acid (Neu5,9Ac2).Because most sialic acids are bound to other sugars, including other sialic acids, as part of the oligosaccharide chains on glycoconjugates, either microbial or endogenous (host) sialidases (NanH, or N-acylneuraminate hydrolases) are needed to release free sugar, which is then transported by NanT in E. coli (15, 16, 26, 31). Once internalized, sialic acid is cleaved by an nanA-encoded aldolase or lyase to yield the 6-carbon hexosamine, N-acetylmannosamine (ManNAc), and pyruvate, with the latter entering the tricarboxylic acid cycle or gluconeogenesis. ManNAc is converted to its 6-phosphate derivative by a specific kinase encoded by nanK and epimerized by NanE to yield N-acetylglucosamine 6-phosphate, which is converted to fructose 6-phosphate by products of the nag operon (15, 17, 31, 32). The functions of the coregulated yjhS, yjhB, yjhC, and yhcH gene products are unknown but are not required for growth on Neu5Ac (15). However, YjhA (NanC) is an outer membrane porin required for diffusion of Neu5Ac in the absence of the major porins (7), while YjhT (NanM) is a mutarotase that catalyzes the conversion of the alpha sialic acid isomer to the more thermodynamically stable beta form (21). Neither nanC nor nanM is required for growth on Neu5Ac (15), suggesting that yjhS, yjhBC, and yhcH are involved in reactions that convert alternative sialic acids to Neu5Ac (22, 23). YhcH was crystallized and has been suggested to be an isomerase or epimerase involved in processing N-glycolylneuraminic acid (Neu5Gc) (25), but deletion of yhcH did not affect growth on this sialic acid as a sole carbon source (16).Computer-assisted analysis indicated that YjhB is a permease similar to NanT (16) whereas YjhC is a likely oxidoreductase or dehydrogenase. Orthologs of yhcH, nanC, nanM, and yjhBC are found in most bacterial species with intact Neu5Ac utilization systems, while yjhS is confined to E. coli and shigellae, either as part of the chromosomes in these strains or integrated with phages or phage remnants. However, a significant match (E value = 0.0007) was found between YjhS and AxeA in Rhodopirellula baltica, where AxeA is an acetyl xylan esterase (11), suggesting YjhS might be a sialate esterase. We propose that YjhS should be designated NanS to indicate its direct participation in utilization of an alternative sialic acid.  相似文献   

17.
18.
FimH, the adhesive subunit of type 1 fimbriae expressed by many enterobacteria, mediates mannose-sensitive binding to target host cells. At the same time, fine receptor-structural specificities of FimH from different species can be substantially different, affecting bacterial tissue tropism and, as a result, the role of the particular fimbriae in pathogenesis. In this study, we compared functional properties of the FimH proteins from Escherichia coli and Klebsiella pneumoniae, which are both 279 amino acids in length but differ by some ∼15% of residues. We show that K. pneumoniae FimH is unable to mediate adhesion in a monomannose-specific manner via terminally exposed Manα(1-2) residues in N-linked oligosaccharides, which are the structural basis of the tropism of E. coli FimH for uroepithelial cells. However, K. pneumoniae FimH can bind to the terminally exposed Manα(1-3)Manβ(1-4)GlcNAcβ1 trisaccharide, though only in a shear-dependent manner, wherein the binding is marginal at low shear force but enhanced sevenfold under increased shear. A single mutation in the K. pneumoniae FimH, S62A, converts the mode of binding from shear dependent to shear independent. This mutation has occurred naturally in the course of endemic circulation of a nosocomial uropathogenic clone and is identical to a pathogenicity-adaptive mutation found in highly virulent uropathogenic strains of E. coli, in which it also eliminates the dependence of E. coli binding on shear. The shear-dependent binding properties of the K. pneumoniae and E. coli FimH proteins are mediated via an allosteric catch bond mechanism. Thus, despite differences in FimH structure and fine receptor specificity, the shear-dependent nature of FimH-mediated adhesion is highly conserved between bacterial species, supporting its remarkable physiological significance.The most common type of adhesive organelle in the Enterobacteriaceae is the type 1 fimbria, which has been most extensively studied in Escherichia coli. The corresponding structures of Klebsiella pneumoniae are similar to those of E. coli with regard to genetic composition and regulation (15). Type 1 fimbriae are composed primarily of the structural subunit FimA, with minor amounts of three ancillary subunits, FimF, FimG, and the mannose-specific adhesin FimH. The FimH adhesin is an allosteric protein that mediates the catch bond mechanism of adhesion where the binding is increased under increased shear stress (48).It has been demonstrated in E. coli that FimH has two domains, the mannose-binding lectin domain (from amino acid [aa] 1 through 156) and the fimbria-incorporating pilin domain (from aa 160 through 279), connected via a 3-aa-long linker chain (6). A mannose-binding site is located at the top of the lectin domain, at the opposite end from the interdomain linker (17).Several studies have demonstrated that type 1 fimbriae play an important role in E. coli urinary tract infection (UTI) (7, 21, 23, 35). In addition, in urinary E. coli isolates, the FimH adhesin accumulates amino acid replacements which increase tropism for the uroepithelium and various components of basement membranes (21, 30, 35, 37, 49). Most of the replacements increase the monomannose binding capability of FimH under low shear, by altering allosteric catch bond properties of the protein (48). The mutated FimH variants were shown to provide an advantage in colonization of the urinary tract in the mouse model (35) and correlate with the overall extraintestinal virulence of E. coli (16). Thus, FimH mutations are pathoadaptive in nature.Klebsiella pneumoniae is recognized as an important opportunistic pathogen frequently causing UTIs, septicemia, or pneumonia in immunocompromised individuals (29). It is responsible for up to 10% of all nosocomial bacterial infections (18, 41). K. pneumoniae is ubiquitous in nature, and it has been shown that environmental isolates are phenotypically indistinguishable from clinical isolates (22, 26, 27, 29, 33). Furthermore, it has been demonstrated that environmental isolates of K. pneumoniae are as virulent as clinical isolates (28, 45).K. pneumoniae possesses a number of known virulence factors, including a pronounced capsule, type 3 fimbriae, and type 1 fimbriae (29, 44). Type 1 fimbriae produced by K. pneumoniae are described as functionally and structurally similar to type 1 fimbriae from E. coli (25) and have been shown to play a significant role in K. pneumoniae UTI (32, 43).We have previously shown that mature FimH from 54 isolates of K. pneumoniae (isolated from urine, blood, liver, and the environment) is represented by seven protein variants due to point amino acid replacements. (42) When K. pneumoniae FimH was aligned with the FimH of E. coli, they showed ∼85% similarity at the amino acid level. Furthermore, a majority (14 out of 21 isolates) of the K. pneumoniae strains isolated from patients with UTI grouped into a single clonal group based on multilocus sequence typing, but fimH in one isolate in the group differed from the others by a single nucleotide mutation resulting in an amino acid change, serine to alanine, in position 62 (42). The same mutation has been found in FimH of a highly uropathogenic clone of E. coli and significantly increases the adhesin''s ability to adhere to monomannose under low or no shear (19, 39, 50).In this study, we describe the extent and pattern of structural variability of the FimH protein from K. pneumoniae and perform comparative analyses of the functional properties of FimH from both K. pneumonae and E. coli.  相似文献   

19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号