2 LIBRARY Zoo 8 M ichigan State mversity This is to certify that the dissertation entitled GENOMIC ANALYSIS OF PATHOGEN EVOLUTION: VIRULENCE GENE ACQUISITION AND GENETIC EROSION IN ESCHERICHIA COLI presented by Adam Michael Nelson has been accepted towards fulfillment of the requirements for the Ph.D degree in Genetics SAL wow/1 15%” Major Professor’s Signature 5/2? / 5%” Date MSU is an affirmative-action, equal-opportunity employer _._.-.-.---.-.-.-.--o--.-.-.-.-.-.---.-.-.-.-.—.-.-.-.-.-.- ----.--.- PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 5l08 KIProj/Aoc8Pres/ClRC/Dateoueindd GENOMIC ANALYSIS OF PATHOGEN EVOLUTION: VIRULENCE GENE ACQUISITION AND GENETIC EROSION IN ESCHERICHIA COLI By Adam Michael Nelson A DISSERTATION Submitted to Michigan State University ‘ in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Genetics 2008 ABSTRACT GENOMIC ANALYSIS OF PATHOGEN EVOLUTION: VIRULENCE GENE ACQUISITION AND GENETIC EROSION IN ESCHERICHIA COLI By Adam Michael Nelson Escherichia coli is a gram-negative, rod-shaped bacterium that lives naturally and commensally in the intestinal tract of humans and other mammals. However, some types of E. coli have acquired genetic elements encoding virulence factors that contribute to disease. Different disease phenotypes are caused by strains carrying a variable array of virulence factors. Many of these virulence factors are largely acquired through horizontal transfer. Pathogenic E. coli are grouped into at least twelve classes, or pathotypes, based on the type of disease they cause. Observing the acquisition of virulence elements as a means to study evolution has been previously used to determine the genetic ancestry of E. coli and other pathogenic microorganisms. Conversely, gene loss can also be important in enhancing pathogenicity by removing genes that encode proteins that may hinder increased virulence, or are no longer functional because they are mutated or otherwise incomplete. The research presented here is intended to enhance the current understanding of the roles of genetic acquisition and genetic loss to virulence changes in pathogenic E. coli by applying three paths of inquiry. To address correlations between virulence profiles and disease incidence, 392 E. coli isolates from 115 pediatric patients were screened for virulence gene content using fluorescently labeled PCR amplicons in a capillary based sequencing system. Virulence profiles were compared to a phylogenetic framework to determine virulence distribution, and correlations between presence of specific genes and the incidence of disease in patients. The screening of this pediatric population lead to the discovery of a variant of E. coli (sequence type 29), which was found very frequently (19%) in all of the samples examined. This population of ST-29 isolates was further characterized by PCR to determine the frequency of a panel of attachment-related loci, and by RFLP to determine the capsular polysaccharide type. To track genetic erosion events, PCR-based screening of individual components in a pathogenicity island, called ETT2, showed examples of gene loss that rids possibly non-functional genetic material from the genome. When a strain acquires new genetic material by horizontal transfer, the new material must provide some fitness benefit or it will be under selection to be removed. Type III secretion systems are complicated structures that require a large number of genes to encode the proteins necessary for proper assembly. If critical components of the type III assembly are missing, the assembly will not function properly, and is likely to be under selective pressure for deletion. A streamlined genome may result in a more efficient pathogen. Here, the loss of all or portions of ETTZ is shown in a variety of pathogenic isolates of E. coli. At least six different deletion variants were discovered in the 57 strains examined. These results help advance our understanding of evolution in E. coli by both acquisition and loss of virulence elements and further demonstrate the dynamic and diverse nature of the E. coli genome. Copyright By Adam Michael Nelson 2008 ACKNOWLEDGEMENTS I would like to thank Thomas Whittam, my mentor, for his invaluable support and guidance, and for the opportunity to work on a project I really enjoyed during my time at Michigan State University. I also thank my guidance committee members, Drs. Michael Bagdasarian, Linda Mansfield and Vincent Young for their suggestions and support. I have met so many people in our lab who were both friends and helpful co- workers that were crucial to the completion of this research. I wish to thank Galeb Abu- Ali, Dr. Teresa Bergholz, Sivapriya Kailasan, Sara Kienzle, Dr. David Lacher, Albert Lee, Dr. Shannon Manning, Lindsey Ouellette, Dr. Weihong Qi, Dr. James Riordan, Dr. Hans Steinsland, Dr. Cheryl Tarr, Dr. Seth Walk, and Dr. Lukas Wick. I owe a great debt for the support of my family during my graduate career. I wish to thank my parents, Bruce and Sue Nelson, my brothers, Josh and Jordan, my wife, Suzanne, and my daughter Emily for their encouragement, patience and love. TABLE OF CONTENTS viii LIST OF TABLES ........................................................................... LIST OF FIGURES ......................................................................... ix KEY TO ABBREVIATIONS .............................................................. x Chapter 1. Literature Review ............................................................... 1 PATHOGENIC TYPES OF E. coli ............................................... 2 PARALLEL EVOLUTION OF VIRULENCE ................................. 4 EVOLUTION BY HORIZONTAL TRANSFER OF GENETIC MATERIAL .......................................................................... 5 GENE LOSS AND THE EFFECT ON VIRULENCE — GENETIC EROSION OF E. coli TYPE III SECRETION SYSTEM 2 ................... 5 DETERMINATION OF VIRULENCE PROFILES ........................... 8 PURPOSE ............................................................................ 9 HYPOTHESES TO BE TESTED ................................................ 9 Chapter 2. Genetic and virulence characterization of Escherichia coli isolated from pediatric patients in Seattle, WA ..................................................... 12 SUMMARY ........................................................................ 13 INTRODUCTION .................................................................. 14 MATERIALS AND METHODS ................................................ 20 Bacterial Strains and DNA Isolation .................................... 20 PCR Primer Design ....................................................... 21 PCR of MLST Genes ...................................................... 24 Phylogenetic Analyses of Sequence Data .............................. 24 Virulence Genes and Protein Functions ................................ 25 MVGP Procedure Summary ............................................. 27 PCR of MVGP Genes .................................................... 28 Controls for MVGP ....................................................... 28 MVGP Data Confirmation ............................................... 29 RESULTS ........................................................................... 30 Genetic Diversity of MLST Loci ........................................ 30 Sequence Type Diversity and Clonal Groups .......................... 3O MVGP Analysis ............................................................ 33 Comparison of Virulence Gene Profiles and Phylogenetic Relationships ............................................................... 33 DISCUSSION ....................................................................... 39 Diversity of Virulence Gene Profiles ................................... 39 Frequency of ST-29 Isolates in this Sample Set ....................... 40 ACKNOWLEDGEMENTS ....................................................... 42 Vi Chapter 3. Characterization of a common meningitis-associated clone (ST-29) of Escherichia coli found in pediatric patients in Seattle, WA SUMMARY ........................................................................ 43 INTRODUCTION ................................................................. 44 MATERIALS AND METHODS ................................................ 45 PCR Primer Design ....................................................... 49 Strain Growth Conditions and DNA Extractions ..................... 49 MVGP PCR ................................................................ 49 Controls ..................................................................... 49 MLST PCR ................................................................. 49 Phylogenetic Analyses .................................................... 50 PCR for Attachment Loci ................................................ 50 Capsular Typing — PCR ................................................... 50 Capsular Typing — Restriction Digestion .............................. 50 RESULTS ........................................................................... 53 Attachment Genes ......................................................... 55 Capsular Typing .............................................................. 55 DISCUSSION ....................................................................... 57 ACKNOWLEDGEMENTS ....................................................... 59 Chapter 4. Genetic erosion of Escherichia coli type III secretion system 2 (ETT2) in 0157:H7 isolates ............................................................... 60 SUMMARY ........................................................................ 61 INTRODUCTION ................................................................. 62 MATERIALS & METHODS .................................................... 68 PCR Primer Design ....................................................... 68 Strain Growth Conditions and DNA Extractions... .................. 68 PCR ......................................................................... 68 RESULTS ........................................................................... 7O ETT2 in EHEC 0157 ..................................................... 71 ETT2 in NON-0157 STEC .............................................. 71 ETT2 in 055:H7 ........................................................... 71 DISCUSSION ....................................................................... 79 ACKNOWLEDGEMENTS ....................................................... 82 Chapter 5. Summary and Synthesis ........................................................ 83 Future Considerations ...................................................... 85 REFERENCES ............................................................................... 89 vii Table 1. Table 2. Table 3. Table 4. Table 5. Table 6. Table 7. Table 8. Table 9. LIST OF TABLES Patients in the Seattle pediatric study grouped based. on the diagnosis of microorganisms .............................. , .............................. Oligonucleotide primers used in the Seattle pediatric study. . . . . . . . Virulence genes associated with diarrhea] disease ........................ Sequence variation among alleles of seven MLST genes. . . . . . . . . . Summary of virulence gene detection frequency in sequence types and in patients from the Seattle pediatric population study.. . Gene and protein data for attachment-related loci used in PCR assays of ST-29 isolates ........................................................ Comparison of genes from the Salmonella Pathogenicity Island 1 (SPI-l) of Salmonella enterica serovar Typhimurium and their corresponding homologous counterpart in ETT2 ......................... Primer sequences used to amplify individual genes within ETT2. . Results from the PCR assay of individual ETT2 genes within a variety of E. coli serotypes and control strains. ........ ' .................... viii 19 22 26 31 34 52 65 67 73 Figure 1. Figure 2. Figure 3. Figure 4. Figure 5. Figure 6. Figure 7. Figure 8. Figure 9. Figure 10. Figure 11. Figure 12. LIST OF FIGURES Fluorescent chromatogram output from the multilocus virulence gene profiling (MVGP) procedure ........... . ................................ 18 Phylogenetic tree showing the genetic relatedness of pediatric E. coli in Seattle, WA ......................................................... 32 Virulence type diversity in two representative patients ................... 35 Non-metric dimensional scaling (NMDS) analysis of MVGP results of 94 patients based on the presence or absence of 29 known or putative virulence factors ........................................ 38 Organization of the kps operon in E. coli .................................. 47 Representative banding pattern from a Sau96I restriction endonuclease digest of the amplified kpsCM region in E. coli .......... 48 Frequency distribution of each gene from MVGP analysis of 392 isolates from 94 pediatric patients. ......................................... 54 Organization of E TT 2 gene cluster in E. coli .............................. 66 Pairing of the phylogenetic analysis and ETT2 deletion profiles in many pathogenic types of E. coli ............................................ 76 Map of the region of ETT2 that was examined by PCR, and the results of that screen. ......................................................... 77 Stepwise evolution model for the emergence of the modern pathogenic clone of EHEC 0157:H7 from ancestral 055:H7 isolates. 78 The EIP island from enteroaggregative strain 042 ........................ 81 ix KEY TO ABBREVIATIONS: E. coli Pathotypes: EPEC EHEC ETEC EIEC STEC EAggEC Phenotypes: A/E HC HUS Enteropathogenic E. coli - cause watery diarrhea in infants, can form microcolonies and NE phenotype. EPEC contain the LEE island and a large adherence-related plasmid. Enterohemorrhagic E. coli - cause severe bloody diarrhea, HC, developing sometimes into HUS. EHEC encode Shiga toxins, a large virulence plasmid, and the LEE island. Enterotoxigenic E. coli - cause travelers diarrhea, similar to cholera, but not as severe. ETEC can contain heat stable and/or heat labile toxins. Enteroinvasive E. coli - can invade epithelial cells through proteins encoded on a virulence plasmid. EIEC have a similar invasive phenotype to Shigella. Shiga toxin-producing E. coli - any strain possessing one or both of the Shiga toxins (Stxl/Stx2). EHEC are a subset of STEC. Enteroaggregative E. coli - cause mild but prolonged watery diarrhea. Attaching/Effacing phenotype seen in EPEC and EHEC. Characterized by effacement of microvilli, intimate attachment, and pedestal formation due to actin polymerization. Characteristic of strains harboring the LEE island. Hemorrhagic colitis - severe bloody diarrhea caused by EHEC. Hemolytic uremic syndrome - life-threatening disease caused by EHEC - characterized by acute renal failure, the destruction of platelets and destruction of red blood cells. Virulence Factors: sth/stxZ Shiga toxin 1 or Shiga toxin 2 genes - both encoded on integrated bacteriophages in EHEC and STEC — encode cytotoxins which inhibit protein synthesis and cause characteristic HC and HUS symptoms. ETT2 EIP TTSS LEE Effector KEY TO ABBREVIATIONS (cont’d): E. coli type III secretion system 2 - cryptic island, widely distributed, but usually deleted, possibly because it no longer functions as a TTSS. ETT2 may encode structural proteins for typeIII secretion, but likely not effectors. Largely uncharacterized island encoding possible effectors for ETT2. Found in EAggEC strain 042, but not in 0157. Type III secretion system - contact dependent method for directly injecting bacterial proteins into the host cell by use of a needle-like complex that spans both the inner and outer bacterial membranes. Locus of enterocyte effacement - pathogenicity island encoding both structural components and effector molecules for a functional type III secretion system in EHEC and EPEC. The actual proteins injected into the host cell using a TTSS. Effectors are responsible for eliciting the pathological changes seen during A/E. Strains of E. coli: ECOR K12 E. coli reference collection - diverse set of strains of E. coli maintained by Dr. Thomas Whittam at Michigan State University. Domesticated, non-pathogenic strain of E. coli lacking many virulence factors characteristic of pathogenic strains (LEE island, Shiga toxins, etc). Molecular Techniques: MLST MVGP CEQ ORF Multi locus sequence typing - a DNA sequencing technique using sequencing of 450-600 nucleotides of conserved genes to generate phylogenetic comparisons. Multi locus virulence gene profiling - typing system to examine 29 virulence genes & 1 positive control in uncharacterized E. coli isolates. Genetic analysis machine developed by Beckman Coulter for use with DNA sequencing (MLST) and virulence profiling (MVGP). Open reading frame - DNA sequences containing a putative gene. xi CHAPTER 1 LITERATURE REVIEW Commensal strains of E. coli exist naturally in the gut of humans and other mammals without causing disease in healthy individuals. However, a small subset of E. coli have acquired additional genetic elements encoding the production of toxins and other virulence factors, such as invasion ability, secretion systems, and increased iron utilization. Pathogenicity islands are horizontally-acquired genetic elements (plasmids, phages, chromosomal islands) that increase pathogenicity in recipient strains. DNA fragments of various sizes containing genes or groups of genes can be transferred between strains or even between bacterial species, and are a force driving rapid evolution. Pathogenic types of E. coli. At least 12 different classes, or pathotypes, of E. coli have been identified based on distinct disease phenotypes [1, 2], and often contain different pathogenicity islands encoding virulence factors responsible for conferring the different disease phenotypes. This allows grouping of strains based on similar disease- causing ability, rather than serological similarities based On typing of the somatic 0- antigen and flagellar H antigens. So while two strains may share the same serotype (OlO3:H2, for example), they may not necessarily share the same virulence profile. Although the more closely related two strains are genetically, the more likely they will have the same array of virulence factors and cause the same disease. The major pathotypes of E. coli used in this work, and common virulence-related genes found within those pathotypes, will be discussed below. In the gut, enteropathogenic E. coli (EPEC) strains have the ability to form microcolonies through localized adherence by genes encoded on a large virulence plasmid [3]. EPEC also contain numerous pathogenicity islands, including the EspC island, which encodes a serine protease autotransporter toxin (espC) that aids in virulence, [4, 5] and the locus of enterocyte effacement (LEE) island which encodes a type III secretion system and the ability to form attaching-effacing (A/E) lesions on intestinal epithelial cells [6, 7]. NE lesions are characterized by brush border microvilli destruction, intimate attachment to the host cell membrane, and the formation of a pedestal structure in the host cell through the polymerization of actin filaments [6, 8]. Enterohemorrhagic E. coli (EHEC) strains also encode LEE and can elicit A/E lesions as well. Unlike EPEC, they contain Shiga toxins (stxl and/or stx2) and the EHEC virulence plasmid. This plasmid contains hemolysin (ehx) for iron piracy, adhesion factors (toxB, iha) for attachment to the intestinal epithelia, iron utilization genes (chuA) to process iron, and a protease autotransporter (espP) to cause proteolytic damage and inhibit blood coagulation [9, 10]. Enterotoxigenic E. coli (ETEC) strains can encode two distinct toxins responsible for increased fluid accumulation in the gut lumen. The heat-stable (estA) toxin, which binds to and activates guanylate cyclase, causes the increase of chloride ion secretion resulting in the rapid accumulation of intestinal fluid. The heat-labile (elt) toxin causes a disease similar to cholera (heavy fluid and electrolyte accumulation in the gut), because it shares the same structure and mode of action as cholera toxin from Vibrio cholerae [11- 13]. Enteroinvasive E. coli (EIEC) strains cause dysentery and have invasive phenotypes similar to Shigella, possessing a plasmid essential for invasion ability [14, 15]. Enteroaggregative E. coli (EAggEC) are able to adhere to HEp-2 cells in an aggregative or ‘stacked brick’ pattern of autoagglutination. These strains encode some enterotoxins, including a plasmid-encoded serine autotransporter toxin (pet) [16, 17], Shigella enterotoxin 1 (setIA) [18-20], and the EAST heat-stable toxin (astA), which is also found in some EPEC strains [21]. EAggEC can cause mild, but prolonged watery diarrhea lacking fever or blood loss, however symptoms are frequently non-uniform [22- 25]. Shiga-toxin producing E. coli (STEC) are strains grouped by one common factor; they contain one or both of the Shiga toxin genes (sth, stx2) [26-28]. Since sth and stx2 are encoded on separate, mobile bacteriophages, STEC strains are often not closely related, but rather have only acquired and incorporated the 5th and/or stx2-containing phages into their genome. Parallel evolution of virulence. Strains from each pathotype are often not evolutionarily related, but rather have acquired similar pathogenic ability through parallel acquisition of the same or similar virulence factors. Despite the harm caused to the host, selection favors increased virulence when it allows for better spread or transmission of the pathogen [29]. Often better transmission equals more severe disease symptoms. The discovery of a new pathogenic variant of E. coli capable of causing widespread foodbome illness occurred in 1982 [30, 31]. Since then, this pathogenic variant, 0157:H7, has been responsible for an estimated 74,000 infections each year in the US. [32], causing bloody diarrhea and hemolytic uremic syndrome, a life threatening illness characterized by acute renal failure, hemolytic anemia, and reduced ability to form blood clots through the destruction of platelets [28, 33]. Since 0157:H7 strains also contain the LEE island, they have the ability to form the characteristic A/E lesions on epithelial cells, just as seen with EPEC. Evolution by horizontal transfer of genetic material. The genome of E. coli is extremely heterogeneous, having undergone remarkable sequence divergence between the common pathotypes that cause disease. Horizontaltransfer of genetic material is common, with striking variability between 0157:H7 isolates and the commensal K12 strain. Nearly 1400 genes in 0157:H7 are unique compared to K12, including many that encode virulence functions. The genome is nearly 20% larger in 0157:H7 strains vs. K12 (5.5MB vs. 4.6MB) [34]. Other pathotypes are also highly divergent. For example, the recent genome sequencing of the uropathogenic strain CFT073 revealed only 39.2% of the total protein content in CFT073, the laboratory K12 strain MG1655, and EHEC 0157:H7 strain EDL933 is conserved between all 3 strains. CF T073 is as different genetically from the commensal K12 strain as it is from EDL933 [35]. Gene loss and the effect on virulence - Genetic erosion in E. coli type III secretion system 2 (ETT2). The publication of genomic DNA sequences has facilitated rapid comparisons between strains to identify unique genes or pathogenicity islands. EPEC and EHEC encode a well-characterized TTSS called LEE, which is known to be functional [7, 36, 37]. Comparative genomic scans were used to discover a cryptic pathogenicity island named ETT2 (for E. coli Type III Secretion System 2) in 0157:H7 strains [3 8]. ETT2 encodes a putative TTSS in select serotypes of E. coli, including 0157:H7 [39]. Type III secretion systems need two basic sets of genes necessary for proper secretion. The structural genes encode proteins that assemble into a needle complex, with a basal component spanning both the inner and outer bacterial membrane and a needle-like component that forms a molecular syringe that allows for direct transfer of bacterial proteins into the host cell [40]. It is thought that the structural genes for a second TTSS in E. coli 0157 could be encoded in ETT2 [41]. However, the other crucial component needed for TTSS are the effector proteins, which are the injected proteins that travel through the type III needle complex into the host cell. Effectors then elicit pathological changes in the host [40]. Some effectors that work with ETT2 may be encoded on a separate pathogenicity island called EIP. Putative effectors with sequence homology to effectors that are secreted through two distinct TTSS in Salmonella (SPI-l and SPI-2) are found in some strains of E. coli [41]. ETT2 shares sequence homology and genetic organization with the SPI-l TTSS [39]. This suggests EIP-encoded effectors may work with the structural framework encoded by ETT2, since both loci may share a common origin from in Salmonella. In LEE, effectors cause pathological changes, such as actin polymerization and pedestal formation. LEE has both the structural genes and the effectors encoded together in the same island. In contrast, ETT2 and EIP are separate islands and rarely found together [4 1 ]. The secretion framework is an energetically expensive structure to assemble, and likely provides little or no fitness benefit without the accompanying effectors necessary for pathogenic effects on the host cell. Genes encoding structural components are conserved throughout various gram-negative bacterial pathogens encoding TTSS. However, the effector molecules secreted into the host cell are oflen highly divergent and cause distinctly different effects [42, 43]. In EPEC and EHEC, these changes are characterized by massive actin polymerization causing structural rearrangement of the eukaryotic cell surface into a pedestal shape that is essential for a close association between the E. coli and the eukaryotic cell. The intimate association is necessary for the onset of diarrhea] disease characteristic of EPEC and EHEC infection [7, 36, 37]. It has not been demonstrated that ETT2 & EIP confer a similar phenotype. However, ETT2 & EIP may be important disease determinants when found together in the same strain. When they are not together, the numerous deletions observed in ETT2 may be an example of genetic refinement through erosion of factors not enhancing virulence. Ren and colleagues showed a widespread distribution of individual genes or groups of genes in ETT2 [41]. This included parts of ETT2 in K12, which does not cause disease. Despite the diverse distribution, most strains examined had deletions that eliminated major sections of ETT2. Only a few strains contained all of the island, including all 0157 isolates, EAggEC strain 042, and several strains from the E. coli Reference Collection (ECOR), a diverse set of 72 strains assembled from a variety of locations and hosts [44]. All strains showing deletions in ETT2, including K12, are apparently missing the EIP island [41]. ETT2 likely had a distant acquisition point and has been selectively eroded in strains where it was lacking EIP, and was possibly non-functional. A variety of deletions of ETT2 are evident in all strains that do not also encode EIP, except for 0157 strains, which curiously have retained an intact set of structural TTSS components, despite containing several frameshift mutations [41]. This may represent the first stages of genetic erosion, preceding the larger deletions seen in K12 and other isolates. Only a few strains have been identified that contain both ETT2 and EIP. One example is an EAggEC strain called 042. To date, no strains have been identified as encoding only EIP but without ETT2. Conversely, all strains with deletions in ETT2 are missing EIP, except for 0157:H7 isolates, as previously mentioned. Determination of virulence profiles. A large-Scale comprehensive assessment of the prevalence of many virulence genes in clinical samples has not been done for pathogenic E. coli. Existing studies have only examined smaller numbers of specific strains. These studies were not randomized, nor did they examine more than a handful of virulence factors [45-47]. The genetic elements necessary for enhanced virulence in the various pathotypes are not fully understood, despite the importance of these pathogens. In addition, the effect of strain mixtures providing complementing virulence factors to cause disease is poorly understood. Multiple samples from the same patient may show the patient is infected with more than one distinct pathogenic clone of E. coli. Two techniques were used to address these issues: the identification of phylogenetic frameworks and the generation of virulence profiles in each strain. Multi-locus sequence typing (MLST) compares the internal sequences of housekeeping genes (n=7), which are considered to be under stabilizing selection resulting in minimal variation in functionally essential proteins. Phylogenetic trees are constructed by statistical software that compares genetic relatedness of nucleotide sequences. This technique has been used successfully in a variety of microorganisms [48- 51], including E. coli. Multi-locus virulence gene profiling (MVGP) was been developed to rapidly characterize outbreak strains and new serotypes of pathogenic E. coli, and to discover novel virulence profiles in each pathotype. MVGP works by combining fluorescently labeled PCR amplicons with capillary-based sequencing tools to generate chromatograms displaying virulence profiles for individual strains. Each profile identifies individual virulence factors as peaks of defined size and dye label, and provides a fingerprint of the virulence content of each strain. When using samples isolated from patients displaying disease symptoms of unknown etiology, MVGP can help identify virulence factors important for causing or contributing to disease. Purpose. The primary objective of this research is to identify examples of evolution through gene loss and gene acquisition in a variety of genes associated with pathogenesis. In patients with unexplained cases of diarrheal disease, analysis of the E. coli collected from them was screened for a panel of known and putative virulence genes that may have contributed to illness. It is important to note that the presence of a virulence-related gene in an isolate of E. coli is not automatically correlated to either the expression of that gene or the ability of that strain to cause disease symptoms, but it at least makes that strain suspect, and worth investigating further. The phenomenon of gene loss enhancing virulence or genetic fitness has been examined with the erosion of a cryptic pathogenicity island in E. coli called ETT2. This island is believed to contain the structural genes to assemble a type III secretion system. ETT2 is found in 0157:H7 isolates, but has numerous deletions in many other serotypes of E. coli. The numerous deletions may be due to selection acting on a non—functional pathogenicity island to remove unnecessary genetic material and streamline the genome. Hypotheses to be tested. We will test the following hypotheses from chapter 2: 1. There will be a correlation between the number of virulence genes per patient and an increase in disease duration. We will use chi-square tests to determine if there is a correlation between virulence genes (0-5 genes) vs (6 or more) and the duration of illness (0-3 days), (4-7 days), and (8 or more days). 2. When patients are put into two categories, a diagnosis group consisting of all the possible diagnoses (Salmonella, rotavirus, etc) combined, there will be more virulence genes in those patients with a diagnosis than in patients without a diagnosis. We will also use chi-square to test this hypothesis using groups for genes (0-5) and (6 or more) per patient). From chapter 3, we will test these hypotheses: 1. Because the clinical isolate R8218 is a ST-29 strain and has a K1 capsule type, we predict all or nearly all pediatric ST-29 isolates from Seattle will also have the K1 capsule type. A capsule typing method using restriction fragment length polymorphism will be used to determine capsule profiles. 2. We also predict the prevalence of ST-29 isolates is related to an enhanced ability to adhere to the intestine. Therefore we predict we will find some attachment- related genes common in extra-intestinal E. coli to be in very high frequency (near 100%) in the ST-29 strains. A panel of genes encoding proteins related to attachment will be screened by individual pairs of PCR primers. From chapter 4, we will test these hypotheses: l. The region of ETT2 we are examining will be found intact and undeleted in all strains from the ancestral stepwise lineage related to modern 0157:H7 isolates. Strains that represent the stepwise changes from ancestral EPEC-like 055:H7 to the modern EHEC 0157:H7 will be screened by PCR for 11 individual genes within a 17-kb region encoding the structural apparatus of the type III secretion system. 10 2. We predict other non-0157 isolates will have a variety of deletion profiles, including some newly described deletion types. In addition to strains from the 0157 stepwise model, additional strains from other pathogroups will be examined and we predict new deletion profiles unseen in previous studies will be found. 11 CHAPTER 2 GENETIC AND VIRULENCE CHARACTERIZATION OF Escherichia coli ISOLATED FROM PEDIATRIC PATIENTS IN SEATTLE, WA 12 SUMMARY A collection of 392 E. coli isolates were examined for virulence content using multi-locus virulence gene profiling. These strains were taken from children hospitalized with diarrheal disease of unknown cause. More than half of the E. coli strains were from patients where no identifiable pathogenic organism was isolated (Campylobacterjejuni, Clostridium diflicile, rotavirus, Shigella, etc). The E. coli from these patients was examined for the presence of 29 known and putative virulence genes that serve as both general markers for all major pathotypes of E. coli and specific markers for certain pathogenicity islands or plasmids. These strains were also sequenced for 7 housekeeping genes to determine clonal relationships and generate phylogenetic trees. The virulence results from MVGP were paired with the phylogenetic framework generated by MLST to infer patterns of acquisition in specific virulence genes and to study their distribution across different clonal groups of E. coli. 13 INTRODUCTION Escherichia coli are Gram-negative bacteria that include harmless, useful, and harmful organisms within this genospecies [1]. There are a variety of pathogenic variants of E. coli that can not only be group based on sequence similarities, but also by the characteristics of the diseases that they cause. Often virulence factors are frequently or exclusively found in specific pathotypes, and are necessary to cause the specific disease associated with each pathotype [2]. E. coli can frequently acquire new virulence abilities by horizontal transfer of genetic material [34, 35]. Despite the possible role of a number of these putative virulence loci in disease, many have not been extensively studied. Correlations between specific virulence profiles in clinical samples of E. coli and the incidence of disease in patients have not yet been established. In addition to well-defined pathogenic clones, the effect of strain mixtures that provide complementing virulence factors is poorly understood. Analysis of multiple isolates from the same patient might show if a patient is infected With more than one distinct pathogenic clone of E. coli. Two techniques have been used to address these issues: the generation of phylogenetic frameworks by multi-locus sequence typing (MLST) [52-54], and the identification of virulence profiles in each strain by multi-locus virulence gene profiling (MVGP) [47, 53, 55, 56]. MLST is a procedure that involves the sequencing of genes under stabilizing selection to generate phylogenies that show the evolutionary history of the strains under examination. Here, we have used a panel of seven “housekeeping” genes encoding proteins that serve basic biological functions in the cell. Once the sequences are assembled, they are concatenated and analyzed for single nucleotide variations to allow 14 for the generation of phylogenies. Based on the pattern of alleles for each gene, an allele assignment can be made. When all of the allele assignment numbers are put together for each of the genes sequenced, a “barcode” of that particular strain is generated. The barcode represents a particular sequence type (ST), which is unique. So when two strains share the same sequence type, every nucleotide sequenced is exactly the same in all seven genes examined. MLST has been used repeatedly to characterize E. coli populations [56- 58]. One major advantage of this system is that the data is portable and can easily be compared with results from other laboratories, provided the same primers were used to sequence the same genes in both labs. MVGP uses similar reagents and equipment as MLST, but generates data in a different manner. It works by combining fluorescently labeled PCR amplicons with capillary-based sequencing tools to generate chromatograms displaying virulence profiles for individual strains. This technique allows rapid generation of the virulence profile of bacterial isolates. Each profile identifies individual virulence factors as peaks of defined size and dye label, and “fingerprints” the virulence content of each strain. When using isolates from patients with illnesses of unknown etiology, MVGP can help identify virulence factors that may be important for causing or contributing to disease. This technique also can be applied to any set of strains of E. coli and can be customized by adding or removing virulence loci, so newly identified virulence genes can be incorporated, if desired. MVGP has already been used to characterize a number of populations of E. coli, including clinical STEC isolates of serotype 0121:H19 [47], E. albertii, which is closely related to Shigella boydii type 13 [59] , and E. coli serogroups 0174 [60]. 15 The MVGP procedure consists of the following steps: 1) collection of bacterial strains for study; 2) extraction of DNA; 3) PCR using one regular and one Beckman dye- labeled primer; 4) pooling of labeled post-PCR products together; 5) running the sample on a Beckman CEQ sequencing machine; and 6) analysis of the results to determine virulence profiles. The Beckman dyes are available in 3 colors: D2 (black), D3 (green), and D4 (blue). The red dye (D1) is reserved for a size standard and should not be used to label PCR products. Post-PCR amplified product can be pooled together (we used 14 and 16 genes in two sets). The data output is shown as a chromatogram with peaks of ' increased fluorescent intensity at a specific size representing each amplicon (Figure 1). Here we used MLST and MVGP to examine 392 pediatric E. coli from patients with diarrheal disease of unknown cause. Samples were collected for this study at the University of Washington and Children’s Hospital & Regional Medical Center in Seattle, WA. The E. coli isolates were collected from stool samples from patients from a number of categories based on the identification of other microorganismsi(Table 1). Up to five isolates from each patient were taken for comparison to improve the statistical power of the results, and to determine if individual patients have mixtures of strains with distinct, and possibly synergistic, virulence profiles. So, for example, in patients diagnosed with specific pathogens such as Salmonella, five E. coli isolates were also collected from those patients for use in this study. Each isolate was confirmed to be E. coli by PCR testing for the presence of the uidA gene, which encodes beta—glucuronidase, an indicator of E. coli. To blind the study from any potential bias, all samples were identified only by a unique number. It was not until after typing was finished that epidemiological info (diagnosis, illness duration, fever, vomiting, and bloody diarrhea) for specific strains was revealed. 16 We hypothesized that there would be a correlation between the number of virulence genes per patient and an increase in disease duration. Also, when patients are put into two categories, a diagnosis group consisting of all the possible diagnoses (Salmonella, rotavirus, etc) there will be more virulence genes in those patients than in those patients without a diagnosis. 17 Dye Signal 225000“; 5 135 200000 : ureA 1.85 . 45 175000 : 112 ma 207 440a 491 1; 163 f aStA ehxl terc 53ng 39‘“ fiyuA 223 toxB 150000 - ; E l 125000-5 l 100000-E ‘ 75000-5 50000-5 l 304 : ll md” . I llllll lll lf 0 I . [I ll ll . 0 100 200 300 400 500 500 700 Size (nt) Figure 1. Fluorescent chromatogram output from the multilocus virulence gene profiling (MVGP) procedure. The Beckman CEQ genetic analysis system displays each dye- labeled PCR amplicon as a peak of fluorescence. Peaks are labeled with nucleotide size, and a locus tag that specifies the gene name. 18 Table 1. Patients in the Seattle pediatric study grouped based on the diagnosis of microorganisms Diagnosis Numbe Numbe Average Averag Vomiti a Bloody Severit Severit Fever r of r of Illness e a . Diarrhea b y and Patient Isolates Duration Numbe ng a ' Duratio s (in days) r of C Stools n 1/221 C. jejuni 3 13 5.3 1 1/3 3/3 1.8 9.5 3/10“1 C. difficile 12 37 5.5 1.2 6/12 8/12 1.5 8.3 STEC (0157) 5 23 2.5 1 4/5 3/5 3/5 2.0 5.0 STEC (non-0157) 5 24 6.3 1.2 2/5 1/5 4/5 1.4 8.8 0/7a Rotavirus 8 35 4.1 1.5 7/8 4/8 1.4 5.7 Salmonella 4 15 1.8 1 2/4 4/4 3/4 2.3 4.1 Shigella 5 25 2 1 4/5 5/5 3/5 2.4 4.8 Negative 32/49a 27/50a 9/50a (all) 51 219 5.1 1 1.3 6.6 Negative a (Groupl) 25 100 4.3 1 17/2412/23 2/24 1.3 5.6 Negative 14/21a 13/20a (Group2) 22 99 6.2 1 5/22 1.5 9.3 a Number of responded patients Calculated as the sum of fractions from vomiting, fever and bloody diarrhea c Index is calculated from the severity index multiplied by the duration in days 19 MATERIALS AND METHODS Bacterial Strains and DNA Isolation. E. coli isolates were cultured from the stools of children with diarrhea who presented to a pediatric Emergency Department at the Children’s Hospital and Regional Medical Center (CHRMC) in Seattle, WA. All patients who presented with diarrhea to the CHRMC emergency department during the period from November 1998 through October 2001 were considered eligible and appropriate for enrollment in the study. This population has also been reported in other studies [61-63]. Parents were given an information form that described the study and also informed them of its voluntary nature. If they agreed to enroll their child, they were instructed to complete a questionnaire in English, Russian, Spanish, Somali, or Vietnamese. Questions asked addressed illness history and patient demographic characteristics. If the patient was unable to provide stool during the visit, a swab specimen from the rectum was obtained, but only if the family consented. The procedures fOllowed were performed with the approval of the CHRMC Institutional Review Board. A panel of diagnostic tests was done to determine a possible cause of illness on all stools submitted in sufficient quantity. These including inoculating the specimen onto sheep's blood, MacConkey, sorbitol-MacConkey, Hektoen, and Salmonella-Shigella—, Campylobacter-, and Yersinia-selective (Prepared Media Laboratories) agars and inoculation into Selenite F (BBL; Becton Dickinson) and MacConkey (Binax) broths. Campylobacter plates were incubated under microaerophilic conditions at 42°C, while all other incubations were at 35°C. These techniques are standard for the isolation of Aeromonas species, Campylobacter species, E. coli 0157:H7, Pleisiomonas shigelloides, and Salmonella, Shigella, Vibrio, and Yersinia species. Also, the MacConkey broth was 20 tested after overnight incubation by EIA (Meridian Biosciences) for the Shiga toxin antigen and Clostridium diflicile toxin was tested using a cytotoxicity assay with cultured human diploid fibroblasts. Cytotoxicity was confirmed by neutralization with Clostridium sordellii antitoxin. We sought parasites using trichrome stains and formalin/ethyl acetate sedimentation. Fluorescence antibody testing was also used to detect Giardia and Cryptosporidium species (Techlab). On frozen specimens, we sought rotavirus, adenoviruses, and astrovirus using Rotaclone (Meridian Biosciences), Adenoclone (Meridian Biosciences), and astrovirus (IDEIA Astrovirus EIA kit, DAKO) EIAs, respectively. All E. coli were grown on LB agar overnight at 37 °C from freezer stocks created with single-colony picks of each strain. Strains were inoculated into 10 ml of sterile LB broth for overnight growth at 37 °C, with moderate shaking. DNA isolation was performed using the PureGene (Minneapolis, MN) DNA isolation kit protocol for Gram- negative bacteria. Extracted DNA was quantified by a Nanodrop ND-1000 (Wilmington, DE), then diluted with sterile water to ~100 ng/ml. DNA samples were stored at —20 °C until use. PCR Primer Design. DNA sequences of MVGP components were obtained from GenBank. Homologous genes were aligned, with PCR primers designed from the conserved regions of each gene (Table 2). A11 primers were synthesized by Integrated DNA Technologies (Coralville, LA) and were stored at a concentration of 100 mM in ddeO. Working concentrations of each primer were 1 mM for MVGP, and 20 mM for MLST. 21 Table 2. Oligonucleotide primers used in the Seattle pediatric study. Gene Name 5' - 3' sequence Reference astA EAST-1P1 GGTCGCGAGTGACGGCTTTGT [64] EAST- 1P2 CCATCAACACAGTATATCCGA [64] bpr bpr-F 11 GTCTGCGTCTGATTCCAATA This study bpr-Rl TCAGCAGGAGTAATAGC This study cth cthl-R2 TGCCGCTCTGACAGGTGGACTTA This study cthl-FZ GCCTTTAAAAACGGGGTGATACA This study chuA chuA-63 6F TGAAACCGCGCCGAATGACGAGT This study chuA-1171R GGGTTCCGCCAAGCAGGGTAATC This study eae eae-F626 ATTATGGAACGGCAGAGGTTAAT This study eae-R1166 ATCCCCATCGTCACCAGAGG This study ehx MFSlFb GTTTATTCTGGAGCAGGCTC [65] MFSIR CTCCACGTCACCATACATAT [65] elt TW20 GGCGACAGATTATACCGTGC [66] J W1 1 CGGTCTCTATATTCCCTGTT [66] espC 601F GTTGGGGCTCGGACGACTTAT This study 1151R CCGGCACCCTTGAATGTTAATT This study espP 2859F CGCGCCAAAAGACACCAATGAA This study 3321R CAGGCCAGCCCCCACAGACTT This study estA JW 14 ATTTTTMTTTCTGTATTRTCTT [66] W 7 CACCCGGTACARGGCAGGATT I [66] fizuA fyuA-924F GCAGCAGCAGCATTATTCG [47] fyuA-RP CGCAGTAGGCACGATGTTGTA [67] iha iha-Fl ACGCAGCCGCCAGTGTT This study iha-R1 CCATCAATCAGTATCAGCGTGTAA This study invG invG-481 F TGACCTGGTCGTTAATGCTG This study invG-572R CGCCACGTAACTCATAAGTCC This study irpZ irp2-FP AAGGATTCGCTGTTACCGGAC [67] irp2-RP TCGTCGGGCAGCGTTTCTTCT [67] mdh 269F GGTATGGATCGTTCCGACCT This study p10 GGCAGAATGGTAACACCAGAGT This study pet F2227 GTTACGGCCAGCAGTTCCCTTTTC This study R2596 AATTGCCGGTCACTTTCCAGAGC This study pic 921F CGATGCCCCCGTAGACTTTGTTTC This study 1333R TACCGTCTCCCCTTTTCAGTCCTC This study saa saa-1442F CGTGATGAACAGGCTATTGC [68] saa-1522R ATGGACATGCCTGTGGCAAC [68] 22 Table 2 (cont’d) sat 1083F TGGTAGCGGTGGTATTATCTTTGA This study 1525R CGGCTTCTTTCGTTGTATCTGAGT This study senA F486 GGGGGATTTTGTCATTCAGC This study R975 CATTCCTTCCGCAGTTAGTAGTTC This study sepA 1672F GGAGCGCCGGGAGACCT This study 2093R GCCGCATCGAGTTTCAGTT’I‘TTC This study setIA F46 ACGGTTTTCCCAGTCTTTCT This study R481 TATATCCCCCTTTGGTGGTA This study shuA F285 CATCGCGGCGTGCTGGTTCTTG This study R657 CTCGTCATTCGGCGCGGTTTCAC This study sigA 30F GCCCAGGGAAAAATGTATGTAGAT This study 434R AAGACTGTCGCGGGTTTTTA This study spaP SpaP-143F GGACTTCAGCAAGTGCCATC This study SpaP-275R ACCACTCATGCCTGTCTCAA This study 3th 1A-251F GGGATAGATCCAGAGGAAGG This study 1A-832R CCGGACACATAGAAGGAAACTC This study stx2 2A-506F CTGGCGTTAATGGAGTTCAG This study 2A-848R CCTGTCGCCAGTTATCTGAC This study terC terC-106F TATGCACCGTGATGACAAGC This study terC-275R GGCGAACCAGGAGAAGATTG This study toxB toxB-911F ATACCTACCTGCTCTGGATTGA . This study toxB-l468R TTCTTACCTGATCTGATGCAGC [10] ureA ureA-109F TAACTATCCCGAATCCGTGG This study ureA-213R GGGATCATTTCTGGTATGCCT This study 23 PCR of MLST Genes. E. coli isolates were characterized using MLST for sequence and phylogenetic characterization of seven conserved housekeeping loci, and MVGP for virulence characterization using a panel of 29 known and putative virulence loci. Aliquots, (1 pl) of each DNA sample were amplified in a 25-pl reaction mixture using the AmpliTaq Gold system (Applied Biosystems, Foster City, Calif). Each reaction contained 2.5 p1 10X Gold buffer (150 mM Tris-HCl, pH 8.3, 500 mM KCl), 2.5 pl dNTPs (2 mM each dATP, dCTP, dGTP, and dTTP), 2.0 pl 25 mM MgC12, 0.5 pl of each primer (lmM), 1.5 units AmpliTaq Gold, and 15.7 pl sterile ddeO. Amplification in a Hybaid PCR Express therrnocycler (Hybaid Limited, Middlesex, England) utilized an initial denaturing step at 94 °C for 10 min., followed by 35 cycles of 92°C for 1 min., 58 °C for 1 min., and 72 °C for 30 see. A final step of 72 °C for 5 min. was used for final completion of any partially extended product. The PCR products were purified using QIAquick PCR Purification Kit (QIAGen Inc., Valencia, CA). Purified PCR amplicons were sequenced using a Beckman CEQ 2000 XL DNA sequencer (Beckman, Fullerton, CA) according to the manufacturer’s suggested protocol. Phylogenetic Analyses of Sequence Data. Sequences were aligned with the ClustalW algorithm using the computer software MegAlign (Lasergene), and allelic sequences were determined. Neighbor-joining trees of the concatenated internal regions of 7 conserved housekeeping genes (aspC, cle,fadD, ich, lysP, mdh, and uidA) were constructed using p-distance of nucleotide substitution with the computer software 24 MEGA version 2.1 [69] and the inferred phylogenies were each tested with 500 bootstrap replications. Virulence Genes and Protein Functions. KnovVn and putative virulence genes used for MVGP screening are listed in Table 3. A summary of gene functions are described below and grouped by the major pathotype in which the genes are found. The gene mdh, encoding malate dehydrogenase, was used as a positive control for all E. coli. A brief description of the virulence genes listed in Table 3 is described below. Two iron-related genes were used to identify enteroaggregative E. coli,fi)uA [67, 70, 71] and irp2 [67]. These are both markers for the High Pathogenicity Island [67], which is believed to have been horizontally acquired by E. coli from Yersinia. EAggEC also contain the serine protease autotransporter, pet in strain 042. This toxin is associated with mucosal damage, increased mucus release, exfoliation of cells, and development of crypt abscesses [16, 17]. Three markers were used for the p0157 plasmid: hemolysin (encoded by ehx) for iron piracy [72], EPEC secreted protein P (encoded by espP) [73-76] and an adherence enhancing protein [10] (encoded by toxB). In addition, strains were screened for an autoagglutinating adhesin (encoded by saa), a marker for a subset of LEE-negative STEC strains capable of causing severe gastrointestinal disease [68]. EHEC also have other genes encoding a variety of possible virulence functions, including toxins (cth) [77-80], and iron utilization (chuA) [81, 82]. EHEC also contain the TAI island (O-island 43). We screened for 3 markers from this island, iha [83], terC [84], and ureA [85, 86]. 25 Table 3. Virulence genes associated with diarrheal. disease Major pathogensa Gene Protein Locationb Control mdh Malate dehydrogenase C EAggEC fizuA Yersinabactin receptor F yuA C irp2 Iron repressor HMWP2 C pet Secreted autotransporter toxin P EHEC cth Cytolethal distending toxin C chuA Heme utilization C iha IrgA homologue adhesin C terC Tellurite resistance protein C C ureA Urease C ehx Enterohemolysin P (p0157) saa Autoagglutinating adhesin P (p0157) espP EPEC secreted protein P P (p0157) toxB Toxin B P (p0157) EHEC/EPEC eae Intimin C spaP Surface protective antigen P C EHEC/STEC stxl Shiga toxin 1 C stx2 Shiga toxin 2 C EPEC astA EASTl heat stable toxin C, P bfoA Bundlin P (EAF) espC Secreted protein C C ETEC elt Heat labile toxin (LT-A) P estA Heat stable toxin (STI) P S. dysenteriae shuA Heme receptor C S. flexneri pic Secreted autotransporter toxin C sat Secreted autotransporter toxin C senA Enterotoxin ShET2 P sepA Secreted protein A P set] a Enterotoxin ShETl C sigA Secreted autotransporter toxin C Shigella Sp. invG Transport C a EAggEC, entero-aggregative E. coli; EHEC, enterohemorrhagic E. coli; EPEC, enteropathogenic E. coli; ETEC, enterotoxigenic E. coli; STEC; Shiga toxin-producing E. coli; S., Shigella b C, chromosome; P plasmid; p0157, plasmid found in E. coli 0157:H7 strains; P (EAF), plasmid found in EPEC strains 26 EHEC and STEC also contain one or both Shiga toxins (encoded by sth and stx2). The toxins’ ability to stop peptide elongation during translation has been well characterized [87-89]. Three genes used in this study are common to bOth EHEC and EPEC: eae (encoding the adhesin intimin and serving as a marker for the LEE island), and two markers for O-island 115, spaP [90], and invG [91]. Three additional markers for EPEC used in this study include astA, the heat—stable enterotoxin EASTl [64], bpr, which encodes bundlin and is a marker for the EAF plasmid [92], and the serine protease autotransporter toxin gene espC [4, 5]. ETEC were identified by detection of two distinct toxins responsible for the disease symptoms in ETEC infection. ST, the heat-stable toxin (encoded by estA) causes the rapid accumulation of intestinal fluid [66], while LT, the heat-labile toxin (encoded by elt), is a secretory toxin with a function similar to that of cholera toxin [66, 93, 94]. Our marker for S. dysenteriae was the heme receptor gene ‘shuA [95, 96], which is closely related to chuA from EHEC. We used six markers for S. flexneri, including 4 serine protease autotransporter genes (pic [85, 97], sat [98, 99], sepA [100], and sigA [101, 102]), and enterotoxins ShETl encoded by set] a [19, 20, 101] and ShET2 encoded by senA [24, 103]. MVGP Procedure Summary. Multi-locus virulence gene profiling involves PCR amplification of extracted DNA with a primer pair that has one regular (unlabeled) and one Beckman fluorescently-labeled primer with the dye attached to the 5’ end of the primer. Up to 16 of the labeled amplicons are then pooled together, and then separated on 27 the CEO Genetic Analysis System using capillary based sequencing analysis. Visual determination of virulence profiles resulted from analyses with pre-set parameters for each locus for size and dye label. PCR of MVGP Genes. This technique allows rapid generation of virulence profile data using proprietary technology developed by Beckman Coulter (Fullerton, CA). We analyzed raw data with a collection of pre-set parameters, including data indicating locus tags specifing gene name, amplicon size (i 3 nucleotides), and dye color. A size marker estimates the size of unknown fragments within 1-2 nucleotides. Each profile identifies individual virulence factors as peaks of defined size and dye label, to provide a fingerprint of the virulence content of each strain. PCR for virulence loci examined by MVGP used the same protocol as the MLST PCR except for the following changes: the 20 p1 cocktail contained 2.0 pl of 10X Buffer, 2.0 p1 dNTPs, 2.4 p1 MgClz, 1.0 pl each of forward and reverse primer, and 6.4 pl sterile ddeO; cycling conditions included: 40 cycles of 92 °C for 30 sec., 50 °C for 30 sec., and 72 °C for 45 sec., with a final extension step at 72 °C for 7 min. 5 pl of DNA template was used per reaction. PCR primers and amplicon sizes for virulence genes are listed in Table 2. Controls for MVGP. Positive controls consisted of the malate dehydrogenase gene (mdh) for all strains and a pooled positive reaction containing a mixture of DNA from all strains serving as positive controls for the other 29 loci (Table 3) in the study. Negative controls contained H20 and DNA from E. coli K12, which was negative for 28 most genes. In some cases, two sets of primers with the same dye-label were multiplexed to save reagents and space. In that case, the H20 per reaction was reduced from 6.4 to 4.4 ml. The mdh gene was used as the positive control strains for each gene in the MGVP set. Each sample run contained both a positive control for each individual strain tested (mdh), but also a pooled positive control that contained positive labeled amplicons for each gene tested in that set. MVGP Data Confirmation. Inconsistencies in the data were clarified by standard PCR using 1.5% agarose gel electrophoresis. Approximately 900 individual PCR reactions were performed to confirm cases where a gene was not uniformly positive in all strains of the same ST in a patient. PCR was used to also check inconsistencies involving results when one, but not both genes on a specific element were detected, such as irp2 and fiuA, markers for the High Pathogenicity Island from Shigella. In all cases, every strain with variable results was re-tested to be as thorough as possible. Images in this dissertation are presented in color. 29 RESULTS Genetic Diversity of MLST Loci. All 392 strains from 94 patients in this study were sequenced by MLST for seven loci. The nucleotide diversities of these loci were between 4.2% (lysP) to 14.2% (fadD). The overall variability across all loci was 9.4% (353 variable sites in 3753 total nucleotides sequenced) (Table 4). The lowest and highest rates of synonymous substitutions, as measured by the Nei-Gojobori method of nucleotide substitution, were also found in lysP (3.00 :l: 0.90) and fadD (9.74 :t 1.62) respectively. However, for non-synonymous substitutions, IysP also had the lowest rate (0.00 i 0.00), but uidA had the highest rate (0.41 d: 0.12). The overall rates of synonymous and non-synonymous substitution were 5.69 d: 0.41 and 0.10 i 0.02, respectively. Sequence Type Diversity and Clonal Groups. A phylogenetic framework was assembled to display the diversity of different STs. The 392 strains contained a total of 67 unique ST s (Figure 2). This sample set was widely distributed across the diversity of all known E. coli and included some strains representing specific classes of disease- associated E. coli that were highly overrepresented in the sample. For example, clonal group 35 (CG-35) contains strains associated with septicemia and meningitis, including clinical isolate R8218 [104, 105]. Our sample set contained 74 strains of ST-29, which belongs to CG-35. ST-29 is the sequence type of 19% of all strains in this study, and was found four times more frequently than the next most common type, ST-27 (n=1 8). 30 Table 4. Sequence variation among alleles of seven MLST genes. Shown here are the number of nucleotides sequenced per gene, the number of variable nucleotides, the number of parsimoniously informative sites (nucleotide variations in more than one strain), the number of alleles identified for that gene in this sample set, and the p- distance. Locus" No. sites No. variable No. PI No. alleles % p distance sites i SE aspC 513 45 25 33 1.20 :l: 0.24 cle 567 68 53 34 2.36 :1: 0.31 fadD 492 70 57 36 2.70 :l: 0.41 ich 567 52 42 33 1.84 :1: 0.29 lysP 477 20 14 20 0.84 i 0.23 mdh 549 36 29 27 1.11 d: 0.22 uz'dA 588 62 42 39 1.83 i 0.31 Total 3753 353 262 222 1.70 3: 0.12 31 No. No. NJ tree 8 T isolates patients 295 3 298 3 1 69 1 2 1 71 1 4 31 3 1 68 297 300 281 282 Distance 1-——I 0.002 .3 d-I —l V AGM‘UDOIAOIhNUI-‘UI-‘UIQO’IAhwdd-‘AmN-‘OJ-‘UIOJ-‘CD-‘ddAV-‘UINMO'IN-‘UI-‘NNUN‘(DU-hmméé-‘UIA .3hawmgmagbaQAJAgfidgaa4.:ANNQQAudnanaaaaudanawmaaaaaaaA—LAANNAAANJUIQAA _|-I J-l A Figure 2. Neighbor-joining tree showing the genetic relatedness of 67 sequence types 392 E. coli from 94 patients. Each branch tip represents a unique sequence type (ST) based on multilocus sequence typing of 7 genes. Distance is measured by p distance with bootstrap support greater than 70% given by italics. The number of isolates and number of patients in which each ST was isolated are given in the right hand columns. The circles denote STs that were recovered from more then 1 patient (n = 24). The most frequently detected ST (ST-29) is indicated by the gray arrow. 32 The second and third most common clonal groups included CG-23 (ECOR A) with 34 isolates, and CG-38 (UTI-l associated strains from ST-27 and ST-28), with 25 isolates. We found more than one representative for 24 of the 67 STs (35.8%). MVGP Analysis. A total of 392 isolates from 94 patients were examined via MGVP, with an average of 8.1 unique virulence genes detected per patient. The overall virulence type (VT) diversity is 71%, indicating that two strains chosen at random have a 71% chance of displaying different profiles in virulence genes. Ten patients had strains with zero diversity (all isolates with the same VT), whereas twenty patients had strains with 100% diversity (all isolates with different VTs). Genes encoding iron acquisition or utilization ability were the most frequently detected in this sample set (Table 5), including shuA (72.5% positive), irp2 (72.0%), fizuA (68.4%), and chuA (65.5%). In contrast, other genes restricted to specific pathotypes were found less frequently, including six] and stx2 (6.0% detection each). Comparison of Virulence Gene Profiles and Phylogenetic Relationships. Virulence profiles were compared with phylogenetic frameworks generated by MLST. Three examples of virulence diversity within individual patients include the following: Patient 1 had 5 individual isolates with the same ST (ST-29) and VT. Patients 34 and 35 (Figure 3) had isolates that showed variability in their profiles. Patient 34 contained 5 strains with the same ST (ST-292), but had 5 slightly different VTs. Patient 35 had strains with 4 different STs (STs 119, 272, 293, and 294) and 5 different VTs. Collectively, patient 35 contained strains with 17 of 29 Virulence genes examined in this study (59%). This patient shows the 33 Table 5. Summary of virulence gene detection frequency in sequence types and in patients from the Seattle pediatric population study. Gene % STs (67) % Patients Gene % STs (67) % Patients £4) (94) shuA 70.1 79.8 set] a 17.9 12.8 irpZ 70.1 81.9 pet 16.4 23.4 fi/uA 67.2 85.1 sepA 16.4 9.6 spaP 64.2 50.0 espP 11.9 11.7 chuA 62.7 74.5 stxl 11.9 12.8 terC 62.7 74.5 toxB 10.4 11.7 invG 59.7 47.9 pic 10.4 8.5 eae 44.8 40.4 stx2 7.5 10.6 iha 37.3 40.4 espC 7.5 6.4 ureA 35.8 25.5 senA 6.0 6.4 saa 32.8 30.9 bfiDA 3.0 2.1 sat 28.4 29.8 cdt 3.0 2.1 ehx 22.4 17.0 elt 3.0 1.1 astA 20.9 11.7 estA 1.5 1.1 sigA 20.9 10.6 34 63865 8? 2 £8:me :5qu .8505 go: 0.8 SEES etaawoz .ocow 832th 35 new :38 m0>2 0353 m E8058 235mm 3.8—8 2.2. .3538 egg—com 3335838 95 E gauze cab coco—35 .m charm o O O o OO OO O O O OO 00. O O O o OO O OO 8 as O O ooo 59: V23. Vase .63 Va.» Dame V36 one E9. NR2 new 9: 3o V56 to VQQ lean Noam V6.5 952 D2: Knee Vase Be Vs O O O O O O O O O O O Vat. Vase :8 V3}. O O O O O O O O O O $2: 00 as swam Hm» 33959 S: Hwy Eng? 3: km» thoozfit ANNN Hwy mummokfi. Ammm Hm» Shook/H uflumcfiO ~39 .N "mmmc:wamfi I mm 33:?— OO O O 000 OO O 0 GOO 000 O O @O OO O O OO 00 0 000 O O O V23. V23. NE V36 USO Vac one :3. .38» e5. 5: 33 V56 mdh Vat. V93. 3m V5}. to VQS meme Noam Von: Uh: bsfi meme Vase En SE 3am Hm» smack/H Ammm pm; @302? Sam Hwy $3023. Amom Hwy $395k Amam Hwy 833:. :— m=:own=m e-e=e£~eh Ewe—.35 I vm 285.5 35 greatest diversity of STs and VTs in this study, and may be an example of how different strains can contribute to unique disease by complementation of virulence ability. A number of genes related to type III secretion (eae, invG, and spaP) and iron utilization (chuA, shuA, fizuA, and irp2) are seen in 30 or more different STs, and 38 or more patients (Table 5). Detection of autotransporter genes was highly variable. Another analysis technique called non-metric multidimensional scaling (NMDS) analysis was used to place most patients into two clustered groups based on the virulence content of results from each patient, not per isolate. This allowed for distinctions based on the detection of virulence genes in any strain found in a patient, while keeping the individual patient as the unit of measurement, and not individual isolates. Since we used 29 virulence loci, NMDS combines these 29 dimensions into a two-dimensional image. NMDS showed two distinct groups (Figure 4). Groups 1 and 2 contain 42 and 43 total patients respectively. All 10 STEC patients were within group 1, whereas 9 of 11 patients with C. difficile were in group 2. Group 1 also had 25 patients with no diagnosis and an average of 9.0 virulence genes per patient, while group 2 had 23 patients with no diagnosis and an average of 6.6 virulence genes. Statistical analysis. Our first hypothesis was that a longer illness duration (0-3 days vs. 4-6 days vs. 7 or more days) was significantly correlated with more virulence genes found per patient. We accepted this hypothesis based on a chi-square test (2:7.5, p-value =0.02). The second hypothesis was that we would find a significant correlation between the number of virulence genes (0-5) vs. (6 or more) and diagnosis group. This was based on the idea that the E. coli from the no diagnosis group were not likely the cause of the illness seen in the patient and therefore would tend to have less virulence 36 genes. We rejected this hypothesis based on a chi-square test (2:2.07, p-value =0.150), so there is no significifcant difference between the number of virulence genes in the diagnosis and the no diagnosis groups. 37 1.20— ' 0 N 0.80- .2 . >< Group 1 . Group 2 (0 \ 0.40" m o .e 0 E 0.00 ' Q o -040— 1 o -0.80 Tl'l'l'l'l'l'l’ -1.00 -0.60 -020 0.20 0.60 -0.80 -0.40 0.00 0.40 Principal axis 1 o No Diagnosis 0 Rotavirus o C. difficile 0 Salmonella O C. jejuni O Shigella 0 E. coli Figure 4. Non-metric dimensional scaling (N MDS) analysis of the MVGP results of 94 patients based on the presence or absence of 29 known or putative virulence factors. Two clusters are indicated by the circles, named group 1 and group 2. 38 DISCUSSION Diversity of Virulence Gene Profiles. Virulence profiles in this study are highly variable. Individual patients often contain multiple STs with minor to major variation in virulence diversity. At least 70 STs were found from 392 isolates sequenced, including 45 newly discovered STs. Some individual patients contained strains with different STs and VTs. Even in patients with strains that all have the same ST, variability in virulence exists, such as was seem in the analysis of strains from patient 34 (Figure 3). Other patients had diversity in both STs and VT. The greatest example of ST and VT diversity was in patient 35, with four unique sequence types (STs 119, 272, 293 and 294) and five unique virulence profiles. This patient harbored E. coli with 17 of 29 (59%) virulence- related genes used in our study. Genes encoding virulence factors responsible for iron acquisition and utilization (shuA, fizuA, chuA, and irpZ) were the most prevalent in the samples examined. Although iron is important for the survival of E. coli [106, 107], the role of iron—related genes in pathogenesis is not entirely clear. Such widespread distribution in this sample set may indicate a role in survival, as a means for virulence enhancement, rather than overt pathogenic ability. Some of the variation seen among isolates with the same ST may be a result of natural rates of gene flux. Because many of the virulence genes in this study are found on mobile elements, it is possible the plasmid or phage containing certain genes could be lost. There is also a possibility of mutations at the primer site affecting proper amplification and the affect of false-negative or false-positive results. Extensive rechecks 39 by standard PCR and gel electrophoresis helped clarify initially inconsistent results in the data set. Frequency of ST-29 Isolates in this Sample Set. The detection of a high frequency of ST-29 isolates in this study was surprising. These may represent a group of strains with enhanced colonization ability. ST-29s appear to be largely lacking the genes encoding many virulence factors typically associated with disease. They do, however contain, an even higher frequency of iron-related genes than the non-8T-29 strains in this study. The detection of such a high number of ST-29 isolates was unexpected. This sequence type is the same as the meningitis-associated isolate RS218. Nearly all (68/70) of the ST-29’s in this set also contained the same capsule type (K1) as R8218 [105]. This isolate may represent a widespread meningitis associated clone that has been under- investigated. The reasons for the high frequency of detection may indicate this strain has an ability to cause disease that is yet uncharacterized, or it simply might colonize and survive better in the gut. Why a strain with potential extra-intestinal disease-causing ability is so frequent in the stool of children is also unclear. Since neonatal meningitis caused by strains like R8218 is present only in very young children [108], perhaps ST- 29’s can cause disease, but only when a child has an underdeveloped immune system. Perhaps, in older children, ST-29 isolates are not able to cause disease, but still persist in the gut through a mechanism of attachment or immune avoidance. A total of 10 of 51 patients without a diagnosis (19.6%) had ST-29 isolated from their stool. However, ST-29’s were isolated from 5 of 12 (41.7%) of patients diagnosed 40 with C. difiicile and 3 of 5 (60%) of patients diagnosed with Shigellaflexneri. No ST-29 isolates were found in patients with a diagnosis of C. jejuni, rotavirus, Salmonella paratyphi B, Salmonella subgenus 1 (Salmonella D) or E. coli serotypes 01032H2, O111:H8, 01 l 1 :NM, or O] 182H16. The presence of C. jejuni, rotavirus, Salmonella, etc, may change the local environment of the gut in ways that inhibit the survival of 8T-29 isolates. Virulence profiling of E. coli has been used previously to detect a large number of mostly attachment-related loci in strains from the E. coli Reference collection (ECOR) [109]. We hypothesized ST-29 isolates would contain a gene or genes found in nearly every strain that is associated with attachment in extra-intestinal variants of E. coli. We further investigated ST-29 strains by examining a select number of attachment-related loci used by Johnson, et al. [109], which were not included in our original MVGP gene set. This work is detailed in a subsequent study. It is also unclear if 8T-29 isolates have a differential geographic distribution. The Seattle, Washington state area has a high incidence of illness due to E. coli 0157:H7 [110-114], but also has the research and laboratory facilities for efficient detection of this serotype. The closely related strains of 8T-29 may in fact exist in large numbers worldwide, but the low detection frequency may be due to a lack of appropriate surveillance. 41 ACKNOWLEDGEMENTS I would like to thank Lindsey O'uellette for her tireless work sequence typing the strains used in this set, as well as Drs. Eileen Klein and Phillip Tarr for collecting and sending us the strains examined here. Thanks are also due to Albert Lee for his help with virulence gene data confirmation and to Drs. David Lacher and Cheryl Tarr for technical assistance. This research was supported by funds from the Centers for Disease Control and Prevention (Cooperative Agreement CCU015040) and NIAID, NIH, DHHS, under NIH research contract # N01-AI-30058 (TSW). 42 CHAPTER 3 CHARACTERIZATION OF A COMMON MENINGITIS-ASSOCIATED CLONE (ST-29) FOUND IN PEDIATRIC PATIENTS IN SEATTLE, WA 43 SUMMARY The intrinsic virulence of individual Escherichia coli strains is determined in part by the ability to encode toxins, enhanced nutrient utilization, attachment, or other abilities important for disease. However the distribution and relationship of virulence factors to disease-causing ability is not entirely understood. As seen in Chapter 2, we reported the sequence types and virulence profiles of 392 E. coli isolates from pediatric patients with diarrheal disease of unknown cause. In this strain set, a high prevalence (19%) of isolates of a single sequence type (ST-29) were detected by MLST analysis. 8T-29 belongs to a larger clonal group that includes 8Ts of extraintestinal pathogenic E. coli (ExPEC) including the meningitis-associated strain R8218. We hypothesized that 8T-29 strains have a fitness advantage related to adherence to the intestine. To address this hypothesis we assayed for the presence of additional genes associated with ExPEC adherence (aafA, aggR, bmaE, fimH, iutA, focG, ibeA, nfaA, pap/1H, sfaA, sfaS) in 70 8T-29 strains using PCR-based assays. MVGP screening showed iron acquisition and iron utilization genes were the most frequently detected, including shuA (97%), ith (97%),fi/uA (90%), and chuA (90%). Toxin genes like six] (10%), and stx2 (3%) were found less frequently. The prevalence of attachment factors was mixed, with some detected frequently, fimH (97%) and nfaA (90%), whereas others were low or absent (aggR, focG, and aafA were not detected in any 8T-29 strain). An RFLP-based capsule typing method showed 68 of 70 (97%) ST-29 isolates to be K1, the same type as R8218. ST-29 warrants further study because it appears to mark a common E. coli intestinal clone, related to ExPEC, which lacks virulence genes associated with diarrheagenic E. coli. 44 INTRODUCTION Escherichia coli is a gram negative bacterial resident of the intestinal flora within a healthy human host. However, some strains of E. coli have acquired the ability to cause a variety of illness, ranging from mild diarrhea to a severe bloody diarrhea that can be life threatening. Certain strains are also associated with disease outside the intestine, including meningitis, septicemia, urinary tract infections, and kidney damage [1, 115]. The different pathogenic abilities are seen because some strains have acquired genes or pathogenicity islands that allow for the production of toxins and other factors necessary for disease onset [1, 2]. E. coli is capable of frequent acquisitions of new virulence ability by horizontal transfer of genetic material. One way to group E. coli is based on the type of disease they cause. Strains of the same disease class are not necessarily evolutionarily related, but may have acquired the same virulence factors encoded in their genome to produce similar disease phenotypes. We are interested in determining the prevalence of the genes encoding many common virulence factors in specific populations. In Chapter 2, we examined 392 E. coli isolates from children hospitalized in Seattle, WA with unexplained cases of diarrheal disease. These strains were characterized by two methods: multi-locus sequence typing (MLST), and multi-locus virulence gene profiling (MVGP). DNA sequencing revealed a predominant sequence type (ST-29) found in 18.9% of strains (n=74). MLST detected 67 unique sequence types in this population that were distributed throughout the diversity of E. coli. However, 8T-29 was discovered at an exceptionally high frequency compared to other sequence types. 8T-29 isolates do not contain many of the classical virulence factors associated with enteric disease. We hypothesized ST-29 45 isolates may be more frequently detected because they are better able to attach to the intestinal tract. The majority of 8T-29 isolates, 60 (81%) were examined for the presence of 11 attachment associated genes by PCR. In addition, 70 (95%) of the ST-29 strains were capsular typed using an RFLP-based method (Figure 5) to determine the specific allele of their polysaccharide capsule. The banding pattern of the ~10,000-12,000 base pair kpsC—kpsM amplified region after digestion with the Sau96I restriction endonuclease is seen in Figure 6. Determination of allele type is easily made based on the distinct banding pattern for each allele standard. 8T-29 is important for further study because it is the same sequence type as R8218, a strain implicated in extra-intestinal infections, including septicemia and meningitis [116, 117]. We also propose that ST-29 marks a clone that has a role in intestinal pathogenesis that has yet to be determined. We wish to test the following hypotheses regarding 8T-29 isolates examined here. 1. Because the clinical isolate R8218 is a 8T-29 strain and has a K1 capsule type, we predict all or nearly all pediatric 8T-29 isolates from Seattle will also have the K1 capsule type. A capsule typing method using restriction fragment length polymorphism will be used to determine capsule profiles. 2. We also predict the prevalence of ST-29 isolates is related to an enhanced ability to adhere to the intestine. Therefore we predict we will find some attachment- related genes common in extra-intestinal E. coli to be in very high frequency (near 100%) in the 8T-29 strains. Individual pairs of PCR primers will screen a panel of genes encoding proteins related to attachment. 46 central variable region Keeeel'e ' "hfih‘ ' enamels eewneahfi 122450bp wmflfiwlfiw @ifiifléfl L_---_---——-_I Figure 5. Organization of the kps operon in E. coli. Primers were designed in the conserved flanking regions (kpsC and kpsM genes) and amplified a central variable region that differs greatly between strains of E. coli. The area amplified is indicated by the boxed region. 47 High Mass DNA Ladder ”-000 5 042 R8218 CFTO73 T18 100 bp — 19999 _ Ladder _ K1 Type K2 Type K5 Type _. A - 6000 - 5123 _ 4469 _ 4159 ._ — 4000 — 2019 — - 2995 - 3000 — - 2420 — 2072 - 2000 -' 1333 —1s49 _ 1500 - 1505 — 1563 -1522 = 1333 - “19 — 1331 - 33 1,000 — - 1000 _ - 979 _ 900 - 1°00 -— - 834 - 783 - 800 fi — 729 - 729 — 729 _ 700 — 527 - 679 __ — 502 — 502 - 500 — 551 _ 500 _ - 463 _ ’77 - 403 — 403 — 400 — 357 _ — 300 — 205 — - 200 — 155 J _ 100 100 Figure 6. Representative banding pattern from a Sau96I digest of the amplified kpsCM region in E. coli. Shown here are the specific types used as standards when comparing the 8T-29 isolates used in this study. 97% of the 70 8T—29s examined by RF LP were K1, the same as the R8218 standard pattern. 48 MATERIALS & METHODS MVGP PCR primer design: DNA sequences of MVGP components were obtained from GenBank. PCR primers were designed from the conserved regions of each gene sequence available. All primers (Table 2) were synthesized by Integrated DNA Technologies (Coralville, IA.) and were stored at a concentration of 100 mM in ddH20. Working concentrations of each primer were 1 mM for MVGP, and 20 mM for MLST. Strain growth conditions and DNA extractions: All E. coli strains were grown on LB agar plates overnight at 37 “C from freezer stocks created with single-colony picks of each strain. Strains were inoculated into 10 m1 of sterile LB broth for overnight grth at 37 °C, with moderate shaking. DNA isolation was performed according to the PureGene (Minneapolis, MN) DNA isolation kit protocol for gram-negative bacteria. After extraction, DNA was quantified by a Nanodrop ND-1000 (Wilmington, DE), and then diluted with water to ~100 ng/ml. MVGP PCR: PCR of virulence-related loci were performed as described in Chapter 2. Controls: Positive controls consisted of the malate dehydrogenase gene (mdh) for all strains, and a pooled positive reaction containing a mixture of DNA from all strains serving as positive controls for the other 29 genes in the study. Negative controls contained H20 and DNA from K12, which was negative for most genes. In some cases, two sets of primers with the same dye-label were multiplexed to save reagents and space. In that case, the H20 per reaction was reduced from 6.4 pl to 4.4 pl. MLST PCR: MLST PCR, purification, and sequencing were performed as described in Chapter 2. 49 Phylogenetic analyses: Phylogenetic analyses of sequencing data was performed as described in Chapter 2. PCR for attachment loci: 5 pl of DNA template was used with 2.5 pl of 10X Buffer, 2.5 pl dNTPs, 2.0 pl MgClz, 1.0 pl each of forward and reverse primer, 1.5 units AmpliTaq Gold, and 15.7 pl sterile ddH20. Cycling conditions included: 10 minute initial soak at 94 °C, 35 cycles of 92 °C for 1 min., 55-60 °C for 30 sec., and 72 °C for 30 sec., with a final extension step at 72 °C for 5 min. Specific annealing temperatures varied, depending on the gene (Table 6). Capsular typing - PCR: Primers were used to amplify a region (~11,000 bp) in the kps operon (Figure 5). PCR was performed using Takara LA Taq polymerase (1 .5U), Takara 10x buffer (2.5 pl), Takara 2.5 mM dNTPs (4.0 p1), 1.5 pl of each primer (10 mM), and 14.2 pl sterile ddH20. Primer sequences are: Forward: kpsC-F558 5’ — AGCCGAAATTTGGGTGAAGGTG — 3’ and Reverse: kpsM-R120 5’ — YGCGCATTTGCTGATACTGTTGG — 3’. Amplicons (5 p1) were visualized on 1.5% agarose gels with 6 ml ethidium bromide stain to confirm successful amplification prior to digestion. Controls included strains R8218 (K1 capsule standard), CFT073 (K2 standard), T18 (K5 standard), and 042 (042 standard), along with a negative control of water. Capsular typing - digest: Amplicons were digested using restriction enzyme Sau96I. Each 30 pl reaction contained 3.0 pl of 10x buffer, 1.0 U of enzyme (0.2 pl), 16.8 pl sterile ddHZO, and 10 pl template. Reactions were digested overnight at 37 °C then separated on a 1.5% agarose gel stained with 12 pl ethidium bromide. Ladders included 100 bp standard (2 pl), and a high DNA mass ladder (4 pl). Gels were loaded with 15 pl 50 of digested PCR product, and run at 80 volts of constant current for about 2.5 hours to allow for sufficient separation. An example of representative banding patterns for four different capsule types is seen in Figure 6. 51 .8: 0800 05250500055005.0008 :05 0&4 H8: 0&0: om 8550 0 co 0: 05050:_ :00: N3 e00t<000050000505<<0 080 . 069 :8: NS 0 855.: <00<5<0050<000<0<00<0 .50 000 080 08: OK 05006:. a: 025 481: Soto <5000050000500<050<00 028 1:5 50: 2800 8 855.0 :80 .5858 00500008200025.0505 mm 08 039 IE: 0500000005050 3% 0800 05500000514205.0000? :80 085 so: 0800 00 58028 05002000015052.5800: mm 50 “.2 085 05000012 0305 0003 x020 5:: N3 <0<0<00C5 3' Size (bp) eer 144F TCAAGGCAAAACTGGATTC 387 530R TGTACGCTTGCAAAACTATTA eer 60F TGATCCGCAGCTGAGACATA 416 475R CAATAGCCTGACTTTCCAATACAT epaS 54F AAAAGGGCAAATTCTAAAAAGTA 475 528R CTCTCCCCAATCTGATAATAAAG epaP193F ATGCCGGTTGGGAAGGAA 366 558R TACCGGACTCATCATCATCATACC eivJ 187F TACCGCGATGGAGACAAAATAA 210 396R TCCCCAACGCTGAAACTGA eivl 44F TACTCGGCGAAGTTGTATTT 201 244R TTT'lTAGTCGGTCGTATTTTTCTC eivC 689F GGTGTAATGCCGCCCTGATG 355 1 052R ATTGCCGGATAGTGACCTTGACC eivA 88F ATTCCATTGCCCACCTATIT 340 410R AGCGGGCAGCAACTTCAG eivE 416F ATGTCCGGCTAAACGCTGAAG 357 772R CTGCCGACTGAAGAGACAATAG eivG 417F GGACGGCAATGGTACTTTCTAT 380 796R AACGCGCTCCTGCTGTCG eivF 27OF TGAATCAACGGGGAGTGTGG 277 546R ACGGCGAAAATGTGAATACGATA 67 MATERIALS & METHODS PCR Primer Design. DNA sequences of ETT2 components eer, eer, epaS, epaP, eivJ, eivI, eivC, eivA, eivE, eivG, and eivF were obtained from GenBank. Homologous genes were aligned, with PCR primers designed from the conserved regions of each gene. All primers were synthesized by Integrated DNA Technologies (Coralville, IA.) and were stored at a concentration of 100 mM in ddH20. Working concentrations of each primer were 10 mM. Strain Growth Conditions & DNA Extractions. All E. coli strains were grown overnight at 37 °C in 10 ml of sterile LB broth with moderate shaking, prior to PCR. DNA isolation was performed by suspending a single large colony of cells in 200 ml of sterile TE before heating at 95 °C for 10 min and centrifuging at 13,200 rpm for 10 min. All colonies were grown on LB agar plates overnight at 37 °C from freezer stocks created with single-colony picks of each strain. PCR. Aliquots, (5 ul) of each DNA sample, were amplified in a 25-ul reaction mixture using the AmpliTaq Gold system (Applied Biosystems, Foster City, Calif). Each reaction contained 2.5 ul 10X Gold buffer (150 mM Tris-HCl, pH 8.3, 500 mM KCl), 2.5 ul dNTPs (2 mM each dATP, dCTP, dGTP, and dTTP), 4.0 [1125 mM MgC12, 1.25 ul of each 10 mM primer, 1.5 units AmpliTaq Gold, and 8.2 ul sterile ddH20. Amplification in a Hybaid PCR Express therrnocycler (Hybaid Limited, Middlesex, England) utilized an initial denaturing step at 94 °C for 10 min., followed by 40 cycles of 92 °C for 1 min., 50 °C for 45 sec., and 72 °C for 45 see. A final step of 72 °C for 5 min. was used for completion of any partially extended product. The positive control strain for all ETT2 genes was Sakai, an E. coli 0157:H7 strain that the primers were designed 68 from sequence available on GenBank (accession number NC_002128). Negative controls included the laboratory strain K—12, an EPEC 1 strain E2348/69 and one reaction that contained no template DNA. Visualization of PCR products (5 ul) was achieved on ethidium bromide-stained 1.5% agarose gels via illumination with UV light. Primer sequences and amplicon sizes for the individual ETT2 genes examined in this study are shown in Table 8. 69 RESULTS ETT2 in EHEC 0157. Phylogenetic analysis involving MLST of 7 housekeeping genes in a set of pathogenic strains allowed for the determination of the evolutionary history of the 57 strains tested (Table 9). The MLST loci sequenced here were the same loci applied to the Seattle pediatric population in Chapter 2 and the ST-29 isolates in Chapter 3. Patterns of ETT2 gene detection were then overlaid onto the phylogenetic framework developed via MLST (Figure 9). This combination allowed for the discovery of possible points of acquisition of these virulence factors in evolutionary time. The detection of 11 genes distributed across a 17-kb segment of ETT2 (Fig. 10) provides evidence for the existence of an intact island in all of the OlS7:H7/H- strains examined in this study. These finding agree with previous work [3 8, 39] on the distribution of ETT2 in these EHEC isolates. While Makino et al. characterized the presence of gene groups or regions of ETT2, this study characterizes ETT2, and the associated erosion of many genes within it, on a finer scale. It is unclear if this secretion system is able to function in the cell. Because of the level of genetic attrition seen in this island, it seems unlikely that it is able to encode a fully functional needle complex [41]. However, some level of genetic expression or control apparently still remains within some genes in ETT2 [143, 144]. So even large- scale deletions seen in some strains my not entirely abolish expression of the remaining components of ETT2. We predicted ETT2 would remain intact in all of the strains representing the stepwise evolution of modern 0157. This hypothesis must be rejected based on the 70 deletion profiles observed in three different 055:H7 isolates. ETT2 is retained, in part, in these strains, and is intact in all known 0157:H7 strains, but some isolates are undergoing erosion of this island by a variety of different deletions. ETT2 in non-0157 STEC. As seen in previous work [38, 39], Shiga toxin- encoding non-0157 isolates had a truncated copy of ETT2 that was lacking the eiv gene cluster, containing 7 genes. All 026, 01 11 and other non-0157 serotypes contained a truncation at or near the epa-eiv gene cluster joint (Figure 10F). Although no experiments have been published on transformation of this truncated ETT2 island into another E. coli strain to see if it is functional, the lack of many critical genes needed for needle formation and proper assembly makes the possibility of functional type III secretion using ETT2 unlikely. Complementation experiments to restore a full copy of the island also have not been attempted. However, residual expression of some genes may still be possible, as was seen in 0157:H7 [144]. We predicted to see new deletion types in non-0157 isolates. We accept this hypothesis due to the discovery of 3 new deletion profiles (types 4-6). Other existing deletions seen in strains like K-12 and CFTO73 were also confirmed by this work. ETT2 in 055:H7. A different truncation of ETT2 was seen when examining representatives of the 055:H7 serotypes. While other serotypes seem to be consistent in the presence or absence of ETT2, there appears to be some variation within 055:H7 strains. A stepwise model for the evolution of modern 015 7:H7 from ancestral 055:H7 isolates has been described previously [145]. Here, this model has been modified to show an ancient point of acquisition of ETT2 in modern 0157:H7, and indicate a variety of deletions (Figure 11). Two of three isolates on the stepwise model after the acquisition of 71 the Stx2 phage (5905 and MDCH-lO) were positive for all individual ETT2 genes tested, and presumably have an intact island (Figure 10B). These were the only 055:H7 strains examined in the study to be positive for all ETT2 genes examined. A third stx2+ 055:H7 strain, 97-3256, contained a unique type of deletion not seen in any other strain (Figure 10C). This strain was negative for eer, eer and epaS. However, this strain was positive for epaP, but may be missing the entire epr gene cluster and part of the adjoining epa cluster. All four isolates representing more ancestral 055:H7 strains from before the acquisition of the Stx2 phage showed two distinct types of deletions of ETT2. Both DEC 5d and TB 1 82A had a major portion of the eiv gene cluster, as well as all of the epr and epa genes that were not detectable via PCR (Figures 9-10). Microarray analysis involving ORF comparisons in the genetic content of 0157:H7 strain Sakai, K-12 and 055:H7 isolates DEC 5d and TB182A have confirmed this deletion in ETT2 [146]. This deletion also extends to 8 ORFs upstream of the epa genes in ETT2. The other two 055:H7 isolates on this stepwise model (DEC 5e and LTOSS-62) contained another unique type of deletion (Figure 10E). All genes examined by PCR were positive, except for epaS. This may represent either sequence diversion in the primer site, or partial or total loss of the gene. 72 73 + + + + + + + + + + + _ Ommm m0 - 02 SE3 + + + + + + + + + + + _ Him 2. - 02 02%: + + + + + + + + + + + _ Ommm 2. - mm 2 00220 + + + + + + + + + + + _ Ummm m0 - mm 2 32003 + + + + + + + + + + + _ Umzm 2. - 2h _ 8202”? + + + + + + + + + + + _ Ummm mm - hm _ $2002 + + + + + + + + + + + _ Ummm n0 - 2.2 00223 + + + + + + + + + + + {BO _ Him 2. - 2.2 3230 + + + + + + + + + + + 005.50 _ 08.5 2. - 2.2 00203 + + + + + + + + + + + _ Ummm 00 22 02 £385 + + + + + + + + + + + _ 00:0 8 22 5 3005 + + + + + + + + + + + _ Ummm 2. - 02 mQNZm + + + + + + + + + + + _ 8:5 mm - 02 $230 + + + + + + + + + + + _ 09$ 00 0 02 $-00 + + + + + + + + + + + _ 0.005 00 .22 02 2350 + + + + + + + + + + + 2 Ommm m0 2. 02 02: + + + + + + + + + + + _ Ummm m0 2. 2.2 wmg + + + + + + + + + + + _ UmIm 00 2. 2.2 7M0 + + + + + + + + + + + _ Ummm m0 0 02 5220 + + + + + + + + + + + 00m _ Ummm 00 n 02 :73 + + + + + + + + + + + 000.23% _ Ommm 00 2. 2.2 080m + + + + + + + + + + + 2.29 _ UB5 00 0 02 mma -Aom k3» 93.20 @320 V3.20 93.20 :20 2.3.20 $50 940% 1.2%..» 0250 9.9.0 9309 Hm I O Eabm .0020 0003800 0:0 850% 0:20 .3 00:00 000 058% 02:03 20 0:30.00 0>00w0: At 0 .3 00000205 0:0 0:30: 0300052 0000302 .2200 .m 2:05:00: .20 000 0 :80 00:0w Nut—.m— _0=0_>_0:_ .20 000000 M00 .20 33.00% .0 030,—. 74 + + + + N 08$ 02: 22 222 m2N2v + + + + N 05$ 82 S22 2 2 2 3.33. + + + + N 085 82 w 222 N220 + + + + 20m N 09.5 82 w 2 2 2 36 2Nm + + + + 222C N 05$ 82 w 22 2 mm 0mm + + . + + N ommm 02: $22 0N 2.23 + + + + N Ummm 02: - 0N $220 + + + + N Dmmm 82 22 0N N-mam + + + + 00m N Him 82 N 0N <22th + + + + 08 N 00220 82 E 8 $-88 + + + + + + + + + + + 200%0025 EN N mm 3 -30.: + + + + + + + + + + 2 Ummm mN N mm No -395 + + + + + + + + + + 2 USE 0N N mm on 85 + + + + 2 Ummm mN 2. mm 20m Umo + + + + 2 ommm mN N mm on own + + + + + + + + + + + 2 USE MN N mm pm 0mm + + + + + + + + + + + 2 09.5 MN N mm an own + + + + 2 Ummm «N N mm t><-«<«-<-« <1=<=<1za<=<1= epr eivH epa ezv 5905,MDCH-10 __ t A B (055:H7’s) KJIH 9:1 S RQPO J I C A E G F S‘Xmsmve <12<1<1<2 c>1><22<1<1<1<1i> %{¢‘¢‘_¢u44 W epr eivH epa » ezv 026and0111 — etrA F Strains KJIH vi SRQPOJICA EGF N°“'°‘57STEC «new mwmm W Figure 10. Results of PCR screen of individual genes in ETT2 showing different types of deletion patterns. All solid gray shaded arrows represent genes with positive PCR detection results. Genes negative for PCR screening are shaded in a grey checkerboard pattern, while genes not tested are shown in white. (A) Profile of positive results for 0157:H7 and 0157:H- isolates. All 0157’s were positive for all genes tested. (B) Two of three Stx2 positive 055:H7 strains were also positive for all genes tested, just like 0157:H7 isolates. (C) A single stx2 positive 055:H7 strain had a distinct deletion type, missing the epr gene cluster and part of the epa cluster. (D) Stx2 negative 055:H7’s results showing a different truncation than seen in non-0157 STEC strains, which is missing a majority of the genes examined. (E) Some Stx2-negative 055:H7 isolated showed the same profile as 0157 strains, except for a single deletion in epaS. (B) The truncation seen in previous work with ETT2 in non-0157 STEC isolates was also found in this study. 77 6020282 220022 822 .2ow2mmvv =o2m20> 02288-5: 0222 25020208020 05 422:0 020282 E02008 .2220 c000 2042 8: 09$ 0 3:000:28 m< 25222022 3 5qu +30 :18 £50 82% 50 EN .005 $5.25: 2.5... 202 0:622: .228 5° 8&0 5,28 has a 25:32:22. 2000:2823 32 :0>2w 2:020:20 mass 00 20258 0222 so 223920 3 20:02.02 N .2. Hm 0222 .20 E222 282220232000 022,2. 222$ 0020282 Nmnmmo 2022000220 E02 22 EB NINEO Ummm mo 0:020 02=0wo222ma 2:0on 0222 .8 0802080 05 28 202008 28223230 0022:202w .22 0.22—M202 AAU nvzvfiv kbmwo2somo~2 :2N02 . >20 0&0 420' nvflk oVflmvflwV (Nani. Um Um: m Hm 222.030 00.20 2.0085 30 Nw1mmOH1— .0m DMD 3:030 «0.20 2.885 cmNmuba I I I I _ .2 o .0 w ,0 2 N 0292 Wm. I: N éomézo 30 0mm Ihwml 65:30.25: 2-:0:E.3% m .5 \ NIUWmO / Nair—m 32m: .20 PM 2.0025 SEC 0w02222 2., m E New _ _ m5. .0 new»: at +~20w 23200 D N Fm ed. 20 M02320 3 M20. yaww 02E 2 8:5 8222-00.20 202200052. 2.3—um 2km 78 DISCUSSION Recent analysis of the genetic composition of ETT2 has shown a number of small deletions, frameshift mutations and premature stop codons [41]. This previous work, and the data presented here demonstrate a degree of genetic erosion in relatives of modern 0157:H7 isolates, which contain various deletions. Despite this level of genetic attrition, some residual expression of ETT2 genes has been detected, which apparently plays a role in enhancing virulence [143, 144]. Since ETT2 may serve as a backup copy of a functional TTSS in E. coli, it is possible these isolates had, at one time, an entire ETT2 island then lost a major portion, along with a number of ORFs upstream, as a result of a deletion. This work demonstrates the effect of genetic erosion on an apparently non— functional secretion system in strains representing the genetic ancestry of modern 0157:H7. This erosion may be ongoing in a number of 055:H7 isolates. It is unclear why 0157:H7 has retained an apparently intact ETT2 island, but this element may play a role in regulating the expression of other genes, rather than producing its own secretion system. The type III secretion system encoded by LEE contains a number of effector molecules that are secreted through the needle apparatus [36]. Initial genomic analysis of ETT2 indicates this island may lack any apparent effectors [41]. This may explain, in part, why ETT2 is undergoing selective pressure to be deleted. Because it is complicated and requires a number of genes, the type III secretion machinery is energetically- expensive to assemble. Without a collection of effector molecules to be secreted into host cells, any needle apparatus that may have been encoded by ETT2 would have allowed for 79 the assembly of a complicated structure without any function. This is roughly akin to a house or other building without any occupants - it was expensive to build, but is now without firnction because of a lack of occupants. The search for potential effector molecules for ETT2 is still underway. With the LEE-encoded TTSS, there have been a number of effectors identified that are encoded outside of the LEE island, but use LEE for secretion [147, 148]. Previous work with ETT2 identified potential effector molecules in an island dubbed EIP [41]. This island has curiously only been found in the enteroaggregative E. coli strain 042, which also contains a complete ETT2 (Figure 12). It is unclear why 0157:H7 strains have an intact ETT2, but lack EIP. The potential effectors encoded in EIP, or other yet-uncharacterized effectors elsewhere may in fact work in conjunction with ETT2 to form a functional TTSS unit in some strains, although this has not yet been demonstrated. Despite the possibility of non-ETT2 encoded effectors, none have yet been found in any strain of E. coli except 042. It is possible only strain 042 and a small number of other strains acquired EIP, leaving most strains with ETT2 without any functional effectors. This could be one reason for the frequent cases of genetic erosion outlined in this work and elsewhere. Obviously a cell copying large regions of DNA spanning 20-40 kb that encode no functional system or tools for survival is wasting valuable energy and resources. There must be strong selective pressures on these types of regions for either conversion to a functional form by recombination or most likely removal by deletion. While acquisition of genetic material can be a means of evolution, the removal of non- functional DNA may also help streamline the genome and enhance fitness. 80 eip eicA B X D eiIA ean 042 (EAggEC) QEDW Ex: 11% C: >{><:J 0157:H7 (Sakai) Figure 12. The EIP island from Enteroaggregative (EAggEC) strain 042, encodes possible ETT2 effectors, including an intimin homologue (ean). EIP has been identified only in a handfirl of strains, and is absent in all 0157:H7 strains, including Sakai. 81 ACKNOWLEDGEMENTS Special thanks to Dr. David Lacher for assistance in the design of the ETT2 primers, and for initial optimization advice. Thanks are also due to Lindsey Ouellette for MLST sequencing of the 0157:H7 isolates used in this study. This work was supported by a USDA National Needs Fellowship, NIH, and the Microbial Research Unit at Michigan State University. 82 CHAPTER 5 SUMMARY AND SYNTHESIS 83 The genetic diversity of Escherichia coli is dynamic, complex, and driven by a variety of external forces that help shape its evolution. The overall purpose of the research presented here is to characterize a small portion of that dynamic genome, to allow for the more efficient tracking and characterizing of unknown E. coli isolates, and to show in greater detail how gene flux is both a process of horizontal acquisition of mobile elements, and occasionally a loss of genetic material. The need for rapid characterization of isolates becomes important with increasing threats to food safety, and ever-larger outbreaks of foodbome disease. MVGP is one technique that can be used to give an initial “fingerprint” of an isolate’s virulence gene content, and characterize that strain to a particular pathotype within the diversity of E. coli. As shown in Chapter 2, this technique is also useful for characterizing strains isolated from a clinical setting. Because any gene can be added or removed easily, the set of virulence loci examined can be tailored exactly to specific strain sets, or to specific and investigative groups. The reagents and machinery used for MLST are similar or identical to what is necessary to perform MVGP, so labs already doing phylogenetic analysis can adapt their existing reagents and laboratory equipment to virulence profiling without much of a learning curve. Also, while the initial order of Beckman dye-labeled primers is costly, the amount used with each reaction is so small, so a primer may last for hundreds or even thousands of reactions. Aside from the practical aspects of MVGP for use as a clinical tool, the work presented here also illustrates why it is important to screen for E. coli from specific groups of individuals. The microbiota of the gut is extremely diverse and E. coli plays only a small component of the flora under most circumstances. However, it is an 84 important component that can often cause or contribute to disease. It also appears to be highly variable with respect to virulence gene content, partially due to the mobile nature of those elements. Determining the basal level of virulence diversity within patients under different circumstances is important to understand not only the diversity of individual virulence genes, but also to link, statistically and epidemiologically, some of these virulence loci to specific clinical symptoms. Many of the virulence loci examined here are listed as putative, and given the risks posed by many of the pathotypes in which they are found, they are likely to remain so indefinitely. The level of regulation and approvals needed for human trials, especially involving potentially pathogenic organisms is understandably complex. There are also obvious ethical considerations, as many of the virulence genes included here may produce proteins that could significantly affect human health. Despite these necessary restrictions, the molecular characterization of virulence-related loci is important. One of the first steps towards characterization is the determination of the distribution of these loci in both the general population and a variety of specific sub-populations, including those with a variety of diseases. Future considerations. There are a number of future projects that could help expand the work started here. One includes a more detailed look at ST-29 loci, including microarray analysis to compare multiple isolates for genomic content, and also for expression differences. This may help identify any loci that could play a role in survival through enhanced attachment or immune system avoidance. Animal models could be 85 used to ascertain hypotheses that ST—29 strains do indeed have enhanced survival, because as of now this is only an epidemiological association. Recent work with ETT2 has shown that despite many examples of gene loss within this region, and no apparent ETT2-specific effectors identified, some genes within this island are still actively transcribed [144]. Curiously, ETT2 may have an antagonistic effect, on the LEE-encoded TTSS in 0157:H7. Strains with mutational inhibition of ETT2 loci etrA and eivF had both “greatly increased secretion of proteins by the LEE” and “increased adhesin to human epithelial cells” [144]. Exactly why ETT2 appears to be encoding proteins that inhibit LEE-related gene expression is still unclear. ETT2 would be another good candidate for expression microarray analysis to further characterize this phenomenon. It is possible that other loci within ETT2 are expressed, either in different strains, aside from 0157’s, or under different conditions. This expression may have a dramatic effect on other loci, as seen with etrA and eivF in ETT2 [144]. A further characterization of the genes encoded within the EIP island may also shed some more light on this situation. In vivo studies involving a fluorescent actin staining (FAS) assay [149] could show if ETT2 does in fact function as a type III secretion system. The assay works by detecting the polymerized actin filaments characteristic of the close adherence of strains attaching to epithelial cells. This polymerization is a hallmark of the attaching/effacing phenotype of EPEC and EHEC, and indicates type III effectors have entered the host cell to induce physiological changes {6,125,150} An ideal candidate for this would be strain 042 since it has both an intact ETT 2 and EIP and does not contain the other type III island encoded by LEE. Using a strain 86 without LEE is critical, so the actin polymerization phenotype could be attributed to ETT2, and not LEE. Finally, MVGP could easily be either expanded or adapted for use on other microbial pathogens. Simply by selecting a set of PCR primers for specific genes of interest, a labeled pair can be ordered and incorporated into a collection to generate a profile. This technique may be especially useful in screening for virulence profiles of other species with high genomic diversity that are not well characterized regarding virulence genes. Organisms with frequent horizontal gene exchange would likely benefit most from virulence profiling because this technique allows for the tracking of numerous virulence loci across the diversity of the species, as it has with E. coli. Evolution in all types of E. coli appears to be a dynamic process, with many documented cases of horizontal acquisition of bacteriophages [27, 151], plasmids [152- 154], and even large chromosomal pathogenicity islands [155, 156]. Even amongst strains of the same serotype, such as the 0157:H7’s, increased Virulence is seen between distinct groups, or clades [157]. The 0157:H7 strains responsible for the recent outbreaks on contaminated spinach [158, 159] and lettuce [158, 160] in late 2006 had enhanced virulence compared to 0157:H7 isolates from past outbreaks (Sakai from 1996 or EDL- 933 from 1982). Virulence was measured by the rate of hospitalization and development of HUS. Both of these markers were significantly higher in the spinach and lettuce outbreak strains, which belong to clade 8, than previous 0157:H7 outbreak strains Sakai (clade 1) and EDL-933 (clade 3) [157]. This is evidence for ongoing evolution that appears to be favoring enhanced virulence, possibly as a means to spread more effectively. It is also possible that other serotypes of E. coli that are not frequently 87 detected in the United States may become more prevalent here in the future. Geographical preferences appear to exist in the distribution of certain serotypes, but this may change given today’s global environment. In the future, it seems reasonable to expect new outbreaks of highly virulent serotypes of E. coli, like 0157:H7, to result in elevated levels of HUS, hospitalization, and death in affected individuals. Therefore, the urgency for rapid and coordinated identification of outbreaks should grow in the future, whether applied to food safety for the consumer and industry or to bioterrorism. 88 REFERENCES 89 10. 11. Kaper, J .B., J .P. Nataro, and H.L. Mobley, Pathogenic Escherichia coli. Nat Rev Microbiol, 2004. 2(2): p. 123-40. Donnenberg, MS. and TS. Whittam, Pathogenesis and evolution of virulence in enteropathogenic and enterohemorrhagic Escherichia coli. J Clin Invest, 2001. 107(5): p. 539-48. Tobe, T., et al., Complete DNA sequence and structural analysis of the enteropathogenic Escherichia coli adherence factor plasmid. Infect Immun, 1999. 67(10): p. 5455-62. Navarro-Garcia, F., et al., The serine protease motif of EspC from enteropathogenic Escherichia coli produces epithelial damage by a mechanism different from that of Pet toxin from enteroaggregative E. coli. Infect Immun, 2004. 72(6): p. 3609-21. Mellies, J .L., et al., espCpathogenicity island of enteropathogenic Escherichia coli encodes an enterotoxin. Infect Immun, 2001. 69(1): p. 315-24. Donnenberg, M.S., J .B. Kaper, and BB. Finlay, Interactions between enteropathogenic Escherichia coli and host epithelial cells. Trends Microbiol, 1997. 5(3): p. 109-14. Donnenberg, M.S., L.C. Lai, and K.A. Taylor, The locus of enterocyte eflacement pathogenicity island of enteropathogenic Escherichia coli encodes secretion functions and remnants of transposons at its extreme right end. Gene, 1997. 184(1): p. 107-14. Cleary, T.G., The role of Shiga-toxin-producing Escherichia coli in hemorrhagic colitis and hemolytic uremic syndrome. Semin Pediatr Infect Dis, 2004. 15(4): p. 260-5. Burland, V., et al., The complete DNA sequence and analysis of the large virulence plasmid of Escherichia coli 015 7.°H7. Nucleic Acids Res, 1998. 26(18): p. 4196-204. Tatsuno, I., et al., toxB gene on p015 7 of enterohemorrhagic Escherichia coli 015 7.'H7 is required for full epithelial cell adherence phenotype. Infect Immun, 2001. 69(1 1): p. 6660-9. Merritt, E.A., et al., Crystal structure of cholera toxin B-pentamer bound to receptor GM] pentasaccharide. Protein Sci, 1994. 3(2): p. 166-75. 90 12. 13. 14. 15. l6. 17. 18. 19. 20. 21. 22. 23. Lencer, W.I., et al., Targeting of cholera toxin and Escherichia coli heat labile toxin in polarized epithelia: role of COOH-terminal KDEL. J Cell Biol, 1995. 131(4): p. 951-62. Rodighiero, C., et al., Structural basis for the diflerential toxicity of cholera toxin and Escherichia coli heat-labile enterotoxin. , Construction of hybrid toxins identifies the A2-domain as the determinant of differential toxicity. J Biol Chem, 1999. 274(7): p. 3962-9. Sansonetti, P.J., et al., Molecular comparison of virulence plasmids in Shigella and enteroinvasive Escherichia coli. Ann Microbiol (Paris), 1983. 134A(3): p. 295-318. Harris, J .R., et al., High-molecular-weight plasmid correlates with Escherichia coli enteroinvasiveness. Infect Immun, 1982. 37(3): p. 1295-8. Navarro-Garcia, F ., et al., Cytoskeletal effects induced by pet, the serine protease enterotoxin of enteroaggregative Escherichia coli. Infect Immun, 1999. 67(5): p. 2184-92. Eslava, C., et al., Pet, an autotransporter enterotoxin from enteroaggregative Escherichia coli. Infect Immun, 1998. 66(7): p. 3155-63. Vila, J ., et al., Enteroaggregative escherichia coli virulence factors in traveler's diarrhea strains. J Infect Dis, 2000. 182(6): p. 1780-3. F asano, A., et al., Effect of Shigella enterotoxin I (ShE TI) on rabbit intestine in vitro and in vivo. Gut, 1997. 40(4): p. 505-11. F asano, A., et al., Shigella enterotoxin 1: an enterotoxin of Shigella flexneri 2a active in rabbit small intestine in vivo and in vitro. J Clin Invest, 1995. 95(6): p. 2853-61. Savarino, S.J., et al., Enteroaggregative Escherichia coli heat-stable enterotoxin is not restricted to enteroaggregative E. coli. J Infect Dis, 1996. 173(4): p. 1019- 22. Itoh, Y., et al., Laboratory investigation of enteroaggregative Escherichia coli 0 untypeablesHI 0 associated with a massive outbreak of gastrointestinal illness. J Clin Microbiol, 1997. 35(10): p. 2546-50. Cobeljic, M., et al., Enteroaggregative Escherichia coli associated with an outbreak of diarrhoea in a neonatal nursery ward. Epidemiol Infect, 1996. 117(1): p. 11-6. 91 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. Nataro, J .P., et al., Heterogeneity of enteroaggregative Escherichia coli virulence demonstrated in volunteers. J Infect Dis, 1995. 171(2): p. 465-8. Mathewson, J .J ., et al., Pathogenicity of enteroadherent Escherichia coli in adult volunteers. J Infect Dis, 1986. 154(3): p. 524-7. Herold, S., H. Karch, and H. Schmidt, Shiga toxin-encoding bacteriophages-- genomes in motion. Int J Med Microbiol, 2004. 294(2-3): p. 115-21. Schmidt, H., Shiga-toxin-converting bacteriophages. Res Microbiol, 2001. 152(8): p. 687-95. O'Brien, A.D., et al., Shiga toxin: biochemistry, genetics, mode of action, and role in pathogenesis. Curr Top Microbiol Immunol, 1992. 180: p. 65-94. Ewald, P.W., Guarding against the most dangerous emerging pathogens. Emerg Infect Dis, 1996. 2(4): p. 245-57. Wells, J .G., et al., Laboratory investigation of hemorrhagic colitis outbreaks associated with a rare Escherichia coli serotype. J Clin Microbiol, 1983. 18(3): p. 512-20. Riley, L.W., et al., Hemorrhagic colitis associated with a rare Escherichia coli serotype. N Engl J Med, 1983. 308(12): p. 681-5. Mead, P.S., et al., F ood-related illness and death in the United States. Emerg Infect Dis, 1999. 5(5): p. 607-25. ' Karmali, M.A., Infection by Shiga toxin-producing Escherichia coli: an overview. Mol Biotechnol, 2004. 26(2): p. 117-22. Pema, N.T., et al., Genome sequence of enterohaemorrhagic Escherichia coli 015 7:H7. Nature, 2001. 409(6819): p. 529-33. Welch, R.A., et al., Extensive mosaic structure revealed by the complete genome sequence of uropathogenic Escherichia coli. Proc Natl Acad Sci U S A, 2002. 99(26): p. 17020-4. Elliott, S.J., et al., The complete sequence of the locus of enterocyte effacement (LEE) from enteropathogenic Escherichia coli E2348/69. Mol Microbiol, 1998. 28(1): p. 1-4. Sperandio, V., et al., Characterization of the locus of enterocyte effacement (LEE) in different enteropathogenic Escherichia coli (EPEC) and Shiga-toxin producing Escherichia coli (ST EC) serotypes. FEMS Microbiol Lett, 1998. 164(1): p. 133-9. 92 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. Hartleib, S., et al., Prevalence of the new, SPII-like, pathogenicity island E TT 2 among Escherichia coli. Int J Med Microbiol, 2003. 292(7-8): p. 487-93. Makino, S., et al., Distribution of the secondary type III secretion system locus found in enterohemorrhagic Escherichia coli 01 5 7:H 7 isolates among Shiga toxin-producing E. coli strains. J Clin Microbiol, 2003. 41(6): p. 2341-7. Galan, J .E. and H. Wolf-Watz, Protein delivery into eukaryotic cells by type III secretion machines. Nature, 2006. 444(7119): p. 567-73. Ren, C.P., et al., The E TT 2 gene cluster, encoding a second type III secretion system from Escherichia coli, is present in the majority of strains but has undergone widespread mutational attrition. J Bacteriol, 2004. 186(11): p. 3547- 60. Galan, J .E. and A. Collmer, Type III secretion machines: bacterial devices for protein delivery into host cells. Science, 1999. 284(5418): p. 1322-8. Pallen, M.J., R.R. Chaudhuri, and LR. Henderson, Genomic analysis of secretion systems. Curr Opin Microbiol, 2003. 6(5): p. 519-27. Ochman, H. and R.K. Selander, Standard reference strains of Escherichia coli from natural populations. J Bacteriol, 1984. 157(2): p. 690-3. Welinder-Olsson, C., et al., Genetic profiling of enterohemorrhagic Escherichia coli strains in relation to clonality and clinical signs of infection. J Clin Microbiol, 2002. 40(3): p. 959-64. ' Obata-Yasuoka, M., et al., Vaginal Escherichia coli share common virulence factor profiles, serotypes and phylogeny with other extraintestinal E. coli. Microbiology, 2002. 148(Pt 9): p. 2745-52. Tarr, C.L., et al., Molecular characterization of a serotype 0121.°H19 clone, a distinct Shiga toxin-producing clone of pathogenic Escherichia coli. Infect Immun, 2002. 70(12): p. 6853-9. F eil, E.J., et al., Estimating recombinational parameters in Streptococcus pneumoniae from multilocus sequence typing data. Genetics, 2000. 154(4): p. 1439-50. Kotetishvili, M., et al., Multilocus sequence typing for characterization of clinical and environmental salmonella strains. J Clin Microbiol, 2002. 40(5): p. 1626-35. Nicolas, P., et al., The 1998 Senegal epidemic of meningitis was due to the clonal expansion of A :4.'P1 .9, clone 111-], sequence type 5 Neisseria meningitidis strains. J Clin Microbiol, 2000. 38(1): p. 198-200. 93 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. Zhou, J ., M.C. Enright, and B.G. Spratt, Identification of the major Spanish clones of penicillin-resistant pneumococci via the Internet using multilocus sequence typing. J Clin Microbiol, 2000. 38(3): p. 977-86. Kriz, P., J. Kalmusova, and J. Felsberg, Multilocus sequence typing of Neisseria meningitidis directly from cerebrospinal fluid. Epidemiol Infect, 2002. 128(2): p. 157-60. Urwin, R. and MC. Maiden, Multi-locus sequence typing: a tool for global epidemiology. Trends Microbiol, 2003. 11(10): p. 479-87. Bizzarri, M.F., A. Prabhakar, and AH. Bishop, Multiple-Locus Sequence Typing Analysis of Bacillus thuringiensis Recovered from the Phylloplane of Clover ( T rifolium hybridum) in Vegetative Form. Microb Ecol, 2007. Tarr, CL. and TS. Whittam, Molecular evolution of the intimin gene in 01]] clones of pathogenic Escherichia coli. J Bacteriol, 2002. 184(2): p. 479-87. Lacher, D.W., et al., Molecular evolution of typical enteropathogenic Escherichia coli: clonal analysis by multilocus sequence typing and virulence gene allelic profiling. J Bacteriol, 2007. 189(2): p. 342-50. Walk, S.T., et al., Genetic diversity and population structure of Escherichia coli isolated from freshwater beaches. Environ Microbiol, 2007. 9(9): p. 2274-88. Johnson, J .R., et al., Molecular comparison of extraintestinal Escherichia coli isolates of the same electrophoretic lineages from humans and domestic animals. J Infect Dis, 2001. 183(1): p. 154-9. Hyma, K.E., et al., Evolutionary genetics of a new pathogenic Escherichia species: Escherichia albertii and related Shigella boydii strains. J Bacteriol, 2005. 187(2): p. 619-28. Tarr, C.L., et al., Molecular characterization reveals similar virulence gene content in unrelated clonal groups of Escherichia coli of serogroup 01 74 (0X3). J Bacteriol, 2008. 190(4): p. 1344-9. Stapp, J.R., et al., Comparison of Escherichia coli 015 7.'H 7 antigen detection in stool and broth cultures to that in sorbitol-MacConkey agar stool cultures. J Clin Microbiol, 2000. 38(9): p. 3404-6. Klein, E.J., et al., Diarrhea etiology in a Children's Hospital Emergency Department: a prospective cohort study. Clin Infect Dis, 2006. 43(7): p. 807-13. 94 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. Klein, E.J., et al., Shiga toxin-producing Escherichia coli in children with diarrhea: a prospective point-of-care study. J Pediatr, 2002. 141(2): p. 172-7. Monteiro-Neto, V., et al., Virulence properties of Escherichia coli 0111:H12 strains. FEMS Microbiol Lett, 1997. 146(1): p. 123-8. F eng, P., S.D. Weagant, and SR. Monday, Genetic analysis for virulence factors in Escherichia coli 0104:H21 that was implicated in an outbreak of hemorrhagic colitis. J Clin Microbiol, 2001. 39(1): p. 24-8. Stacy-Phipps, S., J .J . Mecca, and J .B. Weiss, Multiplex PCR assay and simple preparation method for stool specimens detect enterotoxigenic Escherichia coli DNA during course of infection. J Clin Microbiol, 1995. 33(5): p. 1054-9. Schubert, S., et al., Prevalence of the "high-pathogenicity island " of Yersinia species among Escherichia coli strains that are pathogenic to humans. Infect Immun, 1998. 66(2): p. 480-5. Paton, A.W. and J .C. Paton, Direct detection and characterization of Shiga toxigenic Escherichia coli by multiplex PCR for sth, stx2, eae, eth, and saa. J Clin Microbiol, 2002. 40(1): p. 271-4. Kumar, S., et al., ME GA2: molecular evolutionary genetics analysis software. Bioinformatics, 2001. 17(12): p. 1244-5. Okeke, I.N., et al., Molecular epidemiology of the iron utilization genes of enteroaggregative Escherichia coli. J Clin Microbiol, 2004. 42(1): p. 36-44. Karch, H., et al., A genomic island, termed high-pathogenicity island, is present in certain non-015 7 Shiga toxin-producing Escherichia coli clonal lineages. Infect Immun, 1999. 67(11): p. 5994-6001. Oswald, E., et al., Typing of intimin genes in human and animal enterohemorrhagic and enteropathogenic Escherichia coli: characterization of a new intimin variant. Infect Immun, 2000. 68(1): p. 64-71. Brunder, W., et al., The large plasmids of Shiga—toxin-producing Escherichia coli (ST EC) are highly variable genetic elements. Microbiology, 1999. 145 ( Pt 5): p. 1005-14. Velarde, J .J . and J .P. Nataro, Hydrophobic residues of the autotransporter EspP linker domain are important for outer membrane translocation of its passenger. J Biol Chem, 2004. 279(30): p. 31495-504. 95 75. 76. 77. 78. 79. 80. 81. 82. 83. 84. 85. 86. McNally, A., et al., Differences in levels of secreted locus of enterocyte eflacement proteins between human disease-associated and bovine Escherichia coli 015 7. Infect Immun, 2001. 69(8): p. 5107-14. Brunder, W., H. Schmidt, and H. Karch, EspP, a novel extracellular serine protease of enterohaemorrhagic Escherichia, coli 01 5 7:H 7 cleaves human coagulation factor V. Mol Microbiol, 1997. 24(4): p. 767-78. Bielaszewska, M., et al., Characterization of cytolethal distending toxin genes and expression in shiga toxin-producing Escherichia coli strains of non-015 7 serogroups. Infect Immun, 2004. 72(3): p. 1812-6. Bielaszewska, M., et al., Cytolethal distending toxin from Shiga toxin-producing Escherichia coli 015 7 causes irreversible G2/M arrest, inhibition of proliferation, and death of human endothelial cells. Infect Immun, 2005. 73(1): p. 552-62. Nesic, D., Y. Hsu, and CE. Stebbins, Assembly and function of a bacterial genotoxin. Nature, 2004. 429(6990): p. 429—33. Pickett, C.L., et al., Patterns of variations in Escherichia coli strains that produce cytolethal distending toxin. Infect Immun, 2004. 72(2): p. 684-90. Torres, AG. and SM. Payne, Haem iron-transport system in enterohaemorrhagic Escherichia coli 01 5 7:H 7. Mol Microbiol, 1997. 23(4): p. 825-33. Nagy, G., et al., Expression of hemin receptor molecule ChuA is influenced by RfaH in uropathogenic Escherichia coli strain 536. Infect Immun, 2001. 69(3): p. 1924-8. Tarr, P.I., et al., Iha: a novel Escherichia coli 015 7:H7 adherence-conferring molecule encoded on a recently acquired chromosomal island of conserved structure. Infect Immun, 2000. 68(3): p. 1400-7. Taylor, D.E., et al., Genomic variability of 0 islands encoding tellurite resistance in enterohemorrhagic Escherichia coli 015 7:H7 isolates. J Bacteriol, 2002. 184(17): p. 4690-8. Heimer, S.R., et al., Urease of enterohemorrhagic Escherichia coli: evidence for regulation by fur and a trans-acting factor. Infect Immun, 2002. 70(2): p. 1027- 3 1 . D'Orazio, SE. and CM. Collins, The plasmid-encoded urease gene cluster of the family Enterobacteriaceae is positively regulated by UreR, a member of the AraC family of transcriptional activators. J Bacteriol, 1993. 175(11): p. 3459-67. 96 87. 88. 89. 90. 91. 92. 93. 94. 95. 96. 97. 98. Wang, G., C.G. Clark, and F.G. Rodgers, Detection in Escherichia coli of the genes encoding the major virulence factors, the genes defining the 01 5 7 :H 7 serotype, and components of the type 2 Shiga toxin family by multiplex PCR. J Clin Microbiol, 2002. 40(10): p. 3613-9. De Baets, L., et al., Genetic typing of shiga toxin 2 variants of Escherichia coli by PCR-restriction fragment length polymorphism analysis. Appl Environ Microbiol, 2004. 70(10): p. 6309-14. Eklund, M., K. Leino, and A. Siitonen, Clinical Escherichia coli strains carrying stx genes: stx variants and stx—positive virulence profiles. J Clin Microbiol, 2002. 40(12): p. 4585-93. Collazo, CM. and J .E. Galan, Requirement for exported proteins in secretion through the invasion-associated type III system of Salmonella typhimurium. Infect Immun, 1996. 64(9): p. 3524-31. Crago, A.M. and V. Koronakis, Salmonella InvG forms a ring-like multimer that requires the Ian lipoprotein for outer membrane localization. Mol Microbiol, 1998. 30(1): p. 47-56. Hwang, J ., et al., Structural and topographical studies of the type IV bundle- forming pilus assembly complex of enteropathogenic Escherichia coli. J Bacteriol, 2003. 185(22): p. 6695-701. Green, B.A., et al., Evidence that a new enterotoxin of Escherichia coli which activates adenylate cyclase in eucaryotic target cells is not plasmid mediated. Infect Immun, 1983. 41(1): p. 383-90. Yamamoto, T. and T. Yokota, Release of heat-labile enterotoxin subunits by Escherichia coli. J Bacteriol, 1982. 150(3): p. 1482-4. Wyckoff, E.E., et al., Structure of the Shigella dysenteriae haem transport locus and its phylogenetic distribution in enteric bacteria. Mol Microbiol, 1998. 28(6): p. 1139-52. Mills, M. and SM. Payne, Identification of shuA, the gene encoding the heme receptor of Shigella dysenteriae, and analysis of invasion and intracellular multiplication of a shuA mutant. Infect Immun, 1997. 65(12): p. 5358-63. Henderson, I.R., et al., Characterization of pic, a secreted protease of Shigella flexneri and enteroaggregative Escherichia coli. Infect Immun, 1999. 67(11): p. 5587-96. Guyer, D.M., et al., Identification of sat, an autotransporter toxin produced by uropathogenic Escherichia coli. Mol Microbiol, 2000. 38(1): p. 53-66. 97 99. 100. 101. 102. 103. 104. 105. 106. 107. 108. 109. 110. 111. Guyer, D.M., et al., Sat, the secreted autotransporter toxin of uropathogenic Escherichia coli, is a vacuolating cytotoxin for bladder and kidney epithelial cells. Infect Immun, 2002. 70(8): p. 4539-46. Dutta, P.R., et al., Functional comparison of serine protease autotransporters of enterobacteriaceae. Infect Immun, 2002. 70(12): p. 7105-13. Al-Hasani, K., et al., Distribution and structural variation of the she pathogenicity island in enteric bacterial pathogens. J Med Microbiol, 2001. 50(9): p. 780-6. Al-Hasani, K., et al., The sigA gene which is borne on the she pathogenicity island of Shigella flexneri 2a encodes an exported cytopathic protease involved in intestinal fluid accumulation. Infect Immun, 2000. 68(5): p. 2457-63. Venkatesan, M.M., et al., Complete DNA sequence and analysis of the large virulence plasmid of Shigella flexneri. Infect Immun, 2001. 69(5): p. 3271-85. Rode, C.K., et al., T ype-specific contributions to chromosome size differences in Escherichia coli. Infect Immun, 1999. 67(1): p. 230-6. Huang, S.H., et al., Escherichia coli invasion of brain microvascular endothelial cells in vitro and in vivo: molecular cloning and characterization of invasion gene ibe10. Infect Immun, 1995. 63(11): p. 4470-5. Litwin, CM. and SB. Calderwood, Role of iron in regulation of virulence genes. Clin Microbiol Rev, 1993. 6(2): p. 137-49. Grandjean, D., et al., Iron uptake is essential for Escherichia coli survival in drinking water. Lett Appl Microbiol, 2006. 43(1): p. 111-7. Xie, Y., et al., Identification and characterization of Escherichia coli RS218- derived islands in the pathogenesis of E. coli meningitis. J Infect Dis, 2006. 194(3): p. 358-64. Johnson, J .R., et al., Phylogenetic distribution of extraintestinal virulence- associated traits in Escherichia coli. J Infect Dis, 2001. 183(1): p. 78-88. Ostroff, S.M., et al., A statewide outbreak of Escherichia coli 015 7:H7 infections in Washington State. Am J Epidemiol, 1990. 132(2): p. 239-47. Tarr, P.I., et al., Escherichia coli 015 7:H7 and the hemolytic uremic syndrome: importance of early cultures in establishing the etiology. J Infect Dis, 1990. 162(2): p. 553-6. 98 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124. Tarr, P.I., N.T. Tran, and RA. Wilson, Escherichia coli 0157:H7 in retail ground beef in Seattle: results of a one-year prospective study. J Food Prot, 1999. 62(2): p. 133-9. Haack, J .P., et al., Escherichia coli 015 7 exposure in Wyoming and Seattle: serologic evidence of rural risk. Emerg Infect Dis, 2003. 9(10): p. 1226-31. Denno, D.M., et al., Explaining unexplained diarrhea and associating risks and infections. Anim Health Res Rev, 2007. 8(1): p. 69-80. Nataro, J .P. and J .B. Kaper, Diarrheagenic Escherichia coli. Clin Microbiol Rev, 1998. 11(1): p. 142-201. Achtman, M., et al., Six widespread bacterial clones among Escherichia coli K 1 isolates. Infect Immun, 1983. 39(1): p. 315-35. Bloch, CA. and CK. Rode, Pathogenicity island evaluation in Escherichia coli K1 by crossing with laboratory strain K-12. Infect Immun, 1996. 64(8): p. 3218- 23. Hochhut, B., et al., Role of pathogenicity island-associated integrases in the genome plasticity of uropathogenic Escherichia coli strain 53 6. Mol Microbiol, 2006. 61(3): p. 584-95. Johnson, J.R., et al., Phylogenetic and pathotypic similarities between Escherichia coli isolates from urinary tract infections in dogs and extraintestinal infections in humans. J Infect Dis, 2001. 183(6): p. 897-906. Johnson, J .R., et al., Virulence genotypes and phylogenetic background of Escherichia coli serogroup 06 isolates from humans, dogs, and cats. J Clin Microbiol, 2008. 46(2): p. 417-22. Oelschlaeger, T.A., U. Dobrindt, and J. Hacker, Virulence factors of uropathogens. Curr Opin Urol, 2002. 12(1): p. 33-8. Martin, H.M., et al., Enhanced Escherichia coli adherence and invasion in Crohn 's disease and colon cancer. Gastroenterology, 2004. 127(1): p. 80-93. Adlerberth, I., et al., Adhesins of Escherichia coli associated with extra-intestinal pathogenicity confer binding to colonic epithelial cells. Microb Pathog, 1995. 18(6): p. 373-85. Jores, J ., L. Rumer, and L.H. Wieler, Impact of the locus of enterocyte eflacement pathogenicity island on the evolution of pathogenic Escherichia coli. Int J Med Microbiol, 2004. 294(2-3): p. 103-13. 99 125. 126. 127. 128. 129. 130. 131. 132. 133. 134. 135. 136. Mota, LI. and GR. Comelis, The bacterial injection kit: type III secretion systems. Ann Med, 2005. 37(4): p. 234-49. Clarke, SC, et al., Virulence of enteropathogenic Escherichia coli, a global pathogen. Clin Microbiol Rev, 2003. 16(3): p. 365-78. Zaharik, M.L., et al., Delivery of dangerous goods: type III secretion in enteric pathogens. Int J Med Microbiol, 2002. 291(8): p. 593-603. Elliott, S.J., J. Yu, and J .B. Kaper, The cloned locus of enterocyte effacement from enterohemorrhagic Escherichia coli 01 5 7 :H 7 is unable to confer the attaching and eflacing phenotype upon E. coli K—12. Infect Immun, 1999. 67(8): p. 4260-3. Galan, J .E., Salmonella interactions with host cells: type III secretion at work. Annu Rev Cell Dev Biol, 2001. 17: p. 53-86. Klein, J .R., T.F. Fahlen, and ED. Jones, Transcriptional organization and function of invasion genes within Salmonella enterica serovar T yphimurium pathogenicity island 1, including the prgH, prgI, prgJ, prgK, orgA, orgB, and orgC genes. Infect Immun, 2000. 68(6): p. 3368-76. Kubori, T., et al., Molecular characterization and assembly of the needle complex of the Salmonella typhimurium type III protein secretion system. Proc Natl Acad Sci U S A, 2000. 97(18): p. 10225-30. Collazo, C.M., M.K. Zierler, and J .E. Galan, Functional analysis of the Salmonella typhimurium invasion genes invl and ian and identification of a target of the protein secretion apparatus encoded in the inv locus. Mol Microbiol, 1995. 15(1): p. 25-38. Russmann, H., et al., Molecular and functional analysis of the type III secretion signal of the Salmonella enterica Ian protein. Mol Microbiol, 2002. 46(3): p. 769-79. Eichelberg, K., C.C. Ginocchio, and J .E. Galan, Molecular and functional characterization of the Salmonella typhimurium invasion genes ian and invC: homology of InvC to the FOF1 AT Pase family of proteins. J Bacteriol, 1994. 176(15): p. 4501-10. Ehrbar, K., et al., Ian is required for type III-dependent secretion of SopA in Salmonella enterica serovar T yphimurium. J Bacteriol, 2004. 186(4): p. 1215-9. Ginocchio, CC. and J .E. Galan, Functional conservation among members of the Salmonella typhimurium InvA family of proteins. Infect Immun, 1995. 63(2): p. 729-32. 100 137. 138. 139. 140. 141. 142. 143. 144. 145. 146. 147. Galan, J .E. and R. Curtiss, 3rd, Distribution of the invA, -B, -C, and -D genes of Salmonella typhimurium among other Salmonella serovars: invA mutants of Salmonella typhi are deficient for entry into mammalian cells. Infect Immun, 1991. 59(9): p. 2901-8. Ginocchio, C.C., et al., Contact with epithelial cells induces the formation of surface appendages on Salmonella typhimurium. Cell, 1994. 76(4): p. 717-24. Ginocchio, C., J. Pace, and J .E. Galan, Identification and molecular characterization of a Salmonella typhimurium gene involved in triggering the internalization of salmonellae into cultured epithelial cells. Proc Natl Acad Sci U S A, 1992. 89(13): p. 5976-80. Kubori, T. and J .E. Galan, Salmonella type III secretion-associated protein InvE controls translocation of eflector proteins into host cells. J Bacteriol, 2002. 184(17): p. 4699-708. Eichelberg, K. and J .E. Galan, Differential regulation of Salmonella typhimurium type III secreted proteins by pathogenicity island 1 (SP1- I)-encoded transcriptional activators InvF and hilA. Infect Immun, 1999. 67(8): p. 4099-105. Daefler, S. and M. Russel, The Salmonella typhimurium Ian protein is an outer membrane lipoprotein required for the proper localization of InvG. Mol Microbiol, 1998. 28(6): p. 1367-80. Ideses, D., et al., A degenerate type III secretion system from septicemic Escherichia coli contributes to pathogenesis. J BacteriOl, 2005. 187(23): p. 8164- 71. Zhang, L., et al., Regulators encoded in the Escherichia coli type III secretion system 2 gene cluster influence expression of genes within the locus for enterocyte eflacement in enterohemorrhagic E. coli 01 5 7:H 7. Infect Immun, 2004. 72(12): p. 7282-93. Feng, P.C., et al., Genetic diversity among clonal lineages within Escherichia coli 015 7:H7 stepwise evolutionary model. Emerg Infect Dis, 2007. 13(11): p. 1701- 6. Wick, L.M., et al., Evolution of genomic content in the stepwise emergence of Escherichia coli 01 5 7:H 7. J Bacteriol, 2005. 187(5): p. 1783-91. Garcia-Angulo, V.A., et al., Regulation of expression and secretion of NleH, a new non-locus of enterocyte effacement-encoded effector in Citrobacter rodentium. J Bacteriol, 2008. 190(7): p. 2388-99. 101 148. 149. 150. 151. 152. 153. 154. 155. 156. 157. 158. 159. Gruenheid, S., et al., Identification and characterization of NleA, a non-LEE- encoded type III translocated virulence factor of enterohaemorrhagic Escherichia coli 0157:H7. Mol Microbiol, 2004. 51(5): p. 1233—49. Shariff, M., N. Bhandari, and S. Saini, F luorescent—actin staining (FAS) test: a novel assay to detect enteropathogenic E. coli. Indian J Pediatr, 1994. 61(6): p. 737-8. Jarvis, K.G., et al., Enteropathogenic Escherichia coli contains a putative type III secretion system necessary for the export of proteins involved in attaching and effacing lesion formation. Proc Natl Acad Sci U S A, 1995. 92(17): p. 7996-8000. Shaikh, N. and RI. Tarr, Escherichia coli 015 7:H7 Shiga toxin-encoding bacteriophages: integrations, excisions, truncations, and evolutionary implications. J Bacteriol, 2003. 185(12): p. 3596-605. Hopkins, K.L., R.H. Davies, and E.J. Threlfall, Mechanisms of quinolone resistance in Escherichia coli and Salmonella: recent developments. Int J Antimicrob Agents, 2005. 25(5): p. 358-73. Johnson, T.J., et al., Plasmid replicon typing of commensal and pathogenic Escherichia coli isolates. Appl Environ Microbiol, 2007. 73(6): p. 1976-83. Maeda, S., et al., Horizontal transfer of nonconjugative plasmids in a colony biofilm of Escherichia coli. FEMS Microbiol Lett, 2006. 255(1): p. 115-20. Mills, D.M., V. Bajaj, and CA. Lee, A 40 kb chromosOmal fragment encoding Salmonella typhimurium invasion genes is absent from the corresponding region of the Escherichia coliK-12 chromosome. Mol Microbiol, 1995. 15(4): p. 749-59. Pema, N.T., et al., Molecular evolution of a pathogenicity island from enterohemorrhagic Escherichia coli 01 5 7:H 7. Infect Immun, 1998. 66(8): p. 3810-7. Manning, S.D., et al., Variation in virulence among clades of Escherichia coli 015 7:H7 associated with disease outbreaks. Proc Natl Acad Sci U S A, 2008. 105(12): p. 4868-73. Cooley, M., et al., Incidence and Tracking of Escherichia coli 01 5 7 :H 7 in a Major Produce Production Region in California. PLoS ONE, 2007. 2(11): p. @1159. Jay, M.T., eta1., Escherichia coli 01 5 7:H 7 in feral swine near spinach fields and cattle, central California coast. Emerg Infect Dis, 2007. 13(12): p. 1908-11. 102 160. Uhlich, G.A., et al., Characterization of Shiga toxin-producing Escherichia coli isolates associated with two multistatefood—borne outbreaks that occurred in 2006. Appl Environ Microbiol, 2008. 74(4): p. 1268-72. 103 milWiAd . . . Rwlil . . R. B 1 . uiL/ . . Yd . 4 T1 w}. 41 6 . . . RM... 11 5 . . . . . N|4l2 Ui.i0 . . EH . . Tilii . . . . Sul‘lmg , . “"2 . . , will 1 . . .. . . Hi . 1 u . . . i . , .. . . . f . 1 t. . . . . . . . . .. . . . .. . . . $2.3.» . . . .. . . .. . ... . . ... . . . . Iainx? J. . . . . . . . . . .. . .