an“. h 2;“... v. 36 z .3. a? .‘awmmw £3, 3 - 1 s 31‘ — “1%“ .3: fl: I, i. x it I . . i 2 39...... 2-361 L... ..I..( it I}... 511.. Elm“. I. ‘2: a? L I .\..n|..\~.i. } .0291... .fil - 9" ‘. LIBRARY Mich.E . “state Umversuy This is to certify that the dissertation entitled INTERPLAY BETWEEN ACID RESISTANCE AND VIRULENCE IN ESCHERICHIA COLI O157:H7 presented by SIVAPRIYA KAILASAN VANAJA has been accepted towards fulfillment of the requirements for the Ph.D degree in Comparative Medicine and Integrative Biology /.7 Major rofessor’s Si ature g// 21%? 9 Date MSU is an Affirmative Action/Equal Opportunity Employer 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 5/08 K IProj/AccaPres/CIRC/DateDue indd INTERPLAY BETWEEN ACID RESISTANCE AND VIRULENCE IN ESCHERICHIA COLI O157:H7 BY Sivapriya Kailasan Vanaja A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Comparative Medicine and Integrative Biology 2009 ABSTRACT INTERPLAY BETWEEN ACID RESISTANCE AND VIRULENCE lN ESCHERICHIA COLI O157:H7 By Sivapriya Kailasan Vanaja Escherichia coli O157:H7 is a food borne pathogen of zoonotic origin that causes hemorrhagic colitis and hemolytic uremic syndrome in humans. Acid resistance (AR) is an essential characteristic of pathogenic E. coli O157:H7 because during its transit from reservoir host, cattle, to humans, 0157 has to survive acidic environments of different food vehicles and the human stomach. E. coli O157:H7 can survive in extreme acidic conditions such as pH 1.5 - 2.5 of human stomach with the help of four AR mechanisms. Glutamate decarboxylase (GAD) system is the most efficient of the four AR mechanisms and is regulated by a central activator, GadE. It is possible that while integrating the laterally acquired genes into the chromosomal regulatory network of E. coli O157:H7, chromosomal regulators such as GadE has evolved into a global regulator with additional functions besides regulating GAD system. However, role of GadE on a genome wide scale remains poorly defined in E. coli O157:H7. The effect of AR induction on the expression of virulence factors of E. coli O157:H7 is not clear. When exposed to acidic conditions, E. coli O157:H7 may downregulate its virulence mechanisms to conserve energy. Moreover, it is possible that differential expression of AR and virulence related genes may be the reason for the variation in infectivity observed between different genotypes of E. coli O157:H7. In this context, the specific aims of this project were: 1) to elucidate the role of GadE in acid resistance and virulence of E. coli O157:H7, 2) to determine the effect of acidic pH on the expression of the LEE pathogenicity island and to understand the role of GAD system regulators in this effect, and 3) to determine whether the differences in infectivity of clinical and bovine-biased genotypes of E. coli O157:H7 is due to the differential expression of virulence and stress fitness associated genes. To address these goals transcriptional profiling of wild type and mutants of GAD system regulators was conducted. Expression studies were also conducted to identify the genes differentially expressed between clinical and bovine-biased genotypes of E. coli O157:H7. Understanding the molecular mechanisms of AR and their effect on virulence and pathogenesis of E. coli O157:H7 could help in developing new methods for diagnosis, prevention and control of this pathogen. Moreover, identifying the differentially expressed genes between different genotypes of E. coli O157:H7 with varying infectivity would help us to identify new targets for vaccine development. Copyright By Sivapriya Kailasan Vanaja 2009 This work is dedicated to the fond memories of Thomas S. Whittam, my mentor. ACKNOWLEDGEMENTS I wish to thank my mentor, Thomas S. Whittam, for being a great advisor and for his valuable advice and support, and for providing the resources and opportunities for me to acquire competence in a broad range of topics. Linda Mansfield, my acting advisor, has been a great support during the final stages of my PhD and I thank her for her guidance. I thank my committee members, John Linz, Robert Britton and the CMIB program director, Vilma Yuzbasian-Gurkan, for their support, suggestions, and guidance. Many of my current and previous lab mates have provided suggestions, support, and technical assistance that were critical to the completion of this research. I thank Teresa Bergholz, Shannon Manning, Galeb Abu-Ali, James Riordan, Adam Nelson, Nicole Milton, Hans Steinsland, Lindsey Ouellette, Cody Springman, Jillian Tietjen, Mahesh Neupane and Scott Henderson. I wish to thank my parents, Kailasan Krishnan and Vanaja Parameswaran, my sister, Jayapriya Anil, and my best friend, Mahrukh Banday, for their support and encouragement. My husband Vijay Rathinam has provided me with amazing support during my graduate career. I thank him for his love and encouragement. vi TABLE OF CONTENTS LIST OF TABLES .............................................................................................. ix LIST OF FIGURES ........................................................................................... x Chapter 1. Introduction ..................................................................................... 1 Pathogenic Escherichia coli ........................................................................... 2 Virulence factors and pathogenesis ............................................................... 8 Acid resistance (AR) of E. coli O157:H7 ....................................................... 14 Genotypes of E. coli O157:H7 ....................................................................... 28 Rationale for this study .................................................................................. 32 Chapter 2. Characterization of the Escherichia coli O157:H7 Sakai GadE Regulon ............................................................................................................. 35 Summary ........................................................................................................ 36 Introduction ............................................................................... ' ..................... 37 Materials and Methods ................................................................................... 41 Results ........................................................................................................... 50 Discussion ...................................................................................................... 68 Acknowledgements ........................................................................................ 75 Chapter 3. Effect of acidic pH on the expression of virulence factors of Escherichia coli O157:H7: role of glutamate decarboxylase system regulators 76 Summary ........................................................................................................ 77 Text ................................................................................................................ 78 Acknowledgements ........................................................................................ 86 Chapter 4. Differential expression of virulence and stress-fitness genes between clinical and bovine-biased genotypes of Escherichia coli O157:H7 .................. 87 Summary ........................................................................................................ 88 Introduction .................................................................................................... 89 Materials and Methods ................................................................................... 92 Results ........................................................................................................... 98 Discussion ..................................................................................................... 1 13 Acknowledgements ....................................................................................... 1 19 Chapter 5. Summary and future directions ...................................................... 120 Summary ....................................................................................................... 121 Future directions ........................................................................................... 130 APPENDICES .................................................................................................. 137 Appendix 1 .................................................................................................... 138 Appendix 2 .................................................................................................... 140 Appendix 3 .................................................................................................... 143 Appendix 4 .................................................................................................... 149 vii Appendix 5 .................................................................................................... 150 Appendix 6 .................................................................................................... 155 REFERENCES ................................................................................................. 164 viii LIST OF TABLES Table 2.1 Bacterial strains and plasmids used in GadE regulon study ............ 42 Table 2.2 Oligonucleotide primers used for one-step inactivation in GadE regulon study ................................................................................................................ 43 Table 2.3 Effect of gadE inactivation on expression of AFI and GAD genes in exponential and stationary phase .................................................................... 54 Table 2.4 LEE genes significantly upregulated in the AgadE strain ................. 57 Table 2.5 Effect of pH 5.0 on expression of GAD regulators and LEE genes.. 62 Table 3.1 Alteration in expression of acid resistance and virulence genes upon exposure to extreme acidity ............................................................................. 81 Table 3.2 Effect of gadX inactivation on the expression of LEE genes ............ 83 Table 3.3 Effect of eng inactivation on the expression of LEE genes ............ 85 Table 4.1 Virulence-associated genes upregulated in clinical genotype 1 relative to bovine-biased genotype 5 ............................................................................ 99 Table 4.2 Genes upregulated in bovine-biased genotype 5 ............................. 101 Table 4.3 Enrichment of gene sets in clinical and bovine-biased genotypes... 103 LIST OF FIGURES Figure 1.1. Step-wise evolution of E. coli O157:H7 from EPEC-Iike ancestor .4 Figure 1.2. Mechanism of action of GAD system ............................................. 17 Figure 1.3. Regulatory circuits involved in the induction of GAD system ........ 19 Figure 1.4. The AFI region of the E. coli K-12 genome .................................... 22 Figure 1.5. Interactions between GAD regulators and LEE ............................. 27 Figure 1.6. The phylogenetic network applied to 48 parsimoniously informative (Pl) sites using the Neighbor-net algorithm for 528 E. coli O157 strains .......... 30 Figure 2.1. Alignment of putative GAD box sequences .................................... 53 Figure 2.2. Exponential phase expression of LEE genes in AgadE, AgadEAIen:Km, and Alensz .......................................................................... 60 Figure 2.3. AR mechanism assays .................................................................. 65 Figure 2.4. Survival of wild-type, AgadE, and AgadE/pCR2.1gadE strains in the M88 ................................................................................................................. 67 Figure 4.1. Double loop design for the microarray experiment ........................ 94 Figure 4.2. qRT-PCR validation of the microarray data ........... 106 Figure 4.3. Neighbor joining phylogeny of SNP genotypes representing the eight strains used in genotype study ........................................................................ 110 Figure 4.4. Survival of clinical and bovine-biased genotypes in the model stomach system ............................................................................................... 112 CHAPTER 1 Introduction PATHOGENIC ESCHERICHIA COLI Escherichia coli is a genetically diverse group of Gram-negative facultative anaerobe that typically colonizes the large intestine and lower part of the small intestine of mammals (66). Commensal E. coli coexist with the human host without causing any disease whereas pathogenic forms of E. coli can cause a number of clinical illnesses with varying degrees of severity (8, 66). Acquisition of several virulence factors through horizontal gene transfer during evolution has allowed pathogenic E. coli to adapt to new predilection sites in hosts and to successfully cause a broad spectrum of diseases in them. Different combinations of these acquired virulence factors defines specific pathotypes of E. coli, which can cause characteristic clinical symptoms such as enteric disease, renal and urinary tract infections, and meningitis (66). Clinically important enteric pathotypes of E. coli include enterotoxigenic E. coli (ETEC), enteropathogenic E. coli (EPEC), Shiga toxin (Stx)-producing E. coli (STEC), and enterohemorrhagic E. coli (EHEC). Other enteric pathotypes are enteroaggregative E. coli (EAEC), enteroinvasive E. coli (EIEC), and diffusely adherent E. coli (DAEC). ETEC is the leading cause of traveler’s diarrhea in developing countries and it harbors at least one enterotoxin; heat-labile toxin (LT) or heat-stable toxin (ST). EPEC causes diarrhea in children younger than 2 years and the infection is characterized by the attaching and effacing (NE) lesions in the intestinal mucosa. This lesion is mediated by a laterally acquired pathogenicity island, the locus of enterocyte effacement (LEE). E. coli strains harboring one or more variants of Stx are named STEC, and EHEC is a sub population of STEC defined by the presence of Stx, LEE and a pO157 plasmid. EHEC infections are characterized by hemorrhagic colitis and hemolytic uremic syndrome (HUS) (66, 101, 109). E. coli O157:H7 E. coli O157:H7 is the most prevalent EHEC serotype in the United States (153). It is a zoonotic food-bome pathogen and causes hemorrhagic colitis and HUS in humans similar to other EHEC. E. coli O157:H7 was isolated first in 1983 from stool cultures of patients with hemorrhagic colitis associated with ingestion of undercooked hamburgers and from sporadic cases of HUS (153). It has since become an emerging pathogen and currently, 0157 is estimated to cause approximately 73,000 illnesses in the US yearly with an economic burden of 0.2 — 0.6 billion (120). E. coli O157:H7 strains are distinguished from other serotypes by their inability to ferment sorbitol (80R) and to produce B-D—glucuronidase (GUD‘) (101). A SOR” GUD+ EPEC serotype, E. coli 055:H7, is considered as the evolutionary ancestor of E. coli O157:H7 (121). E. coli O157:H7 was evolved from 055:H7 in a step-wise manner through acquisition of Stx2 and the pO157 plasmid; followed by an antigenic shift, gain of Stx1, and loss of motility, SOR“, and GUD+ phenotypes (Fig. 1.1) (170). Whole genome sequence comparison of E. coli O157:H7 Sakai with benign laboratory strain E. coli K-12 revealed that the 0157 genome contains 1.4-Mb of O157:H7-specific sequence most of which is ET1 EDL-933 Ancestral 1 EPEC-like 3:9, ET 2 strain 3 activity 413 GUD+ soa+ GUD+ SOR+ Stx2+ ‘---- ET 5 DEC so Figure 1.1. Step-wise evolution of E. coli O157:H7 from EPEC-like ancestor. Phenotypes of ancestors are shown (38). contributed by foreign DNA elements that are acquired horizontally during evolution (46). Clinical manifestations and treatment The infective dose of E. coli O157:H7 is extremely low: 10 -100 cells are sufficient to cause a clinical infection (68, 147, 163). The average incubation period for E. coli O157:H7 infection is 3.7 days. The disease starts as non-bloody diarrhea, which lasts for 1 — 3 days followed by bloody diarrhea. Bloody diarrhea occurs in 90% of the cases due to severe hemorrhagic colitis. In 0157 infections, most patients remain afebiile and the abdominal pain is more severe compared to other bacterial gastroenteritis (153). The most lethal complication of E. coli O157:H7 infection is HUS, which causes acute renal failure in 15% of the infected children younger than 10 years old. Onset of HUS occurs typically between 5 - 13 days of the infection and generally HUS starts as thrombocytopenia followed by hemolysis and azotaemia (153). HUS is a thrombotic disorder and the characteristic lesions include microvascular thrombi and swollen endothelial cells (57, 160). It is believed that intravenous rehydration and maintenance fluids provide optimal protection against kidney damage and thus, constitute the most common management strategy in patients with bloody diarrhea. Expansion of parenteral volume has been associated with attenuation of renal injury (3). However, standard rehydration protocols are considered inadequate for 0157 infections. lsotonic crystalloid is highly recommended for volume expansion and maintenance in HUS. Antibiotics are not administered to patients with 0157 infections as many studies have shown a strong correlation between antibiotic therapy and increased risk of HUS (128, 172). Reservoir hosts of E. coli O157:H7 Cattle are the primary reservoir hosts of E. coli O157:H7 and 0157 exists as part of normal intestinal microflora of cattle without causing any disease (25, 74). In addition, other food animals such as pigs, sheep, and goats (104) act as reservoir hosts of E. coli O157:H7 (106). E. coli O157:H7 colonizes the gastrointestinal tract of cattle, specifically the lymphoid follicle-dense mucosa of the recto—anal junction (40, 102). Earlier surveys in cattle indicated lower prevalence of E. coli 0157 in feces possibly due to the poor sensitivity of isolation methods (49). Only 1.8% of fecal samples were found to contain E. coli 0157 in one of the largest surveys. However, recent studies have indicated a markedly higher prevalence of 0157 in cattle (51, 62). An investigation on the correlation of E. coli O157 prevalence in feces, hides and carcasses of beef cattle during processing demonstrated that of the 30 lots sampled, 87% had at least one 0157 positive pre-evisceration sample (36). Currently, E. coli O157:H7 is considered ubiquitous in cattle farms and the shedding rate in cattle farms is estimated to be greater than 10%, sometimes approaching 100% (14, 45). Interestingly, there is a seasonal effect in the prevalence of E. coli O157:H7 in cattle with peak prevalence in summer and early fall (45). This corresponds with the peak in outbreaks involving ground beef in summer (120). Transmission of E. coli O157:H7 A factor critical for the pathogenicity of E. coli O157:H7 is its ability to be transmitted from cattle to humans through a variety of food vehicles. In the earlier outbreaks 0157 was typically transmitted through contaminated ground beef (115, 118). Later it was found to be transmitted through vehicles such as apple cider (13, 53), dry salami (116), apple juice (29), and raw milk (114). Most recently, green leafy vegetables such as spinach and lettuce are implicated in 0157 outbreaks (52, 117). Interestingly, many of these vehicles pose multiple environmental stresses and E. coli O157:H7 appears to be successful in surviving under these conditions. Response of 0157 to the acidity of apple juice (pH 3.5) has been investigated using a model apple juice (MAJ) system and it was found that 0157 is induced for multiple stress response regulons including the RpoS, RpoH and CpxRA regulons in response to MAJ exposure (9). VIRULENCE FACTORS AND PATHOGENESIS Important virulent factors of E. coli O157:H7 include Stx, LEE and the pO157 plasmid. Each of these factors and their role in pathogenesis and regulation are described'in detail below. (i) Stx: Stx, also known as verotoxin (VT), is encoded by bacteriophages that are inserted into the 0157 chromosome (66). There are two subgroups in the Stx family: Stx1 and Stx2 with five allelic variants of Stx2 (Stx2, Stx20, Stx2d, Stx29 and Stx2f) (153). There is a 55% amino acid homology between Stx1 and Stx2. Stx is a typical A35 exotoxin with five identical B subunits forming a pentamer with a single A subunit. The A subunit of Stx has two peptide chains A1 and A2. A1 is a N-glycosidase enzyme and A2 connects A1 with the B-pentamer (66). The B-pentamer mediates the binding of Stx to the glycolipid globotriaosylceramide (Gb3) receptors on the hostcell surface. This binding is followed by a receptor-mediated endocytosis and transfer through the Golgi apparatus and endoplasmic reticulum. During this transfer, a membrane bound protease, furin, nicks the A subunit from the holotoxin leaving the A1 fragment attached to A2 by a disulfide bond. This bond is later reduced and subsequently, the A1 enzyme is translocated into the cytoplasm where it removes an adenine residue from the 28s rRNA and inhibits the elongation step of protein synthesis. Inhibition of protein synthesis leads to death of cells that carry Gb3 receptors. There are cellular differences in the expression of Gb3 and renal glomerular endothelial,.mesangial and tubular epithelial cells are characterized by higher concentration of Gb3 receptors, which makes them highly susceptible to the damage by Stx. Direct toxicity by Stx along with induction of cytokines and chemokines cause destruction of renal endothelial cells and microvascular occlusion ultimately leading to HUS. Stx also causes intestinal epithelial cell death and damage in the colon leading to hemorrhagic colitis and bloody diarrhea by a similar mechanism (7). Stx is encoded on a single operon comprising two genes corresponding to the A and B subunits. The 1:5 ratio of the NB subunit synthesis for generating the A35 holotoxin is maintained by a stronger ribosomal binding efficiency of the B-subunit gene resulting in increased translation of B-subunits compared to the A-subunit (109). Stx genes are located on the late gene region of the Stx- producing Iambdoid prophage down stream of the phage lambda Q antiterrninator gene. Increased toxin synthesis occurs upon induction of the prophage due to an increase in copy number as the phage genome replicates and due to upregulation of stx expression from the phage late promoter PR'. Stx1 has its own promoter in addition to the phage late promoter and this promoter can induce toxin gene transcription independent of phage induction. No phage independent promoter. has been identified for Stx2. Genes encoding holin and endolysin are located down stream of the Stx genes and therefore, it is believed that coupling of toxin and lysis gene expression facilitates release of the toxin from the bacterial cell (149). Several environmental factors that regulate Stx expression have been identified. There is a strong correlation between DNA damage and Stx phage induction. Antibiotics such as norfloxacin and ciprofloxacin inhibit cellular DNA gyrase activity resulting in DNA damage and SOS response, which lead to a dramatic increase in Stx production from cells (48). Iron downregulates Stx1 expression as the ferrous iron binding protein, Fur, directly binds to the Fur box upstream of the phage independent Stx1 promoter and represses its transcription (20, 95). (ii) LEE: LEE is a laterally acquired pathogenicity island that encodes a type 3 secretion system (T388) in EHEC and EPEC. The LEE consists of 41 genes that are transcribed as five polycistronic operons, LEE1 through LEE5. LEE1 encodes the key regulator of LEE, the Ler (LEE encoded regulator), which positively regulates the expression of LEE2 through LEE5. LEE also encodes additional regulators such as GrIA (global regulator of LEE activator) and GrlR (global regulator of LEE repressor). The structural components of T3SS are mainly encoded on the 5' end of LEE whereas the outer membrane adhesin intimin (Eae) and the translocated intimin receptor (Tir) are encoded on the central part of LEE. The 3' end of LEE encodes translocators, effectors and additional structural proteins of the T388. LEE-encoded T3SS is involved in translocating effector proteins encoded on the LEE and on different locations in the genome (158), into the host cell cytosol, which is important for the pathogenesis of 0157 infection (41, 95). The T3SS apparatus assembled from the products of LEE genes has the typical multicomponent organelle structure of gram-negative T3SS with outer and 10 inner membrane ring structures and a needle complex (41 ). This TSSS is responsible for the characteristic intimate attachment between E. coli O157:H7 and intestinal epithelial cells. Following initial adherence of E. coIiO157zH7 to the host cell, the T388 needle apparatus is inserted into the host cell plasma membrane and Tir is secreted into the cytoplasm, which acts as a receptor for intimin. Binding between intimin and Tir leads to an intimate attachment of bacteria to host cell. This induces rearrangement of actin cytoskeleton resulting in formation of pedestal-like structures and effacement of intestinal microvilli ultimately causing A/E lesions in the intestine, the hallmark of 0157 infection. Regulation of LEE expression is complex and involves multiple environmental factors and regulators (95). Growth in tissue culture medium such as Dulbecco’s modified Eagle’s medium (DMEM) at host body temperature (37°C), pH 7.0, and physiological osmolarity induces maximal LEE expression (69, 70). Other environmental factors that stimulate LEE expression include presence of iron and sodium bicarbonate. Presence of ammonium chloride and exclusion of calcium from the growth medium can inhibit LEE expression (56). As mentioned above, Ler is the key positive regulator of LEE and it acts by disrupting the silencing of LEE by the histone-like DNA binding protein, H-NS (95). Several regulatory systems such as RcsCDB phosphorelay system and the EHEC-specific GrvA protein regulate expression of LEE by controlling the Ier expression in 0157 (157). In addition, there is a quorum-sensing regulation of LEE, which involves an Al-3-signaling molecule that cross talks with epinephrine hormone and activates LEE expression (141, 142). 11 (iii) p0157 plasmid: pO157 is a laterally acquired 92 kb F-Iike plasmid present in all E. coli O157:H7 strains (19). p0157 contains 100 open reading frames (ORFs) and encodes a number of virulence factors such as enterohemolysin (Eth), toxin B (ToxB) and a type 2 secretion system (T288) encoded by etp genes (19). Enterohemolysin is a repeats in toxin (RTX) family toxin. Although the exact role of enterohemolysin in the pathogenesis of 0157 infections is not clear, it has been shown to cause endothelial injury and induce production 'of cytokine, IL-18, which is a hallmark of HUS (152). eth is the structural gene of enterohemolysin, which is encoded on a single operon containing four genes ethABD. Eth converts Eth into an active form by the addition of a fatty acid group (58) and Eth and D form a secretion machinery for Eth (166). Another virulence factor encoded by pO157, ToxB, is a potential adhesin with sequence similarity to the Chlostridium toxin family. The etp-T288 secretes StcE, a zinc metalloprotease that cleaves the C1 esterase inhibitor of the complement pathway, which may contribute to the tissue damage (76). StcE also has a mucinase activity and is involved in intimate adherence of 0157 to host cells (44). A positive interaction exists between LEE and many of the pO157 encoded virulence genes. GrlA, the LEE encode regulator, positively regulates the transcription of eth. Similarly, Ler upregulates the expression of stcE (44). Furthermore, it has been shown that mutation of the toxB gene results in reduced 12 expression and secretion of proteins encoded by LEE leading to a decrease in adherence to cultured epithelial cells (143). 13 ACID RESISTANCE (AR) OF E. COLI O157:H7 Gastric acidity (pH 1.5—2.5) is one of the first innate defense barriers encountered by enteric pathogens upon entry into the human body (112). Enteric pathogens have adapted several mechanisms to breach this defense barrier and establish infection. One common strategy is the “assault tactic’ as seen in Vibrio cholerae and Salmonella infections, where the infectious dose of the organism is so huge that some organisms will survive and enter the intestine. Another strategy shown typically by Helicobacter pylori Is to mount a counterdefence mechanism such as the urease system to neutralize the extreme acidity (39, 135, 137). E. coli O157:H7 also has to overcome the adverse conditions of the stomach before successfully colonizing its niche, the large intestine. Similar to H. pylori, E. coli strains have been shown to be more AR than other enteric pathogens (83). This extreme AR is governed by four principal mechanisms in E. coli; the oxidative (OXI) system, the GAD system, the arginine decarboxylase (ARG) system and the lysine decarboxylase (LYS) system (39, 83). The OXI system is the least understood AR system in E. coli. It is induced upon entry into the stationary phase in a complex medium at pH 5.5. Once induced, the OXI system helps the bacteria to survive in pH 2.5 in minimal medium. Global regulatory protein CRP (cAMP receptor protein) and the stationary phase sigma factor, RpoS are essential for the functioning of the OXI system (23, 82). Presence of glucose in the medium inhibits CRP and hence, represses the OXI system (23). 14 Functioning of GAD, ARG and LYS systems are dependent on the availability of the amino acids glutamate, arginine and lysine, respectively. These systems typically consist of pairs of amino acid decarboxylases and antiporters. The ARG system comprises the arginine decarboxylase enzyme encoded by adiA and the arginine-agmantine antiporter encoded by adiC (59). Anaerobic conditions in a complex medium at low pH are essential for transcription of adiA and adiC. It is believed that a regulator, CysB, senses these factors and induces their transcription. An AraC-like regulator, AdiY, also activates the expression of adiA and adiC, however, this regulator is not essential for their expression (43, 145). Similarly, RpoS is involved in the functioning of the OXI system, but is not required for the transcription of adiA and adiC genes. The LYS system is a much less efficient system compared to the other three AR systems. It consists of lysine decarboxylases (cadB) and the lysin'e-cadaverine antiporter (cadA) and provides protection against mild acidity (59). GAD system The GAD system is the most efficient AR mechanism in E. coli and provides best protection against pH as low as 2.0 (75). The components of the GAD system include two glutamate decarboxylase isozymes, GadA and GadB, and a membrane-associated glutamate-gamma amino butyric acid (GABA) antiporter, GadC (23, 24, 75). GadA and GadB are pyridoxal 5’-phosphate (PLP)- dependent enzymes and have a hexameric structure with one PLP moiety per monomer (16). The nucleotide sequences of GadA and GadB are 98% similar 15 and they have highly similar peptide structure with differences only in 5 amino acid residues mainly at the N-terminal region (140). The two isozymes also share similar biochemical characteristics such as specific activity and isoelectric points (16). The two genes gadA and gadB map to distinct loci in the E. coli K-12 chromosome with gadB forming an operon with gadC at 33.8 min and gadA at 78.98 min as part of the acid fitness island. Either one of the decarboxylases isozymes is sufficient for GAD dependent AR at pH 2.5 whereas both are required for AR at pH 2.0. Presence of GadC is essential fOr the functioning of GAD system at both conditions (23). The GAD system is induced upon entry into stationary phase of growth as well as when the cells are exposed to acidic conditions. Presence of glutamate in the environment is essential for the functioning of the GAD system. As the extracellular pH decreases, the intracellular pH of E. coli also goes down. At an intracellular pH of 4.2, GadA and GadB get activated and these enzymes catalyze the replacement of the carboxyl group of glutamate with a proton from the cytoplasm producing GABA and C02. This reaction reduces the intracellular H+ ion concentration and stabilizes the pH homeostasis of the cell and maintains the pH at 4.2, which is tolerable to the E. coli physiology (39, 123). GABA is in turn expelled into the extracellular environment via GadC, in exchange for incoming glutamate (Fig. 1.2) (120). 16 pH 2.5 H+ + CI'Z HCI Glutamate GABA COZ Figure 1.2. Mechanism of action of GAD system. GABA= gamma amino butyric acid. Adapted from (39). Regulation of GAD system in E. coli The GAD system is spontaneously induced as cells enter the stationary phase of growth. Stationary phase sigma factor, RpoS, regulates this induction (87). Low pH also induces the GAD system independent of RpoS (161 ), but, the exact mechanism of this induction is not clearly defined. Previous studies in E. coli K-12 have revealed that the genetic regulation of the GAD system is complex and involves at least 14 regulatory genes and a central activator, GadE (39). The GadE activates the transcription of gadA and gadBC genes by directly binding to a conserved 20 bp GadE-binding region upstream of gadA and gadB namely, the GAD box (22, 54, 84). Distinct circuits of regulation are induced depending on the physiological status of the cell and the medium in which the organism is grown. These circuits are focused primarily on regulating the expression of gadE (Fig. 1.3) (24, 39). When cells are grown in minimal medium, the EngS two-component system is activated and it upregulates the expression of YdeO, which in turn triggers the expression of gadE (85, 89, 90). On the other hand, in a complex medium such as LB, during the stationary phase of growth, another circuit comprised of CRP, RpoS and two AraC-like regulators, GadX and GadW comes into play. RpoS induces the expression of GadX/W, which in turn indirectly activates the transcription of gadA/B via GadE (87, 159, 162). GadX and GadW also form a negative feed back loop by binding to the GAD box sequences and repressing the gadA and gadBC promoters. Because of this dual role, GadX/W can act as activators of 18 Acidity at exponential phase (minimal media) Stationary phase (complex media) RpoS Eng —> Eng GadX GadW 1“ 1’ gadB gadC gadA Figure 1.3. Regulatory circuits involved in the induction of GAD system. The growth phase and medium that activate each circuit is given within the circles Modified from (39). glutamate decarboxylase under some conditions and repressors under others (87). GadW forms another feed back loop by inhibiting RpoS, which is necessary for the activation of GadX. Similarly, GadX can repress the transcription of GadW. During exponential growth in complex media GadX and GadW tightly control each other and prevent the activation of the GAD system (159). A third regulatory circuit of the GAD system involving a GTPase protein, TrrnE (MnmE), becomes active during growth in LB containing glucose (39, 42). TrrnE is also required for efficient translation of GAD genes (42). Unlikethe genes in these regulatory circuits, GadE is necessary for the functioning of the GAD system at all stages of growth in any medium (39, 84). In addition to these regulatory circuits, other regulators such as RcsB and RNaseE are also involved in the regulation of GAD system. Basal level expression of RcsB, a component of RcsCDB signal transduction system, is necessary for GAD system functioning whereas, if over-expressed, RcsB represses GAD expression (24). RNaseE, an essential endoribonuclease involved in processing and degradation of RNAs, is also required for induction of the GAD system (150). Similarly, a functional topoisomerase I is essential for full induction of the GAD system (144). PhoP, the response regulator of the Pth/PhoP two-component system, also positively regulates the GAD system by directly binding to upstream sequences of GadE and GadW (176). Besides above-mentioned protein regulators, a small RNA, GadY, is also involved in induction of the GAD system. GadY is encoded by the AFI, on the opposite strand between gadX and gadW. There are three forms of GadY, a 105 20 nucleotide long form and two processed forms of 90 and 59 nucleotides. RpoS controls GadY expression and it is induced at stationary phase of growth. GadY acts by conferring increased stability to gadX mRNA by forming base pairs with its 3’-untranslated region. This leads to accumulation of gadX mRNA, which ultimately results in increased GAD expression (105). One of the major repressors of GAD system is the histon-like protein, H- NS (167). H-NS directly downregulates the GadE expression and affects the GAD system (55). Another repressor of the GAD system is TorR, the response regulator of the TorS/TorR two component system, which negatively regulates gadE, gadABC and many AFI genes (18). It has been recently shown that Ler, the key positive regulator of LEE also negatively regulates GAD expression (1). Post transcriptional regulation of GAD system is not clear. However, the fact that exponential phase cells grown under moderate acidity (pH 5.5) express high level of GAD genes but remain acid sensitive indicates the presence of regulation at translational level. Acid fitness island (AFI) Fitness islands are horizontally transferred genetic elements that provide advantageous traits that are not directly related to the virulence of the organism. One such element in E. coli and Shigella is the AH, the acquisition of which is considered as a crucial step in the early evolution of these organisms. The AFI is located at 78.8 min in E. coli K-12 chromosome and encodes 12 genes that are involved in acid and stress resistance including gadA (Fig. 1.4) (8). Also encoded 21 Acid Fitness Island (AF l) gadA 3665210 yhiFD hdeBAD yhiU V gadX / E. coli K-12 I I I I I I I- I I - 78.8 4,639,675 nt gadE gadW gadA mi" gadB 12.98 kb 1570069 Figure 1.4. The AFI region of the E. coli K-12 genome. Previous studies have shown that the 12-kb AFI is located at 78.8 min on the K-12 chromosome and includes the gadA gene and other metabolic stress-related genes, whereas the duplicate gadB locus occurs across the K-12 chromosome at 33.0 min (8). 22 on AFI are three important regulators of the GAD system, GadE, GadX, and GadW. Many of the other AFI proteins confer different types of AR in E. coli. For example, the periplasmic acid stress chaperones, HdeAB, a putative LuxR family regulator, YhiF, and the lipoprotein, Slp, are required for protection against self- metabolic products at low pH and in spent medium. Similariy, a putative MgtC- family transporter, YhiD, and a predicted inner membrane protein, HdeD, are necessary for high cell density-dependent AR (71, 91). GadE regulates the expression of all the AFI genes in E. coli K-12 and hence, is involved in resistance to metabolic stress and high cell-density AR (91 ). Functions of the remaining AFI genes, thU and thV remain unclear. GadE regulator GadE is a LuxR family regulator with a potential helix-tum-helix DNA- binding domain in the secondary structure. GadE binds to a conserved 20 bp GAD box sequence upstream of gadA and gadBC and activates the transcription of these genes (22). A 798 bp intergenic region between gadE and the upstream gene hdeD is essential for the expression of gadE. The highly conserved sequence between -750 and -1 is considered as the gadE sensory integration region as this region coordinates the environmental and genetic regulators of gadE expression. At least nine activators and repressors of gadE converge at this region including Eng, YdeO, GadE, TorR, Hns, PhoP, TrmE, GadX and GadW (1 31 ). 23 GadE is transcribed as three transcripts T1, T2 and T3. T1 starts at -124 bp, T2 at -324/-317 and T3 at -566 relative to the gadE translational start site with P1, P2 and P3 as corresponding promoters. P1 is the strongest promoter and is typically induced at stationary phase in complex or minimal medium. P3 is highly responsive to induction in minimal medium while P2 is expressed only in combination with P3 or alone when there is an over-expression of GadX or YdeO regulators. P2 is also activated by a TnnE-dependent manner in LB-glucose. Initial induction of GadE occurs through activation of P3P2 promoters, and once induced, GadE represses P3P2 and autoactivates P1 leading to a sustained transcription of gadE from P1 (131). Upon removal of inducing signals, GadE is degraded by the Lon protease, a major protease that is active in cellular protein quality control and degradation of various naturally unstable regulators (50). GAD system and GadE in E. coli O157:H7 A previous study in our laboratory comparing survival rates of E. coli O157:H7 in complex acidic conditions such as a model stomach system (pH 2.0) with that of E. coli 026:H11 and E. coli 011 1 :H8, revealed that E. coli O157:H7 has a superior ability to survive in the simulated gastric environment than the other strains tested. The quantitative PCR data from this study also showed that E. coli O157:H7 expresses higher transcript levels of gadA and gadB genes. This led to the suggestion that the regulation of GAD system may be different in E. coli O157:H7 compared to the laboratory strain E. coli K-12, in which most of the studies about this system have been done (10). Together, these results support 24 the hypothesis that the predominance of E. coli O157:H7 in clinical cases compared to other EHEC strains is due to its enhanced ability to resist stresses encountered in the environment both outside and inside the host. Also, this superior AR is considered a virulence factor because it contributes to the low infective dose of E. coli O157:H7. Furthermore, our laboratory recently showed that gadA and gadB sequences remain divergent in E. coli O157:H7, whereas in other E. coli strains, they have undergone multiple gene conversion events leading to genetic homogenization (8). All of these findings strongly suggest that the regulation and functioning of the GAD system are distinct in E. coli O157:H7. Much is known about the upstream regulatory circuits and downstream effects of GadE in non-pathogenic E. coli (39, 84). However, GadE and its role in AR and virulence are not well characterized in any of the pathogenic E. coli strains. Genome comparisons between two sequenced O157:H7 strains and K12 MG1655 revealed that O157:H7 has approximately 25% additional O157- specific loci compared to E. coli K12 (46). During its evolution, O157:H7 has acquired many mobile elements such as Iambdoid phages carrying virulence and fitness islands (121, 170). As mentioned before, important virulence factors of E. coli O157:H7 including shiga toxins and LEE are encoded by these horizontally transferred phage elements (77). For E. coli, Integrating these acquired elements into the chromosomal regulatory network is critical to becoming a successful pathogen (1 )..Although, the GadE sequence remains unchanged and maps to a homologous location in the 0157 chromosome, it is possible that a chromosomal regulator, such as GadE, has acquired additional functions in O157:H7. 25 However, effects of the GadE regulator on a genome-wide scale in E. coli O157:H7 are still unknown. Interaction of GAD system regulators with LEE The relationship between AR and virulence of pathogenic E. coli, especially the interaction between GAD system and LEE remains largely unknown. A few studies in the past have reported that some of the GAD system regulators negatively effect the LEE expression (24, 99, 138, 157) (Fig. 1.5). In E. coli O157:H7, gadE inactivation was shown to increase the expression of LEE encoded genes espB, espD and tir but not Ier (156). Hence, GadE mediated down-regulation of LEE is considered Ler-independent and the pathway through which GadE affects LEE is not known (95). Two upstream regulators of gadE also negatively affect expression of LEE in EPEC. One of them is GadX, which negatively regulates the expression of LEE through a plasmid encoded regulator, PerA (138). Similarly, Eng represses the LEE expression in EPEC by activating YdeO and YdeP (99). But, the regulation of LEE by GadX and Eng in EHEC has not been studied. It is possible that GadX down-regulates the expression of LEE through Pch, the PerC-homolog in EHEC. Additionally, because YdeO is a positive regulator of gadE, the down-regulation by Eng might be mediated through GadE. However, a detailed study is warranted in this aspect in order to fully understand the mechanism of LEE repression by GAD system regulators in EHEC. 26 a LEE1 LEE2 LEE3 LEE5 LEE4 Figure 1.5. Interactions between GAD regulators and LEE. Solid lines indicate experimentally confirmed mechanisms and dotted lines indicate suggested mechanisms that are not proved experimentally. 27 GENOTYPES OF E. COLI O157:H7 Several genotyping methods such as pulse field gel electrophoresis (PFGE) and multi locus sequence typing (MLST) have been employed to investigate the genetic relatedness of E. coli O157:H7 strains from various sources. One of the most sensitive of these methods is the single nucleotide polymorphism (SNP) typing that was established recently by our laboratory (88). This study typed >500 clinical 0157 strains using a SNP genotyping scheme that targets 96 SNP loci. Phylogenetic analysis classified the strains into 39 SNP genotypes, which formed nine distinct clades (Fig. 1.6) (88). Interestingly, the clades showed differences in the frequency and distribution of Stx genes and in the spectrum of clinical diseases reported. lmportantly, clade 8 was found to be an emergent subpopulation with more chances of causing severe disease with HUS whereas groups such as clade 7 was less frequent in clinical cases and caused less severe disease. Both spinach and lettuce strains that caused fresh produce-associated outbreaks with severe disease and death in 2006 belonged to clade 8. The 0157 genome strain Sakai, that was implicated in the 1996 radish sprouts outbreak in Japan belonged to clade 1 whereas the strain that caused a Hamburger outbreak in northwest regions of the United States grouped to clade 2 (88). Even though SNP typing has been conducted extensively on clinical strains of E. coli O157:H7, strains of bovine origin are yet to be classified by this method. 28 Figure 1.6. The phylogenetic network applied to 48 parsimoniously informative (Pl) sites using the Neighbor-net algorithm for 528 E. coli O157 strains. The ellipses mark clades supported in the minimum evolution phylogeny. The numbers at the nodes denote the SNP genotypes (SGs) 1—39, and the white circle nodes contain two SGs that match at the 48 Pl sites. The seven SGs found among multiple continents are marked with squares (88). 29 ilormallve s. The The a while 865 found I. Clade 2 4 Clade 6 24 26 p distance 0331 30 A recent study by Besser et al.(2007) classified E. coli O157:H7 strains of bovine (n=80) and clinical (n=282) origin into various genotypes based on the Stx-encoding bacteriophage insertion sites (14). A greater diversity of Stx- encoding bacteriophage insertion sites was observed in strains from bovine sources compared to the strains from clinical sources and these two groups were classified into different genotypes based on this method. Subsequently, strains that were predominant in clinical cases were typed as clinical genotypes whereas strains that were isolated mostly from bovine sources were considered bovine- biased genotypes (14). The basis for the variation in prevalence of clinical and bovine-biased genotypes in human clinical cases is not clear. 31 RATIONALE FOR THIS STUDY E. coli O157:H7 has to survive a number of environmental stresses during transmission from the reservoir host, typically cattle, to the human gastrointestinal tract. Surviving acid stress is a critical component of transmission, as the typical human stomach pH ranges from 1.5-3.0 (112). However, E. coli O157:H7 has a greater average level of survival in complex acidic conditions such as the gastric environment than other groups of EHEC (10). This superior AR contributes to the low infective dose of E. coli O157:H7. The GAD system is the most effective AR system in E. coli and GadE is the central activator of the GAD system (39). Even though the GAD system and GadE are well characterized in non pathogenic laboratory strains of E. coli, they remain poorly defined in E. coli O157:H7. It is possible that a chromosomal regulator, such as GadE, has acquired additional functions in O157:H7 to integrate the mobile virulence genes acquired during its evolution into the chromosomal regulatory network. In order to successfully colonize the host, pathogenic E. coli must conserve energy by orchestrating their gene expression profiles in such a way that only necessary genes are expressed at each step of the infection process. It is possible that during passage through the human stomach, E. coli O157:H7 down-regulates the virulence mechanisms needed for later life in colon to conserve energy and activates the AR mechanisms to survive the extreme acidic pH, thus leading to a negative interaction between AR and virulence. However, the effect of acidic pH on virulence factors such as the locus of enterocyte effacement (LEE) in E. coli O157:H7 has not been investigated. 32 Based on the insertion sites of Shiga toxin-encoding bacteriophages, E. coli O157:H7 strains isolated from cattle and human sources are classified into clinical and bovine-biased genotypes (14). Clinical genotypes are isolated from both cattle and humans and have been shown to be associated with human disease whereas bovine-biased genotypes are isolated mostly from bovine sources (14). The varying infectivity of clinical and bovine-biased genotypes could be due to the differences in expression of vimlence and stress fitness- associated genes. This proposal aims to investigate the relationship between AR and virulence of E. coli O157:H7 and to understand whether differences in expression of virulence and stress fitness-associated genes can explain the variation in infectivity of bovine-biased and clinical genotypes, with the following specific aims: Specific Aim 1: To elucidate the role of GadE in acid resistance and virulence of E. coli O157:H7. Hypothesis: GadE regulates the expression of the GAD system and LEE pathogenicity island in E. coli O157:H7, thus contributing to its acid resistance and virulence. Specific Aim 2: To determine the effect of acidic pH on the expression of the LEE pathogenicity island and to understand the role of GAD system regulators in this effect. 33 Hypothesis: Acidic pH upregulates the expression of GAD system regulators, which in turn repress the expression of LEE genes. Specific Aim 3: To determine whether the differences in infectivity of clinical and bovine-biased genotypes of E. coli O157:H7 is due to the differential expression of virulence and stress fitness associated genes. Hypothesis: Virulence and stress fitness associated genes are differentially expressed between clinical and bovine-biased genotypes of E. coli O157:H7. Chapter 2 describes the findings of specific aim 1 and parts of results from specific aim 2. Chapter 3 is a follow up to chapter 2 that describes remaining results from specific aim 2 and chapter 4 describes results from specific aim 3. Summary of this study and future directions are described in chapter 5. 34 CHAPTER 2 Characterization of the Escherichia coli O157:H7 Sakai GadE regulon Kailasan Vanaja, 8., T. M. Bergholz, and T. S. Whittam. 2009. Characterization of the Escherichia coli O157:H7 Sakai GadE regulon. Journal of Bacteriology 191:1868—77 35 SUMMARY Integrating laterally acquired virulence genes into the backbone regulatory network is important for the pathogenesis of Escherichia coli O157:H7, which has captured many virulence genes through horizontal transfer during evolution. GadE is an essential transcriptional activator of the glutamate decarboxylase (GAD) system, the most efficient acid resistance (AR) mechanism in E. coli. The full contribution of GadE to the AR and virulence of E. coli O157:H7 remains largely unknown. We inactivated gadE in E. coli O157:H7 Sakai and compared global transcription profiles of mutant with that of wild type in exponential and stationary phases of growth. Inactivation of gadE significantly altered the expression of 60 genes independent of growth phase and 122 genes in a growth phase-dependent manner. Inactivation of gadE markedly down-regulated the expression of gadA, gadB, gadC and many acid fitness island genes. Nineteen genes encoded on the locus of enterocyte effacement (LEE), including Ier, showed a significant increase in expression upon gadE inactivation. Inactivation of Ier in AgadE reversed the effect of gadE deletion on LEE expression, indicating that Ler is necessary for LEE repression by GadE. GadE is also involved in down-regulation of LEE expression at moderately acidic pH. Characterization of AR of AgadE revealed that GadE is indispensable for a functional GAD system and for survival of E. coli O157:H7 in a simulated gastric environment. Altogether, these data indicate that GadE is critical for AR of E. coli O157:H7 and that it plays an important role in vimlence by down-regulating expression of LEE. 36 INTRODUCTION Escherichia coli O157:H7 is the prevalent variant of enterohemorrhagic E. coli (EHEC) associated with hemorrhagic enteritis and hemolytic uremic syndrome (HUS) in humans in the United States (3, 67). E. coli O157:H7 has to survive a number of environmental stresses during transmission from cows to humans. Surviving acid stress is critical during transmission, as the typical human stomach pH ranges from 1.5-3.0 (112). E. coli strains are more AR than other enteric pathogens and this AR is considered a vinilence factor in E. coli O157:H7 as it contributes to the low infective dose (23, 112). E. coli have four distinct AR mechanisms, the oxidative (OXI) system, glutamate decarboxylase (GAD) system, arginine decarboxylase (ARG) system and lysine decarboxylase (LYS) system (39, 85) that are phenotypically distinct and provide protection against low pH dependent on the type of acidic environment encountered (119). In addition to the defined mechanisms, other factors of the general stress response, including the stress response sigma factor RpoS and the DNA binding protein Dps, also contribute to the AR of E. coli (28, 169). The GAD system is the most effective system in protecting E. coli cells against low pH compared to other known AR mechanisms (23, 24, 75, 161 ). The GAD system has three components, two GAD isozymes, GadA and GadB, and the GABA-glutamate antiporter, GadC (39, 84). The gadA is a member of the acid fitness island (AFI), which is located at 78 min, whereas gadB and gadC form a separate operon located at 33 min in the E. coli K12 chromosome (55, 91, 37 140). Environmental signals that induce the GAD system include entry into stationary phase and acidic pH (84). Regulation of the GAD system is complex, involving multiple regulatory circuits that influence the expression of GAD components through the central activator, GadE (39). GadE, a LuxR family regulator, is transcribed as two transcripts of sizes 0.68 kb and 1.06 kb and its secondary structure contains a potential helix-tum-helix DNA-binding domain (54, 84). However, a recent study demonstrated that GadE is possibly transcribed as three transcripts of sizes 0.9 kb, 1.1 kb and 1.38 kb (J. W. Foster and A. Sayed, presented at the 108'h General Meeting of American Society of Microbiology, Boston, MA, 1 to 5 June, 2008). GadE binds to a conserved 20 bp GAD box sequence upstream of gadA and gadBC in E. coli K12 and activates the transcription of these genes (22, 54, 84). Although the GAD system and GadE are well-characterized in E. coli K12, they remain pooriy defined in E. coli O157:H7. A study from our lab demonstrated that E. coli O157:H7 strains have a greater average level of survival in complex acidic conditions, such as a simulated gastric environment, compared to other serogroups of EHEC (11). O157:H7 also expresses higher transcript levels of gadA and gadB genes than other EHEC strains in minimal medium containing glucose (11). Also, we recently showed that gadA and gadB sequences remain divergent in E. coli O157:H7 compared to other E. coli strains (8). Taken together, these findings suggest that the regulation and function of the GAD system may be distinct in E. coli 01 57: H7. 38 Genome sequence comparisons revealed that the O157:H7 Sakai strain has approximately 1650 O157-specific loci compared to E. coli K12 (47). Through evolution, the 0157 population has acquired many mobile elements such as Iambdoid phages carrying virulence and fitness islands (121, 171 ). Some of the important virulence factors of E. coli O157:H7, including Shiga toxins and the LEE, are encoded by horizontally transferred phage elements (78). The LEE, which is encoded on an acquired pathogenicity island, encodes a type three secretion system that mediates intimate adherence of bacteria to the intestinal mucosa through formation of attaching and effacing (AE) lesions (93). Integrating these acquired elements into the chromosomal regulatory network is critical for a pathogen to be successful (1). An example of this is one of the GAD system regulators, GadX, which has been shown to influence the expression of the LEE (138). Hence, it is possible that a chromosomal regulator such as GadE has acquired additional functions, though the effects of the GadE regulator on a genome-wide scale are still unknown for E. coli O157:H7. Comparison of GadE amino acid sequences among pathogenic and non pathogenic E. coli strains revealed no significant divergence. Recently, a study by Tatsuno et al., found that inactivation of gadE increases the expression of many LEE genes in O157:H7 (156). However, gadE inactivation did not affect the expression of the LEE encoded regulator, Ier, and hence the pathway through which this LEE down-regulation occurs was not identified. Recently, Ler was found to negatively regulate expression of gadE and it was suggested that there is a reciprocal negative interaction between Ler and the GadE regulators (1). 39 The objectives of this study were to identify the genes regulated by GadE in E. coli O157:H7 and to gain insight into the mechanism underlying the negative regulation of the LEE by GadE. By comparing whole genome transcription profiles of E. coli O157:H7 Sakai and isogenic AgadE, we found that gadE positively influences expression of the GAD system genes and other AF l genes, whereas it negatively impacts the expression of the LEE genes, including Ier. Expression of gadE was markedly increased in stationary phase, thereby affecting the expression of numerous genes in a growth phase-dependent manner. In addition, we also demonstrate that a functional Ler is necessary for the down-regulation of LEE by GadE. Characterization of the AR phenotype of AgadE revealed that GadE is indispensable for a functional GAD system and for survival of E. coli O157:H7 in a simulated gastric environment. 40 MATERIALS AND METHODS Bacterial strains and growth conditions. Bacterial strains and plasmids used in this study are summarized in Table 2.1. All strains were stored at -70°C in LB broth containing 10% glycerol, inoculated into 10 ml LB broth, and grown to an 00600 of ~01 to recover cells. To minimize the confounding effect of acidic pH that would develop in stationary phase of growth in un-buffered medium, the strains were grown in MOPS minimal medium buffered to pH 7.4. Cells recovered in LB were then grown twice to stationary phase in MOPS-buffered minimal medium before a final transfer at 1:30 dilution into 100 ml MOPS medium for RNA isolation and model stomach assay (12). Genetic manipulations. E. coli O157:H7 Sakai AgadE and Aler strains were constructed by the modified one-step gene inactivation method for EHEC developed by Murphy et al (32, 98). Briefly, recombinant PCR products containing a kanamycin (Km) resistance marker flanked by 4550 bp sequences homologous to the upstream and downstream regions of target genes were generated using the primers listed in Table 2.2 from plasmid pKD4 (32). PCR products were electroporated into red recombinase-producing E. coli O157:H7 (TW15901) as described (98) and the transfon'nants were identified on LB agar plates with 25 pg/ml Km at 37°C. The Km resistance marker was removed from the AgadE by introducing plasmid pCP20 that encodes FLP recombinase (32). Subsequently, the double mutant, AgadEA/er was constructed by one step inactivation of Ier in AgadE using the method described above. 41 TABLE 2.1 Bacterial strains and plasmids used in this study Strain or Genotype Reference or plasmid Source Strains TW08264 E. coli O157:H7 RIMD0509952 (Sakai) wild (96) type TW15901 TW08264 harboring pKM208 plasmid This study TW15902 A gadE This study TW15903 AgadEA/ertsz This study TW15904 Aler.:Km This study TW15905 AgadE/pCR2.1gadE This study Plasmids pKM208 Red recombinase expression plasmid, Apr (32) pKD4 Template plasmid for lambda Red (32) recombination system, Kmr pCP20 Flp recombinase expression plasmid, Cmr (32) pCR2.1 Cloning vector lnvitrogen pCR2.1gadE gadE cloned into pCR2.1 This study 42 TABLE 2.2 Oligonucleotide primers used for one step inactivation Primer“ Sequence (5’ to 3’)b gadE-H-F ATCAATTCCCTGTCAGAGATCAAAAAAGTAGGCAATAAACCCTTCAAGG Tgtgtaggctggagctgcttc gadE-H-R CTCGTCATGCCAGCCATCAATTI'CAG'ITGCT'I'ATGTCCTGACTAAAAATA catatgaatatcctccttag Ier-H-F" T'ITCATCTI'CCAGCTCAGTTATCGTTATCAT'ITAA'ITATTI'CATthgtaggct ggagctgcttc Ier-H-R* GTTGGTCCTTCCTGATAAGGTCGCTAATAGCTTAAAATA'ITAAAGcatatga atatcctccttag aThe homology regions for the Ier primers are from lyoda et al., (60) b Priming sites for pKD4 are in lowercase letters 43 For complementation of the AgadE strain, DNA fragments of 2,630 bp containing the Sakai gadE coding region and additional flanking regions of gadE were amplified from E. coli O157:H7 Sakai chromosomal DNA using TaKaRa LA TaqT" polymerase (Takara Bio USA, Madison, WI). The resulting PCR products were cloned into pCR®2.1-TOPO® vector (lnvitrogen, Carlsbad, CA) to make pCR2.1gadE plasmid, which was transformed into the AgadE strain creating the AgadE/pCR2.1gadE strain. RNA isolation and cDNA labeling. For RNA isolation, the wild type Sakai and AgadE strains were grown to early exponential phase (2.25 h, ODeoo ~0.25) and stationary phase (5.5 h, ODsoo ~1.5) in MOPS medium as described above. RNA was isolated from five independent cultures using hot acid- phenolzchloroform extraction. At each growth phase, 5 ml of culture was mixed with equal volume of hot acid-phenolzchloroforrn (pH 4.5 with Iso Amyl Alcohol (IAA), 125:2521) (Ambion, Austin, TX) and incubated at 65°C with periodic shaking for 10 min. The samples were centrifuged at 3220 x g for 20 min and the supernatant was subjected to further extractions with phenolzchlorofomi and chloroform:lAA (12). RNA was precipitated overnight at -70°C in 2.5 volume 100% ethanol and 1/10 volume 3 M sodium acetate pH 5.2. RNA purification and DNase treatment of RNA samples were done with the RNeasy kit (Qiagen, Valencia, CA) and RNA quality was assessed on a fonnaldehyde-agarose gel. Six pg of RNA was used for reverse transcription reactions containing 3 pg random primers (lnvitrogen), 1x first strand buffer (lnvitrogen), 10 mM DTT, 400 U Superscript II (lnvitrogen), 0.5 mM of dATP, dCTP and dGTP, 0.3 mM 44 d‘I‘l'P and 0.2 mM amino-allyl dUTP (12). After incubation at 42°C overnight, the cDNAs were purified with Qiagen PCR clean up columns with phosphate wash buffer (5 mM K2HPO4, pH 8.0, 80% ethyl alcohol) and phosphate elution buffer (4 mM K2HPO4, pH 8.5) and were coupled with either Cy3 or Cy5 dyes (Amersham Biosciences, Piscataway, NJ) as described previously (12). cDNA hybridizations. Hybridizations were performed according to a loop design, which included between strain (wild type vs. mutant at same growth phase) and within strain (same strain at two growth phases) comparisons. Five biological replicates were included for each comparison resulting in 20 arrays. As described (12), the cDNAs were hybridized onto microarray slides printed with 6088 ORFs including 110 ORFs from pO157 plasmid, and representing E. coli strains K12, EDL933 and Sakai. Arrays were scanned with an Axon 4000b scanner (Molecular Devices, Sunnyvale, CA) followed by image analysis using GenePix 6.0 (Molecular Devices) (12). Data analysis. The microarray data were analyzed using R (v. 2.2.1) and the MAANOVA (v. 0.98.8) package. Raw intensity values from replicate probes were averaged and logz transformed after normalization with the pin-tip LOWESS method. The normalized intensity values were fitted to a mixed model ANOVA considering array and biological replicates as random factors and dye, strain and growth phase as fixed factors (30). The linear model tested was Y (intensity) = array + dye + strain (wild type or mutant) + growth phase (exponential or stationary) + strain*growth phase + sample (biological replicate) + error. Each main effect had 2 levels: mutant and wild type for strain and exponential and 45 stationary phases for growth phase. The design included between and within strain comparisons using 5 biological replicates. Significant differences in expression due to strain, growth phase and strain*growth phase were determined using the Fs test in MAANOVA, which uses a shrinkage estimator for gene- specific variance components that makes no assumption about the variances across genes (31) with 500 random permutations to estimate the p-values. ANOVA with a mixed linear models have been used to analyze microarray experiments with repeated measures where transcript levels of the strains at two different growth phases are measured (6, 63, 80). The q-value package in R was used for determining the false discovery rate (FDR) (146). Overrepresentation of gene sets with a common biological function in the wild type or mutant strain were determined using the GSEA Preranked analysis in Gene Set Enrichment Analysis (GSEA v2.0) program (Broad Institute, Massachusetts Institute of Technology) (148). The gene sets were designated based on the TIGR annotation for the Sakai genome (http:l/cmr.jcvi.org/tigr- scripts/CMR/GenomePage.cgi?org=ntec03). Additionally, two gene sets, the LEE and the AFI and GAD, also were included in the analysis (33, 41, 91, 158). The pattern search tool in coIrBASE (http://xbase.bham.ac.uk/colibase/pattem.pl?id;1073) was used to identify GAD box (22, 85) sequences (5'-'I'l'AGGA'l‘I'I"I'G'l"l'AT'I'I'AAA-3') in the putative promoter regions of genes differentially regulated between wild type and AgadE. A cut off of 70% similarity to the query sequence was set to apply higher stringency because experimental confirmation of GadE binding was not 46 conducted. A sequence logo for the consensus sequence was created at http://weblogo.berkeley.edu/logo.cqi. Quantitative real-time PCR. RNA isolations were conducted at both exponential and stationary phases of growth. For assays with AgadE/pCR2.1gadE, AgadEA/er and AIer, RNA was isolated only at the exponential phase. Taqman assays (11) were used for quantifying the expression of gadA, gadB and Ier with mdh as reference for normalization. For the remaining genes, SYBR green chemistry was used for measuring expression levels. Primers were designed using the Primer3 server (125) based on the published reference genome sequence of E. coli O157:H7 strain Sakai (Table 2.83 in appendix). cDNA synthesis was conducted using iScript Select cDNA synthesis kit (BioRad, Hercules, CA) with 1 pg of total RNA according to the manufacturer’s instructions. After reverse transcription, 5-fold serial cDNA dilutions were used for Q-PCR assays containing 12.5 pl 2x i0 SYBR green supermix (BioRad), 0.63 pl of each primer (10 pM stock), 9.24 pl of H20 and 2 pl cDNA with cycle conditions of 95°C for 2 min followed by 40 cycles of 10 sec at 95°C and then 20 sec at the specific annealing temperature (12). The expression levels of the 168 rRNA gene were used for normalization of data and the relative expression levels were quantified using Pfaffl’s method (113). The results presented are averages from at least three biological replicates :l: standard error of mean (SEM). Expression studies in E6 minimal medium. Wild type and AgadE cells were grown in EG minimal medium at pH 7.0 and pH 5.0 (83) to late exponential 47 phase (ODeoo ~0.5). RNA extractions were conducted using a modified hot- phenol extraction protocol (15) that utilized 5% acidic phenol in ethanol. cDNA synthesis and Q-PCR methods are described above. AR mechanism assays. AR mechanism assays for the GAD, ARG and OXI systems were conducted as described previously (75, 83). Briefly, for the GAD system, strains grown in LB broth with 0.4% glucose (LBG) were challenged at pH 2.0 in a test environment (EG + glutamate) and in a control environment (EG), whereas for the ARG system, after growth in BHI broth with 0.6% glucose, strains were tested in test (EG + arginine) and control (EG) environments at pH 2.5. For testing the OXI system, strains were grown in LBMES (pH 5.0) and EG (pH 7.0) and challenged at pH 2.5 in EG. Samples were withdrawn at specific time points (30 min or 1 h intervals) and plated on LB agar plates using an Autoplate 4000 Spiral Plater (Spiral Biotech, Bethesda, MD). Colonies were counted after overnight incubation at 37°C using the 0- Count (Spiral Biotech). Assays were conducted for at least two biological replicates, each with two technical replicates. CFU/ml from technical replicates were averaged and converted to log10 CFU/ml. The results reported are averages 21: SEM for at least three experiments. Model stomach assay. The model stomach system (M88) (64) was prepared as described previously (11). Gerber Turkey Rice Dinner© baby food (30 g) was mixed with 120 ml of synthetic gastric fluid (pH 1.75) yielding a final acidity of pH 2.5. Contents of the M88 were stomached for 30 sec, sampled, diluted, and plated onto LB agar plates every 30 min for 1.5 h to enumerate 48 viable cells. CFU/ml from duplicate plates were averaged and converted to log10 CFU/ml. The results reported are averages 1: SEM for three experiments. Microarray data accession. Microarray data are available at NCBI GEO (httpzllwww.ncbi.nlm.nih.qov/geo), accession number GSE13132. 49 RESULTS Identification of genes regulated by GadE. Inactivation of gadE did not cause a significant difference in the growth rate of E. coli O157:H7; the generation time of wild type was 43.7 min and that of AgadE was 44.1 min. A two-factor ANOVA with two main effects (strain and growth phase) and the interaction effect (strain*growth phase) was used to determine the impact of gadE inactivation on the transcriptome of E. coli O157:H7 at both exponential and stationary phase. Genes with FDR < 0.1 for strain effect and FDR < 0.05 for interaction effect were considered as regulated by GadE. Significant strain effects were identified for 60 genes, indicating differential expression between the wild type and AgadE (FDR < 0.1) (Table 2.81 in appendix). Of these, 58 genes had higher transcript levels in AgadE demonstrating that GadE has a negative influence on their transcription, and 2 genes had lower transcript levels in AgadE, indicating that GadE has a positive influence on their transcription. Among the 2 genes with lower transcript levels in AgadE, ECs3904 had significantly greater transcript levels in exponential phase compared to stationary phase, and ECs2294 had greater levels in stationary phase. Among the 58 genes with higher transcript levels in AgadE, 33 were significantly higher in exponential phase, including LEE genes tir, espF and cesT, and 25 were higher in stationary phase, including genes vgrE, iIvG, and treR (Table 2.81 in appendix). A significant interaction effect, which indicates that inactivation of gadE affects expression of genes differently at each growth phase, was identified for 122 genes, including the AFI and GAD genes gadA, gadB and gadC (FDR < 50 0.05) (Table 2.82 in appendix). GSEA analysis identified significant enrichment of the AFI and GAD genes (FDR < 0.05) in wild type and LEE genes (FDR < 0.05) in AgadE. In summary, the array data demonstrated that inactivation of gadE had an effect on several genes, including members of the LEE pathogenicity island, in addition to the GAD and AF I genes. Interaction effects of gadE inactivation and growth phase. Expression of gadE was 84.2 t 13.4-fold higher in stationary phase compared to exponential phase in wild type as measured by Q-PCR. Since the expression of gadE is growth phase-dependent, it affected the expression of 122 genes in a growth phase-dependent manner leading to a significant interaction effect (FDR <0.05) (Table 2.82 in appendix). Genes with a significant interaction effect included several genes belonging to the AFI and GAD system, as well as a number of genes involved in energy metabolism and genes encoding transcriptional regulators. In exponential phase, GadE exhibited a positive effect on transcript levels of genes such as cyoDC, sdhCDAB, and sucAB involved in energy metabolism, while in stationary phase, GadE exhibited a negative effect on transcript levels of these genes (Table 2.82 in appendix). Transcript levels of genes encoding sigma factors and transcriptional regulators such as rpoS, IysR, and pspF were slightly elevated in AgadE in exponential phase, but were 1.3- to 2.9-fold higher in the wild type compared to AgadE in stationary phase, indicating that these genes are positively regulated by GadE at stationary phase (Table 2.82 in appendix). 51 Putative GadE binding site upstream of GadE regulated genes. To determine whether differentially expressed genes identified by the microarray analysis are regulated directly by GadE, a pattern search against the E. coli O157:H7 Sakai genome was conducted in coerASE to identify potential GAD boxes upstream of these genes. A conserved GAD box sequence described previously in E. coli K12 (22, 85) was used for the pattern search. Typically one GAD box is observed upstream of GadE regulated genes, gadA and gadBC, in E. coli K12 (22). Sequences with 270% similarity to the query sequence were considered as putative binding sites of GadE. Matching sequences were detected upstream of 8 genes that were significantly differentially expressed in AgadE. As expected, GAD box sequences preceding gadA and gadB showed 100% similarity to the conserved sequence in E. coli K12. There were two putative GadE binding regions identified upstream of hdeD. Matching sequences also were identified in the upstream regions of vgrE and 3 LEE genes (sepZ, 9300, and Ier) (Fig. 2.1). Expression of AFI and GAD genes in AgadE. Inactivation of gadE resulted in a decrease in expression of many of the AFI and GAD genes and the magnitude of decrease was dependent on the growth phase. Six AFI and GAD genes, including gadA, gadBC and hdeBAD had a significant interaction effect (Table 2.3). These results were verified with Q-PCR. At exponential phase, for gadA, gadB, gadC, hdeA and hdeB, microarray data revealed higher expression in the mutant, whereas Q-PCR detected higher expression in wild type. This discrepancy could be due to the negligible expression of these genes at 52 Gene % identity Sequence gadA 100 TTAGGATI'I'I‘GTTAT'ITAAA gadB 100 TI‘AGGATTI‘TG’ITA'ITI‘AAA pspF 7o TI‘ATCTT'I’ITGATTTATAAA hdeD1* 75 'ITAGGAAATI'I‘TTATTAAAT hdeDz * 7o ATCAGATA’I'I'ITI‘AT'ITCAA vgrE 7o TGA'I‘TA'I'ITTGTTGACTAAA sepZ 7o GGATAA'I'ITGG'ITATTTATA escC 7o ”I'I‘CGACTCT'I'I‘TTAA'I'I‘AAA Ier 7o ATATGATTITI'I'TGTI‘GACA M iiilanAA Figure 2.1. Alignment of putative GAD box sequences. Sequences upstream of 8 gadE- regulated genes with greater than 70% identity to the conserved GAD box sequence were identified by the pattern search tool in coIrBASE. Letters in boldface in the table are unmatched bases. Asterisks indicate the two matching sequences found upstream of hdeD. 53 TABLE 2.3 Effect of gadE inactivation on expression of AFI and GAD genes in exponential and stationary phase Exponential phase Stationary phase expression ratio expression ratio (WT/ (wrr AgadE) AgadE ) ECS :1 Micro Micro number Gene Function array QPCRb array QPCRb E03 4389 M98 Periplasmic 0.6 5.2120 2.8 31.411.5 chaperone ECs 4390 hdeA Protection from 0.3 3.2103 11.1 39.3130 organic acid metabolites E05 4391 hdeD Acid resistance at 0.6 2.5 high cell densities ECs 4397 gadA Glutamate 0.4 2210.8 6.4 46.3113] decarboxylase isozyme E03 2097 gadC* Glutamate-GABA 0.5 2310.5 4.1 1610.9 antiporter ECs 2098 gadB* Glutamate 0.4 2211.5 8.0 22.511.67 decarboxylase isozyme 3 Genes marked with an asterisk are GAD system genes not encoded in AF I, genes in bold face have putative GAD boxes upstream of their sequence. ° Fold change1SEM as determined by Q-PCR 54 exponential phase in minimal medium at neutral pH (84). It has been shown that microarrays are less sensitive than Q-PCR in detecting changes in expression when the transcript levels are low (21 ). At stationary phase, Q-PCR supported the microarray results for these 5 genes, though a greater difference was detected in transcript levels compared to the microarray (Table 2.3). This underestimation of fold changes by microarrays has been reported in many of the previous studies (27, 174). The fold increase in expression of gadA and gadB in wild type compared to AgadE was approximately 10-20 times higher at stationary phase than at exponential phase. Similarly, the increase in expression of hdeAB in wild type was 6-12 times higher at stationary phase than at exponential phase (Table 2.3). In summary, both microarray and QPCR data demonstrated that the difference in expression of AFI and GAD genes between wild type and AgadE was minimal at exponential phase whereas at stationary phase there was a marked decrease in expression of gadABC and hdeAB in the mutant. Moreover, in the wild type, the expression of AFI and GAD genes increased markedly from exponential to stationary phase whereas in the AgadE their expression decreased minimally or remained unchanged as the cells entered stationary phase. This demonstrates that inactivation of gadE abrogates the growth phase regulation of AFI and GAD genes in E. coli O157:H7. Six AF I genes did not show differential expression between the mutant and wild type, including two AraC-Iike regulators of the GAD system, gadX and gadW. A significant increase in expression of multi drug resistance-related efflux 55 pump gene, yhiU, was observed in the mutant (Table 2.81 in appendix). Consistent with previous observations (12), all of the AFI and GAD genes had significantly higher expression at stationary phase than at exponential phase. GSEA analysis confirmed the enrichment of the gene set representing AF I and GAD genes in the wild type. GadE represses expression of LEE. Inactivation of gadE significantly elevated the expression of 19 LEE genes independent of the growth phase, among which 8 genes showed 21.35-fold increase in expression in the mutant (Table 4). Fold changes in expression between mutant and wild type detected by microarray and Q-PCR were highly correlated for the LEE genes (Table 2.4). Contrary to previous findings (156), the LEE encoded regulator, Ier, was up- regulated by 2.2-fold in the mutant as determined by Q-PCR (1 .35-fold in the microarray). Other LEE genes such as fir and espD were up—regulated by 1.8- and 1.6-fold, respectively (microarray detected 1.4-fold for both) (Table 2.4). Most of the LEE genes with significant differential expression between AgadE and the wild type also had higher expression in exponential phase compared to stationary phase. Six LEE genes, including sepZ (espZ), had higher expression at stationary phase and sepZ showed a 1.4-fold (1 .36-fold in microarray) increase in expression in the mutant. Two non-LEE encoded effectors, nIeGZ-2 and nIeG2-3, also had significant increase in expression in the mutant at stationary phase. In the array data, eae, espA and espB were not significantly up-regulated, whereas Q-PCR identified an Increase in expression of 1.3-1.6-fold 56 TABLE 2.4 LEE genes significantly up-regulated in AgadE (FDR<0.1) ECs Genea Function Micro Q-PCR Microarray number array ( (AgadE/ (Stationary/ AgadE WT) b Exponential) NVT) C ECs 4550 espF type III secretion 1.30 1.34100 0.6 system, secreted 8 effector E05 4553 cesD2 type III secretion 1.36 0.5 system chaperone E05 4554 espB* type III secretion 1.20 13810.0 system, secreted 9 translocator ECs 4555 espD type III secretion 1.38 1.63100 system, secreted 4 translocator E05 4555 espA" type III secretion 1.10 1.58102 system, secreted translocator ECs 4558 9300 type III secretion 1.18 1.3 system, structural protein ECs 4558 eae’ gamma intimin 1.20 1.3101 ECs 4560 cesT type III secretion 1.38 0.3 system, chaperone E03 4561 tir translocated 1.39 18410.0 0.5 intimin receptor 6 protein ECs 4562 map type III secretion 1.26 system, secreted effector ECs 4563 cesF type III secretion 1.31 1.6 system, chaperone E05 4564 espH type III secretion 1.33 3.1 system, secreted effector ECs 4565 sepQ type III secretion 1.27 2.9 system, structural protein 57 Table 2.4 continued... E05 4567 orf15 orf of unknown 1.19 0.9 function ECs 4571 sepZ type III secretion 1.36 1.44100 6.4 system, secreted 6 effector ECs 4572 rorf8 orf of unknown 1.26 3.0 function ECs 4575 escC type III secretion 1.21 0.3 system, structural protein E03 4584 orf5 orf of unknown 1.34 0.5 function E05 4585 orf4 orf of unknown 1.41 0.6 function E03 4586 orf3 orf of unknown 1.37 0.6 function ECs 4587 cesAB type III secretion 1.35 0.6 system, chaperone ECs 4588 Ier type III secretion 1.35 2.23102 0.8 system, regulator 6 E05 4590 espG type III secretion 1.27 0.6 system, secreted effector ECs 1994 nIeGZ-Z non LEE-encoded 1.26 2.5 effector ECs 2156 nIeGZ-3 non LEE-encoded 1.28 2.9 effector 8 Genes underlined are non LEE-encoded effectors and genes with asterisks are not significant in microarray, but detected as up-regulated by Q-PCR. Genes in bold face have putative GAD boxes upstream of their sequence. b Exponential phase fold change1SEM as determined by Q-PCR c Ratios are reported only for genes with a significant growth phase effect (FDR<005) 58 in the mutant, similar to other significant LEE genes (Table 2.4). Significant enrichment of the LEE gene set in AgadE was detected by GSEA. To further confirm the negative influence of GadE on LEE expression, exponential phase transcript levels of select LEE genes were measured in the complement strain, AgadE/pCR2.1gadE, which over-expresses gadE. Expression of Ier decreased by 6-fold in the complement, whereas espD and sepZ decreased by 9.9- and 9.5- fold, respectively. Altogether, these data demonstrate that GadE is a repressor of LEE genes, including Ier, in E. coli O157:H7. Repression of LEE by GadE is mediated through Ler. Expression data from AgadE and gadE-over expressing strains demonstrated that GadE negatively regulates the expression of Ier. Moreover, a pattern search in coIrBASE to find GAD box sequences in the LEE island region revealed a putative GAD box upstream of Ier (-199 to -180bp) with 6 mismatches (70% identity). To determine whether repression of LEE genes by GadE is mediated by Ler, we inactivated Ier in both the AgadE and wild type strains and compared the expression of select LEE genes. If GadE down-regulates LEE expression independent of Ier, then an increase in expression of LEE genes in the double mutant, AgadEA/er, similar to AgadE was expected. In this study, however, expression of tir, sepZ, espA, espB and espD decreased by 11.6-, 20.1 -, 38.2-, 17.9- and 33.4—fold, respectively in AgadEA/er (Fig. 2.2). A similar decrease in LEE expression was observed in Mar. These data demonstrate that the positive effect of gadE inactivation on LEE expression is reversed by Ier inactivation, suggesting that Ier is essential for the repression of LEE by GadE. 59 Fold change in expression '50 T fi j I I T tir sepZ espA espB espD Ier Gene Figure 2.2. Exponential phase expression of LEE genes, tir, sepZ. espA, espB, espD and Ier in AgadE (grey bars), AgadEA/en:Km (black bars) and Aleme (white bars) compared to the wild type. Results shown are average fold change in expression measured by QPCR with standard error of mean (SEM) from at least three biological replicates. 6O Effect of acidic pH on expression of LEE genes. To determine the influence of acidic pH on expression of the LEE genes, wild type and AgadE cultures were grown to exponential phase (0D600~0.5) in EG minimal medium adjusted to pH 7.0 (control) and pH 5.0 (moderately acidic) for comparing gadE and LEE expression. Growth in EG pH 5.0 resulted in a 34-fold increase in the wild type expression of gadE. Expression of three LEE genes, Ier, espD and sepZ, was down-regulated in EC pH 5.0 compared to EG pH 7.0 in the wild type (Table 2.5). To determine if the down-regulation of LEE genes in response to moderate acidity was directed exclusively by GadE, the expression of LEE genes in AgadE grown in EG pH 5.0 was assessed. There was a 3.9- and 6.5-fold decrease in expression of espD and sepZ, respectively, in AgadE, however, this decrease was lower relative to the wild type, where expression of espD and sepZ decreased by 7.5- and 9.4-fold, respectively. Interestingly, the pattern of Ier expression at acidic pH was different from the other 2 LEE genes tested. In the wild type, there was a 6-fold decrease in Ier expression, whereas in AgadE, Ier expression increased 4.5-fold at pH 5.0 (Table 2.5). Together, these observations indicate that down-regulation of the major LEE regulator, Ier, is mediated through GadE in response to moderate acid stress, whereas repression of other LEE genes under the same conditions involve additional GadE/Ler-independent factors. Two regulators that may affect the expression of LEE genes at acidic pH, independent of GadE and Ler, are GadX and Eng, which have a negative effect on LEE expression in EPEC (99, 138). Hence, we measured the expression of 61 TABLE 2.5 Effect of pH 5.0 on expression of GAD regulators and LEE genes Fold change in expression (pH 5.0/pH7.0) Gene WT AgadE gadE 34.4169 no expression gadX 10412.5 4311.0 eng 2.8103 3.4107 Ier 0.16100 4.5104 espD 0.161006 02610.02 sepZ 01210.04 01710.05 62 these regulators in EG pH 7.0 and EG pH 5.0 in wild type and AgadE and observed a strong induction of both genes in both strainsat pH 5.0. The gadX gene had > 2-fold higher induction in wild type, whereas eng induction was similar in both wild type and AgadE (Table 2.5). Functional GadE is necessary for optimal performance of the 3 principal AR mechanisms in E. coli O157:H7. To functionally confirm the microarray data, which revealed a marked decrease in expression of GAD and AFI genes in AgadE, we conducted AR mechanism assays for the GAD, ARG and 0Xl systems. The ARG and OXI systems were included in the study since GadE has been shown to influence their function in E. coli K12 (84). The AgadE strain could not survive in the test environment for the GAD system (pH 2.0 with glutamate) even for 30 min, indicating a non-functional GAD system (Fig. 2.3A). The wild type and complement showed a log reduction of 0.20 1 0.08 and 0.17 1 0.02 CFU/ml, respectively, after 6 h of exposure to the test environment. In the ARG system test environment (pH 2.5 with arginine), survival of AgadE was similar to the wild type and complement for up to 2 h. However, at 4 h there was reduction in viable cell numbers and the mutant showed high variation in cell numbers up to 5.5 h, and by 6 h, no viable mutants were recovered. The wild type and complement showed a log reduction of 1.07 1 0.08 and 1.22 1 0.16 CFU/ml, respectively, after 6 h (Fig. 2.3B). The OXI system was less effective in protecting all three strains compared to the ARG system. The mutant survived for only 3 h at pH 2.5 and the log reduction in CFU/ml after 4 h was 2.27 1 0.1 for wild type (Fig. 2.30). Interestingly, the complement also did not survive after 3 h, 63 Figure 2.3 A-C. AR mechanism assays. Survival of wild type (white bars), AgadE (grey bars) and AgadE/pCR2.1gadE (black bars) strains in the 3 AR systems. (A) Survival for the GAD system test at pH 2.0 (B) Survival for the ARG system test at pH 2.5 (0) Survival for the OXI system test at pH 2.5. The results presented are average CF U/ml with SEM from 3 experiments for each AR system. 64 w m D .1. 3 as: N Fmdo 1 Y E 2:: omfimmév deo L L J J. A Y [\(OIDVC‘ONi—O NCDIOVOONw—O 3 see c A 1 NCDlDV'OONv-O IUJ/ndO 60I 65 indicating that gadE in trans does not reconstitute the phenotype for the OXI system. It is possible that flooding the cell with multiple copies of gadE, as in the complement, adversely effected the functioning of OXI system. These findings demonstrate that inactivation of gadE abolished the functioning of the GAD system and rendered ARG and OXI systems less effective in protecting the cells against low pH. Survival of AgadE in a simulated gastric environment. Since the 3 principal AR systems were defective in protecting the AgadE from acidic stress in defined minimal test conditions, the ability of AgadE to survive in a complex acidic environment was assessed using the M88 (pH 2.5). The wild type and complemented cells showed an average log reduction of 1.05 1 0.06 and 0.32 1 0.04 CFU/ml, respectively, after 1.5 h in the M88 (Fig. 2.4). Viable cells could not be recovered from the M88 inoculated with AgadE, which indicates that functional gadE is necessary for survival in the simulated gastric environment. 66 log”) CFU/ml N U) A LII 0 \l _. 0 30 60 90 Time (min) Figure 2.4. Survival of wild type (white bars), AgadE (grey bars) and AgadE/pCR2.1gadE (black bars) strains in M88. The average log CFU/ml with SEM from 3 experiments is plotted for each time point. 67 DISCUSSION Although the upstream regulatory circuits and downstream effects of GadE in non-pathogenic E. coli has been examined (39, 84), little is known about GadE and its role in AR and virulence among pathogenic E. coli strains. Here, the role of the GadE regulator in AR and virulence of E. coli O157:H7 was investigated by constructing an isogenic AgadE strain and comparing its expression profiles with that of the wild type strain. Our findings demonstrate that besides being a positive regulator of GAD and many AFI genes, GadE acts as a negative regulator of the LEE pathogenicity island, an important factor in the virulence of E. coli O157:H7. GadE, along with additional regulators, is involved in the down-regulation of LEE expression at moderately acidic pH. In addition, the characterization of AR phenotypes of AgadE revealed that GadE is indispensable for a functional GAD system and plays a vital role in the survival of E. coli O157:H7 in a simulated gastric environment. In this study, the microarray data demonstrated that inactivation of gadE in E. coli O157:H7 altered expression of 60 genes independent of growth phase and 122 genes in a growth phase-dependent manner. The genes with altered expression included both AR and virulence genes, indicating that the regulatory function of GadE is not restricted to AR, but has a more global effect on the transcriptome of E. coli O157:H7. Over-expression of gadE in non-pathogenic E. coli was shown to affect the expression of ~40 genes, including GAD genes (54). Most of these genes, however, differed from those identified following inactivation 68 of gadE in O157:H7, suggesting that apart from its effect on the GAD system, GadE has additional regulatory functions in E. coli O157:H7. Expression of gadA, gadB and gadC were not completely abolished in E. coli O157:H7 AgadE similar to E. coli K12 where minimal expression of GadAB proteins was observed in AgadE (84). Expression of GAD system components in the absence of GadE could be induced by the GadX regulator, which has been shown to bind to and activate gadA and gadBC transcription directly under in vitro conditions, but not during in vivo growth (132, 159, 162). Another interesting observation was that at stationary phase, the magnitude of increase in expression of gadA and gadB in the wild type compared to AgadE were different, indicating that the inactivation of gadE affects these duplicated genes in distinct ways. This corroborates the recent finding that the sequences of gadA and gadB are divergent in E. coli O157:H7 in contrast to other E. coli strains where gene conversion events between gadA and gadB have led to genetic homogenization (8). In contrast to the wild type, no increase in expression of GAD genes was observed in AgadE as the cells entered stationary phase demonstrating that GadE is required for the growth phase regulation of GAD genes. This study demonstrates that gadE inactivation has differential effects on the expression of AFI genes in E. coli O157:H7. In non-pathogenic E. coli strains, gadE induces the expression of AFI genes such as hdeB, hdeA, hdeD, gadX and thF in addition to gadA (54, 91, 126). In this study, expression of gadA and hdeBAD showed growth phase-dependent down-regulation in AgadE. However, 69 gadX and yhiF were not differentially expressed in AgadE indicating that at pH 7.0, loss of gadE does not influence the expression of these two genes in E. coli O157:H7. These differences between O157:H7 and non pathogenic strains could also be due to the differences in growth conditions, since strains were grown in rich or minimal media at acidic pH in most of the previous studies. The relationship between AR and virulence of pathogenic E. coli, particularly the interactions between the GAD system and LEE, remains poorly defined. Few studies in the past have shown that some 'of the GAD system regulators negatively affect LEE expression (24, 99, 138, 157). In E. coli O157:H7 Sakai, gadE inactivation was found to increase the expression of LEE encoded espB, espD and tir genes, but not Ier (156). Hence, Mellies et al. considered that GadE mediated down-regulation of LEE was independent of Ler and the pathway through which GadE affects LEE was undetermined (95). Moreover, the extent to which GadE inhibits the transcription of LEE genes has not been demonstrated quantitatively before. The data presented here demonstrate that GadE has a global effect on LEE genes: GadE influences the expression of at least 19 LEE encoded genes belonging to all five LEE operons and 2 non-LEE encoded effectors. These data also provide insight about the mechanism underlying GadE-mediated LEE down-regulation. In contrast to the previous study (156), there was a significant increase in the expression of Ier in AgadE, which may be due to differences in growth medium used. The previous study used DMEM containing glycerol (156) whereas in this study MOPS minimal medium was used for growing the cells. This discrepancy could also be due to 70 difference in sensitivity of the assays used (northern blotting vs. QPCR and microarray). The negative correlation between expression of gadE and Ier was marked in the gadE-over expressing strain in which Ier expression was substantially down-regulated. Furthermore, the identification of a putative GAD box sequence upstream of Ier with 70% identity to the conserved GAD box sequence provides additional evidence of direct regulation, as GadE has been shown to bind to box sequences with as low as 60% identity to the conserved sequence (85). Additionally, inactivation of Ier in AgadE led to a marked decrease in LEE expression, confirming that Ier is essential for the up-regulation of LEE in AgadE. Taken together, these findings illustrate that GadE indirectly down-regulates LEE expression most likely through down-regulation of Ler. However, additional putative GAD boxes were observed upstream of other LEE genes, sepZ and escC, also and therefore, it is possible that GadE directly regulates these LEE genes independent of Ler. Because GadE negatively influences LEE expression, we hypothesized that environmental conditions that induce gadE may down-regulate the expression of LEE. Two conditions that lead to induction of gadE are entry into stationary phase and acidic pH (39). Stationary phase expression of LEE genes has been described previously (12). Similarly, influence of several environmental factors such as temperature, bicarbonate ion concentration and membrane stress on the expression of LEE has been investigated (2, 100, 157, 164). However, the effect of pH on LEE expression and the factors regulating that effect in EHEC remain largely unknown. Our experiments demonstrated that 71 exposure to moderately acidic pH strongly induces gadE and has a substantial negative effect on LEE expression in E. coli O157:H7. This inhibitory effect could be more profound in extreme acidic conditions such as the gastric environment. To determine whether the acidic pH-induced down-regulation of LEE is exclusively regulated by GadE, we measured the expression of LEE genes in the AgadE following growth at pH 5.0. Except for Ier, a partial down-regulation at acidic pH was observed in the expression of the LEE genes in AgadE, indicating that GadE is not the only regulator responsible for the pH induced down- regulation of LEE. This partial down-regulation of LEE is not mediated through Ler, as expression of Ier was increased in the AgadE at pH 5.0. To understand this phenomenon further, we analyzed the expression of other AR regulators that could act on LEE, independent of GadE, and found that gadX and eng were also strongly induced at acidic pH in both wild type and AgadE. GadX has been shown to negatively regulate the expression of LEE genes through the plasmid encoded regulator, Per, in EPEC (138). However, the effect of GadX on expression of LEE in EHEC has not been determined. It is possible that GadX regulates LEE through an unknown regulator in EHEC. The decrease in LEE gene expression in an acidic environment in the AgadE is likely to be mediated by Eng also. Previously, Eng has been shown to repress LEE, independent of Ler, by activating ydeO and ydeP (99). Because YdeO is a positive regulator of gadE, the partial decrease in LEE expression in AgadE may occur through YdeP. Collectively these experiments suggest that GadE, GadX and Eng may cooperatively repress the expression of LEE genes at acidic pH and that GadE is 72 the sole regulator responsible for the changes in expression of Ier in the acidic conditions used in this study. The AR of AgadE also was characterized by assessing survival at pH 2.0 and 2.5 in the minimal AR mechanism environments and in the complex acidic conditions of the M88. The AgadE failed to survive the acid challenge at pH 2.0 + glutamate, indicating lack of a functional GAD system. Inactivation of gadE in E. coli O157:H7 negatively impacted the protective ability of the ARG and OXI AR systems. The survival of AgadE was tested in the M88, which evaluates the ability of bacterial strains to survive in a gastric environment after ingestion of food (64). The GadE central regulator, and thus a functional GAD system, is a critical component for survival, as inactivation of gadE abrogated the ability of E. coli O157:H7 to survive in the M88. Hence, GadE most likely plays a protective role during the passage of 0157 through the gastric environment. This assumption is supported by a previous study by Price et al., (119) which demonstrated that gadC is required for the survival of E. coli O157:H7 in calves. In summary, this study shows that GadE is an important regulator that modulates the expression of AR and virulence genes in E. coli O157:H7 in response to environmental conditions similar to those that are found in various food matrices and the human gastrointestinal tract. GadE has acquired additional functions in E. coli O157:H7 and it acts as a link between AR and virulence: it activates the GAD system of AR and at the same time down- regulates the expression of LEE genes, which are important for the adhesion of the organism to intestinal mucosa and development of AE lesions. 73 Consequently, we propose that during passage through the human stomach GadE protects E. coli O157:H7 by inducing the GAD system and aids in energy conservation by inhibiting the unnecessary expression of the LEE genes. As the organism reaches the intestine, environmental changes including alkaline pH and high NaHCO3 concentration induce the LEE regulator, Ler, which negatively regulates expression of gadE (1) leading to inhibition of the GAD system. 74 ACKNOWLEDGMENTS This work was supported by Food Safety NRI #2005-35201-16362 from the United States Department of Agriculture, and in part by funds from the NIAID, NIH, DHHS, under the Food and Waterbome Diseases Integrated Research Network (NIH Research Contract N01-Al-30058). The authors wish to thank Vanessa Sperandio, University of Texas South Western Medical center, for helpful suggestions for creating genetic complements and Galeb Abu-Ali, James T. Riordan and Shannon Manning for critical review of the manuscript. 75 CHAPTER 3 Effect of acidic pH on the expression of virulence factors of Escherichia coli O157:H7: role of glutamate decarboxylase system regulators 76 SUMMARY Effect of gastric acidity on the expression of vimlence factors remains less investigated in an important enteric pathogen, Escherichia coli O157:H7. We hypothesized that exposure to extreme acidity downregulates the expression of virulence factors in E. coli O157:H7. Stationary phase cells of E. coli O157:H7 were exposed to pH 2.0 and pH 7.0 for 10 min to analyze the expression of virulence and acid resistance genes. Supporting the hypothesis, exposure to pH 2.0 resulted in a marked decrease in the expression of the major virulence factors of E. coli O157:H7 such as Shiga toxins, locus of enterocyte effacement (LEE) and p0157 plasmid encoded genes. 0n the other hand, acid resistance genes such as gadA, gadB, and gadC showed increased expression at pH 2.0. Furthermore, we investigated whether two acid resistance regulators that negatively regulate LEE expression in enteropathogenic E. coli (EPEC), GadX and Eng, have a similar effect on LEE expression in E. coli O157:H7. Contrary to EPEC, inactivation of gadX had only a minimal effect on the expression of LEE genes in E. coli O157:H7 whereas inactivation of eng resulted in a marked increase in expression of the LEE genes, tir, espD, eae, and sepZ similar to EPEC. 77 E. coli O157:H7 is the most prevalent enterohemorrhagic E. coli (EHEC) serotype in the United States (67). It is a food borne zoonotic pathogen and causes hemorrhagic colitis and hemolytic uremic syndrome (HUS) in human beings (153). HUS is a thrombotic disorder that causes kidney damage in 15% of the children infected with E. coli O157:H7 (153). Before colonizing the lower intestinal tract of humans, E. coli O157:H7 has to pass through the stomach, which presents one of the first host defense barriers. Gastric acidity (pH 1.5 - 2.5) is lethal to most bacteria including many enteric pathogens (112). E. coli strains are extremely acid resistant and can survive gastric acidity efficiently (39). At least four mechanisms contribute to this acid resistance of E. coli, among which glutamate decarboxylase (GAD) system is the most efficient (39). The effect of acidity on the expression of virulence factors of E. coli O157:H7 is largely unknown. Major virulence factors of E. coli O157:H7 include one or more Shiga toxins (Stx), a pathogenicity island namely locus of enterocyte effacement (LEE) and a plasmid, p0157 (170). Several environmental factors such as growth phase, temperature, presence of iron, osmolarity, and presence of antibiotics have been shown to affect the expression of these virulence factors (12, 20, 95). However, the effect of acidity on the expression of Stx, LEE and p0157 genes remains less clear. Therefore, a study on the expression of virulence factors at acidic pH is critical as it could provide important insights about how E. coli O157:H7 synchronizes the expression of acid resistance genes and virulence genes when exposed to extreme acidic environments such as the human stomach. In this context, we hypothesized that upon exposure to extreme 78 acidity such as pH 2.0, E. coli O157:H7 downregulates its virulence factors and upregulates the acid resistance genes in order to facilitate survival and conserve energy. In a previous study, we determined the effect of acidity on LEE expression by growing the E. coli O157:H7 cells at pH 5.0 and demonstrated a marked decrease in expression of Ier, espD, and sepZ, at pH 5.0 (65). However, the expression pattern of genes observed at pH 5.0, although interesting, may not be representative of the expression at gastric pH as pH 5.0 represents only moderate acidic conditions. Moreover, the change in expression of other major virulence factors such as Stx and p0157-encoded genes at acidic pH was not investigated. Therefore, to determine the effect of gastric pH on the expression of virulence and acid resistance genes, we exposed stationary phase (00600 ~ 1.5) cells of E. coli O157:H7 Sakai strain to EG minimal medium containing glutamate (65, 83) at pH 2.0 and pH 7.0 for 10 min and compared the expression of selected virulence genes. The virulence genes tested were LEE genes, Ier, tir, sepZ (espZ), and espD that are involved in causing attaching and effacing lesions in the intestine; Stx genes, stx1 and stxZ, which encode two variants of Shiga toxins; p0157 genes, eth, which encodes enterohemolysin, and etpC, a type II secretion system gene involved in adhesion to the epithelial cells. Additionally, to determine the effect of pH 2.0 on acid resistance genes, the expression of three GAD system genes gadA, gadB and gadC was analyzed. For all the experiments described in this report, RNA was extracted using modified hot phenol extraction method as described previously (9). One microgram of 79 RNA was converted into cDNA, which was then used for quantitative real time PCR (qRT-PCR) as described before (65). The fold change in expression of genes was calculated by Pfaffi’s method (113). Exposure to pH 2.0 altered the expression of all the genes tested. There was a 2 - 5-fold upregulation of the three components of GAD system, gadA, gadB and gadC. This is an expected result as the GAD system has been shown to provide maximum protection to the cells at extreme acidity (39, 75). In contrast, all of the virulence genes tested showed decrease in expression of different magnitudes at pH 2.0 with some of them showing a dramatic decrease in expression. stx1 had a 28.2-fold decrease in expression whereas stx2 was down regulated by only 2.3-fold upon exposure to pH 2.0. Similarly, the LEE genes tested demonstrated decreased expression at pH 2.0, among which the highest repression of 74.1- and 60.9-fold were observed for fir and Ier, respectively. espD had a 23.6-fold decrease in expression whereas sepZ showed only a 2-fold decrease in expression at pH 2.0. The two p0157 plasmid-encoded genes eth and etpC also showed decrease in expression of 16.2- and 9.6-fold, respectively (Table 3.1). In our previous study (65), the central activator of the GAD system, GadE (39), was involved in the decrease in expression of the LEE genes at acidic pH in E. coli O157:H7. However, inactivation of gadE only partially relieved the acid- induced repression of LEE and therefore, in this study, we hypothesized that additional regulators such as GadX and Eng are also involved in this 80 TABLE 3.1 Alteration in expression of acid resistance and virulence genes upon exposure to extreme acidity Gene Fold change (pH 2.0/pH 7.0)* gadA 2.6108 gadB 2.3107 gadC 5211.4 sepZ -210.2 tir -74.1114.5 espD -23.613.3 Ier -60.9129.2 thA -16.217.8 etpC -9.610.9 stx1 -28.2116.3 stx2b -2.310.3 * Positive values indicate increased expression at pH 2.0 and negative values indicate decreased expression at pH 2.0. 81 phenomenon as GadX and Eng have been previously shown to negatively regulate LEE expression in EPEC (99, 139). To determine whether GadX and Eng regulate the expression of LEE in EHEC, AgadX and Aeng E. coli O157:H7 Sakai strains were constructed through a one step inactivation method (32, 98) and the LEE expression was analyzed by qRT-PCR. Wild type (WT) and AgadX strains were grown in previously described gadX-inducing conditions (39, 139) to analyze the expression of LEE genes. This included growth in morpholino propane sulfonic acid (MOPS) minimal medium at pH 7.4 and Dulbecco’s modified Eagle’s medium (DMEM) at pH 5.5 (0.2% glucose, 3.7% NaHCO3, pH adjusted with morpholino ethane sulfonic acid (MES)) upto an 00600 of 0.5. Expression analysis was not conducted at pH 2.0 as AgadX is not viable at this pH. Inactivation of gadX resulted in a minimal increase in expression of some of the LEE genes at neutral pH (Table 3.2). Interestingly, this effect was not observed at pH 5.5 (Table 3.2). This result is in contrast with that of EPEC, where inactivation of gadX had a significant effect on LEE expression at pH 5.5 but not at neutral pH. GadX negatively regulates LEE expression in EPEC by downregulating a plasmid encoded positive regulator of LEE, PerA (139). No PerA homolog has been identified in EHEC (61) and that could be the reason for the observed differences in GadX-mediated LEE regulation in E. coli O157:H7. Subsequently, the expression of LEE genes in Aeng was analyzed in previously described eng-inducing conditions (39). The cells were grown to exponential phase (00600 = 0.5) in minimal medium (E minimal medium with 82 TABLE 3.2 Effect of gadX inactivation on the expression of LEE genes Gene AgadX/WT AgadX/WT (pH 7.4) (pH 5.5) tir 1.1102 0.5101 espD 1.8101 0.8101 eae 1.5102 0810.2 sepZ 1.5103 0.8102 83 0.2% glucose and 12mM M9804) at pH 7.0 and pH 5.0. There was a marked increase in expression of LEE genes in the Aeng (Table 3.3) at neutral pH indicating that a negative interaction between Eng and LEE, similar to that in EPEC, exists in E. coli O157:H7 as well. Interestingly, this effect was not as pronounced at pH 5.0 as at neutral pH (Table 3.3). As in EPEC (99), this negative regulation of LEE by Eng could be mediated through YdeO and GadE regulators. The ability to orchestrate gene expression in response to environmental conditions is an important requirement for being a successful pathogen. Expression of virulence factors in inappropriate environments, particularly where they are not needed, can negatively affect energy conservation and markedly impair survival of the pathogen in the host. It has been shown in BordeteIIa bronchiseptica that ectopic expression of flagellar regulon interferes with the tracheal colonization by the pathogen (4). During passage through human stomach, the main objective of E. coli O157:H7 is its survival and therefore, it is expected that genes that provide protection at extreme acidity are positively selected for upregulation and at the same time, expression virulence factors that are not required in that environment is suppressed. The results from this study support this argument; upon exposure to gastric acidity E. coli O157:H7 upregulated acid resistance genes such as GAD genes and downregulated major virulence factors such as Stx, LEE and p0157 genes. This may be beneficial to the organism as decreased LEE expression prevents attaChement of the bacteria to the gastric mucosa and allows for faster transit through stomach. 84 TABLE 3.3 Effect of eng inactivation on the expression of LEE genes Aeng/WT Aeng/WT Gene (pH 7.0) (pH 5.0) tir 3.7105 1.4105 espD 4.4105 1.2103 eae 3.8106 1.0101 sepZ 5311.5 1.2104 85 Additionally, this pattern of gene expression may help in maximum energy conservation by the organism. We hypothesize that GAD system regulators such as GadE and Eng are involved in the decrease in LEE expression at extreme acidity. It is possible that the EngS two-component system senses the acidity in the environment and upregulates the expression of GadE, which in turn induces GAD expression and suppresses LEE expression. We do not know the factors that are involved in the repression of Stx and p0157 genes at extreme acidity at this time because none of the GAD system regulators investigated in this study have been shown to regulate their expression. Further work is needed in this aspect to identify the genes controlling this response and the mechanism of downregulation. This work was supported by Food Safety NRI #2005—35201-16362 from the United States Department of Agriculture, and in part by funds from the NIAID, NIH, DHHS, under the Food and Waterbome Diseases Integrated Research Network (NIH Research Contract N01-AI-30058). 86 CHAPTER 4 Differential Expression of Virulence and Stress Fitness Genes between Clinical and Bovine-biased Genotypes of Escherichia coli O157:H7 This chapter is submitted to Applied and Environmental Microbiology 87 SUMMARY Escherichia coli O157:H7 strains can be classified into different genotypes based on the presence of specific Shiga toxin-encoding bacteriophage insertion sites. Certain O157:H7 genotypes predominate among human clinical cases (clinical genotypes), while others are more frequently found in bovines (bovine- biased genotypes). To determine whether inherent differences in gene expression explains the variation in infectivity of these genotypes, we compared the expression patterns of clinical genotype 1 strains with those of bovine-biased genotype 5 strains using microarrays. Important O157:H7 virulence factors including locus of enterocyte effacement genes, the enterohemolysin, and several p0157 genes, showed increased expression in the clinical versus bovine-biased genotype. In contrast, genes essential for acid resistance (e.g., gadA, gadB, and gadC) and stress fitness were upregulated in bovine-biased genotype 5 strains. Increased expression of acid resistance genes was confirmed functionally using a model stomach assay, in which strains of bovine- biased genotype 5 had a 2-fold higher survival rate than strains of clinical genotype 1. Overall, these results suggest that the increased prevalence of O157:H7 illness caused by clinical genotype 1 strains is due in part to the overexpression of key virulence genes. The bovine-biased genotype 5 strains, however, are more resistant to adverse environmental conditions, a characteristic that likely facilitates asymptomatic O157:H7 colonization of bovines. 88 INTRODUCTION Escherichia coli O157:H7, a food-bome zoonotic pathogen that causes hemorrhagic colitis and hemolytic uremic syndrome (HUS) in humans, is the most prevalent type of enterohemorrhagic E. coli (EHEC) in the United States (67, 153). Cattle are the primary reservoir of E. coli O157:H7 and the fecal shedding rate on cattle farms can be up to 100% (45). While colonized cattle do not exhibit clinical disease (127), it has been reported that only 10-100 cells of E. coli O157:H7 are sufficient to induce overt disease in humans (163). Although it has been suggested that bovine-derived E. coli O157:H7 strains vary in their ability to cause human disease (14), the basis behind this variation is not known. E. coli O157:H7 possesses unique virulence properties that facilitate disease development including Shiga toxins (Stx), the locus of enterocyte effacement (LEE) pathogenicity island, and the p0157 vimlence plasmid (171). The LEE encodes a type 3 secretion system (T388) that mediates the formation of attaching and effacing lesions (94), while the p0157 plasmid encodes several putative virulence factors such as an enterohemolysin (Eth or EHEC-HlyA) (78) and a type 2 secretion system (T288) (133). Both the LEE and p0157 have shown to be critical for disease pathogenesis (66, 81). Shiga toxins, which are the cytotoxins responsible for renal damage in HUS (66), are encoded by genes located on lysogenic Iambdoid phages that are inserted into the O157:H7 chromosome at specific locations (136). A prior study of 80 bovine isolates and 282 clinical isolates from humans with 0157-associated disease demonstrated that the distribution of Stx insertion sites varies between isolate types (14). 89 Furthermore, these isolates were classified into different genotypes based on the insertion sites of Stx-encoding bacteriophages and genotypes 1 - 3, although isolated from cattle also, were predominant in clinical isolates and were considered as clinical genotypes. 0n the other hand, genotypes such as 5 and 7 were overrepresented among the cattle isolates and therefore, considered as bovine-biased genotypes (14). Similarly, octamer-based genome scanning of bovine and clinical isolates of E. coli O157:H7 identified two genetically distinct lineages, of which lineage I was isolated mostly from humans and lineage ll, mostly from bovines (72). Comparing the presence of virulence genes between E. coli O157:H7 isolates from various sources using DNA microarrays also has revealed that 0157 isolates from beef cattle and humans are genetically distinct (79). In addition to intrinsic differences, it is possible that there are differences in the expression of important virulence genes as well as variation in the degree of resistance to adverse environmental conditions between clinical and bovine- biased genotypes. To investigate this hypothesis, the exponential phase transcriptomes of four clinical genotype 1 strains were compared to the transcriptomes of four bovine-biased genotype 5 strains using microarrays. All the strains used in this study were from bovine sources and they belonged to either genotype 1 or 5. Therefore the Genotype 1 strains used in this study are in fact bovine-derived clinical genotype strains. The goal of this study was to identify specific genes that are differentially expressed between the two genotypes to better understand why genotype 1 strains cause more clinical 90 disease than genotype 5 strains. The identification of genes that are upregulated in clinical versus bovine-biased genotypes is important for detecting and controlling those strains that are more likely to cause E. coli O157:H7 infections in humans. 91 MATERIALS AND METHODS Bacterial strains. The eight bacterial strains used in this study were selected based on the Stx-encoding bacteriophage insertion site genotypes determined in a prior study (14). Strains representing genotypes 1 (clinical genotype) and 5 (bovine-biased genotype) were selected among 80 bovine strains originally isolated between 1991 and 2004 as described (14). Although the clinical genotype 1 strains in this study were bovine-derived, their genotype was identical to those genotype 1 strains isolated in a prior study from humans with 0157 infections (14). Four strains of each genotype were included in the microarray analyses. A previously described (88) stx2/stx2c RFLP demonstrated that all the genotype 1 strains used in this study harbored stx2 alone, whereas the genotype 5 strains contained only stx20. Growth conditions. Each strain was stored at -70°C in LB broth containing 10% glycerol, was inoculated into 10 ml LB broth, and grown to an ODeoo of ~01 to recover cells. Cells were grown twice to stationary phase in MOPS-buffered minimal medium (pH 7.4) before transferring at a 1:30 dilution into 100 ml Dulbecco’s Modified Eagle Medium (DMEM) (0.45% glucose) for both RNA isolation and the model stomach assay. To minimize the confounding effect of acidic pH that develops in stationary phase of growth in un-buffered medium, the DMEM was buffered with MOPS to pH 7.4. Microarray design. To compare global gene expression profiles between genotypes 1 and 5, microarrays were hybridized in a double loop design, thereby allowing strains from one genotype to be directly compared to strains from the 92 other genotype (Fig. 4.1). The four strains from each genotype represent biological replicates and therefore, significant differences in gene expression are representative of the two genotypes. RNA isolation and cDNA labeling. For RNA isolation, the strains were grown to exponential phase (~2.25 h, 00600 ~05) in DMEM and RNA extractions were performed using a modified version of the previously described hot-phenol method (15). Briefly, 5 ml of the culture was mixed with 1/10V of 10% phenolzethanol buffer to stabilize the RNA, and centrifuged at 4° C (4300 x g) for 30 min to pellet cells. The supernatant was decanted and cell pellets were suspended in 5 ml of buffer (2 mM EDTA, 20 mM NaOAc, pH 5.2) before RNA extraction with hot-phenol. Reverse transcription reactions and the coupling of cDNA with Cy3 or Cy5 dyes were conducted as described elsewhere (12). cDNA hybridizations. Hybridizations were performed according to the double loop microarray design (Fig. 1). As described (12), the cDNAs were hybridized onto microarray slides printed with 6,088 ORFs representing E. coli genome strains K12 (17), EDL933 (110) and Sakai (46);110 ORFs from the p0157 plasmid were included. Arrays were scanned with an Axon 4000b scanner (Molecular Devices, Sunnyvale, CA) followed by image analysis using GenePix 6.0 (Molecular Devices). Data analysis. Microarray data were processed as previously described (65) and fitted to a mixed ANOVA model (30). The linear model tested was Y (intensity) = array + dye + strain (clinical or bovine-biased) + sample (biological replicate) + error. Significant differences in expression were determined using the 93 / \\.. \ /< BZA C4¢ Figure 4.1. Double loop design for the microarray experiment. B1-B4 represent four different strains of the bovine-biased genotype 5, whereas C1-C4 represent four different strains of the clinical genotype 1. Each arrow indicates a hybridization with the arrow head representing Cy3 and tail Cy5 dyes. 94 Fs test in MAANOVA with 500 random permutations to estimate the p-values. This test uses a shrinkage estimator for gene-specific variance components that makes no assumption about the variance across genes (31 ). In addition, the q- value package in R was used to determine the false discovery rate (FDR) (146). Additionally, significance analysis of microarrays (SAM) was used to analyze data with a FDR of 0.05. Overrepresentation of gene sets with a common biological function in the two genotypes was determined using the Gene Set Enrichment Analysis (GSEA) Preranked analysis program (GSEA v2.0; Broad Institute, Massachusetts Institute of Technology) (148). The gene sets were designated based on the annotation for the Sakai genome (46) available through the J. Craig Venter Institute (http:l/cmr.jcvi.org/tigr-scripts/CMR/GenomePage.cgi?org=ntecO3). Additionally, genes for the LEE and the AFI-GAD, which represents the glutamate decarboxylase (GAD) system and acid fitness island (AF I), also were included in the analysis. Quantitative real-time PCR (qRT-PCR). Select genes that had significantly different levels of expression between genotypes in the microarray analysis were confirmed by qRT-PCR. Taqman assays (11) were used to quantify the expression of gadA, gadB, and Ier, with mdh as a reference for normalization. For all other genes, SYBR green was used as described elsewhere (65); methods for cDNA synthesis and qRT-PCR also were described previously (65). The expression level of the 16S rRNA gene was used for normalization of data and the relative expression levels were quantified using 95 modified Livak method (134). The results presented are averages from four biological replicates 1 standard error of mean (SEM). Single Nucleotide Polymorphism (SNP) Genotyping. Genomic DNA was extracted with the Purgene DNA extraction kit (Gentra systems, Minneapolis, MN) for use with the GenomeLab SNPstreamT" system (Beckman Coulter, Fullerton, CA). SNP genotyping via the SNPstreamTM was performed using a modified version of a previously described protocol (88) according to the manufacturer’s instructions. Briefly, PCR was conducted using four panels of 48- plex primers targeting 192 distinct SNP loci identified previously via the comparison of three O157:H7 genomes (88). The primers, which differed from the original protocol, were designed using the Autoprimer program (175). After cleaning, the PCR products were subjected to single base primer extension reactions that add a labeled nucleotide to the SNP site followed by hybridization onto a 384-well SNP microarray plate. Detection and processing were performed via the SNPstreamT" lmager (version 2.3; GenomeLab). A total of 52 of the 192 SNPs were found to be informative; these SNPs were concatenated in MEGA4 (151) to construct a neighbor-joining tree (130) for examining the phylogenetic relationships between the eight strains. SNP data from reference strains representing each of the nine O157:H7 clades (88) were included in the analysis. Model stomach assay. The model stomach system (M88) (64) was prepared as described previously (11). Gerber Turkey Rice Dinner© baby food (30 g) was mixed with 120 ml of synthetic gastric fluid (pH 1.70), yielding a final pH of 2.5. Strains, grown to an 00300 ~2.5 in DMEM, were inoculated into the 96 MSS at the rate of 10° cells/ml. Contents of the M88 were stomached for 30 sec, sampled, diluted, and plated onto LB agar plates every 30 min for 1.5 h to enumerate viable cells. CFU/ml from duplicate plates were averaged and converted to Iog10 CFU/ml. Survival rates were calculated as the log decrease in viable cell count per 30 min, and the average from two experimental replicates were reported. Microarray data accession. Microarray data are available at NCBI GEO (http://www.ncbi.nlm.nih.gov/geo), accession number GSE15783. 97 RESULTS Differentially expressed genes between clinical and bovine-biased genotypes of E. coli O157:H7. The F3 test identified significant differential expression of 191 genes between the two genotypes, of which 71 were upregulated in the clinical genotype and 120 were upregulated in the bovine- biased genotype (FDR<0.1, fold change 2 1.5) (Table 4.81 in appendix). Additionally, SAM identified more genes to be significantly differentially expressed between the two genotypes; 154 were upregulated in the clinical and 238 were upregulated in the bovine-biased genotype (FDR<005, fold change >1 .5) (Table 4.82 in appendix). One hundred and sixteen genes were found. to be differentially expressed by both MAANOVA and SAM. Differentially expressed genes included those involved in virulence, response to stress, acid resistance and metabolism. Overall, important O157:H7 virulence factor genes were upregulated in the clinical genotype (Table 4.1), whereas genes related to acid resistance and stress fitness were upregulated in the bovine-biased genotype (Table 4.2). GSEA also identified enrichment of 8 gene sets in the clinical genotype and 6 gene sets in the bovine-biased genotype (Table 4.3), thereby providing additional support for the MAANOVA and SAM results. The LEE genes. There was an overall increase in expression of the LEE genes in clinical genotype 1 relative to bovine-biased genotype 5 strains; expression was significantly different in12 genes (Table 4.1). For example, secreted proteins encoded by espF and espG, T388 proteins encoded by escF, sepQ, escT, and escR, and the cesD chaperone were upregulated in genotype 1 98 TABLE 4.1 Virulence-associated genes upregulated in clinical genotype 1 relative to bovine- biased genotype 5 Expression ECs no.‘I Gene Function ratio (CIB) ° Test ° LEE genes ECs4550 espF Effector protein 2.0 S ECs4552 escF T388 EscF protein 2.1 S ECs4565 sepQ T388 structural protein 1.6 M ECs4566 orf16 Secretion of translocators 2.2 S ECs4570 orf12 T388 2.1 S ECs4574 sepD T3SS SepD protein 2.0 S ECs4576 cesD T388 chaperone 1.8 M ECs4579 rorf3 T388 1 .9 S ECs4581 escT T388 structural protein 2.0 S ECs4583 escR T388 structural protein 1.9 S ECs4590 espG Effector protein 1.9 S ECs4591 rorf1 Unknown 1 .6 M Genes encoded by p0157 plasmid p01 57p02 etpC T288 2.8 M p01 57p05 eth T288 1 .9 M+S p01 57p07 etpH T288 1 .9 M p01 57p09 eth T288 2.0 M+S p01 57p10 etpK T288 2.2 M+S p0157p1 1 etpL T288 1 .8 M p01 57p12 etpM T288 1 .9 M+S p0157p14 etpO T288 1.7 M p01 57p18 eth enterohemolysin 1 .7 M p0157p24 repFIB replication protein 1.7 M p0157p58 toxB toxin B 1.6 M p0157p79 unknown 2.4 M+S p0157p80 unknown 2.1 S ' p0157p81 unknown 2.1 M+S a Locus ID for E. coli O157:H7 Sakai strain (Genbank # BA000007) ° Expression ratio between two genotypes; C=clinical genotype, B=bovine-biased genotype 99 °Test that identified a gene as statistically significant; M=MAANOVA, S=SAM 100 TABLE 4.2 Genes upregulated in bovine-biased genotype 5 ECs no. a Gene Function (BIC) b Test ° GAD and AFI genes ECs2097 gadC GABA-glutamate antiporter 2.5 M+S ECs2098 gadB Glutamate decarboxylase 4.4 M+S isozyme ECs4377 slp Outer membrane protein 2.5 M+S ECs4389 hdeB Periplasmic chaperone 3.4 M+S ECs4390 hdeA Protection from organic acid 5.6 M+S metabolites ECs4391 hdeD Acid resistance at high cell 2.7 M+S density ECs4392 gadE Central activator of the GAD 2.6 8 system ECs4394 yhiV lVIulti drug efflux pump protein 2.2 M+S ECs4395 gadW ARAC-type GAD system 1.8 M+S regulator ECs4397 gadA Glutamate decarboxylase 1.9 M+S isozyme Stress fitness-associated genes ECsO890 dps DNA protection during 3.6 M+S starvation ECsOQ66 cspD Cold shock protein 2.5 M+S ECs1722 chaB Cation transport regulator 2.4 M+S ECs4871 katG Catalase-peroxidase 1 .7 M+S ECs2086 osmC Osmotically inducible protein 2.1 M+S EC35334 osmY Osmotically inducible protein 3.0 M+S ECs4367 uspA Universal stress protein 2.6 S EC30968 prA Degradation of abnormal 2.0 8 proteins ECs1723 chaC Cation transport regulator 2.2 S RpoN regulated (involved in nitrogen metabolism) ECs0169 glnD uridylyltransferase acts on 1.8 S regulator of gInA ECsO504 gInK nitrogen regulatory protein P-ll 11.6 S 2 ECsO505 amtB probable ammonium transporter 14.9 S 101 Table 4.2 continued... ECsO692 gItK glutamate/aspartate transport 1.9 8 system permease E050693 gltJ glutamate/aspartate transport 2.2 8 system permease ECs0889 glnH permease of periplasmic 2.1 S glutamine-binding protein ECs2784 nac nitrogen assimilation control 12.7 8 protein ECs3194 argT periplasmic arginine binding 1.5 M protein ECs4091 gltB glutamate synthase, large 1.9 S subunfi ECs4790 gInG nitrogen regulator l 2.6 8 Mac regulated ECs1743 oppA oligopeptide transport 2.1 M+S ECs1744 oppB oligopeptide transport 1 .6 M+S ECs1746 oppD oligopeptide transport 1 .8 M+S ECs1747 oppF oligopeptide transport 1 .7 M+S E053522 gabD succinate-semialdehyde 3.7 S dehydrogenase E053523 gabT 4-aminobutyrate 3.2 S aminotransferase activity ECs4424 dppA dipeptide transport protein 1.9 S RpoN-regulated (not involved in nitrogen metabolism) EC53582 hypA pleiotrophic effects on 3.0 S hydrogenase isozymes ECs3583 hypB hydrogenase isoenzyme HypB 2.6 S ECs5061 fth subunit of formate 2.1 S dehydrogenase H a Locus ID for E. coli O157:H7 Sakai strain (Genbank # BA000007) ° Expression ratio between two genotypes; B=bovine-biased genotype, C=clinical genotype cTest that identified a gene as statistically significant; M=MAANOVA, S=SAM 102 TABLE 4.3 Enrichment of gene sets in clinical and bovine-biased genotypes FDR-q Gene set NES' value” Enriched in clinical genotype 1 LEE 2.58 0.000 Folic acid biosynthesis 1.88 0.008 tRNA rRNA base modification 1.87 0.006 Pyrimidine ribonucleotide biosynthesis 1.82 0.013 RNA processing 1.74 0.026 Nucleotide and nucleoside interconversions 1 .69 0.037 Toxin production and resistance 1.68 0.036 Ribosomal protein synthesis and modification 1 .65 0.040 Enriched in bovine-biased genotype 5 AFI-GAD -2.09 0.000 Adaptations to atypical conditions -1.94 0.004 Glutamate family -1.90 0.004 Pyruvate family -1.89 0.004 Gcholysis/gluconeogenesis -1 .89 0.003 Fermentation -1.78 0.010 a NES= normalized enrichment score indicating the degree to which a gene set is overrepresented in the top or bottom of the ranked list of genes. ° F DR-q value 5 0.05, indicates false discovery rate. 103 strains (Table 4.1). Although the remaining 29 LEE genes were not significantly different between genotypes, 27 were upregulated in genotype 1 relative to genotype 5. Members of all the five LEE operons showed change in expression in the same direction. The insignificant result is possibly due to inter-strain variation within the genotypes. To confirm expression differences, qRT-PCR was used to examine the expression of four important LEE genes including Ier, espB, espD and tir that were not significantly different by microarrays (Fig. 4.2). More than a two-fold increase in expression was observed for espB, espD and tir by qRT-PCR in the clinical genotype (Fig. 4.2), a level that was similar to the microarray data for the 12 significant genes. Expression of Ier was slightly lower than the other genes, though it still exhibited a 1.4—fold increase in clinical strains (Fig. 4.2). Additionally, GSEA confirmed the enrichment of the entire set of 41 LEE genes in the clinical genotype (Table 4,3). p0157 plasmid encoded genes. The p0157 plasmid encodes a number of virulence associated genes in E. coli O157:H7 strains. Fourteen of these genes, which includes eth (EHEC-hlyA; enterohemolysin), toxB (toxin B), and eight of the thirteen genes that encode the T288 were significantly upregulated in the clinical genotype 1 (Table 4.1). The T288 etp cluster (133) showed a 1.8- to 2.8-fold increase in expression in the clinical genotype, which was confirmed by qRT-PCR (Fig. 4.2). The eth and toxB also were confirmed to have a 2.2- and 2.1-fold increase in expression by qRT-PCR (Fig. 4.2). Expression of some virulence genes encoded by p0157 such as ethBD and stcE, were not significantly different between the two genotypes. 104 Figure 4.2. qRT-PCR validation of microarray data. The expression ratio between clinical and bovine-biased genotypes as calculated by microarrays and qRT-PCR are given. Results shown are average fold change in expression with standard error of mean (SEM) from four biological replicates (strains) per genotype. 105 mace m8: <8: 88 meme 328 x98 mums one 9.9 SE s 38 meme 5 L p F p p F mNI «61ch _H_ . om- 28890:). I H . m..- I OFI . .II—i 106 (euera/ieoiuuo) efiueuo Pl0:I Acid resistance and stress fitness-associated genes. Numerous genes that are essential for acid resistance in E. coli were significantly upregulated in the bovine-biased genotype 5 strains. This included genes that encode all three components of the GAD system, gadA, gadB, and gadC (39). In addition, the twelve AF I genes (91) had increased expression in the bovine- biased genotype, with eight of the 12 having significantly different levels (Table 4.2). The increased expression of GAD system genes was confirmed by qRT- PCR, which showed a more than 10-fold increase relative to clinical genotype 1 (Fig. 4.2). Similarly, there was a 3.6-, 5.6-, 6.6- and 7.5-fold increase in expression of gadX, gadE, hdeA and hdeB, respectively, by qRT-PCR (Fig. 4.2). The upregulation of gadX, however, was not statistically significant in the microarray analysis, though the direction was the same. This discrepancy is possibly due to high inter-strain variation in expression within the genotype. Expression of dps, which is involved in protecting DNA during starvation and acid stress (28), was upregulated by 3.6-fold in the bovine-biased genotype strains. Similarly, prA, a chaperone necessary for protein degradation by the ClpAP protease (73), showed a 2-fold increase in expression. Other stress fitness-associated genes with increased expression in the bovine-biased genotype included the cold shock protein, cspD (173), cation transport regulators, chaBC (107) and the universal stress protein, uspA (26, 111) (Table 4.2). Moreover, expression of katG (108), osmC and osmY (168), the genes involved in resistance to peroxide and osmotic stress, also were upregulated 107 (Table 4.2). Interestingly, the general stress sigma factor, rpoS (169), was not differentially expressed between the two genotypes. The RpoN regulon. Several metabolic genes, including nine genes involved in the nitrogen regulatory response that are regulated by the sigma factor RpoN (122), were upregulated in the bovine-biased genotype. The nitrogen regulatory protein, gan, and the ammonium transporter, amtB, were both upregulated by 11.6- and 14.9-fold respectively. glnD, which is involved in the post transcriptional modification of gan, also was upregulated in bovine- biased strains as were the nitrogen regulator l (gInG), a permease of the periplasmic glutamine binding protein (glnH), and genes associated with glutamate biosynthesis (gltB, gItK and gltJ) (Table 4.2). Furthermore, nac, which encodes the nitrogen assimilation control protein, was upregulated by 12.7-fold in the bovine-biased genotype. Consequently, a number of Nae-regulated genes including oppA, oppB, oppD, oppF, gabD, gabT, and dppA, had higher expression levels in bovine-biased strains. Other RpoN- regulated genes such as hypA, hypB and fth, also had increased expression in the bovine-biased genotype (Table 4.2). SNP genotyping and re-analysis of microarray data. Because the eight strains in this study were only characterized by the distribution of Stx insertion sites and represented the same multilocus sequence type (14), a more sensitive SNP genotyping method (88) was used to better understand the phylogenetic relationships of strains within and between the two genotypes. Among the four strains representing clinical genotype 1, three grouped together with a clade 8 108 control strain and one with a clade 1 control strain (Fig. 4.3). By contrast, all four strains representing the bovine-biased genotype 5 belonged to clade 7 (Fig. 4.3). Since one clinical genotype 1 strain was part of a phylogenetically distinct lineage (clade 1) relative to the other three clinical genotype 1 strains (clade 8), the microarray data was re-analyzed after excluding the data generated from the clade 1 strain. The PS test identified significant upregulation of 400 genes in the clinical genotype and 349 genes in the bovine-biased genotype (FDR <01, fold change >1 .5) in this re-analysis. All but three genes (bioB, dps and ycaL) identified to be differentially expressed in the first analysis were also differentially expressed in the second analysis. Further, 561 additional genes were identified in the second analysis, as elimination of the clade 1 strain likely reduced the within-genotype variation. Twenty eight LEE genes including the genes encoding intimin (eae), the translocated intimin receptor (fir), and a positive regulator of LEE (gr/A), were significantly upregulated in the three clinical genotype strains. Similarly, p0157-encoded genes such as eth, toxB, and 11 genes of the etp polycistron that encodes a T288, also were upregulated in the clinical genotype. As expected, the bovine-biased genotype strains showed increased expression of GAD and AF I genes relative to the three clinical genotype strains. One additional gene was gadX, which was significantly upregulated in the second analysis, but not the first analysis. The stationary phase sigma factor, rpr, and adiY, which encodes an ARAC-like regulator of the arginine decarboxylase acid resistance system, also were upregulated in the bovine-biased genotype as were a number of stress fitness-associated and RpoN-regulated genes. 109 I Clinical genotype1 Sakai control Clade 1 A Bovine biased genotype 5 95 .10916 84 99 10045 control Clade 2 97 - 10008 control Clade 3 ~ 14618 control Clade 5 14283 control Clade 6 01663 control A 10917 94 A 10938 Clade 7 94 A 10957 A 10948 90 08635 control 99 II13:19:7 Ciade 8 I-—-I 97 0.02 TI—I10950 Figure 4.3. Neighbor joining phylogeny of single nucleotide polymorphism (SNP) genotypes representing the eight strains examined in the study. Three of the four clinical genotype 1 strains (black squares) belong to clade 8, whereas one clinical genotype 1 strain is part of clade 1. All four bovine-biased genotype 5 strains (black triangles) are members of clade 7. 110 Model stomach assay. To determine whether the increased expression of acid resistance genes in the bovine-biased genotype translates to a phenotypic difference, model stomach assays were conducted. These assays were used to directly compare the survival of both genotypes in a complex acidic environment that simulates the human stomach. Consistent with the microarray expression data, there was a significant difference (P = 0.003) between the survival rates of clinical and bovine-biased genotypes, as bovine-biased strains had a 2-fold increase in survival in the M88 (Fig. 4.4). The average survival rate per 30 min for the clinical genotype was 05510.04, whereas the bovine-biased genotype was 0271004. 111 -0.2 ‘ -O.3 ~ -0.4 - -0.5 - ~0.6 - -0.7 - Survival rate (log decrease per 30 min) -08 , j Clinical Bovine-biased Figure 4.4. Survival of clinical and bovine-biased genotypes in the model stomach system. The average survival rate (log decrease in CFU/ml per 30 min) with the standard error of mean (SEM) from two independent experiments is plotted for each genotype. 112 DISCUSSION Genotyping based on Stx-encoding bacteriophage insertion sites has demonstrated that the E. coli O157:H7 strains present in the bovine reservoir are considerably more diverse when compared to strains that cause human infections (14). Furlhennore, it was suggested that some bovine-biased genotypes have reduced virulence and hence, cause disease less frequently relative to those bovine-derived clinical genotypes that are commonly isolated from patients (14). This variation could be due to gene content differences, including allelic variation in key genes among genotypes or due to expression differences in critical genes. Here, we describe differences in the expression of important genes that provide an explanation for the variation in infectivity between bovine-biased and clinical genotypes. Specifically, genome-wide expression profiling revealed differential expression of key virulence and stress fitness genes between the two genotypes, which was confirmed by qRT-PCR and a phenotypic assay. Because microarrays and qRT-PCR target different regions per gene, we suspect that the differential expression of genes identified in this study is not due to differences in gene content or allelic variation between genotypes, as both microarrays and qRT-PCR detected similar levels of expression. One of the most important differences identified was the upregulation of the LEE in clinical versus bovine-biased genotypes. The LEE is considered a critical factor in E. coli O157:H7 disease pathogenesis, as it encodes a T388 that mediates adherence to the intestinal mucosa (94). Because strains of the clinical 113 genotype expressed key LEE genes at a higher level, it is likely that these strains have an enhanced ability to adhere to the intestinal epithelium and cause the attaching and effacing lesions that initiate the disease process. By contrast, it is possible that increased expression of negative LEE regulators suppresses the expression of important LEE genes in the bovine-biased genotype, thereby reducing adherence and subsequent disease. The increase in gadE expression in the bovine-biased genotype supports this hypothesis, as our prior study determined that GadE, the central activator of the GAD system, negatively regulates LEE in O157:H7 strains (65). Similariy, GadX, a negative regulator of LEE in enteropathogenic E. coli (EPEC) (138), also was upregulated in the bovine-biased genotype as was yhiF, an AFI-encoded regulator that suppresses LEE expression in EHEC (156). Similar to the LEE, another important factor in E. coli O157:H7 disease pathogenesis is the possession of the p0157 plasmid. The 92 kb plasmid carries genes for many putative virulence factors including eth (enterohemolysin) (129), toxB (toxin B) (19), and a number of etp genes necessary for a T288 that secretes factors such as the StcE protease (44); all were upregulated in the clinical genotype. eth is a cell-associated, pore forming toxin from the repeats- in-toxin (RTX) family (129). Although the exact role of the enterohemolysin in O157:H7 pathogenesis is not clear, it has been shown to induce production of the interleukin-1 beta (IL-1B) proinflammatory cytokine, which is a serum marker of HUS (152), and to cause injury to microvascular endothelium (5). The increased expression of ngA, a LEE-encoded positive regulator of eth (129), 114 may partly explain why eth was upregulated in the clinical genotype. StcE, a protease that is involved in the intimate adherence of EHEC to host cells, is secreted by the T288 encoded by the etp gene cluster (44); eight of the etp genes were significantly upregulated in the clinical genotype. Similarly, toxB, which was upregulated in clinical genotype strains, also has been shown to be important for complete adherence to human epithelial cells (154). Together, these data demonstrate that the clinical genotype 1 strains are expressing factors important for adherence, suggesting that genotype 1strains may have an enhanced ability to adhere to human host cells and thus, are inherently more vinilent than bovine-biased genotype 5 strains. A comparison of Stx expression was not possible as the genotypes used in this study do not have the same stx profiles. During passage through the bovine gastrointestinal tract, E. coli O157:H7 has to survive a number of adverse environmental conditions, including extreme acidity in the abomasum (165), organic acid stress from volatile fatty acids in the nimen and colon, and occasional hyperosmolarity. Therefore, the capacity to withstand acidity and other environmental stresses is critical for O157:H7 strains to successfully persist in cattle. Consistent with this, we observed upregulation of acid resistance and stress fitness-associated genes in the bovine-biased genotype relative to the clinical genotype. The GAD system is the most efficient acid resistance system in E. coli (23, 75) and is essential for the survival of E. coli O157:H7 in bovines (119). All three GAD system genes (gadA, gadB and gadC (39)) had increased expression in the bovine-biased genotype. Moreover, AFI 115 genes, which are also involved in acid resistance (91), showed increased expression in the bovine-biased genotype. This increased expression of GAD and AFI genes is possibly induced by the upregulation of GadE, an essential activator of GAD and many AFI genes in E. coli O157:H7, as shown recently (65). DMEM buffered with MOPS maintains a neutral pH at exponential phase and therefore, it appears that under non-inducing conditions, the expression level of gadE and the GAD system genes is markedly higher in the bovine-biased genotype, which may enhance survivability in the bovine gastric environment (pH 2.1) (165). This inference was confirmed by the model stomach assays, where bovine-biased strains survived better than clinical genotype strains. The model stomach represents a complex acidic environment and hence, the increased survivability of bovine-biased strains may also be attributable to the increased expression of stress fitness genes such as dps, prA and uspA, in addition to the acid resistance genes. Together, these findings indicate that strains of the bovine-biased genotype 5 are more resistant to adverse environmental conditions, which likely facilitates survival and the subsequent colonization of bovines. A prior study observed similar results by comparing resistance to acetic acid (pH 3.3) among E. coli O157:H7 strains isolated from environmental sources and humans (103). Specifically, those isolates from humans were found to be less resistant to acetic acid than isolates obtained from bovine feces (103). Our study suggests that this difference in acetic acid resistance could be due to differences in expression of important acid and stress resistance genes. The variation in resistance to acidity among different genotypes within the bovine 116 population is interesting, and we demonstrate that those bovine-derived strains that can infect humans are actually less resistant to acid than those strains that colonize the bovine. This characteristic enables less resistant strains, or genotypes, to express alternative factors, particularly those involved in adherence, when in contact with human cells mainly due to the negative interaction between acid resistance regulators and adherence genes. Although bovine-biased genotypes that are highly resistant to acid can get transmitted to humans, these strains cause disease much less frequently, which could possibly be due to the decreased expression of important virulence factors. As the nitrogen source in DMEM is glutamine, strains face an ammonia limiting condition during growth in DMEM. This condition leads to the induction of the nitrogen regulatory response (122), and expression data from this study indicates that bovine-biased strains are more efficient at mounting this response at the transcriptional level. Ten RpoN-regulated genes involved in the response were upregulated in bovine-biased strains, and important regulators and assimilation proteins such as gan, amtB and nac, had more than a 10-fold increase in expression. These observations indicate that strains of the bovine- biased genotype are equipped with a more efficient nitrogen regulatory response system, which can enhance survival in ammonia limiting conditions. Such conditions are likely encountered in the bovine gastrointestinal tract or external environment. A recent study from our laboratory classified clinical O157:H7 strains into nine clades based on a highly sensitive SNP genotyping method (88). The 117 severity of disease was shown to vary between the clades, with clade 8 being associated with HUS. Interestingly, three of the four clinical genotype 1 strains used in this study belonged to clade 8, thereby demonstrating the presence of this clade in the bovine reservoir. After excluding the phylogenetically distinct clinical genotype 1 strain, the three clade 8 strains still had increased expression of several LEE and p0157 genes. In contrast, all of the bovine-biased strains used in this study belonged to clade 7, a lineage that was associated with less severe clinical disease (88). Decreasing the prevalence of 0157 colonization in cattle has become increasingly significant in the current strategies to control 0157 infections, as it is less feasible to completely prevent fecal contamination of food vehicles such as vegetables, fruit juices and beef (34). In this context, identifying the genes that are critical for 0157 persistence in cattle is important for developing novel prevention strategies against this pathogen. This study, along with previous studies (119), indicates that acid and stress resistance genes encompass an important set of genes that are crucial for survival of E. coli O157:H7 in cattle and hence, can be used as targets for prevention measures. In addition, there is considerable variation in the expression of different genes between genotypes of E. coli O157:H7 isolated from cattle, with strains of clinical genotype 1 having increased expression of important virulence factors. The upregulation of key virulence components provides a possible explanation for the predominance of this genotype in clinical cases. 118 ACKNOWLEDGEMENTS This project has been supported by Federal funds from the National Institute of Allergy and Infectious Diseases, National Institutes of Health, Department of Health and Human Services, under Contract No. N01-AI-30055. The authors wish to thank Lindsey M. Ouellette and Hans Steinsland for technical assistance, Teresa M. Bergholz for helpful suggestions for the microarray design, and Linda 8. Mansfield, Galeb S. Abu-Ali and James T. Riordan for critical review of the manuscript. 119 CHAPTER 5 Summary and future directions 120 SUMMARY Acid resistance (AR) is a critical factor in the pathogenicity of E. coli O157:H7 as it confers the ability to be transmitted through a variety of food vehicles and to breach one of the first host defense barriers—the gastric acidity—while passing through the human stomach. E. coli O157:H7 has been shown to be more efficient in resisting complex acidic conditions when compared to other strains of E. coli (10). This superior AR is considered as the main contributing factor for the low infective dose of E. coli O157:H7. Four distinct mechanisms are known to contribute to the AR of E. coli including the OXI system, GAD system, ARG system, and LYS system, of which GAD system is most efficient (39, 75). The GAD system is comprised by two glutamate decarboxylase isozymes, GadA and GadB, and a GABA-glutamate antiporter, GadC. The GAD system is one of the most complexly regulated systems in E. coli, which indicates the importance of this system in the maintenance of normal physiology of the organism (39). Furthermore, animal experiments have shown that a functional GAD system is necessary for the survival of E. coli O157:H7 in cattle (1 19). Maximum GAD expression is induced at stationary phase of growth or when the cells are exposed to acidic pH (12, 39). At least 14 regulators are involved in the regulation of GAD system, among which GadE functions as the central activator of the system. A functional GadE is essential for the expression of GAD components at all growth phases in rich or minimal media (54, 84). GadE is a DNA binding LuxR family regulator and it induces GAD expression by directly 121 binding to the 20 bp GAD box region upstream of gadA and gadBC genes. Even though the GAD system and its regulation are well characterized in benign laboratory strains of E. coli, these aspects remain largely unknown in pathogenic strains of E. coli. Most importantly, the function of GadE at a genome wide scale is not known for any of the pathogenic strains of E. coli. E. coli O157:H7 genome is markedly different from the laboratory strains of E. coli as it has acquired many foreign DNA elements through horizontal gene transfer during evolution. It is possible that during the process of integrating laterally acquired DNA elements into the chromosomal regulatory network, a chromosomal regulator such as GadE has evolved into a global regulator with multiple functions. Therefore, I hypothesized that GadE has additional functions besides the positive regulation of GAD system and the purpose of the first part of this project was to identify the GadE regulon and to characterize the AR phenotype of a gadE mutant E. coli 01 57: H7. The E. coli O157:H7 Sakai strain isolated from the radish sprout outbreak in Japan was used in this part of study (46, 96). Here, the experimental approach for identifying the GadE regulon was to compare the transcription profiles of wild type (WT) and AgadE strains at different growth phases to detect the genes whose expression is affected by the inactivation of gadE. An E. coli O157:H7 AgadE strain was constructed by a one step inactivation method (32, 98). Inactivation of gadE did not affect the growth patterns of E. coli O157:H7. WT and mutant strains were grown in MOPS minimal medium up to exponential and stationary phase and the transcriptomes were compared using whole genome 122 microarrays. Inactivation of gadE affected the expression of 60 genes independent of growth phase and 122 genes in a growth phase-dependent manner. Expression of gadE has been shown to be growth phase-dependent (12) and hence, this growth phase effect on genes was expected. Additionally, putative GadE binding sequences or GAD boxes were observed upstream of eight genes that had significant differential expression due to gadE inactivation indicating a direct regulation of these genes by GadE. Genes involved in AR such as the GAD and AFI genes comprised an important part of the GadE regulon. Expression of all the three components of the GAD system, gadA, gadB and gadC, was markedly decreased by the inactivation of gadE. Similarly, hdeBAD, the AF I genes involved in resistance to self metabolites and AR at high cell density showed decrease in expression in AgadE. These findings were similar to that in the laboratory strain, E. coli K-12 in which inactivation of gadE completely abolished the expression of GAD and markedly reduced the expression of AFI genes (84, 91). Here, we also found that the effect of gadE inactivation on GAD and AFI genes was significantly higher at stationary phase than at exponential phase. The regulation of GAD and AFI genes by GadE appears to be direct because GAD boxes with 100% similarity were detected upstream of gadA and gadBC and two putative Gad boxes were identified upstream of hdeD. The acid resistance phenotype of AgadE was characterized using functional assays such as AR mechanism assays and a model stomach assay. AR mechanism assays for GAD, ARG and OXI systems assess the effect of 123 gadE inactivation on the functioning of the three systems. As expected, there was complete abrogation of GAD system functioning in the AgadE. Interestingly, inactivation of gadE also affected the ability of ARG and OXI systems to protect the strains from acidic challenge. Complementation with wild type gadE restored the phenotype for GAD and ARG systems, but not for the OXI system. We believe that flooding the cells with wild type gadE through a multi copy plasmid might have negatively affected the OXI system. In the model stomach system, inactivation of gadE abolished the ability of E. coli O157:H7 to survive in this complex acidic environment. The AgadE did not survive for even 30 min in the model stomach whereas wild type and complement survived up to 90 min with minimal decrease in cell counts. This demonstrates that gadE is essential for the survival of E. coli O157:H7 in a simulated gastric environment. A previous study in E. coli O157:H7 Sakai strain has shown an increased adherence phenotype for AgadE. Expression analysis of WT and AgadE from the same study found that inactivation of gadE resulted in increased expression of LEE4 genes (155). However, the mechanism underlying this negative regulation was not identified as the expression of Ler, the key positive regulator of LEE, remained unchanged in the AgadE (95). Our study found that GadE negatively regulates genes of all five of the LEE operons including Ier. To determine the mechanism by which GadE negatively regulates LEE, a AgadEA/er strain was constructed and the expression of LEE genes was analyzed. Inactivation of Ier reversed the effect of gadE inactivation on LEE indicating that GadE acts through 124 Ler. Presence of a putative GAD box upstream of Ier also suggested that GadE directly suppresses Ier expression resulting in decreased LEE expression. Expression of LEE is affected by several environmental factors such as temperature, osmolarity and bicarbonate ion concentration in the medium (2, 157, 164). However, the effect of acidity on the LEE is not well investigated. It is important to understand how acidic pH affects a major virulence factor of E. coli O157:H7 such as LEE as it may provide insights into the temporal expression of virulence factors as the organism passes through the stomach. As GadE is a negative regulator of LEE and as acidic pH induces GadE, LEE expression possibly decreases at low pH. Supporting this, there was a 6 — 9-fold decrease in expression of LEE genes including, Ier, when E. coli O157:H7 cells were grown at pH 5.0, a moderate acidic condition. GadE appears to control this decrease in expression of Ier as there was no reduction in Ier expression in the AgadE at pH 5.0. However, the other LEE genes tested still showed a partial decrease in expression in the AgadE at acidic pH indicating that other regulators may also be involved in this mechanism. Other GAD regulators such as GadX and Eng have been shown to negatively regulate LEE expression in EPEC (99, 139). GadX negatively regulates LEE expression through the PerA regulator at pH 5.5 in a complex medium (139). To determine whether the same mechanism exists in E. coli O157:H7, a AgadX strain was constructed and the LEE expression was analyzed at different growth conditions. Unlike in EPEC, inactivation of gadX did not have a marked positive effect on the expression of LEE genes at acidic conditions in 125 E. coli O157:H7. The PerA regulator is not present in E. coli O157:H7 (61) and that may be the reason for the lack of GadX regulation of LEE. On the other hand, eng inactivation in E. coli O157:H7 resulted in a marked upregulation of the LEE genes tir, espD, eae, and sepZ demonstrating that similar to EPEC (99), a negative interaction between Eng and LEE exists in EHEC as well. Subsequently, to determine the effect of gastric acidity on the expression of the major virulence factors of E. coli O157:H7, stationary phase cells were exposed to pH 2.0 and the expression of virulence factors were analyzed. Exposure to pH 2.0 resulted in a marked decrease in the expression of virulence genes, stx1, ser, LEE genes and p0157 genes. 0n the other hand, there was an increase in expression of the three GAD system genes, gadA, gadB and gadC. These results indicate that upon exposure to extreme acidity, E. coli O157:H7 upregulates the expression of AR genes that are required for protection from acidity and concurrently downregulates the expression of virulence genes whose expression is not required in such an environment. This appears to be an efficient survival strategy because E. coli O157:H7 is not a gastric pathogen and therefore, expressing the virulence factors in stomach is unnecessary and may cause a heavy burden on the energy conservation and survival of the organism. 126 Identification of differentially expressed genes between E. coli O157:H7 genotypes E. coli O157:H7 strains isolated from clinical cases and from cattle can be broadly divided into two genotypes, clinical and bovine-biased. Clinical genotypes predominate in clinical cases, but are also isolated from bovine sources. 0n the other hand, bovine-biased genotypes are isolated mostly from bovine sources and rarely from clinical cases (14). We hypothesized that this variation in infectivity of the two genotypes is due to differential expression of virulence and stress fitness-associated genes. To test this hypothesis, four strains were selected from each genotype and their exponential phase expression profiles were compared. A large number of genes were differentially expressed between the two genotypes. lmportantly, major virulence factors of E. coli O157:H7 such as LEE and p0157 plasmid encoded genes had increased expression in clinical genotype strains. On the other hand, genes that are essential for acid and stress resistance had higher expression in bovine-biased genotype strains. All of the three components of the GAD system, gadA, gadB and gadC, and many of the AFI genes were upregulated in strains of the bovine-biased genotype. The central activator of GAD system, gadE, also had higher expression in bovine-biased genotype, which could be the reason for increased expression of GAD and AFI genes. The increased expression of gadE could have contributed to the decreased expression of LEE in the bovine-biased genotype because as we demonstrated in these studies GadE is a negative regulator of LEE in E. coli O157:H7 (65, 127 155). This increased expression of AR genes in strains of the bovine-biased genotype was confirmed functionally by a model stomach assay in which bovine- biased strains demonstrated at least twice as much survival rate as the clinical genotype strains. Also upregulated in strains of the bovine-biased genotype were the genes of the RpoN regulon. These genes are involved in the nitrogen regulatory response in ammonia limiting conditions (122) and hence, their increased expression could contribute to better survival in a nutrient limiting environment. Single nucleotide polymorphism (SNP) genotyping of the clinical and bovine-biased genotype strains used in this study confirmed that they belong to distinct genetic populations. Three of the clinical genotype strains belonged to clade 8, which has been shown to be associated with more severe disease including a higher incidence of HUS (88) and one strain belonged to clade 1, which has also been frequently isolated from clinical cases (88, 96). Bovine- biased genotype strains on the other hand belonged to a clade that is less frequently isolated from clinical cases, clade 7 (88). Together, the findings from this study suggest that one of the reasons for the variation in distribution of E. coli O157:H7 genotypes between clinical cases and cattle could be the differences in expression levels of genes involved in virulence and resistance to environmental conditions. Presence or absence of genes could also lead to variation in infectivity; however, we have not addressed that aspect in this study. Based on the results from this study, we hypothesize that increased expression of virulence factors such as LEE and p0157 genes is 128 the basis for the predominance of clinical genotype strains in human cases. In contrast, bovine-biased strains are more resistant to acid and other environmental stress, which helps them persist in the bovine reservoir. At the same time, due to the negative interactions between acid resistance and virulence genes, bovine-biased genotype strains express virulence factors at a lower level, which may be the reason for their decreased prevalence in clinical cases. It is possible that these highly environmental stress-resistant bovine- biased strains reach the humans through various food vehicles, but dUe to decreased expression of important adherence and virulence factors do not cause disease in humans. This study also demonstrated that clinical and bovine-biased genotype strains belong to distinct SNP genotypes or clades, which vary in their ability to cause disease in humans. Presence of clade 8 strains among the clinical genotype strains isolated from cattle supports the argument that clade 8 outbreak strains such as the spinach strain could have originated from a cattle farm. Moreover, it demonstrates that SNP genotyping could be used as a technique for detecting highly virulent strains in the reservoir host itself, which may help in prevention and control of E. coli O157:H7 outbreaks. 129 FUTURE DIRECTIONS Effects of gadE inactivation on the survival and virulence of E. coli O157:H7 In vlvo An important finding from this study was that inactivation of gadE in E. coli O157:H7 resulted in increased expression of important adherence factors that are needed for colonization of the host. Furthermore, a functional gadE was found to be necessary for the survival of E. coli O157:H7 in a simulated gastric environment. Naturally, the next step is to assess the survival and virulence of AgadE in vivo using the best animal models of the human disease. Based on the expression results, we hypothesize that even though E. coli O157:H7 AgadE expresses virulence factors such as LEE at a higher level, it may not be able to cause disease because of its inability to cross the gastric acid barrier in the absence of GadE. However, lack of a good animal model has been an issue hindering in vivo studies for E. coli O157:H7. Several animal models such as rabbits, conventional mice, specific pathogen free (SPF) mice and streptomycin-treated mice have been used as disease models. However, in most of them development of clinical signs and pathological lesions are only minimal (97, 124). Recently, germ free Swiss-Webster mice have been shown to develop HUS after EHEC infection even with infectious dose as low as 100 cells. In these germ free mice, the bacteria were found adherent to the cecal and ileal mucosa and caused renal lesions unlike other animal models (35). Therefore, germ free mice could be used 130 for assessing the pathogenicity of AgadE E. coli O157:H7. A caveat in using a mouse model for determining the survival of AgadE is that the gastric pH of mice is significantly higher than that of the human stomach (92). The mouse stomach pH varies between 3.0 (fed) and 4.0 (fasted) (92) and hence, the AgadE may be able to survive better and the results may not be representative of the conditions when bacteria must pass through the human stomach. An alternative approach is to use cattle as animal models for assessing the survival of AgadE. Cattle are the reservoir hosts of E. coli O157:H7 and the pH of cattle abomasum (the true stomach) is similar to that of the human stomach (165). Cattle could provide an ideal environment to determine the role of GadE in survival in gastric acid in vivo. However, the pathogenicity of AgadE cannot be assessed in cattle as 0157 does not cause any clinical disease in them. Also the presence of a four chambered stomach including the rumen in cattle may affect the results. Altemately, the rabbit animal model could be used, which has a stomach pH of 1 — 2 (37). However, similar to some of the mouse models, rabbit is not a well-established model for clinical symptoms for E. coli O157:H7. Therefore, assessing the survival and pathogenicity of AgadE E. coli O157:H7 in the same animal model appears to be a less feasible option currently. 131 GAD system in E. coli O157:H7 This study, along with some of the previous studies from our lab (8, 10) demonstrates that there are many differences in regulation and functioning of GAD system in E. coli O157:H7 compared to the laboratory strain E. coli K-12, in which most of the studies on GAD system have been conducted. However, many aspects of the GAD system in E. coli O157:H7 still remain unclear. Most importantly, the contribution of GadA and GadB to the AR of 0157 has not been investigated. Specifically, it is not known whether one of these genes is sufficient or both are necessary for AR of 0157. In E. coli K-12, at pH 2.0 both gadA and gadB are required for AR whereas at higher pH such as 2.5 either one of them can provide protection (23). This aspect could be different in 0157 because unlike other E. coli strains gadA and gadB sequences remain divergent with distinct regulatory regions in 0157 (8), which might allow them to function more independently. Also, E. coli O157:H7 has been shown to have superior AR compared to other E. coli strains (10). Hence, a study comparing the survival of AgadA and AgadB E. coli O157:H7 strains at pH 2.0 — 2.5 in minimal and complex environments is needed to understand the functioning of GAD system in 0157. l hypothesize that under a minimal acidic environment where the only stress is acidity, presence of either gadA or gadB is sufficient for survival of E. coli O157:H7. However, in a complex acidic condition such as model stomach, which presents multiple stresses, both these genes may be essential for maximum survival. 132 Two conditions that induce the GAD system are stationary phase of growth and acidic pH. Stationary phase induction of GAD system is well investigated and it is known that the alternative sigma factor, RpoS regulates the GAD expression through GadXW at the stationary phase (39). However, the signaling pathways and regulators that induce the GAD system at acidic pH, especially at exponential phase, remain unknown. It has been shown that low pH conditions lower the physiological concentration of CAMP and CRP. This reduction in CAMP-CRP levels can initiate increased transcription of RpoS, which can then induce gadX and gadE transcription leading to increased expression of gadABC (86). However, the expression of RpoS is minimal at exponential phase (39) and therefore I hypothesize that other unidentified regulators are also involved in the acidic induction of GAD system at exponential phase. To address this, a transposon mutagenesis approach could be used in which the survival of mutants at acidic pH during exponential phase is assessed. This might help to identify the regulators that are involved in the induction of the GAD system at acidic pH. 133 Interactions between GadE and Ler in E. coli O157:H7 Comparison of E. coli O157:H7 wild type and AgadE expression profiles showed that GadE is a negative regulator of LEE and a functional Ler is essential for this negative regulation. Presence of a putative GAD box region upstream of Ier implied that GadE directly binds to this GAD box and repress the Ier expression. However, this binding was not confirmed experimentally. An electrophoretic mobility shift assay (EMSA) with tagged-GadE protein and a DNA fragment containing the putative GAD box upstream of Ier could confirm whether the Ier GAD box is a functional GadE binding region. Further, once confirmed as a GadE binding region, this GAD box could be inactivated in a gadE-over expressing strain to demonstrate that in the absence of the Ier GAD box, GadE mediated repression of Ier and LEE does not occur. 134 Differences between clinical and bovine-biased genotypes of E. coli O157:H7 Numerous virulence and acid/stress resistance genes were found to be differentially expressed between strains of the clinical and bovine-biased genotypes of E. coli O157:H7 in this study. Increased expression of AR genes in the bovine-biased genotype was confirmed functionally by model stomach assays. However, phenotypic assays have not been done to confirm that the statistically significant upregulation of LEE genes and p0157 encoded virulence genes in the clinical genotype is biologically significant. l hypothesize that increase in expression of adherence factors such as LEE in clinical genotype can lead to an increase in adherence of these strains to the epithelial cells. This could be confirmed by epithelial cell association assays, which measure the adherence and invasion of bacteria to the epithelial cell. In our lab, bovine mammary epithelial cells (MAC-T) have been used extensively to assess the association of E. coli O157:H7 with host cells and the same could be used for comparing clinical and bovine-biased genotypes. Another approach would be to quantify the association of bacteria with epithelial cells using flow cytometry, in which the bacteria would be labeled with CF DA-SE dye and the labeled bacteria would be used to infect the MAC-T cells. The level of fluorescence from MAC-T cells, as assessed by flow cytometry, will indicate levels of adherence/invasion of E. coli O157:H7. Enterohemolysin (Ehx) assay, which quantifies the hemolytic activity of E. coli O157:H7 in defibrinated sheep blood (129) could be used to confirm the 135 increased expression of p0157-encoded hemolysin (eth) in clinical genotype strains. Finally, animal model experiments comparing the pathogenicity of clinical and bovine-biased genotypes are needed to confirm the results of expression studies and phenotypic assays. Germ free mice, which have been recently established as a model for E. coli O157:H7-induced disease, could be used for this purpose. SNP genotyping of the strains used in this study showed that clinical and bovine-biased strains belong to distinct genetic populations of E. coli O157:H7. Clinical genotype strains belonged to hyper virulent clade 8 and clade 1 whereas bovine-biased strains were all clade 7, a group that causes human disease less frequently (88). An extension of these findings would be to conduct SNP genotyping of a larger number of bovine E. coli O157:H7 isolates that belong to clinical and bovine-biased genotypes and analyze the distribution of clades between them. If the clinical genotype strains isolated from bovines belong to groups that cause severe disease such as clade 8 and clade 2, that could provide further explanation to the predominance these genotypes in human cases. This will also validate SNP genotyping as a superior method for identifying virulent E. coli O157:H7 strains in clinical settings and cattle farms. 136 APPENDICES 137 APPENDIX 1 RIMAANOVA: steps in expression data analysis 1. Preparing the data file Assemble the data from all the arrays into one MS excel file that contains the following information; i) location information for a spot including grid, metarow, metacolumn, row and column ii) probe lD iii) data from each array in the order, Cy3, Cy5 and flag. Save the file in TAB delimited text format. 2. Preparing the design file This file contains information about the microarray experimental design. It is prepared in MS excel. The number of rows depends on the number of arrays and number of columns depends on the number of factors that are included in.the design. The file should include information about array, dye and sample. Sample indicates the biological replicate. Also included in design file is information about the factors included in the experimental design such as strain, growth phase, and treatment. Save the file in TAB delimited text format. 3. Launch R and load MAANOVA 4. Load the data Use read.madata function. 5. Normalize the data Use transfonnmadata function to normalize the data by spatial-intensity joint lowess (rlowess). 6. Create the data 138 The final data set used for statistical analysis is created by createData function by collapsing the replicates and log transforming the values. 7. Make the model Make a statistical model for the data analysis using makeModel function. Include all the factors in the formula such as array, dye, sample, strain, and growth phase in the same order as in the design file and mention which are the random factors. 8. Fit the model into the data Use fitmaanova function. Here the statistical model is fitted into the data for each gene. 9. Test for differential expression Use matest function. Significant differences in expression are identified by F test. Four types of F tests are available in maanova among which Fs test is considered to provide the best results as it does not assume variance is homogenous across genes. P-values are calculated by permutation test. 10. Correction for multiple comparisons Use adevaI function. This step calculates FDR adjusted p-values for the test results by Benjamini-Hochberg linear step-up correction. 139 APPENDIX 2 TABLE 2.81: Genes showing significant strain effect (FDR < 0.1) ECs Geneb Function Expression Expre number'3 ratioc ssion (AgadE/ ratiod WT) (Stat/e nil Genes significantly down-regulated in AgadE ECs 2294 putative oxidoreductase, major subunit 0.75 1.8 E05 3904 putative transport periplasmic protein 0.82 0.5 Genes significantly up-regulated In AgadE ECs 0323 yagZ hypothetical protein 1.32 2.0 ECs 1114 nusA partial putative tail component of prophage 1.22 1.8 CP-933R ECs 1305 ”CT putative helicase 1.19 3.2 ECs 1471 fabG 3—oxoacyl-[acyl-carrier-protein] reductase 1.25 1.14 ECs 1830 yciF putative structural proteins 1.31 6.2 ECs 1994 nIeG2-2 non-LEE effector protein 1.26 2.5 ECs 2060 vgrE * unknown protein associated with Rhs 1.24 2.2 element, VgrE protein E03 2156 nIeGZ-3 non-LEE effector protein 1.28 2.9 ECs 4262 hypothetical protein 1.26 0.5 ECs 4270 cysG hypothetical membrane protein 1.29 0.6 ECs 4271 recG putative ATP binding protein of ABC 1.29 1.3 transporter E05 4272 pde hypothetical membrane protein 1.32 3.5 ECs 4280 iIvG putative DNA processing chain A 1.30 3.3 ECs 4300 hcaT putative membrane protein 1.43 7.9 E05 4307 bass Signal transduction mechanisms 1.43 5.4 E09 4326 yeeA unknown function 1.22 0.2 ECs 4327 putative phospholipid biosynthesis 1.17 0.3 acyltransferase ECs 4334 rhaB hypothetical protein 1.23 0.6 ECs 4335 yegB putative membrane! transport protein 1.43 0.6 140 ECs Geneb Function Expression Expre number‘ ratioc ssion (AgadE/ ratioCl WT) (Stat/e xp) E05 4336 equ hypothetical membrane protein 1.32 0.6 E03 4337 ynhE hypothetical lipoprotein 1.28 0.6 ECs 4338 MM putative 3-oxoacyl-synthase II 1.27 0.8 E03 4339 putative beta-hydroxydecanoyl-ACP 1.22 0.8 dehydrase ECs 4382 chuT putative hemin binding protein 1.39 7.6 ECs 4393 thU member of acid fitness island, component of 1.46 1.5 the MthF multidrug transporter ECs 4550 espF type III secretion system, secreted effector 1.30 0.6 protein ECs 4553 cesDZ type III secretion system chaperone 1.36 0.5 E05 4555 espD type III secretion system, secreted 1.38 translocator protein ECs 4558 escD type III secretion system, structural protein 1.18 1.3 ECs 4560 cesT type III secretion system, chaperone 1.38 0.3 ECs 4561 tir translocated intimin receptor protein 1.39 0.5 ECs 4562 map type III secretion system, secreted effector 1.26 protein ECs 4563 cesF type III secretion system, chaperone 1.31 1.6 ECs 4564 espH type III secretion system, secreted effector 1.33 3.1 protein E03 4565 sepQ type III secretion system, structural protein 1.27 2.9 ECs 4567 orf15 orf of unknown function 1.19 0.9 E03 4571 sepZ * type III secretion system, secreted effector 1.36 6.4 protein ECs 4572 rorf8 orf of unknown function 1.26 3.0 ECs 4575 ’ escC * type III secretion system, structural protein 1.21 0.3 ECs 4584 orf5 orf of unknown function 1.34 0.5 E03 4585 orf4 orf of unknown function 1.41 0.6 ECs 4586 orf3 orf of unknown function 1.37 0.6 ECs 4587 cesAB type III secretion system, chaperone 1.35 0.6 ECs 4588 Ier * type III secretion system, regulator 1.35 0.8 141 ECs Gene” Function Expression Expre number” ratio” ssion (AgadEl ratio” WT) (Stat/e xp) ECs 4590 espG type III secretion system, secreted effector 1.27 0.6 protein ECs 5072 putative carbohydrate ABC transport system 1.23 0.5 permease E05 5074 IysA putative histidine protein kinase 1.30 0.5 ECs 5110 yde lysine tRNA synthetase, inducible; heat 1.39 0.8 shock protein ECs 5218 treR repressor of treA,B,C 1.28 4.8 ECs 5242 putative integrase 1.31 0.3 ECs 5248 orf of unknown function 1.28 1.2 ECs 5249 putative resolvase 1.23 1.2 E03 5256 orf of unknown function 1.28 0.5 ECs 5257 orf of unknown function 1.23 0.6 ECs 5305 orf of unknown function 1.21 ECs 5532 orf of unknown function 1.26 0.7 p0157p38 plasmid gene 1.23 1.4 p0157p80 plasmid gene 1 .23 0.09 ” Locus ID for E. coli 0157:H7 Sakai strain (Genbank # BAOOOOO7). ” The genes marked with asterisks have putative GAD boxes upstream of their sequence. ° Expression ratio = 2 [”32 (”gadE ) “ WWW ” Expression ratio between two growth phases = 2 “”22 (“at )’ “Map” determined by microarray, ratios are reported only for genes with a significant growth phase effect (FDR<005) 142 APPENDIX 3 TABLE 2.82: Genes with significant Interaction between growth phase (exponential and stationary) and strain (wild type and A gadE) effects (FDR< 0.05) E05 Gene” Function WT/ WT/ number AgadE AgadE Exp Stat phase phase ECs0017 nhaA putative cell division protein 1.2 0.7 ECsO120 lpdA lipoamide dehydrogenase (NADH); 1 .1 0.8 component of 2-oxodehydrogenase and pyruvate complexes ECsO282 unknown protein from prophage CP-933H 1.1 0.8 5030283 Iny unknown protein from prophage CP-933H 1.1 0.9 ECsO345 yng hypothetical protein 1.2 0.8 ECsO464 tsx nucleoside channel; receptor of phage T6 1.1 0.8 and colicin K ECs0483 cyoD cytochrome c ubiquinol oxidase subunit IV 1.2 0.8 ECs0484 cyoC cytochrome c ubiquinol oxidase subunit III 1.2 0.8 ECs0681 ybeL putative alpha helical protein 0.8 1.3 ECsO702 erB putative RNA 0.8 1.4 ECs0746 sdhC succinate dehydrogenase, cytochrome 1.2 0.8 b556 ECsO747 sth succinate dehydrogenase, hydrophobic 1.2 0.8 subunn ECs0748 sdhA succinate dehydrogenase, flavoprotein 1.2 0.8 subunfi ECsO749 sth succinate dehydrogenase, iron sulfur 1.2 0.7 protein ECsO750 - hypothetical protein 1.1 0.7 ECs0751 sucA 2-oxoglutarate dehydrogenase 1.2 0.7 (decarboxylase component) 143 ECs Gene‘ Function WT/ WTI number AgadE AgadE Exp Stat phase phase ECs0752 sucB 2-oxoglutarate dehydrogenase 1.1 0.7 (dihydrolipoyltranssuccinase E2 component) ECs0759 htrE unknown function 1.1 0.8 ECsO804 - unknown protein encoded by prophage CP- 1.1 0.9 933K ECs0881 ybiI hypothetical protein 0.7 1.5 ECs1012 ompF outer membrane protein 1a (la;b;F) 1 .1 0.8 ECs1028 ych putative chaperone 1.1 0.9 ECs1122 - partial putative outer membrane protein 1.1 0.9 Lom precursor encoded by prophage CP- 933R ECs1257 - putative synthetase 0.9 1 .3 ECs1271 ych hypothetical protein 1.2 0.8 ECs1355 terD putative tellurium resistance protein TerD 1.1 0.8 ECs1356 terE putative phage inhibition, colicin resistance 1.1 0.8 and tellurite resistance protein ECs1377 yjiM unknown function 1 .1 0.9 ECs1490 ych hypothetical protein 0.8 1.3 ECs1663 ompT outer membrane protein 3b (a), protease 1.2 0.7 VII ECs1668 minE cell division topological specificity factor, 1 .1 0.8 reverses MinC inhibition of ftsZ ring formation ECs1669 minD cell division inhibitor, a membrane ATPase, 1.1 0.8 activates minC ECs171 1 ychM hypothetical protein 0.7 2.1 ECs1722 chaB cation transport regulator 0.6 3.2 ECs1744 oppB oligopeptide transport permease protein 1.2 0.8 ECs1825 yicL unknown function 1 .1 0.8 144 ECs Genea Function WT/ WT/ number AgadE AgadE Exp Stat phase phase E031829 yciE hypothetical protein 1.2 0.7 E031879 goaG 4-aminobutyrate aminotransferase 0.8 1 .4 E031880 pspF* psp operon transcriptional activator 0.6 2.9 E031902 tyrR putative acyltransferase 0.5 3.4 E032018 ygiB unknown function 1 .1 0.9 E032061 - unknown protein associated with Rhs 1.1 0.8 element E032097 gadC" Glutamate-GABA antiporter 0.5 4.1 E032098 gadB* glutamate decarboxylase isozyme 0.4 8.0 E032099 pqu putative peptidase 0.8 1.5 ECs2100 yddB hypothetical protein 0.7 2.0 E032260 unknown protein encoded by cryptic 0.5 4.3 prophage 0P-933P E032279 mdoB putative repressor protein encoded by 0.8 1.2 cryptic prophage 0P-933P E032432 - hypothetical protein 0.9 1.2 ECs2453 - hypothetical protein 0.8 1.4 E032514 fadD acyl-CoA synthetase, long-chain-fatty-acid— 0.8 1 .4 00A ligase E032613 fin cytoplasmic ferritin (an iron storage protein) 1.2 0.8 E032692 - hypothetical protein 0.8 1.3 E032839 - GDP-mannose dehydratase 1.2 0.8 E032840 cchA glycosyl transferase 1.1 0.8 E032841 ydeB perosamine synthetase 1.2 0.8 E032887 baeR hypothetical protein 0.9 1.3 E033027 rfi‘H putative salicylate hydroxylase 1.2 0.8 145 E03 Gene” Function WT/ WT/ number AgadE AgadE Exp Stat phase phase E033041 mgIA ATP-binding component of methyl- 1.2 0.8 galactoside transport and galactose taxis ECs3174 - putative aminotransferase 1 .1 0.8 ECs3213 aroC chorismate synthase 0.8 1.8 ECs3221 - putative fimbrial usher 1.1 0.8 E033223 hypothetical protein 0.8 1.4 E033224 - putative enzyme 07 2.4 E033228 - hypothetical protein 1.2 0.7 5033306 amiA N-acetylmuramoyl-l-alanine amidase l 1.2 0.8 5033307 hemF coproporphyrinogen III oxidase 1.2 0.8 E033340 dapA dihydrodipicolinate synthase 1.1 0.8 ECs3389 pepB putative peptidase 1.2 0.8 E033390 yfl'rJ hypothetical protein 1.1 0.8 E033391 fdx [2FE-2S] ferredoxin, electron carrer protein 1.1 0.8 E033393 yfliE hypothetical protein 1.1 0.8 E033395 NifU-Iike protein 1.2 0.7 E033396 yth cysteine desulfurase 1.1 0.8 E033397 - hypothetical protein 1.2 0.8 E033416 yth hypothetical protein 1.2 0.8 E033422 yfhK putative prophage integrase 0.8 1.6 E033540 proV ATP-binding component of transport 0.9 1.2 system for glycine, betaine and proline E033595 rpoS RNA polymerase, sigma S (sigma38) factor 0.8 1.5 ECs3596 nIpD lipoprotein 0.9 1 .3 146 503 Gene” Function WT/ WT/ number AgadE AgadE Exp Stat phase phase 5033696 IysR positive regulator for lys 0.9 1.3 5033704 yqu putative sensory transducer 0.6 2.4 5033720 yraJ putative transcriptional regulator 1.1 0.9 5033721 phnE putative integral membrane protein- 1.1 0.8 component of type III secretion apparatus 503391 1 ygiN hypothetical protein 0.9 1 .3 5033929 ribB 3,4 dihydroxy-2-butanone-4-phosphate 1.1 0.8 synthase 5033969 ygiR hypothetical protein 0.9 1.3 5034000 yhaB hypothetical protein 0.9 1.4 5034188 hopD leader peptidase HopD 1.1 0.8 5034213 yth hypothetical protein 0.9 1.3 5034294 yhhA hypothetical protein 0.8 1.4 5034363 yhiM hypothetical protein 0.5 2.8 5034389 hdeB member of acid fitness island 0.6 2.8 5034390 hdeA member of acid fitness island 0.3 11.1 5034391 hdeD* member of acid fitness island 0.6 2.5 5034392 gadE member of acid fitness island 0.5 4.4 5034397 gadA" glutamate decarboxylase isozyme 0.4 6.4 5034483 IIdD L-Iactate dehydrogenase 0.5 4.8 5034484 yibK hypothetical protein 0.4 5.4 5034485 cysE serine acetyltransferase 0.8 1.6 5034490 yibO putative 2,3-bisphosphoglycerate- 0.4 5.9 independent phosphoglycerate mutase 5034491 yibP putative membrane protein 0.6 3.4 147 503 Gene” Function WT/ WT/ number AgadE AgadE Exp Stat phase phase 5034579 yfil hypothetical protein 0.9 1.3 5034608 - unknown function 0.9 1.2 5034636 dnaN DNA polymerase III, beta-subunit 0.6 2.8 5034683 gidA glucose-inhibited division; chromosome 1.1 0.8 replication? 5034737 - hypothetical protein 0.5 5.4 5034738 cyaY, Iron binding frataxin homolog 0.6 2.6 5034739 dapF diaminopimelate epimerase 0.6 2.8 5034745 - hypothetical protein 0.8 1.6 5034854 mch unknown function 0.7 1.9 5034891 udhA soluble pyridine nucleotide 1.1 0.8 transhydrogenase 5035030 yij hypothetical protein 1.2 0.8 5035080 phnK hypothetical protein 1.2 0.8 5035092 ych hypothetical protein 1.1 0.8 50351 16 n’dC hypothetical protein 0.7 2.5 5035124 mopA GroEL, chaperone Hsp60, heat shock 1 .1 0.8 protein 5035163 aidB putative acyl coenzyme A dehydrogenase 0.8 1.5 5035215 nrdD anaerobic ribonucleoside-triphosphate 0.9 1 .2 reductase 5035253 hemN partial putative integrase 0.8 1.2 5035334 osmY hyperosmotically inducible periplasmic 0.8 1.3 protein 5035496 - unknown function 0.9 1.3 148 a The genes marked with asterisks have putative GAD boxes upstream of their sequence. APPENDIX 4 T ABLE 2.83: Primer sequences used for Q-PCR (SYBR Green chemistry). Primer Sequence (5’ to 3’) Annealing temperature 163-982 CGATGCAACGCGAAGAACCT 55°C 163-1143 CCGGACCGCTGGCAACAAA tir-664 ACTTCCAGCCTTCGTTCAGA 57°C tir-869 TTCTGGAACGCTTCT'ITCGT espZ-46 GCGACCTCACTCAGTGGAA 55°C espZ-193 CCGCTGCAATACCTGTACCT espD-F GGTTACAAGTCGCACTGAGGA 59°C espD-R CCAGGGATAACAGAGTGACCA espF-F AGCAGCCAGGTGACTTCATT 54°C espF-R CTGTGCAATGGGCGGTAAAG eae-2188 GCCGGTAAAGCGACTGTTAG 55°C eae-2325 ATTAGGCAACTCGCCTCTGA espA-128 AGGCTGCGATTCTCATGTTT 57°C espA-310 GAAGTTTGGCTTTCGCA‘I‘I'C espB-319 TCAGCATTGGGGATCTTAGG 57°C espB-487 CTGCGACATCAGCAACAC‘I‘I’ gadC-571 TGCAAGACCTTCTTCCCTGA 55°C gadC-694 GCCCTGGGTTACTCATTTCA hdeA-F GAAGAT'I'I'CCTGGCTGTGGA 59°C hdeA-R ACGGTTGCAATACCCTGAAC hdeB-F CACTGGTGAACGCACAATCT 59°C hdeB-R TTTCTTCATGCAGCATCCAC gadX-F 'I‘I'ACAACCGAACATGCGAAC 59°C gadX-R CAGACTTGGACTCATCAACAGC eng-F GAGTTGACTGAAGGCGGAAG 59°C eva-R GGTCA‘ITI'TTAGCGGAGACG 149 APPENDIX 5 TABLE 4.81: Genes significantly differentially expressed between clinical genotype 1 and bovine-biased genotype 5 as detected by MAANOVAIFs test (FDR < 0.1) 503 no. Gene Function CIB Genes up-regulated in clinical genotype 1 5030124 speD S-adenosylmethionine decarboxylase 1 .72 5030125 speE sperm idine synthase 1 .78 5030228 hypothetical protein 2.39 5030239 hypothetical protein 2.09 5030273 hypothetical protein 4.48 5030274 unknown 1 .91 5030288 hypothetical protein 2.07 5030290 hypothetical protein 1.76 5030291 hypothetical protein 1.71 5030299 putative DNA binding protein 1.71 5030331 putative NADH-dependent flavin oxidoreductase 1.51 5030445 rdgC recombination associated protein 1.60 5030603 Rhs core protein 1.70 5030744 putative fimbrial-Iike protein 3.1 1 5030754 sucD succinyl-CoA synthetase alpha subunit 1.59 5030764 putative glutamate mutase S - hypothetical protein 1.65 5030815 anti-termination protein 1.53 5030829 putative protease/scaffold protein 2.13 5030837 putative tail length tape measure protein precursor 1.52 5030852 bioA 7,8—diaminopelargonic acid synthetase 1.60 5030853 bioB biotin synthetase 1.53 5031053 ych putative sulfite reductase 1.58 5031 127 hypothetical protein 3.1 3 5031 139 ych hypothetical protein 1 .53 5031 166 hypothetical protein 1 .95 5031 1 67 hypothetical protein 2.1 5 5031220 putative terminase large subunit 1 .97 5031222 hypothetical protein 2.91 5031224 hypothetical protein 1.76 5031225 hypothetical protein 1.55 5031226 hypothetical protein 1.57 5031228 putative tail fiber protein 4.50 5031230 hypothetical protein 2.53 5031235 hypothetical protein 1 .66 5031327 ureG 2.60 5031349 hypothetical protein 3.03 5031370 putative glucosyl-transferase 1 .92 5031378 hypothetical protein 1 .79 5031386 hypothetical protein 1.68 150 5031386 hypothetical protein 1.83 5031 389 hypothetical protein 1 .78 5031814 hypothetical protein 3.00 5032116 putative ATP-binding component of a transport system 2.04 50321 56 hypothetical protein 1 .55 5032276 putative replication protein 1.76 5032332 hypothetical protein 1.73 5032813 sch exonucleasel 1.97 5032947 putative minor tail protein 1.67 5033075 rsuA 168 pseudouridylate 516 synthase 1.74 5033078 yejK nucleoid-associated protein 1.59 5033550 gshA gamma-glutamate-cysteine ligase 1.85 5034426 putative fimbrial protein precursor 2.17 5034565 sepQ TTSS 1.61 5034576 cesD Secretion of EspD 1.77 5034591 rorf1 unknown 1 .58 5034653 hypothetical protein 2.92 5035073 putative ATP-binding component of sugar ABC transporter 1.81 5035318 yjiN putative oxidoreductase 1.55 p0157p02 etpC T288 2.75 p0157p05 eth T288 1.86 p0157p07 etpH T288 1.91 p0157p09 eth T288 2.01 p0157p10 etpK T288 2.22 p0157p11 etpL T288 1.75 p0157p12 etpM T288 1.92 p0157p14 etpO T288 1.69 p0157p18 thA hemolysin A 1 .73 p0157p24 repFIB RepFlB 1.67 p0157p58 toxB toxin B 1.59 p0157p79 hypothetical protein 2.44 p0157p81 hypothetical protein 2.12 ECs no. Gene Function BIC Genes up-regulated In bovine-biased genotype 5 5030068 araC transcriptional regulator for ara operon 1.79 5030075 IeuD isopropylmalate isomerase subunit 2.49 5030076 leuC 3-isopropylmalate isomerase (dehydratase) subunit 2.45 5030077 IeuB 3-isopropylmalate dehydrogenase 2.40 5030231 unknown function 1.62 5030419 tauA taurine transport system periplasmic protein 1.65 5030420 tauB taurine ATP-binding component of a transport system 1.74 5030499 ybaE hypothetical protein 1.67 5030538 ybaS putative glutaminase 3.00 5030645 ahpF hypothetical protein 1.51 5030732 hypothetical protein 2.04 5030848 unknown protein encoded by prophage CP-933K 4.20 5030868 hypothetical protein 1.86 5030890 dps DNA protection during starvation 3.56 5030892 ome outer membrane protein X 1.71 151 5030957 5030966 5030992 5031 077 5031 159 5031 181 5031 182 5031 197 5031 201 5031 31 1 5031 428 5031 508 5031 509 5031 546 5031 683 5031 722 5031 743 5031 744 5031 746 5031 747 5031 769 5031 769 5031 800 5031 831 5031 878 5031 975 5032044 5032053 5032086 5032097 5032098 50321 96 5032206 5032236 5032300 5032370 5032384 5032430 5032431 5032546 5032595 5032692 5032758 5032761 5032825 5032900 5032942 5033018 5033061 cspD ycaL ninE ninG ych chaB oppA oppB oppD oppF yciG ordL yncB osmC gadC gadB mlc Ipp tap hisA yehZ fruB pyruvate oxidase cold shock protein putative heat shock protein hypothetical protein hypothetical protein putative anti-termination protein N hypothetical protein hypothetical protein unknown protein encoded by bacteriophage BP-933W unknown in IS600 hypothetical protein unknown protein encoded by prophage CP-933N unknown protein encoded by prophage CP-933N unknown protein encoded by prophage 0P-933N putative sporulation protein cation transport regulator oli90peptide transport; periplasmic binding protein oligopeptide transport permease protein ATP-binding protein of oligopeptide ABC transport system ATP-binding protein of oligopeptide ABC transport system fimbrial minor pilin protein precursor putative fimbrial minor pilin protein precursor putative tail component encoded by cryptic prophage CP- 933M hypothetical protein probable oxidoreductase unknown protein encoded by cryptic prophage CP-933M putative transport protein putative oxidoreductase osmotically inducible protein GABA-glutamate antiporter glutamate decarboxylase isozyme hypothetical protein unknown protein encoded within prophage CP-9330 putative tail assembly protein of cryptic prophage CP- 933P putative NAGC-like transcriptional regulator cyclopropane fatty acyl phospholipid synthase murein lipoprotein hypothetical protein hypothetical protein hypothetical protein methyl-accepting chemotaxis protein IV hypothetical protein hypothetical protein unknown protein encoded within prophage CP-933U imidazolecarboxamide isomerase hypothetical protein hypothetical protein putative transport system permease protein PTS system, fructose-specific IlAlfpr component 152 2.48 2.49 1.59 1.70 2.96 3.06 3.55 2.31 1.74 2.02 1.63 2.02 2.44 1.64 2.96 2.37 2.10 1.56 1.82 1.73 2.60 3.09 1.53 1.55 1.50 2.02 1.54 1.55 2.14 2.48 4.35 2.32 2.05 1.54 1.51 1.96 1.99 1.83 2.47 1.69 1.56 2.52 2.05 2.57 2.14 1.50 1.74 1.73 4.47 5033092 50331 04 503311 1 50331 89 50331 94 5033206 5033326 5033327 5033409 5033533 5033680 5033839 5033973 5033974 5033980 5033981 503421 2 503421 3 5034273 5034275 5034276 5034278 5034377 5034389 5034390 5034391 5034394 5034395 5034397 503441 2 5034427 5034440 5034483 5034490 5034660 5034706 5034734 5034737 5034760 5034871 5034957 5034958 5034959 5034969 5034970 5035038 5035039 50351 20 50351 33 5035253 ompC argT taIA ppdC mItC uxaA uxaC yqiD algP 9’90 ngX asd slp hdeB hdeA hdeD yhrV gadW gadA IIdD yibO phoU iIvA hemD YSQA katG frdC cytochrome c-type protein outer membrane protein 1b hypothetical protein UDP-galactopyranose mutase Iysine-, arginine-, ornithine-binding periplasmic protein putative transport protein transaldolase A transketolase 2 isozyme hypothetical protein hypothetical protein hypothetical protein transport of nucleosides, permease protein altronate hydrolase uronate isomerase hypothetical protein hypothetical protein induced in stationary phase, recognized by rpoS hypothetical protein glycogen phosphorylase glucose-1-phosphate adenylyltransferase part of glycogen operon, a glycosyl hydrolase aspartate-semialdehyde dehydrogenase outer membrane protein periplasmic chaperone protection from organic acid metabolites acid resistance at high cell density multidrug efflux pump protein ARAC-type GAD system regulator glutamate decarboxylase isozyme hypothetical protein hypothetical protein hypothetical protein L-lactate dehydrogenase Putative phosphoglycerate mutase negative regulator for pho regulon threonine deaminase (dehydratase) uroporphyrinogen lll synthase hypothetical protein putative enzyme catalase; hydroperoxidase HPI(I) hypothetical protein hypothetical protein hypothetical protein putative portal protein hypothetical protein hypothetical protein hypothetical protein aspartate ammonia-Iyase (aspartase) fumarate reductase, anaerobic partial putative integrase 153 2.19 1.51 2.52 2.24 1.54 2.43 1.97 2.19 1.69 2.58 2.08 1.57 1.52 1.57 1.97 1.91 2.37 2.21 1.52 1.50 1.55 1.61 2.45 3.37 5.60 2.67 2.24 1.79 1.87 1.77 1.74 1.82 1.70 1.56 1.78 1.80 2.13 3.38 1.81 1.71 2.37 2.89 2.17 1.90 2.61 2.21 2.18 2.83 1.54 1.87 5035303 503531 2 503531 3 5035325 5035326 5035334 yti 1270 bglJ osmY hypothetical protein hypothetical protein putative carbon starvation protein putative regulator 2-component transcriptional regulator hyperosmoticajy inducible periplasmigarotein 154 1.84 2.52 3.58 1.67 2.13 3.01 APPENDIX 6 TABLE 4.82: Genes differentially expressed between clinical genotype 1 and bovine-biased genotype 5 as detected by SAM (FDR < 0.05) 50s no. Gene Function CIB Genes upregulated in clinical genotype 1 5030026 rpsT 308 ribosomal subunit protein 820 2.04 5030063 rapA probable ATP-dependent RNA helicase 2.27 5030201 abc ATP-binding component of a transporter 2.29 5030207 dniR transcriptional regulator for nitrite reductase 3.17 5030228 hypothetical protein 2.27 5030239 hypothetical protein 2.1 5 5030251 yafK hypothetical protein 1.99 5030265 gpt guanine-hypoxanthine phosphoribosyltransferase 3.14 5030273 unknown protein from prophage CP-933H 4.14 5030274 repressor protein CI 1.89 5030288 unknown protein encoded by IS2 1.92 503031 1 putative transcriptional regulator 2.17 5030338 putative dehydrogenase 1.89 5030351 Unknown function 1.96 5030456 queA synthesis of queuine in tRNA 2.19 5030477 yajK putative oxidoreductase 1.90 5030488 yajG putative polymerase/proteinase 1.88 5030510 yIaB hypothetical protein 2.17 5030522 apt adenine phosphoribosyltransferase 2.12 5030584 purK phosphoribosylaminoimidazole carboxylase 1.84 5030586 ybbF hypothetical protein 1.83 5030588 cysS cysteine tRNA synthetase 1.88 5030673 mrdA cell elongation 2.18 5030699 erA hypothetical protein 1.92 5030716 ybe hypothetical protein 2.06 5030744 ybgD putative fimbriae structural protein 3.34 5030771 ybgC hypothetical protein 2.19 5030772 toIQ inner membrane protein 2.17 5030800 yth putative integrase encoded by prophage CP- 2.11 933K 5030829 putative protease encoded in prophage CP-933K 2.11 5030883 putative outer membrane receptor for iron 2.07 transport 5030896 ybiS hypothetical protein 2.26 503091 5 yIiG hypothetical protein 1 .97 5030939 ybjF putative enzyme 1.85 5030944 artM arginine 3rd transport system permease protein 1.53 5030945 artQ arginine 3rd transport system permease protein 1.81 5030946 art! arginine 3rd transport system periplasmic binding 1.94 protein 5030947 artP ATP-binding component of 3rd arginine transport 1.65 155 5030960 5030997 5031 127 5031142 5031166 5031167 5031192 5031 203 5031 220 5031 222 5031228 5031230 5031 233 5031234 5031236 5031 327 5031 349 5031 362 5031 370 5031 378 5031 379 5031 386 5031 387 5031 398 5031433 5031464 5031494 5031 500 5031564 5031605 5031612 ECs1 646 5031 703 . 5031 705 5031 709 5031 733 5031 740 5031814 5031854 5031859 ybjE msbA ymcC ureG yceA yceC ych potC yceC tdk pyrF rnb system putative surface protein ATP-binding transport protein unknown protein encoded by cryptic prophage CP-933M putative regulator hypothetical protein hypothetical protein unknown protein encoded by bacteriophage BP- 933W antitermination protein 0 of bacteriophage BP- 933W partial putative terminase large subunit of bacteriophage BP-933W unknown protein encoded by bacteriophage BP- 933W putative tail fiber protein of bacteriophage BP- 933W unknown protein encoded by bacteriophage BP- 933W unknown protein encoded by bacteriophage BP- 933W putative outer membrane protein putative outer membrane protein of bacteriophage BP-933W Unknown function Unknown function Unknown function putative glucosyltransferase Unknown function hypothetical protein Unknown function Unknown function Unknown function hypothetical protein hypothetical protein hypothetical protein spermidinelputrescine transport system permease I82 hypothetical protein hypothetical protein putative replication protein P of bacteriophage BP-933W putative tail component of prophage CP-933K putative PTS system enzyme l putative dihydroxyacetone kinase (50 2.7.1.2) peptidyl-tRNA hydrolase a protaminelike protein thymidine kinase hypothetical protein orotidine-5'-phosphate decarboxylase RNase II, mRNA degradation 156 2.34 1.94 2.92 1.81 2.01 2.09 1.78 2.60 1.86 2.67 3.90 2.54 2.36 1.73 2.00 2.51 2.97 2.05 1.93 1.67 1.79 1.79 1.86 1.80 1.81 2.00 1.77 1.87 2.16 1.94 1.74 2.82 2.04 1.69 1.68 1.94 1.95 2.90 2.13 1.87 5031872 5031 928 5032021 5032099 50321 00 50321 16 5032229 5032259 5032337 5032367 503251 6 5032653 503281 3 503281 6 5032828 5032873 5032889 5032973 5032974 5032975 5032992 5033002 5033048 5033067 5033079 5033223 5033309 5033362 5033380 5033399 5033530 5033550 5033620 5033673 5033689 5033818 5033948 5034043 5034054 5034069 50341 32 50341 33 5034214 5034238 ymr'A ydaO aIdA pqu yddB purR yeaZ yecF sch yeeF udk yegQ stx1b stx1a IysP SPF yell- purN ndk suhB stpA ych mItA metK deaD secG yrbA yth fis ppiA hypothetical protein hypothetical protein - aldehyde dehydrogenase, NAD-linked putative peptidase hypothetical protein putative ATP-binding component of a transport system unknown protein encoded by prophage CP-9330 putative lysozyme hypothetical protein transcriptional repressor for pur regulon, glyA, glnB, prsA, speA hypothetical protein hypothetical protein exonuclease I, 3' -> 5' specific putative amino acid/amine transport protein regulator of length of O-antigen component of Iipopolysaccharide chains uridine/cytidine kinase hypothetical protein shiga-like toxin 1 subunit B encoded within prophage CP-933V shiga-like toxin 1 subunit A encoded within prephage 0P-933V putative antitermination protein 0 for prophage CP-933V unknown protein encoded within prophage 0P- 933V putative exonuclease of bacteriophage BP-933W lysine-specific permease putative lipoprotein hypothetical protein hypothetical protein hypothetical protein phosphoribosylglycinamide formyltransferase 1 nucleoside diphosphate kinase extragenic suppressor DNA-binding protein unknown protein encoded by bacteriophage BP- 933W putative 6-pyruvoyl tetrahydrobiopterin synthase membrane-bound lytic murein transglycosylase A putative transport protein methionine adenosyltransferase 1 308 ribosomal subunit protein 821 inducible ATP-independent RNA helicase protein export - membrane protein hypothetical protein putative dehydrogenase site-specific DNA inversion stimulation factor peptidyl-prolyl cis-trans isomerase A (rotamase A) peptidoglycan synthetase; penicillin-binding 157 1.92 1.86 2.31 2.51 2.13 2.08 1.79 2.83 1.83 2.19 1 .89 2.42 2.01 3.83 2.39 2.08 2.46 19.82 15.10 1.89 4.28 1.76 2.17 1.63 1.82 2.10 1.94 1.84 2.14 3.30 2.24 1.84 1.80 2.05 2.18 1.91 2.27 2.07 2.17 1.72 2.50 2.18 1.83 1.96 ECs4426 protein 1A putative fimbrial subunit 1.89 5034461 yiaU putative transcriptional regulator LYSR-type 2.35 5034550 espF secreted protein/effector 1.97 5034552 escF T388 2.09 5034566 orf16 secretion of translocators 2.16 5034570 orf12 T388 2.1 3 5034574 sepD 2.04 5034579 rorf3 1.91 5034581 escT T388 1 .98 5034583 escR T388 1 .91 5034590 espG secreted protein/effector 1.88 5034606 uhpA response regulator, positive activator of uhpT 2.24 transcription 5034638 508 ribosomal subunit protein L34 2.38 5034639 rnpA RNase P, protein component; protein 05; 2.73 processes tRNA, 4.58 RNA 5034653 Unknown function 2.97 5034976 hypothetical protein 1.83 5034982 hypothetical protein 1.84 5035074 Unknown function 1.80 5035147 miaA hypothetical protein 2.09 5035246 Unknown function 2.59 5035252 putative transcriptional regulator 1.88 5035320 yjiA putative glycoprotein/receptor 1.72 p0157p02 etpC T288 3.04 p01 57p05 eth T288 1 .98 p01 57p09 etoJ T288 2.21 p0157p10 etpK T288 2.33 p0157p12 etpM T288 2.02 p0157p79 hypothetical protein 2.43 p01 57p80 hypothetical protein 2.07 p0157p81 hypothetical protein ‘ 2.30 pOSAK1_02 hypothetical protein 8.58 pOSAK1_03 hypothetical protein 20.66 ECs no. Gene Function BIC Genes upregulated in bovine-biased genotype 5 5030007 yan inner membrane transport protein 2.01 5030068 araC transcriptional regulator for ara operon 1.84 5030075 IeuD isopropylmalate isomerase subunit 2.77 5030076 IeuC 3-isopropylmalate isomerase subunit 2.68 5030077 IeuB 3-isopropylmalate dehydrogenase 2.62 5030078 IeuA 2-isopropylmalate synthase 1.94 50301 18 9095 pyruvate dehydrogenase (decarboxylase 3.10 component) 50301 19 aceF pyruvate dehydrogenase 3.61 (d ihydrolipoyltransacetylase component) 5030169 glnD uridylyltransferase acts on regulator of glnA 1.83 5030191 yan hypothetical protein 2.19 5030287 yhrV putative AraC-type regulatory protein encoded in 1.99 158 5030357 5030358 5030408 5030504 5030505 5030537 5030538 5030646 5030681 5030692 5030693 5030694 5030732 5030780 5030781 5030848 5030868 5030873 5030889 5030890 5030892 5030916 5030927 5030957 5030966 5030968 5031002 5031 037 5031 077 5031091 5031 159 5031 159 5031 180 5031 181 5031 182 5031 183 5031 184 5031 185 5031 197 5031 201 5031252 5031253 5031 254 5031255 5031 256 5031 257 503131 1 betA betB mhpT gInK amtB ybaR ybaS ybdQ ybeL gItK gltJ ybeJ ybgA ybgR yth glnH dps ome yjiU poxB espD prA rmf gef ych wrbA poIB citB yde figK yiaW ych ych prophage 0P-933H choline dehydrogenase, a flavoprotein NAD+-dependent betaine aldehyde dehydrogenase putative transport protein nitrogen regulatory protein P-ll 2 probable ammonium transporter putative ATPase putative glutaminase hypothetical protein putative alpha helical protein glutamate/aspartate transport system permease glutamate/aspartate transport system permease putative periplasmic binding transport protein hypothetical protein putative transport system permease protein putative homeobox protein unknown protein encoded by prophage 0P-933K hypothetical protein putative membrane protein permease of periplasmic glutamine-binding protein global regulator, starvation conditions outer membrane protein X putative receptor hypothetical protein pyruvate oxidase cold shock protein ATP-binding component of serine protease hypothetical protein ribosome modulation factor hypothetical protein putative transcriptional regulator hypothetical protein trp repressor binding protein hypothetical protein putative anti-termination protein N hypothetical protein hypothetical protein hypothetical protein putative cl repressor protein hypothetical protein unknown protein encoded by bacteriophage BP- 933W putative transport protein hypothetical protein putative enzyme putative acetyltransferase hypothetical protein putative synthetase unknown in l8600 159 2.02 1.70 2.19 11.58 14.94 1.98 3.30 2.50 2.24 1.91 2.16 3.52 2.00 3.00 2.07 4.00 1.88 1.77 2.12 3.98 1.70 9.18 1.99 2.75 2.55 1.99 4.08 3.11 1.71 2.22 3.47 2.94 2.15 3.37 3.79 2.26 2.19 2.49 2.31 1.77 3.56 2.61 6.83 10.01 8.93 5.59 2.03 5031 316 5031 360 5031442 5031 508 5031 509 5031 572 5031683 5031684 5031 71 O 5031 722 5031 723 5031 728 5031 741 5031 743 5031 746 5031 747 5031 769 5031 775 5031 91 4 5031975 5032084 5032086 5032087 5032088 5032089 5032090 5032091 5032092 5032097 5032098 50321 96 5032203 5032206 5032209 503221 2 5032299 5032370 5032384 5032429 5032430 5032431 5032450 5032451 5032452 5032454 5032467 gpr yjiO pepT yegB dadA ychH chaB chaC narK adhE oppA oppD oppF ydaA adhE rpsV osmC gadC gadB mch pstC ydaU ybj T yng pflrB ydiS cstC gdhA putative diacylglycerol kinase putative sensor-type regulator glutaredoxin 2 unknown protein encoded by prophage 0P-933N unknown protein encoded by prophage 0P-933N putative peptidase T putative sporulation protein D-amino acid dehydrogenase subunit hypothetical protein cation transport regulator cation transport regulator nitrite extrusion protein CoA-Iinked acetaldehyde dehydrogenase oligopeptide transport oligopeptide transport oligopeptide transport putative phage replication protein hypothetical protein hypothetical protein unknown protein encoded by cryptic prophage 0P-933M 308 ribosomal subunit protein 822 osmotically inducible protein putative ATP-binding component of a transport system putative ATP-binding component of a transport system putative transport protein putative transport system permease protein putative hemin-binding lipoprotein hypothetical protein GABA-gluatame antiporter glutamate decarboxylase isozyme hypothetical protein hypothetical protein unknown protein encoded within prophage CP- 9330 putative repressor protein encoded within prophage CP-9330 hypothetical protein hypothetical protein cyclopropane fatty acyl phospholipid synthase murein lipoprotein 6-phosphofructokinase ll; suppressor of pka hypothetical protein hypothetical protein hypothetical protein hypothetical protein putative aldehyde dehydrogenase acetylomithine delta-aminotransferase NADP-specific glutamate dehydrogenase 160 2.06 1.92 2.08 2.01 2.55 2.66 3.23 1.94 4.95 2.45 2.23 1.91 3.83 2.18 1.83 1.77 3.17 2.12 2.72 2.04 2.63 2.27 5.17 6.60 2.75 3.55 8.57 9.17 2.50 4.70 2.42 2.14 2.13 1.71 2.11 2.91 2.10 1.95 1.90 1.95 2.69 5.68 6.96 12.07 9.53 2.12 5032492 5032543 5032603 5032604 5032614 5032670 5032692 5032705 503271 2 5032737 5032737 5032758 5032761 5032783 5032784 5032792 5032825 5032882 5032887 5032927 5032942 503301 8 5033035 5033058 5033060 5033061 5033092 5033097 50331 1 1 50331 89 5033206 5033226 5033257 5033259 5033326 5033327 5033346 5033358 5033445 5033460 503351 9 5033522 5033523 5033533 5033543 5033582 5033583 5033643 5033644 5033655 yeaG yecG otsA yecH yedL yedU uxaB araF aas cbl nac hisA baeR yehL ymflll yehZ yeiC fi’uK I‘l’uB napF ng taIA tktB hny yfiD gabD gabT proX hypA hypB chpR reIA ygdH hypothetical protein hypothetical protein putative regulator trehalose-6-phosphate synthase hypothetical protein hypothetical protein hypothetical protein hypothetical protein putative cytochrome putative transcriptional regulator putative transcriptional regulator hypothetical protein unknown protein encoded within prophage CP- 933U transcriptional regulator cys regulon nitrogen assimilation control protein hypothetical protein imidazolecarboxamide isomerase putative membrane protein hypothetical protein hypothetical protein hypothetical protein putative transport system permease protein cytidine/deoxycytidine deaminase putative kinase fructose-1-phosphate kinase PTS system, fructose-specific IIAI‘fpr component cytochrome c-type protein ferredoxin-type protein: electron transfer hypothetical protein UDP-galactopyranose mutase putative transport protein hypothetical protein hypothetical protein putative aminotransferase transaldolase A transketolase 2 isozyme hydrogenase 4 membrane subunit putative DNA replication factor putative formate acetyltransferase putative yhbH sigma 54 modulator hypothetical protein succinate-semialdehyde dehydrogenase 4-aminobutyrate aminotransferase activity hypothetical protein hypothetical protein pleiotrophic effects on 3 hydrogenase isozymes hydrogenase isoenzyme HypB suppressor of inhibitory function of ChpA (p)ppGpp synthetase l (GTP pyrophosphokinase) hypothetical protein 161 6.18 2.46 2.78 2.42 2.10 2.05 2.70 1.95 1.80 2.29 2.32 2.13 2.61 9.76 12.69 2.28 1.90 1.87 2. 54 2.15 1.71 1 .68 2. 79 3.99 2.04 5.30 2.12 2.48 2.46 2.11 2. 34 3. 28 2.52 3.09 2.10 2.29 2.51 3.09 6.10 5.01 6.97 3.69 3.22 2.88 2.47 2.96 2.61 2.10 2.35 1.85 5033669 5033680 5033689 5033799 5033842 5033891 5033904 5033931 5033980 5033981 5034034 5034038 5034089 5034091 5034141 5034142 503421 2 503421 3 503421 6 5034250 5034309 5034323 5034366 5034367 5034377 5034389 5034390 5034391 5034392 5034394 5034395 5034397 5034402 503441 2 5034424 5034427 5034440 5034456 5034477 503461 5 5034624 5034660 5034706 5034708 5034734 5034737 5034760 5034787 5034790 5034803 ybbF ngA gIgS m5 yth yhbT ach gltB fic yth nirB feoA yhiO uspA slp hdeB hdeA hdeD gadE yhrV gadW gadA yth yth dppA yde yiaG yial mtIR yidF yidP phoU iIvA iva hemD ysgA yihA gInG yihS hypothetical protein hypothetical protein hypothetical protein hypothetical protein putative transport protein hypothetical protein putative transport periplasmic protein glycogen biosynthesis, rpoS dependent hypothetical protein hypothetical protein hypothetical protein hypothetical protein aerobic respiration sensor-response protein glutamate synthase, large subunit putative periplasmic binding transport protein putative transport system permease protein induced in stationary phase hypothetical protein nitrite reductase (NAD(P)H) subunit ferrous iron transport protein A high-affinity amino acid transport system hypothetical protein hypothetical protein universal stress protein outer membrane protein periplasmic chaperone protection from organic acid metabolites acid resistance at high cell density central activator of GAD system multidrug efflux pump protein ARAC-type GAD system regulator glutamate decarboxylase isozyme hypothetical protein hypothetical protein dipeptide transport protein hypothetical protein hypothetical protein hypothetical protein repressor for mfl putative transcriptional regulator putative transcriptional regulator negative regulator for pho regulon threonine deaminase (dehydratase) ketol-acid reductoisomerase uroporphyrinogen lll synthase orf, hypothetical protein putative enzyme orf; Unknown function nitrogen regulator I hypothetical protein 162 1.85 2.08 7.67 2.82 2.23 4.13 2.72 2.27 2.11 1.95 1.75 2.44 1.99 1.91 5.55 3.15 2.64 2.45 1.82 2.68 4.99 7.11 2.32 2.64 2.52 3.38 6.12 3.64 2.56 2.31 1.77 1.92 1.88 1.84 1.89 1.75 1.86 1.95 2.13 1.93 2.35 1.86 1.80 3.70 2.20 3.53 1.91 1.98 2.64 2.44 5034847 5034848 5034856 5034893 5034956 5034957 5034958 5034968 5034969 5034970 5034973 503501 2 5035036 5035038 5035039 5035052 5035061 5035100 5035120 50351 34 50351 35 5035184 5035189 5035190 5035253 5035303 503531 1 503531 2 503531 3 503531 5 5035316 503531 7 5035326 5035334 yiiS yii T menG yheN yral zipA purH menC hde 9W3 yij yij yjbR nrfA fth fidB frdA hemN ms fiiA yer yji Y tsr yjiZ yiiM ngJ osmY hypothetical protein putative regulator menaquinone biosynthesis putative citrate permease hypothetical protein hypothetical protein hypothetical protein hypothetical protein putative portal protein hypothetical protein putative protease protein hypothetical protein tyrosine aminotransferase hypothetical protein hypothetical protein periplasmic cytochrome c(552): plays a role in nitrite reduction selenopolypeptide subunit of formate dehydrogenase H regulator of melibiose operon aspartate ammonia-lyase (aspartase) fumarate reductase, anaerobic, iron-sulfur protein subunu fumarate reductase, anaerobic, flavoprotein subunn hypothetical protein hypothetical protein putative oxidoreductase partial putative integrase hypothetical protein hypothetical protein hypothetical protein putative carbon starvation protein methyl-accepting chemotaxis protein I, serine sensor receptor putative transport protein, cryptic, orf, joins former yjiZ and yij hypothetical protein 2-component transcriptional regulator hyperosmotically inducible periplasmic protein 163 2.08 2.15 1.91 1.78 1.91 2.47 2.23 1.81 1.98 2.77 2.11 2.10 1.92 2.45 2.34 2.33 2.08 2.01 3.15 2.22 2.49 2.14 2.05 2.14 1.84 1.93 2.09 2.68 3.84 1.92 2.09 2.37 2.10 3.33 REFERENCES 164 Abe, H., A. Miyahara, T. Oshima, K. Tashiro, Y. Ogura, S. Kuhara, N. Ogasawara, T. Hayashi, and T. Tobe. 2008. Global Regulation by Horizontally Transferred Regulators Establishes the Pathogenicity of Escherichia coli. DNA Res 15:25-38. Abe, H., I. Tatsuno, T. Tobe, A. Okutani, and 0. Sasakawa. 2002. Bicarbonate Ion Stimulates the Expression of Locus of Enterocyte Effacement-Encoded Genes in Enterohemorrhagic Escherichia coli O157:H7. Infect. Immun. 70:3500—3509. Ake, J. A., S. Jelacic, M. A. Ciol, S. L. Watkins, K. F. Murray, D. L. Christie, 5. J. Klein, and P. I. Tarr. 2005. Relative nephroprotection during Escherichia coli D157:H7 infections: association with intravenous volume expansion. Pediatrics 1 1 5:e673-80. Akerley, B. J., P. A. Cotter, and J. F. Miller. 1995. Ectopic expression of the flagellar regulon alters development of the BordeteIIa-host interaction. Cell 80:611-20. Aldick, T., M. Bielaszewska, W. Zhang, J. Brockmeyer, H. Schmidt, A. W. Friedrich, K. S. Kim, M. A. Schmidt, and H. Karch. 2007. Hemolysin from Shiga toxin-negative Escherichia coli 026 strains injures microvascular endothelium. Microbes Infect 9:282-90. Alignan, M., T. Hewezl, M. Petitprez, G. Dechamp-Guillaume, gory, and L. Gentzbittel. 2006. A cDNA microarray approach to decipher sunflower (Helianthus annuus) responses to the necrotrophic fungus Phoma macdonaldii. New Phytologist 170:523-536. Andreoli, S. P., H. Trachtman, D. W. Acheson, R. L. Siegler, and T. G. Obrig. 2002. Hemolytic uremic syndrome: epidemiology, pathophysiology, and therapy. Pediatr Nephrol 17:293—8. Bergholz, T. M., 0. L. Tarr, L. M. Christensen, D. J. Betting, and T. S. Whittam. 2007. Recent Gene Conversions between Duplicated Glutamate Decarboxylase Genes (gadA and gadB) in Pathogenic Escherichia coli. Mol Biol Evol 24:2323-2333. Bergholz, T. M., S. K. Vanaja, and T. S. Whittam. 2009. Gene expression induced in Escherichia coli O157:H7 upon exposure to model apple juice. Appl Environ Microbiol 75:3542-53. 165 10. 11. 12. 13. 14. 15. 16. 17. 18. Bergholz, T. M., and T. S. Whittam. 2007. Variation in acid resistance among enterohaemorrhagic Escherichia coli in a simulated gastric environment. J Appl Microbiol 102:352-362. Bergholz, T. M., and T. S. Whittam. 2007. Variation in acid resistance among enterohaemorrhagic Escherichia coli in a simulated gastric environment. Journal of Applied Microbiology 102:352-362. Bergholz, T. M., L. M. Wick, W. 01, J. T. Riordan, L. M. Ouellette, and T. S. Whittam. 2007. Global transcriptional response of Escherichia coli O157:H7 to growth transitions in glucose minimal medium. BMC Microbiol 7:97. Besser, R. 5., S. M. Lott, J. T. Weber, M. P. Doyle, T. J. Barrett, J. G. Wells, and P. M. Griffin. 1993. An outbreak of diarrhea and hemolytic uremic syndrome from Escherichia coli O157:H7 in fresh-pressed apple cider. Jama 269:2217-20. Besser, T. 5., N. Shaikh, N. J. Holt, P. I. Tarr, M. 5. Konkel, P. Malik- Kale, C. W. Walsh, T. S. Whittam, and J. L. Bono. 2007. Greater diversity of Shiga toxin-encoding bacteriophage insertion sites among Escherichia coli O157:H7 isolates from cattle than in those from humans. Appl. Environ. Microbiol. 73:671-679. Bhagwat, A. A., R. P. Phadke, D. Wheeler, S. Kalantre, M. Gudipati, and M. Bhagwat. 2003. Computational methods and evaluation of RNA stabilization reagents for genome-wide expression studies. Journal of Microbiological Methods 55:399-409. Biase, D. D., A. Tramonti, R. A. John, and F. Bossa. 1996. Isolation, Overexpression, and Biochemical Characterization of the Two lsoforrns of Glutamic Acid Decarboxylase from Escherichia coli. Protein Expression and Purification 8:430-438. Blattner, F. R., G. Plunkett, 3rd, 0. A. Bloch, N. T. Pema, V. Burland, M. Riley, J. Collado-Vides, J. D. Glasner, C. K. Rode, G. F. Mayhew, J. Gregor, N. W. Davis, H. A. Kirkpatrick, M. A. Goeden, D. J. Rose, B. Mau, and Y. Shao. 1997. The complete genome sequence of Escherichia coli K-12. Science 277: 1 453-74. Bordi, 0., L. Theraulaz, V. Mejean, and C. Jourlin-Castelli. 2003. Anticipating an alkaline stress through the Tor phosphorelay system in Escherichia coli. Mol Microbiol 48:21 1-23. 166 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. Burland, V., Y. Shao, N. Pema, G. Plunkett, H. Sofia, and F. Blattner. 1998. The complete DNA sequence and analysis of the large virulence plasmid of Escherichia coli O157:H7. Nucl. Acids Res. 26:4196-4204. Calderwood, S. B., and J. J. Mekalanos. 1987. Iron regulation of Shiga- Iike toxin expression in Escherichia coli is mediated by the fur locus. J Bacteriol 169:4759-64. Canales, R. D., Y. Luo, J. C. Willey, B. Austermiller, C. 0. Barbacioru, C. Boysen, K. Hunkapiller, R. V. Jensen, 0. R. Knight, K. Y. Lee, Y. Ma, B. Maqsodi, A. Papallo, 5. H. Peters, K. Poulter, P. L. Ruppel, R. R. Samaha, L. Shi, W. Yang, L. Zhang, and F. M. Goodsaid. 2006. Evaluation of DNA microarray results with quantitative gene expression platforms. Nat Biotech 24:11 15-1 122. Castanie-Comet, M.-P., and J. W. Foster. 2001. Escherichia coli acid resistance: cAMP receptor protein and a 20 bp cis-acting sequence control pH and stationary phase expression of the gadA and gadBC glutamate decarboxylase genes. Microbiology 147:709-715. Castanie-Comet, M.-P., T. A. Penfound, D. Smith, J. F. Elliott, and J. W. Foster. 1999. Control of acid resistance in Escherichia coli. J. Bacteriol. 181:3525-3535. Castanle-Comet, M.-P., H. Treffandier, A. Francez-Charlot, C. Gutierrez, and K. Cam. 2007. The glutamate-dependent acid resistance system in Escherichia coli: essential and dual role of the His-Asp phosphorelay RcsCDB/AF. Microbiology 153:238-246. Chapman, P. A., C. A. Siddons, D. J. Wright, P. Norman, J. Fox, and E. Crick. 1993. Cattle as a possible source of verocytotoxin-producing Escherichia coli 01 57 infections in man. Epidemiol Infect 1 1 1 :439-47. Chen, J., and M. W. Griffiths. 1999. Cloning and sequencing of the gene encoding universal stress protein from Escherichia coli O157:H7 isolated from Jack-in-a-Box outbreak. Lett Appl Microbiol 29:103-7. Chen, J., J. Lozach, 5. W. Garcia, B. Barnes, S. Luo, I. Mikoulitch, L. Zhou, G. Schroth, and J.-B. Fan. 2008. Highly sensitive and specific microRNA expression profiling using BeadArray technology. Nucl. Acids Res. 36ze87. Choi, S. H., D. J. Baumler, and C. W. Kaspar. 2000. Contribution of dps to Acid Stress Tolerance and Oxidative Stress Tolerance in Escherichia coli O157:H7. Appl. Environ. Microbiol. 66:3911-3916. 167 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. Cody, S. H., M. K. Glynn, J. A. Farrar, K. L. Cairns, P. M. Griffin, J. Kobayashi, M. Fyfe, R. Hoffman, A. S. King, J. H. Lewis, B. Swaminathan, R. G. Bryant, and D. J. Vugia. 1999. An outbreak of Escherichia coli 01572H7 infection from unpasteurized commercial apple juice. Ann Intern Med 130:202-9. Cui, X., and G. A. Churchill. 2003. Statistical tests for differential expression in cDNA microarray experiments. Genome Biol 4:210. Cui, X., J. T. Hwang, J. Qiu, N. J. Blades, and G. A. Churchill. 2005. Improved statistical tests for differential gene expression by shrinking variance components estimates. Biostatistics 6:59-75. Datsenko, K. A., and B. L. Wanner. 2000. One-step inactivation of chromosomal genes in Escherichia coli K-12 using PCR products. Proceedings of the National Academy of Sciences 97:6640-6645. Deng, W., J. L. Puente, S. Gruenheid, Y. Li, B. A. Vallance, A. Vazquez, J. Barba, J. A. lbarra, P. O'Donnell, P. Metalnikov, K. Ashman, S. Lee, D. Goode, T. Pawson, and B. B. Finlay. 2004. Dissecting virulence: Systematic and functional analyses of a pathogenicity island. Proceedings of the National Academy of Sciences 1 01 :3597-3602. Diez-Gonzalez, F., T. R. Callaway, M. G. Kizoulis, and J. B. Russell. 1998. Grain feeding and the dissemination of acid-resistant Escherichia coli from cattle. Science 281 :1666-8. Eaton, K. A., D. I. Friedman, G. J. Francis, J. S. Tyler, V. B. Young, J. Haeger, G. Abu-Ali, and T. S. Whittam. 2008. Pathogenesis of renal disease due to enterohemorrhagic Escherichia coli in gerrn-free mice. Infect Immun 76:3054-63. Elder, R. 0., J. 5. Keen, G. R. Siragusa, G. A. Barkocy-Gallagher, M. Koohmaraie, and W. W. Laegreid. 2000. Correlation of enterohemorrhagic Escherichia coli O157 prevalence in feces, hides, and carcasses of beef cattle during processing. Proc Natl Acad Sci U S A 97:2999-3003. ‘ Fann, M. K., and D. O'Rourke. 2001. Normal bacterial flora of the rabbit gastrointestinal tract: a clinical approach. Seminars in Avian and Exotic Pet Medicine 10:45-47. Feng, P., K. A. Lampel, H. Karch. and T. S. Whittam. 1998. Genotypic and phenotypic changes in the emergence of Escherichia coli O157:H7. J Infect Dis 177:1750-3. 168 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. Foster, J. W. 2004. Escherichia coli acid resistance: tales of an amateur acidophile. Nat Rev Micro 2:898-907. Fox, J. T., X. Shi, and T. G. Nagaraja. 2008. Escherichia coli 0157 in the rectoanal mucosal region of cattle. Foodbome Pathog Dis 5:69-77. Garmendia, J., G. Frankel, and V. F. Crepin. 2005. Enteropathogenic and Enterohemorrhagic Escherichia coli Infections: Translocation, Translocation, Translocation. Infect. Immun. 73:2573-2585. Gong, 8., Z. Ma, and J. W. Foster. 2004. The Era-like GTPase TrrnE conditionally activates gadE and glutamate-dependent acid resistance in Escherichia coli. Molecular Microbiology 54:948-961. Gong, S., H. Richard, and J. W. Foster. 2003. deE (AdiC) Is the ArgininezAgmatine Antiporter Essential for Arginine-Dependent Acid Resistance in Escherichia coli. J. Bacteriol. 185:4402-4409. Grys, T. 5., M. B. Siegel, W. W. Lathem, and R. A. Welch. 2005. The StcE protease contributes to intimate adherence of enterohemorrhagic Escherichia coli O157:H7 to host cells. Infect Immun 73:1295—303. Hancock, D., T. Besser, J. Lejeune, M. Davis, and D. Rice. 2001. The control of VTEC in the animal reservoir. Int J Food Microbiol 66:71-8. Hayashi, T., K. Makino, M. Ohnishi, K. Kurokawa, K. Ishii, K. Yokoyama, C. G. Han, 5. Ohtsubo, K. Nakayama, T. Murata, M. Tanaka, T. Tobe, T. lida, H. Takami, T. Honda, 0. Sasakawa, N. Ogasawara, T. Yasunaga, S. Kuhara, T. Shiba, M. Hattori, and H. Shinagawa. 2001. Complete genome sequence of enterohemorrhagic Escherichia coli 0157:H7 and genomic comparison with a laboratory strain K-12. DNA Res 8:11-22. Hayashi, T., K. Makino, M. Ohnishi, K. Kurokawa, K. Ishii, K. Yokoyama, C.-G. Han, 5. Ohtsubo, K. Nakayama, T. Murata, M. Tanaka, T. Tobe, T. lida, H. Takami, T. Honda, 0. Sasakawa, N. Ogasawara, T. Yasunaga, S. Kuhara, T. Shiba, M. Hattori, and H. Shinagawa. 2001. Complete Genome Sequence of Enterohemorrhagic Eschelichia coli O157:H7 and Genomic Comparison with a Laboratory Strain K-12. DNA Res 8:11-22. Herold, S., J. Siebert, A. Huber, and H. Schmidt. 2005. Global expression of prophage genes in Escherichia coli O157:H7 strain EDL933 in response to norfloxacin. Antimicrob Agents Chemother 49:931-44. 169 49. 50. 51. 52. 53. 54. 55. 56. 57. Herriott, D. 5., D. D. Hancock, 5. D. Ebel, L. V. Carpenter, D. H. Rice, and T. 5. Besser. 1998. Association of herd management factors with colonization of dairy cattle by Shiga toxin-positive Escherichia coli 0157. J Food Prot 61:802-7. Heuvellng, J., A. Possling, and R. Hengge. 2008. A role for Lon protease in the control of the acid resistance genes of Escherichia coli. Mol Microbiol 69:534-47. Heuvelink, A. 5., F. L. van den Biggelaar, J. Zwartkruis-Nahuis, R. G. Herbes, R. Huyben, N. Nagelkerke, W. J. Melchers, L. A. Monnens, and 5. de Boer. 1998. Occurrence of verocytotoxin-producing Escherichia coli 0157 on Dutch dairy farms. J Clin Microbiol 36:3480-7. Hilbom, 5. D., J. H. Mennin, P. A. Mshar, J. L. Hadler, A. Voetsch, C. Wojtkunski, M. Swartz, R. Mshar, M. A. Lambert-Fair, J. A. Farrar, M. K. Glynn, and L. Slutsker. 1999. A multistate outbreak of Escherichia coli O157:H7 infections associated with consumption of mesclun lettuce. Arch Intern Med 159:1758-64. Hilbom, E. D., P. A. Mshar, T. R. Fiorentino, Z. F. Dembek, T. J. Barrett, R. T. Howard, and M. L. Cartter. 2000. An outbreak of Escherichia coli O157:H7 infections and haemolytic uraemic syndrome associated with consumption of unpasteurized apple cider. Epidemiol Infect 124:31-6. Hommais, F., E. Krin, J.-Y. Coppee, C. Lacroix, 5. Yeramlan, A. Danchin, and P. Bertin. 2004. GadE (YhiE): a novel activator involved in the response to acid environment in Escherichia coli. Microbiology 150:61- 72. Hommais, F., 5. Krin, C. Laurent-Winter, O. Soutourina, A. Malpertuy, J.-P. Le Caer, A. Danchin, and P. Bertin. 2001. Large-scale monitoring of pleiotropic regulation of gene expression by the prokaryotic nucleoid- associated protein, H-NS. Molecular Microbiology 40:20-36. lde, T., S. Michgehl, S. Knappstein, G. Heusipp, and M. A. Schmidt. 2003. Differential modulation by Ca2+ of type III secretion of diffusely adhering enteropathogenic Escherichia coli. Infect Immun 71 :1725-32. Inward, C. D., A. J. Howie, M. M. Fitzpatrick, F. Rafaat, D. V. Milford, and C. M. Taylor. 1997. Renal histopathology in fatal cases of diarrhoea- associated haemolytic uraemic syndrome. Pediatr Nephrol 11 :556-9. 170 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. Issartel, J. P., V. Koronakis, and 0. Hughes. 1991. Activation of Escherichia coli prohaemolysin to the mature toxin by acyl carrier protein- dependent fatty acylation. Nature 351 :759-61. Iyer, R., C. Williams, and C. Miller. 2003. Arginine-agmatine antiporter in extreme acid resistance in Escherichia coli. J Bacteriol 185:6556-61. lyoda, S., and H. Watanabe. 2004. Positive effects of multiple pch genes on expression of the locus of enterocyte effacement genes and adherence of enterohaemorrhagic Escherichia coli O157 : H7 to HEp-2 cells. Microbiology 150:2357-2571. lyoda, S., and H. Watanabe. 2004. Positive effects of multiple pch genes on expression of the locus of enterocyte effacement genes and adherence of enterohaemorrhagic Escherichia coli 0157 : H7 to HEp-2 cells. Microbiology 150:2357-571. Jackson, S. G., R. B. Goodbrand, R. P. Johnson, V. G. Odorico, D. Alves, K. Rahn, J. B. Wilson, M. K. Welch, and R. Khakhria. 1998. Escherichia coli 0157:H7 diarrhoea associated with well water and infected cattle on an Ontario farm. Epidemiol Infect 120:17-20. Jln, W., R. M. Riley, R. D. Wolfinger, K. P. White, G. Passador-Gurgel, and G. Gibson. 2001. The contributions of sex, genotype and age to transcriptional variance in Drosophila melanogaster. Nat Genet 29:389- 395. Just, J. R., and M. A. Daeschel. 2003. Antimicrobial Effects of Wine on Escherichia coli O157:H7 and Salmonella typhimurium in a Model Stomach System. Journal of Food Science 68:285-290. Kailasan Vanaja, S., T. M. Bergholz, and T. S. Whittam. 2009. Characterization of the Escherichia coli O157:H7 Sakai GadE regulon. J. Bacteriol. 191:1868-1877. Kaper, J. B., J. P. Nataro, and H. L. Mobley. 2004. Pathogenic Escherichia coli. Nat Rev Microbiol 2:123-40. Karch, H., P. l. Tarr, and M. Bielaszewska. 2005. Enterohaemorrhagic Escherichia coli in human medicine. Int J Med Microbiol 295:405-18. Keene, W. 5., J. M. McAnulty, F. C. Hoesly, L. P. Williams, Jr., K. Hedberg, G. L. Oxman, T. J. Barrett, M. A. Pfaller, and D. W. Fleming. 1994. A swimming-associated outbreak of hemorrhagic colitis caused by Escherichia coli O157:H7 and Shigella sonnei. N Engl J Med 331 :579-84. 171 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. Kenny, B., A. Abe, M. Stein, and B. B. Finlay. 1997. Enteropathogenic Escherichia coli protein secretion is induced in response to conditions similar to those in the gastrointestinal tract. Infect Immun 65:2606-12. Kenny, B., and B. B. Finlay. 1995. Protein secretion by enteropathogenic Escherichia coli is essential for transducing signals to epithelial cells. Proc Natl Acad Sci U S A 92:7991-5. Kern, R., A. Malki, J. Abdallah, J. Tagourti, and G. Richanne. 2007. Escherichia coli HdeB Is an Acid Stress Chaperone. J. Bacteriol. 189:603- 610. Kim, J., J. Nietfeldt, and A. K. Benson. 1999. Octamer-based genome scanning distinguishes a unique subpopulation of Escherichia coli O157:H7 strains in cattle. Proc Natl Acad Sci U S A 96:13288-93. Kress, W., H. Mutschler, and E. Weber-Ban. 2007. Assembly pathway of an AAA+ Protein: tracking ClpA and ClpAP complex formation in real time. Biochemistry 46:6183-6193. Laegreid. W. W., R. 0. Elder, and J. E. Keen. 1999. Prevalence of Escherichia coli O157:H7 in range beef calves at weaning. Epidemiol Infect 123:291-8. Large, T. M., S. T. Walk, and T. S. Whittam. 2005. Variation in acid resistance among Shiga toxin-producing clones of pathogenic Escherichia coli. Appl. Environ. Microbiol. 71:2493—2500. Lathem, W. W., T. 5. Grys, S. 5. Witowski, A. G. Torres, J. B. Kaper, P. I. Tarr, and R. A. Welch. 2002. StcE, a metalloprotease secreted by Escherichia coli O157:H7, specifically cleaves 01 esterase inhibitor. Mol Microbiol 45:277-88. Law, D. 2000. Virulence factors of Escherichia coli 0157 and other Shiga toxin-producing E. coli. J Appl Microbiol 88:729-45. Law, D. 2000. Virulence factors of Escherichia coli 0157 and other Shiga toxin-producing E. coli. Journal of Applied Microbiology 88:729-745. Lefebvre, B., M. S. Diarra, H. Moisan, and F. Malouin. 2008. Detection of virulence-associated genes in Escherichia coli 0157 and non-0157 isolates from beef cattle, humans, and chickens. J Food Prot 71 :1774- 1784. . 172 80. 81. 82. 83. 84. 85. 86. 87. 88. Li, H., C. Wood, T. Getchell, M. Getchell, and A. Stromberg. 2004. Analysis of Oligonucleotide array experiments with repeated measures using mixed models. BMC Bioinforrnatics 5:209. le, J. Y., H. Sheng, K. S. Seo, Y. H. Park, and C. J. Hovde. 2007. Characterization of an Escherichia coli O157:H7 plasmid 0157 deletion mutant and its survival and persistence in cattle. Appl Environ Microbiol 73:2037-47. Lin, J., I. S. Lee, J. Frey, J. L. Slonczewski, and J. W. Foster. 1995. Comparative analysis of extreme acid survival in Salmonella typhimurium, Shigella flexneri, and Escherichia coli. J Bacteriol 177:4097-104. Lin, J., M. Smith, K. Chapin, H. Balk, G. Bennett, and J. Foster. 1996. Mechanisms of acid resistance in enterohemorrhagic Escherichia coli. Appl. Environ. Microbiol. 62:3094-3100. Ma, Z., S. Gong, H. Richard, D. L. Tucker, T. Conway, and J. W. Foster. 2003. GadE (Y hi5) activates glutamate decarboxylase-dependent acid resistance in Escherichia coli K-12. Molecular Microbiology 49:1309- 1 320. Ma, Z., N. Masuda, and J. W. Foster. 2004. Characterization of 5ngS- YdeO-GadE Branched Regulatory Circuit Governing Glutamate- Dependent Acid Resistance in Escherichia coli. J. Bacteriol. 186:7378- 7389. Ma, Z., H. Richard, and J. W. Foster. 2003. pH-Dependent modulation of cyclic AMP levels and GadW-dependent repression of RpoS affect synthesis of the GadX regulator and Escherichia coli acid resistance. J Bacteriol 185:6852-9. Ma, Z., H. Richard, D. L. Tucker, T. Conway, and J. W. Foster. 2002. Collaborative Regulation of Escherichia coli Glutamate-Dependent Acid Resistance by Two AraC-Like Regulators, GadX and GadW (YhiW). J. Bacteriol. 184:7001-7012. Manning, S. D., A. S. Motiwala, A. C. Springman, W. Qi, D. W. Lacher, L. M. Ouellette, J. M. Mladonicky, P. Somsel, J. T. Rudrik, S. E. Dietrich, W. Zhang, B. Swaminathan, D. Alland, and T. S. Whittam. 2008. Variation in virulence among clades of Escherichia coli O157:H7 associated with disease outbreaks. Proc Natl Acad Sci U S A 105:4868- 73. 173 89. 90. 91. 92. 93. 94. 95. 96. 97. 98. 99. Masuda, N., and G. M. Church. 2002. Escherichia coli Gene Expression Responsive to Levels of the Response Regulator 5ng. J. Bacteriol. 184:6225-6234. Masuda, N., and G. M. Church. 2003. Regulatory network of acid resistance genes in Escherichia coli. Molecular Microbiology 48:699-712. Mates, A. K., A. K. Sayed, and J. W. Foster. 2007. Products of the Escherichia coli Acid Fitness Island Attenuate Metabolite Stress at Extremely Low pH and Mediate a Cell Density-Dependent Acid Resistance. J. Bacteriol. 189:2759-2768. McConnell, E. L., A. W. Basit, and S. Murdan. 2008. Measurements of rat and mouse gastrointestinal pH, fluid and lymphoid tissue, and implications for in-vivo experiments. Journal of Pharmacy and Pharmacology 60:63-70. . McDaniel, T., K. Jarvis, M. Donnenberg, and J. Kaper. 1995. A Genetic Locus of Enterocyte Effacement Conserved Among Diverse Enterobacterial Pathogens. Proceedings of the National Academy of Sciences 92:1664-1668. McDaniel, T. K., K. G. Jarvis, M. S. Donnenberg, and J. B. Kaper. 1995. A genetic locus of enterocyte effacement conserved among diverse enterobacterial pathogens. Proc Natl Acad Sci U S A 92:1664-8. Mellies, J. L., A. M. S. Barron, and A. M. Carmona. 2007. Enteropathogenic and Enterohemorrhagic Escherichia coli Virulence Gene Regulation. Infect. Immun. 75:4199-4210. Michino, H., K. Arakl, S. Minami, S. Takaya, N. Sakai, M. Miyazaki, A. Ono, and H. Yanagawa. 1999. Massive outbreak of Escherichia coli O157:H7 infection in schoolchildren in Sakai City, Japan, associated with consumption of white radish sprouts. Am J Epidemiol 150:787-96. Mundy, R., F. Girard, A. J. FitzGerald, and G. Frankel. 2006. Comparison of colonization dynamics and pathology of mice infected with enteropathogenic Escherichia coli, enterohaemorrhagic E. coli and Citrobacter rodentium. FEMS Microbiol Lett 265:126-32. Murphy, K., and K. Campellone. 2003. Lambda Red-mediated recombinogenic engineering of enterohemorrhagic and enteropathogenic E. coli. BMC Molecular Biology 4:11. Nadler, 0., Y. Shifrin, S. Nov, 8. Kobi, and I. Rosenshine. 2006. Characterization of Enteropathogenic Escherichia coli Mutants That Fail 174 100. 101. 102. 103. 104. 105. 106. 107. 108. To Disrupt Host Cell Spreading and Attachment to Substratum. Infect. Immun. 74:839-849. Nakanishi, N., H. Abe, Y. Ogura, T. Hayashi, K. Tashiro, S. Kuhara, N. Sugimoto, and T. Tobe. 2006. ppGpp with DksA controls gene expression in the locus of enterocyte effacement (LEE) pathogenicity island of enterohaemorrhagic Escherichia coli through activation of two virulence regulatory genes. Molecular Microbiology 61:194-205. Nataro, J. P., and J. B. Kaper. 1998. Diarrheagenic Escherichia coli. Clin. Microbiol. Rev. 1 1 :142-201 . Naylor, S. W., J. 0. Low, T. E. Besser, A. Mahajan, G. J. Gunn, M. C. Pearce, I. J. McKendrick, D. G. Smith, and D. L. Gally. 2003. Lymphoid follicle-dense mucosa at the terminal rectum is the principal site of colonization of enterohemorrhagic Escherichia coli O157:H7 in the bovine host. Infect Immun 71 :1505-12. Oh, D. H., Y. Pan, 5. Berry, M. Cooley, R. Mandrell, and F. Breidt, Jr. 2009. Escherichia coli O157:H7 strains isolated from environmental sources differ significantly in acetic acid resistance compared with human outbreak strains. J Food Prot 72:503-9. Ollveira, M. G., J. R. Brito, T. A. Gomes, B. E. Guth, M. A. Vieira, Z. V. Naves, T. M. Vaz, and K. Irlno. 2008. Diversity of virulence profiles of Shiga toxin-producing Escherichia coli serotypes in food-producing animals in Brazil. Int J Food Microbiol 127:139-46. Opdyke, J. A., J.-G. Kang, and G. Storz. 2004. GadY, a Small-RNA Regulator of Acid Response Genes in Escherichia coli. J. Bacteriol. 186:6698-6705. Oporto, B., J. I. Esteban, G. Aduriz, R. A. Juste, and A. Hurtado. 2008. Escherichia coli O157:H7 and non-0157 Shiga toxin-producing E. coli in healthy cattle, sheep and swine herds in Northern Spain. Zoonoses Public Health 55:73-81. Osborne, M. J., N. Siddiqui, P. lannuzzi, and K. Gehring. 2004. The solution structure of ChaB, a putative membrane ion antiporter regulator from Escherichia coli. BMC Struct Biol 4:9. Park, S., X. You, and J. A. Imlay. 2005. Substantial DNA damage from submicromolar intracellular hydrogen peroxide detected in pr- mutants of Escherichia coli. Proc Natl Acad Sci U S A 102:9317-9322. 175 109. 110. 111. 112. 113. 114. 115. 116. 117. 118. Paton, J. 0., and A. W. Paton. 1998. Pathogenesis and diagnosis of Shiga toxin-producing Escherichia coli infections. Clin Microbiol Rev 1 1 :450-79. Pema, N. T., G. Plunkett, 3rd, V. Burland, B. Mau, J. D. Glasner, D. J. Rose, G. F. Mayhew, P. S. Evans, J. Gregor, H. A. Kirkpatrick, G. Posfai, J. Hackett, S. Klink, A. Boutin, Y. Shao, L. Miller, 5. J. Grotbeck, N. W. Davis, A. Lim, 5. T. Dimalanta, K. D. Potamousis, J. Apodaca, T. S. Anantharaman, J. Lin, G. Yen, D. C. Schwartz, R. A. Welch, and F. R. Blattner. 2001. Genome sequence of enterohaemorrhagic Escherichia coli O157:H7. Nature 409:529-33. Persson, O., A. Valadi, T. Nystrom, and A. Farewell. 2007. Metabolic control of the Escherichia coli universal stress protein response through fructose-6-phosphate. Mol Microbiol 65:968-78. Peterson, W. L., P. A. Mackowiak, C. C. Barnett, M. Marling-Cason, and M. L. Haley. 1989. The human gastric bactericidal barrier. mechanisms of action, relative antibacterial activity, and dietary influences. J Infect Dis 159:979-83. Pfaffl, M. W. 2001. A new mathematical model for relative quantification in real-time RT-PCR. Nucl. Acids Res. 29:e45. Prevention, 0. f. D. C. a. 2007. Escherichia coli O157:H7 infection associated with drinking raw milk—Washington and Oregon, November- December 2005. MMWR Morb Mortal Wkly Rep 56:165-7. Prevention, 0. f. D. C. a. 1997. Escherichia coli 0157:H7 infections associated with eating a nationally distributed commercial brand of frozen ground beef patties and burgers—Colorado, 1997. Prevention, 0. f. D. C. a. 1995. Escherichia coli 0157:H7 outbreak linked to commercially distributed dry-cured salami—Washington and California, 1994. MMWR Morb Mortal Wkly Rep 44:157-60. Prevention, 0. f. D. C. a. 2006. Ongoing multistate outbreak of Escherichia coli serotype O157:H7 infections associated with consumption of fresh spinach—United States, September 2006. MMWR Morb Mortal Wkly Rep 55:1045—6. Prevention, 0. f. D. C. a. 1993. Update: multistate outbreak of Escherichia coli O157:H7 infections from hamburgers—westem United States, 1992-1993. MMWR Morb Mortal Wkly Rep 42:258-63. 176 119. 120. 121. 122. 123. 124. 125. 126. 127. 128. 129. Price, S. B., J. C. Wright, F. J. DeGraves, M.-P. Castanie-Comet, and J. W. Foster. 2004. Acid Resistance Systems Required for Survival of Escherichia coli O157:H7 in the Bovine Gastrointestinal Tract and in Apple Cider Are Different. Appl. Environ. Microbiol. 70:4792-4799. Rangel, J. M., P. H. Sparling, C. Crowe, P. M. Griffin, and D. L. Swerdlow. 2005. Epidemiology of Escherichia coli O157:H7 outbreaks, United States, 1982-2002. Emerg Infect Dis 11:603-9. Reid, S. D., C. J. Herbelin, A. C. Bumbaugh, R. K. Selander, and T. S. Whittam. 2000. Parallel evolution of virulence in pathogenic Escherichia coli. Nature 406:64-7. Reitzer, L., and B. L. Schneider. 2001. Metabolic context and possible physiological themes of sigma(54)—dependent genes in Escherichia coli. Microbiol Mol Biol Rev 65:422-44. Richard, H., and J. W. Foster. 2004. Escherichia coli Glutamate- and Arginine-Dependent Acid Resistance Systems Increase lntemal pH and Reverse Transmembrane Potential. J. Bacteriol. 186:6032-6041. Robinson, 0. M., J. F. Sinclair, M. J. Smith, and A. D. O'Brien. 2006. Shiga toxin of enterohemorrhagic Escherichia coli type O157:H7 promotes intestinal colonization. Proc Natl Acad Sci U S A 103:9667-72. Rozen, S., and H. Skaletsky. 2000. Primer3 on the WWW for general users and for biologist programmers. Methods Mol Biol 132:365-86. Ruiz, 0., L. M. McMurry, and S. B. Levy. 2008. Role of the Multidrug Resistance Regulator MarA in Global Regulation of the hdeAB Acid Resistance Operon in Escherichia coli. J. Bacteriol. 190:1290-1297. Russell, J. B., F. Diez-Gonzalez, and G. N. Jarvis. 2000. Potential effect of cattle diets on the transmission of pathogenic Escherichia coli to humans. Microbes Infect 2:45-53. Safdar, N., A. Said, R. E. Gangnon, and D. G. Maki. 2002. Risk of hemolytic uremic syndrome after antibiotic treatment of Escherichia coli 0157:H7 enteritis: a meta-analysis. Jama 288:996-1001. Saitoh, T., S. lyoda, S. Yamamoto, Y. Lu, K. Shimuta, M. Ohnishi, J. Terajima, and H. Watanabe. 2008. Transcription of the ehx enterohemolysin gene is positively regulated by GrlA, a global regulator encoded within the locus of enterocyte effacement in enterohemorrhagic Escherichia coli. J Bacteriol 190:4822-30. 177 130. 131. 132. 133. 134. 135. 136. 137. 138. 139. Saitou, N., and M. Nei. 1987. The neighbor-joining method: a new method for reconstructing phylogenetic trees. Mol Biol Evol 4:406-25. Sayed, A. K., and J. W. Foster. 2009. A 750 bp sensory integration region directs global control of the Escherichia coli GadE acid resistance regulator. Mol Microbiol 71 :1435-50. Sayed, A. K., C. Odom, and J. W. Foster. 2007. The Escherichia coli AraC-family regulators GadX and GadW activate gadE, the central activator of glutamate-dependent acid resistance. Microbiology 153:2584- 2592. Schmidt, H., B. Henkel, and H. Karch. 1997. A gene cluster closely related to type II secretion pathway operons of Gram-negative bacteria is located on the large plasmid of enterohemorrhagic Escherichia coli O157 strains. FEMS Microbiol Lett. 148:265-272. Schmittgen, T. D., and K. J. Livak. 2008. Analyzing real-time PCR data by the comparative C(T) method. Nat Protoc 3:1101-8. Scott, D. R., E. A. Marcus, D. L. Weeks, A. Lee, K. Melchers, and G. Sachs. 2000. Expression of the Helicobacter pylori urel gene is required for acidic pH activation of cytoplasmic urease. Infect Immun 68:470-7. Serra-Moreno, R., J. Jofre, and M. Muniesa. 2007. Insertion site occupancy by stx2 bacteriophages depends on the locus availability of the host strain chromosome. J Bacteriol 189:6645-54. Shao, 0., Q. Zhang, W. Tang, W. Qu, Y. Zhou, Y. Sun, H. Yu, and J. Jia. 2008. The changes of proteomes components of Helicobacter pylori in response to acid stress without urea. J Microbiol 46:331-7. Shin, S., M.-P. Castanie-Comet, J. W. Foster, J. A. Crawford, 0. Brinkley, and J. B. Kaper. 2001. An activator of glutamate decarboxylase genes regulates the expression of enteropathogenic Escherichia coli virulence genes through control of the plasmid-encoded regulator, Per. Molecular Microbiology 41 :1 133-1150. Shin, S., M.-P. Castanie-Comet, J. W. Foster, J. A. Crawford, 0. Brinkley, and J. B. Kaper. 2001. An activator of glutamate decarboxylase genes regulates the expression of enteropathogenic Escherichia coli vimlence genes through control of the plasmid-encoded regulator, Per. Mol. Microbiol. 41:1133-1150. 178 140. 141. 142. 143. 144. 145. 146. 147. 148. 149. Smith, D. K., T. Kassam, B. Singh, and J. F. Elliott. 1992. Escherichia coli has two homologous glutamate decarboxylase genes that map to distinct loci. J. Bacteriol. 174:5820-5826. Sperandio, V. 2004. Striking a balance: inter-kingdom cell-to-cell signaling, friendship or war’? Trends lmmunol 25:505-7. Sperandio, V., A. G. Torres, B. Jarvis, J. P. Nataro, and J. B. Kaper. 2003. Bacteria-host communication: the language of hormones. Proc Natl Acad Sci U S A 100:8951-6. Stevens, M. P., A. J. Roe, I. Vlisidou, P. M. van Diemen, R. M. La Ragione, A. Best, M. J. Woodward, D. L. Gally, and T. S. Wallis. 2004. Mutation of toxB and a Truncated Version of the efa-1 Gene in Escherichia coli O157:H7 Influences the Expression and Secretion of Locus of Enterocyte Effacement-Encoded Proteins but not Intestinal Colonization in Calves or Sheep. Infect. Immun. 72:5402-5411. Stewart, N., J. Feng, X. Liu, D. Chaudhuri, J. W. Foster, M. Drolet, and Y.-C. Tse-Dinh. 2005. Loss of topoisomerase I function affects the RpoS- dependent and GAD systems of acid resistance in Escherichia coli. Microbiology 1 51 :2783—2791 . Stim-Hemdon, K. P., T. M. Flores, and G. N. Bennett. 1996. Molecular characterization of adiY, a regulatory gene which affects expression of the biodegradative acid-induced arginine decarboxylase gene (adiA) of Escherichia coli. Microbiology 142 ( Pt 5):1311-20. Storey, J. D., and R. Tibshlrani. 2003. Statistical significance for genomewide studies. Proceedings of the National Academy of Sciences 100:9440-9445. Strachan, N. J., D. R. Fenlon, and l. D. Ogden. 2001. Modelling the vector pathway and infection of humans in an environmental outbreak of Escherichia coli O157. FEMS Microbiol Lett 203:69-73. Subramanian, A., P. Tamayo, V. K. Mootha, S. Mukherjee, B. L. Ebert, M. A. Gillette, A. Paulovich, S. L. Pomeroy, T. R. Golub, E. S. Lander, and J. P. Mesirov. 2005. From the Cover: Gene set enrichment analysis: A knowledge-based approach for interpreting genome-wide expression profiles. Proceedings of the National Academy of Sciences 102:15545- 1 5550. Sung, L. M., M. P. Jackson, A. D. O'Brien, and R. K. Holmes. 1990. Transcription of the Shiga-like toxin type II and Shiga-like toxin type II variant operons of Escherichia coli. J Bacteriol 172:6386-95. 179 150. 151. 152. 153. 154. 155. 156. 157. 158. 159. Takada, A., G. Umitsuki, K. Nagai, and M. Wachi. 2007. RNase E is required for induction of the glutamate-dependent acid resistance system in Escherichia coli. Biosci Biotechnol Biochem 71 :1 58-64. Tamura, K., J. Dudley, M. Net, and S. Kumar. 2007. MEGA4: Molecular Evolutionary Genetics Analysis (MEGA) software version 4.0. Mol Biol Evol 24:1 596-9. Taneike, l., H. M. Zhang, N. Wakisaka-Saito, and T. Yamamoto. 2002. Enterohemolysin operon of Shiga toxin- -producing Escherichia coli: a virulence function of inflammatory cytokine production from human monocytes. FEBS Lett 524: 219-24. Tarr, P. l., C. A. Gordon, and W. L. Chandler. 2005. Shiga-toxin- producing Escherichia coli and haemolytic uraemic syndrome. Lancet 365:1073—86. Tatsuno, l., M. Horie, H. Abe, T. Miki, K. Makino, H. Shinagawa, H. Taguchi, S. Kamiya, T. Hayashi, and C. Sasakawa. 2001. toxB gene on p0157 of Enterohemorrhagic Escherichia coli O157:H7 is required for full epithelial cell adherence phenotype. Infect. Immun. 69:6660-6669. Tatsuno, l., K. Nagano, K. Taguchi, L. Rong, H. Mori, and C. Sasakawa. 2003. Increased adherence to Caco-2 cells caused by disruption of the yhiE and yhiF genes in enterohemorrhagic Escherichia coli O157:H7. Infect Immun 71:2598-606. Tatsuno, l., K. Nagano, K. Taguchi, L. Rong, H. Mori, and C. Sasakawa. 2003. Increased Adherence to Caco-2 Cells Caused by Disruption of the yhiE and yhiF Genes in Enterohemorrhagic Escherichia coli O157:H7. Infect. Immun. 71:2598-2606. Tobe, T., H. Ando, H. lshikawa, H. Abe, K. Tashiro, T. Hayashi, S. Kuhara, and N. Sugimoto. 2005. Dual regulatory pathways integrating the RcsC-RcsD-RcsB signalling system control enterohaemorrhagic Escherichia coli pathogenicity. Molecular Microbiology 58:320-333. Tobe, T., S. A. Beatson, H. Taniguchi, H. Abe, C. M. Bailey, A. Fivian, R. Younis, S. Matthews, 0. Marches, G. Frankel, T. Hayashi, and M. J. Pallen. 2006. An extensive repertoire of type III secretion effectors in Escherichia coli 0157 and the role of Iambdoid phages in their dissemination. Proceedings of the National Academy of Sciences 103:14941-14946. Tramonti, A., M. De Canio, I. Delany, V. Scarlato, and D. De Biase. 2006. Mechanisms of Transcription Activation Exerted by GadX and 180 160. 161. 162. 163. 164. 165. 166. 167. 168. 169. GadW at the gadA and gadBC Gene Promoters of the Glutamate-Based Acid Resistance System in Escherichia coli. J. Bacteriol. 188:8118-8127. Tsai, H. M., W. L. Chandler, R. Sarode, R. Hoffman, S. Jelacic, R. L. Habeeb, S. L. Watkins, C. S. Wong, G. D. Williams, and P. I. Tarr. 2001. von Willebrand factor and von Willebrand factor-cleaving metalloprotease activity in Escherichia coli O157:H7-associated hemolytic uremic syndrome. Pediatr Res 49:653-9. Tucker, D. L., N. Tucker, and T. Conway. 2002. Gene Expression Profiling of the pH Response in Escherichia coli. J. Bacteriol. 184:6551- 6558. Tucker, D. L., N. Tucker, Z. Ma, J. W. Foster, R. L. Miranda, P. S. Cohen, and T. Conway. 2003. Genes of the GadX-GadW Regulon in Escherichia coli. J. Bacteriol. 185:3190—3201. Tuttle, J., T. Gomez, M. P. Doyle, J. G. Wells, T. Zhao, R. V. Tauxe, and P. M. Griffin. 1999. Lessons from a large outbreak of Escherichia coli O157:H7 infections: insights into the infectious dose and method of widespread contamination of hamburger patties. Epidemiol Infect 122:185-92. Umanski, T., l. Rosenshine, and D. Friedberg. 2002. Therrnoregulated expression of virulence genes in enteropathogenic Escherichia coli. Microbiology 148:2735-2744.’ Van Winden, S. 0., K. E. Muller, R. Kuiper, and J. P. Noordhuizen. 2002. Studies on the pH value of abomasal contents in dairy cows during the first 3 weeks after calving. J Vet Med A Physiol Pathol Clin Med 49:157-60. Wagner, W., M. Vogel, and W. Goebel. 1983. Transport of hemolysin across the outer membrane of Escherichia coli requires two functions. J Bacteriol 1 54:200-10. Waterman, S. R., and P. L. 0. Small. 2003. Transcriptional Expression of Escherichia coli Glutamate-Dependent Acid Resistance Genes gadA and gadBC in an hns rpoS Mutant. J. Bacteriol. 185:4644-4647. Weber, A., S. A. Kogl, and K. Jung. 2006. Time-dependent proteome alterations under osmotic stress during aerobic and anaerobic growth in Escherichia coli. J. Bacteriol. 1 88:71 65-71 75. Weber, H., T. Polen, J. Heuveling, V. F. Wendisch, and R. Hengge. 2005. Genome-Wide Analysis of the General Stress Response Network in 181 170. 171. 172. 173. 174. 175. 176. Escherichia coli: {sigma}S-Dependent Genes, Promoters, and Sigma Factor Selectivity. J. Bacteriol. 1 87:1591-1603. Wick, L. M., W. Qi, D. W. Lacher, and T. S. Whittam. 2005. Evolution of genomic content in the stepwise emergence of Escherichia coli O157:H7. J Bacteriol 187:1 783-91. Wick, L. M., W. Qi, D. W. Lacher, and T. S. Whittam. 2005. Evolution of Genomic Content in the Stepwise Emergence of Escherichia coli O157:H7. J. Bacteriol. 187:1783—1791. Wong, 0. S., S. Jelacic, R. L. Habeeb, S. L. Watkins, and P. I. Tarr. 2000. The risk of the hemolytic-uremic syndrome after antibiotic treatment of Escherichia coli O157:H7 infections. N Engl J Med 342:1930-6. Yamanaka, K., W. Zhang, 5. Crooke, Y. H. Wang, and M. Inouye. 2001. CspD, a novel DNA replication inhibitor induced during the stationary phase in Escherichia coli. Mol Microbiol 39:1572-84. Yuen, T., E. Wunnbach, R. L. Pfeffer, B. J. Ebersole, and S. C. Sealfon. 2002. Accuracy and calibration of commercial Oligonucleotide and custom cDNA microarrays. Nucl. Acids Res. 30:e48. Yuryev, A. 2007. PCR primer design using statistical modeling. Methods Mol Biol 402:93-104. Zwir, I., D. Shin, A. Kato, K. Nishino, T. Latifi, F. Solomon, J. M. Hare, H. Huang, and E. A. Groisman. 2005. Dissecting the PhoP regulatory network of Escherichia coli and Salmonella enterica. Proc Natl Acad Sci U S A 102:2862-7. 182