‘n n-aa“ a“... L J“ t MJ,’ 53;. Jim“ ' 2:3? :E Tax-I :rzr A ‘31! ‘. “a .... a.» _4.. fig . a.“ $2“, ”’35. . ‘~ .q 1‘ r31. 1-3:; u I 4m _ W F: . ,. ram 8 200% This is to certify that the dissertation entitled SURVIVAL OF ENTEROHEMORRHAGIC ESCHERICHIA COLI IN FOOD AND THE GASTRIC ENVIRONMENT presented by Teresa Marie Bergholz has been accepted towards fulfillment of the requirements for the Ph.D. degree in Food Science and Human Nutrition WWW Major Professor’s Signature Ar“! lb, 2,007- Date MSU is an affirmative-action, equal-opportunity employer —-v---.—.—.-.—.-,-.-.-a-u-.-u--u-.—.-a—.-t- — g.-.-._.-,-._.-._.-.-.-._.-.-.-.-._‘_.-.-— -.—------.--.-.—-.- LIBRARY Michigan State University 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 6/07 p:/ClRC/DaleDue.indd-p.1 SUR IVi SURVIVAL OF ENTEROHEMORRHAGIC ESCHERICHIA COLI IN FOOD AND THE GASTRIC ENVIRONMENT By Teresa Marie Bergholz A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Food Science and Human Nutrition 2007 SUR‘» been sh eXpecte may alsc account 1 acidic for; strains of fOOCIbOme media hm 0011' 0157: l9presemi IOOd main °bSeWe th Commons IINo differe of three kn ABSTRACT SURVIVAL OF ENTEROHEMORRHAGIC ESCHERICHIA COLI IN FOOD AND THE GASTRIC ENVIRONMENT By Teresa Marie Bergholz Pathogenic strains of Escherichia coli, such as E. coli O157zH7, have been shown to be unusually acid resistant, surviving acidic conditions typically expected to inactivate bacteria. Adaptations for survival in low pH environments may also contribute to risk of foodborne illness. These adaptations could account for the low infectious dose (10 to 100 cells) and ability to contaminate acidic foods, such as apple cider. Recently, Enterohemorrhagic E. coli (EHEC) strains of serotypes other than O157:H7 have been implicated in outbreaks of foodborne disease. Acid resistance (AR) studies in acidic foods and in defined media have revealed considerable variation in survival rates among strains of E. coli O157:H7. The basis of this research is to: 1) measure AR ability of strains representing 3 major clonal groups of EHEC, 2) assess the effect of incubation in food matrices on subsequent AR in a simulated gastric environment, and 3) observe the genome-wide transcriptional response of E. coli O157:H7 to stressful conditions. To address these goals, phenotypic AR assays were conducted in two different environments: a minimal acidic environment designed to isolate one of three known AR mechanisms, and a complex acidic environment that simulates gastric conditions. Whole genome transcription profiling was conducted to assess the response of O157zH7 to growth transitions in a minimal mediL owya envro ackfic asses: EHEC ofEHE medium. The extent to which the known AR mechanisms of E. coli contribute to overall survival and the AR response in naturally acidic foods and the gastric environments was then determined. Variation in survival rates under distinct acidic conditions will provide data for the development of more accurate risk assessment models. Observed differences in survival among the clonal groups of EHEC in this study will allow further studies to characterize the genetic response of EHEC induced by food matrices. Copyright By Teresa Marie Bergholz | wish suppo compe Linz, E Many c asasua Abu-Ali Mannin Sankan My fami Large, 3 ACKNOWLEDGEMENTS I wish to thank my mentor, Thomas S. Whittam, for his valuable advice and support, and for providing the resources and opportunities for me to acquire competence in a broad range of topics. I thank my committee members, John Linz, Elliot Ryser, and Vincent Young for their suggestions and guidance. Many of my fellow lab mates have provided suggestions, support, and technical assistance that were critical to the completion of this research. I thank Galeb Abu-Ali, Katie Hyma, Sivapriya Kailasan, Sara Kienzle, David Lacher, Shannon Manning, Adam Nelson, Lindsey Ouellette, Megan Parsons, Weihong Qi, Janani Sankaralingam, Cheryl Tarr, Seth Walk, and Lukas Wick. My family has provided me with amazing support during my graduate career. I wish to thank my parents, Charles and Johann Large, my sisters, Kim and Kellie Large, and my husband, Peter Bergholz, for their love and encouragement. LIST C LIST C Chapte lnfecl Acid l Purpc Chapte coli Clor Sumr lnfiod fikfier F I Resul \ C F Discus ACKHQI Chapter Qertes (r SUITIIT Introd. Materi G Result P G Chapter COII in a ‘ SUmm IHIrOGL atern Al O ReSUHg TABLE OF CONTENTS LIST OF TABLES .............................................................................................. viii LIST OF FIGURES ............................................................................................ ix Chapter 1. Literature Review ............................................................................ 1 Infectious dose and gastric acidity ................................................................. 9 Acid resistance of Escherichia coli O157:H7 .................................................. 10 Purpose .......................................................................................................... 13 Chapter 2. Variation in acid resistance among Shiga toxin-producing Escherichia coli clones ......................................................................................................... 1 5 Summary ........................................................................................................ 16 Introduction .................................................................................................... 17 Materials and Methods ................................................................................... 19 RpoS status ............................................................................................ 19 AR mechanisms assays ......................................................................... 22 Results ........................................................................................................... 26 Variation in AR among E. coli O157:H7 ................................................. 30 Comparison of STEC clones .................................................................. 31 RpoS+ vs. RpoS- isolates ...................................................................... 35 Discussion ....................................................................................................... 39 Acknowledgements ......................................................................................... 45 Chapter 3. Past gene conversions between duplicated glutamate decarboxylase genes (gadAB) in pathogenic Escherichia coli .................................................. 46 Summary ........................................................................................................ 47 Introduction .................................................................................................... 48 Materials and Methods ................................................................................... 52 GAD sequencing .................................................................................... 52 Results ........................................................................................................... 58 Paralogous gene conversions ................................................................ 60 Gene phylogenies .................................................................................. 64 Promoter region ...................................................................................... 68 Discussion ....................................................................................................... 71 Acknowledgements ......................................................................................... 74 Chapter 4. Variation in acid resistance among Enterohemorrhagic Escherichia coli in a simulated gastric environment ............................................................. 75 Summary ........................................................................................................ 76 Introduction .................................................................................................... 77 Materials and Methods ................................................................................... 80 AR assays in the M88 ............................................................................ 82 Quantitative PCR .................................................................................... 85 Results ........................................................................................................... 86 vi l Disc Ackr Chapt growtl Sunr Intro Mate ReSL Discu Acknc Chapte Survival in the M88 ................................................................................ 90 Quantifying injured cells in the M88 ....................................................... 94 GAD transcription levels ......................................................................... 96 Discussion ...................................................................................................... 100 Acknowledgements ........................................................................................ 106 Chapter 5. Global gene expression patterns in Escherichia coli O157:H7 during growth transitions in glucose minimal medium ................................................. 107 Summary ....................................................................................................... 108 Introduction ................................................................................................... 1 10 Materials and Methods .................................................................................. 114 RNA isolation ......................................................................................... 1 14 cDNA synthesis and hybridization ......................................................... 116 Data analysis ......................................................................................... 118 Results .......................................................................................................... 120 Significant changes in expression in exponential phase ....................... 122 Significant changes in expression in stationary phase .......................... 128 Expression of LEE island genes ............................................................ 136 Expression of AR fitness region genes .................................................. 145 Discussion ...................................................................................................... 151 Acknowledgements ........................................................................................ 160 Chapter 6. Summary and Synthesis ................................................................. 161 Future considerations .................................................................................... 165 References ....................................................................................................... 1 67 vii LIST OF FIGURES Figures in this dissertation are presented in color Figure 1.1. Characteristics and relationship of EHEC and STEC ..................... 4 Figure 2.1. Diagram of assay conditions for acid resistance mechanisms ....... 21 Figure 2.2. Phylogenetic tree of STEC clonal groups ...................................... 29 Figure 2.3. Variation in survival rates of STEC for each acid resistance mechanism ....................................................................................................... 32 Figure 2.4. Box plots of survival rates of STEC groups for each acid resistance mechanism ....................................................................................................... 34 Figure 2.5. Box plots of survival rates for RpoS+ and RpoS- STEC ................ 37 Figure 3.1. Location of gadA and gadB on the K-12 MG1655 chromosome... 51 Figure 3.2. Putative recombination events in gadA and gadB .......................... 59 Figure 3.3. Alignment of variable nucleotide sites in gadA ............................... 62 Figure 3.4. Alignment of variable nucleotide sites in gadB ............................... 63 Figure 3.5. Neighbor-joining phylogenies for gadA and gadB codons 2-81 ..... 66 Figure 3.6. Neighbor-joining phylogeny for gadA and gadB codons 82-349 67 Figure 3.7. Promoter region consensus sequences for gadA and gadB .......... 70 Figure 4.1. Diagram of assay conditions for survival in the model stomach system ......................................................................................................................... 83 Figure 4.2. Box plots of survival rates in the model stomach system ............... 91 Figure 4.3. Box plots of survival rates in the model stomach before and after incubation in apple juice ................................................................................... 93 Figure 4.4. Box plots of survival rates in the model stomach after incubation in apple juice at different pH ................................................................................ 97 Figure 4.5. Box plots of relative transcript levels of gadA and gadB ................ 99 Figure 5.1. Average density of E. coli O157:H7 Sakai in MOPS minimal medium over time .......................................................................................................... 1 15 viii Figuri Figure Figure signifii Figure of E. c Figure Figure Figure Figure Figure Figure 5.2. Signal intensity plots ...................................................................... 121 Figure 5.3. QT clusters for significantly differentially expressed ORFs ............ 123 Figure 5.4. Proportion of backbone and O157-specific ORFs that change significantly over specific time intervals ............................................................ 124 Figure 5.5. Residual dissolved oxygen tension in the MOPS media during growth of E. coli O157:H7 Sakai .................................................................................. 129 Figure 5.6. Heatmap of expression of LEE genes over time ............................ 142 Figure 5.7. Heatmap of expression of TTS effector genes over time ............... 143 Figure 5.8. Heatmap of expression of Acid Fitness Region genes over time... 146 Figure 5.9. Expression plots of the genes encoding Shiga toxin over time ...... 148 Figure 5.10. Expression plots of genes from the TAI over time ........................ 149 ix LIST OF TABLES Table 1.1 EHEC outbreaks since 1982 ............................................................ 7 Table 2.1 STEC strains used in the acid resistance mechanisms study .......... 25 Table 2.2 Nucleotide diversity of STEC strains ................................................ 29 Table 3.1 E. coli strains used for gadA and gadB sequencing ......................... 53 Table 2.2 Primer sequences used for gadA and gadB sequencing ................. 55 Table 3.3 Significant fragments found using Sawyers test .............................. 61 Table 4.1 STEC strains used in the model stomach system study ................... 81 Table 4.2 QPCR primer and probe sequences for gadA, gadB and mdh ......... 86 Table 4.3 Log difference of injured and non-injured cells in the model stomach system .............................................................................................................. 95 Table 5.1 ORFs with significant changes in expression exponential phase ..... 125 Table 5.2 ORFs with significant changes in expression during the transition to stationary phase ............................................................................................... _ 130 Table 5.3 ORFs with significant changes in expression from transition to eariy stationary phase ............................................................................................... 134 Table 5.4 ORFs with transient significant changes in expression .................... 138 Table 5.5 ORFs with significant changes in expression in early stationary phase ......................................................................................................................... 140 Table 5.6 Significant changes in expression of Type III secretion effectors ..... 144 Chapter 1 Literature Review habit route some disea. enterir Sa/mc animal contac rel>resis f00d Wil farm, f0 It 5900 de foodborr COmDIiCa abOFIIOn, virujen Ce and the IV As pan of the eCOIOQI‘Cth adVerSe SEI Enteric bacteria, those that are adapted to and colonize the intestinal habitat of humans and other animals, are typically transmitted via the fecal—oral route from host to host. Although many enteric bacteria are harmless to the host, some strains have acquired or evolved pathogenic mechanisms and can cause disease in a wide range of hosts or be limited to a specific host species. Many enteric bacteria that can be pathogenic to humans, such as Escherichia coli and Salmonella enterica, are generally commensal organisms in animal hosts. Such animal hosts serve as the major reservoir for these pathogen populations; direct contact with these animals or their fecal material through food and water represent the primary vehicles for transmission to humans. Contamination of food with enteric pathogens can occur at many stages; during production on the farm, food processing, and food handling (270) It is estimated that foodborne disease accounts for 76 million illnesses and 5,000 deaths in the United States each year (173). Disease attributed to foodborne pathogens ranges in severity from mild gastroenteritis to severe complications, such as hemolytic uremic syndrome (HUS) and spontaneous abortion, or death. Both disease severity and the risk of infection depend on the virulence and the numbers of the pathogen ingested, susceptibility of the host, and the type of food vehicle. As a bacterial species, E. coli is normally a harmless organism comprising part of the intestinal flora in warm-blooded animals. These bacteria are ecologically diverse, and have evolved the ability to adapt to and survive in many adverse secondary environments, such as the farm environment and in natural watt into . envh an BS wfihEF infectic on the . infectior foodbon Categori; clinical s: Pathova r I0 DTOduc. capabie 01 water sources. This ability to survive environmental stresses is critical for reentry into a host and further propagation. Survival and persistence in external environments and colonization of new hosts in the face of host defenses is also an essential characteristic for transmission of pathogenic E. coli (89). STEC and EHEC. There are a variety of diseases that are associated with E. coli including many extra-intestinal infections such as urinary tract infections and newborn meningitis (267). The primary focus of this research is on the class of pathogenic E. coli that cause intestinal diseases, most notably infectious diarrhea and its complications that have been linked to outbreaks of foodborne illness (79, 267). These diarrheagenic E. coli strains have been categorized into 5 distinct mechanisms or pathovars based on virulence factors, clinical symptoms, mechanisms of pathogenicity, and serotype (79). One pathovar is Shiga toxin-producing E. coli (STEC), which is classified by the ability to produce Shiga toxin(s). Enterohemorrhagic E. coli (EHEC), the pathovar capable of producing the most severe disease, has three characteristic traits; the ability to produce Shiga toxin, to attach and efface epithelial cells, which is encoded by the Locus of Enterocyte Effacement (LEE) pathogenicity island (169), and the presence of the pO157 plasmid (149) (Fig. 1.1). EHEC strains can be further divided into clonal groups based on genetic similarity. These clonal groups contain strains that have similar genetic and phenotypic features, but are not identical, and result from sharing a common ancestor. One feature, for example, is the presence of specific surface antigens, which dictate the different E. coli serotypes, such as O157:H7 and over 200 additional serotypes Fi9Ure 1, SerOIYDes the bOVing Cases of r prodUCtion Cells ellCoc mOst DI’BVa hemorrhaer ozsand 0‘ Figure 1.1. Characteristics and relationship of STEC and EHEC. More than 200 serotypes of E. coli that can produce Shiga toxin (STEC) have been isolated from the bovine reservoir, but only a subset of these serotypes are associated from cases of human disease. A subset of STEC is EHEC, which is classified by the production of Shiga toxin, as well as the ability to attach and efface epithelial cells encoded by the LEE island and the presence of the p0157 plasmid. The most prevalent EHEC serotypes that cause cases of human illness, including hemorrhagic colitis and HUS, are 0157, 0121, 0103, and the closely related 026 and 0111 serotypes. (nor the a (18, 1 with r. 0121; group: consis seroty; Multi-Io I several (non-0157) that also produce Shiga toxins. Many other serotypes of E. coli have the ability to produce Shiga toxin and elicit disease similar to that of O157:H7 (18, 19, 32). Diarrheal disease caused by non-0157 EHEC has been increasing, with recent outbreaks traced to strains of serotypes 026:H11, 0111:H8, and 0121 :H19 (31, 44, 167, 268). These serotypes mark genetically distinct clonal groups, with EHEC 1 consisting of O157:H7 and its relatives and EHEC 2 consisting of serotypes 026:H11, 0111:H8, and relatives (208). EHEC of serotype 0121:H19 belong to a lineage distinct from both 1 and 2 based on Multi—locus sequence typing (MLST) (241 ), and is referred to as EHEC 3. E. coli O157:H7 emerged as a cause of epidemic bacterial diarrhea after several outbreaks of foobome illness were first linked to human disease in 1982 (211). In the US. alone, E. coli O157:H7 is estimated to account for 75,000 illnesses and 600 fatalities per year (173). Infection by E. coli O157:H7 can result in hemorrhagic colitis and serious complications such as HUS (103), which can result in kidney failure and death. In 1999, it was estimated that E. coli serotype O157:H7 is most prevalent in the US, causing 75% of EHEC infections in humans (173). Since then, studies on EHEC prevalence in sporadic cases of human illness have found different rates of prevalence for different locations around the US. A study of diarrheal cases in children in Seattle reported that 72% of STEC infections were caused by O157:H7 (139). In Nebraska, however, non-0157 EHEC were as prevalent in diarrheal cases as O157:H7 (85). A two- year study of diarrheal cases in Montana found 62% of sporadic EHEC infections were due to non-0157 STEC (124). Whether the increasing prevalence of non- O1: incr an II 015 Foor cons. 211) juice 1 outbre with E 0157 EHEC is a result of the development of improved detection methods, or the increasing frequency and geographic spread of these clones is unknown. While an increase in sporadic cases of non-0157 EHEC has been observed, E. coli O157:H7 is still the leading cause of outbreaks in North America. Foodbome transmission. Outbreaks of E. coli O157:H7 have occurred as a consequence of contamination of various foods, including ground beef (46, 56, 211), lettuce (114), alfalfa sprouts (28), raw milk (45), apple cider (16, 115), apple juice (65), dry cured salami (47), and spinach (50). Because initial EHEC outbreaks in the North America were attributed to contamination of ground beef with E. coli O157:H7, most studies have only focused on prevalence of this serotype in beef and dairy cattle. Early studies on prevalence of O157:H7 in dairy and beef cattle estimated that less than 10% of cattle were positive (94). The development of more sensitive culture and isolation techniques has led to the finding that O157:H7 is more prevalent in cattle than previously thought; around 10-50% of cattle are estimated to be positive for O157:H7 at any given time (83, 120, 217). However, in the past decade, several outbreaks of human illness have been attributed to non-0157 EHEC, which led researchers to begin studying prevalence of non-0157 EHEC as well as O157:H7 in cattle. One recent study found that 17/82 (21%) of cull cows were positive for EHEC, and of the positives, 76% were non-0157 EHEC serotypes (121). A second study reported that as many as 40% of beef carcasses were positive for EHEC O157:H7 and 40-65% were positive for non-0157 EHEC, depending on the season (1 1). AAA/\f‘ “AAA!“ nn nnnnn Table 1.1. Confirmed outbreaks of EHEC since 1982. Date Serotype Locale Vehicle Number Fatalities Ref. affected 2006 O157:H7 Multi-state, U.S. Spinach 204 4 (50) 2005 O157:H7 Washington and Raw milk 18 - (45) Oregon 2002 026:H- Austria Raw milk 2 - (5) 2002 O157:H7 Colorado Ground beef 28 - (49) 2002 O157:H7 Japan Beef 28 - (251) 2001 O157:H7 Ohio Airborne 23 - (257) (county fair) 2000 01 571NM California Water (lake) 4 - (84) 2000 O157:H7 Ontario Water 2300 7 (1 19) (drinking) 2000 026:H1 1 Germany Beef 1 1 - (268) 2000 O157:H7 Pennsylvania and Farm visits 56 - (55) Washington 1999 O157:H7 California and Beef tacos 13 - (123) Nevada 1999 O157:H7 British Columbia Dry fermented 135 - (161) salami 1999 O157:H7 Scotland Water 6 - (153) 1999 0121:H19 Connecticut Water (lake) 1 1 - (167) 1999 O157:H7 England Pasteurized 88 - (98) milk 1999 O1 1 1 :H8 Texas Ice 58 - (44) (31 ) 1999 O157:H7 New York Well water 128 - (52) (27) 1998 O157:H7 Wyoming Water 71 - (192) 1998 O157:H7 Wisconsin Cheese curd 55 - (53) 1997 O157:H7 Finland Water (lake) 14 - (196) 1997 O157:H7 Connecticut Apple cider 66 (54) 1997 O157:H7 Michigan and Alfalfa sprouts 85 - (28) Virginia 1997 026:H1 1 Japan unknown 32 - (116) 1997 O157:H7 Japan Radish sprouts 47 - (261) 1997 O157:H7 Colorado Hamburger 14 - (46) 1996 O157:H7 Japan School lunches 8004 5 (174) 1996 O157:H7 Western US Apple juice 28 - (65) (unpasteurized) 1996 O157:H7 Connecticut Apple cider 14 - (1 15) 1996 O157:H7 Illinois and Lettuce 49 - (114) Connecticut 1995 01 1 1 :H8 Australia Sausage 23 (43) 1994 01043H21 Montana Milk 18 - (51) 1994 O157:H7 California and Dry cured 23 - (47) Washington salami 1992 O157:H7 Western US Hamburger 477 3 (56) 1992 O157:H7 Scotland Pool 5 - (29) 1992 01 1 1 Italy Unknown 9 1 (39) 1991 O157:H7 Massachusetts Apple cider 16 - (16) 1991 O157:H7 Oregon Lake 21 - (135) Table Date 1991 1990 1990 1985 C 1982 C Table 1.1, continued Date Serotype Locale Vehicle Number Fatalities Ref. affected 1991 O157:H7 England Yogurt 16 - (176) 1990 O157:H7 North Dakota Roast beef 70 - (48) O157:H7 Missouri Water supply 86 (237) 1990 (non- chlorinated) 1985 O157:H7 Canada Unknown 73 19 (40) 1982 O157:H7 Michigan and Hamburger 47 - (211) Oregon minir meas for di. serov infecti reSpor 0f ~ 5( can co i"active been re (198). ‘ likelihoc Ir and natu- IOOds are carbohyd Effect on . after a me C0r18tantr SUWIVaj 0f Infectious dose and gastric acidity. The infectious dose (i.e., the minimum number of cells ingested to induce illness) is difficult to accurately measure, but appears to vary widely across different pathogens and is important for disease pathogenesis. For example, the infectious dose of Salmonella serovars and Vibrio cholerae ranges from 105 to 1010 cells (24, 191 ), whereas the infectious dose of Shigella strains'can be as low as 200 cells (143). Dose response data have indicated that E. coli O157:H7 is highly infectious and a dose of ~ 50 cells was estimated for the Sakai, Japan outbreak (246). One factor that can contribute to such a low infectious dose is the ability of an organism to resist inactivation by acid. The low pH of gastric secretions (pH 1.5 to 3.0) has long been recognized as the first line of defense against foodborne enteric pathogens (198). The ability of enteric bacteria to survive in stomach acid increases the likelihood of colonizing the intestines and causing infection. In addition to inherent acid resistance (AR) of the microbe, the presence and nature of food can increase survival in the stomach. Proteins and fats from foods are thought to provide more protection against acid stress than carbohydrates (262). Food particles in human gastric fluid also have a protective effect on enteric pathogens, as survival rates are greater in gastric fluid taken after a meal compared to those taken before a meal when the pH is held constant (198). Thus, there are at least two primary factors that influence survival of enteric pathogens through the gastric barrier, presence and nature of food particles in the stomach and the inherent AR ability of the pathogen. surviv coli 0 epider 0157: as 10- haslec patth. additior vanafior demons least 2 r (127) TI Oflginall} Strains h sampje ( rat98 arr present!2 Strains is DUFDOSes It for Sun,“ thIQUght AR of E. coli O157:H7. Acid resistance refers to the ability of bacteria to survive at normally lethal pH levels, (i.e., pH less than or equal to 2.5) (89). E. coli O157:H7 is hypothesized to be highly resistant to acid because of epidemiological data that indicate a low infectious dose. The number of viable O157:H7 detected in foods implicated in disease outbreaks suggests that as few as 10-100 viable cells can initiate disease (59). This characteristic of O157:H7 has led to the development of various in vitro assays to study different aspects of pathogen survival in low pH environments. These assays have provided additional evidence of AR in E. coli O157:H7 (8, 100, 155). For example, variations of complex and minimal media and acidulants have been used to demonstrate that E. coli O157:H7 can survive under various acidic conditions; at least 2 h in Luria Broth (LB) pH 2.5 (100, 263) and over 24 h in LB at pH 3.0 (127). The observed high survival rate of several E. coli O157:H7 strains, originally implicated in different outbreaks, led to the conclusion that all O157:H7 strains have an “unusual AR ability” (175). More recent AR studies on a larger sample of O157:H7, however, have revealed substantial variability in survival rates among O157:H7 strains (34, 35). The basis of this inter-strain variability is presently unknown, and quantifying variation in survival rates among pathogenic strains is important to develop accurate predictive models for risk assessment purposes. It is clear that many enteric bacteria have evolved multiple mechanisms for surviving environmental stresses. Acid resistance in E. coli is achieved through three different types of genetic systems; pH inducible AR systems (often 10 base assor prote condi snafio sflafior altern. acid tc induce idenfifi DH-an asSOCie Chang CYCIOpj I0 Drotc has als Charge an9d1 aSSUmr based on amino acid decarboxylation), general stress response systems associated with stationary phase, and systems involved with macromolecular protection and repair (89). These systems function in distinct and overlapping conditions. The glutamate system can be induced by low pH as well as stationary phase, the arginine and oxidative systems are induced by low pH in stationary phase (210). In addition, many AR responses are induced by the alternate sigma factor RpoS in stationary phase (227, 252, 264). The addition of acid to stationary phase cells will result in increased expression of genes already induced by RpoS, as well as those induced by acid. While not attributed to a specific AR mechanism, other proteins have been identified as being involved with the AR of E. coli; these include RpoS (62), the pH- and stationary phase induced chaperone, HdeA (93), the RNA-polymerase associated protein, SspA (104), and the DNA binding protein, Dps (63). Changes to the cell membrane are also important to successful AR; increases in cyclopropane fatty acids are thought to decrease permeability of the membrane to protons (58, 127). The production of colonic acid as part of exopolysaccharide has also been shown to be critical to AR of E. coli by providing a negatively charged mucoid barrier that repels protons (147, 163, 164). These functions linked to AR have been elucidated in laboratory media conditions, and it is assumed that these functions would also occur in naturally acidic environments, such as encountered in agricultural settings, food items and processing, or the stomach. 11 cond icod E.co when 11-14 While the AR of E. coli O157:H7 can be measured under in vitro conditions in the laboratory, it is also important to understand survival in different food matrices, especially those implicated in prior outbreaks such as apple cider. E. coli O157:H7 is capable of persisting in acidic foods for many days, especially when stored at refrigeration temperatures (4°C). E coli O157:H7 can endure for 11-14 days in apple cider at 8°C (277), 5 days in pineapple juice at 22°C (178), and 36 days in commercial mayonnaise at 7°C (265). E. coli O157:H7 can survive for at least 8 days on lettuce stored at 4°C (20). Studies also have indicated that E. coli O157:H7 exhibits increased survival in acidic foods after exposure to mildly acidic conditions (150). Taken together, it is clear that E. coli O157:H7 is well suited to survival in adverse environments. There is lack of substantial comparisons, however, to survival rates of other EHEC groups to assess whether O157:H7 strains are any different from other nonpathogenic E. coli found in the environment. AR of non-0157 EHEC. Although non-0157 serotypes of EHEC appear to have a low infectious dose similar to O157:H7 strains (103), the AR capability of these strains is not well investgated and has not been clearly characterized. One study measuring survival of O157:H7 and other EHEC at pH 4.0 and pH 2.0 found that O157:H7 strains were more sensitive to acid than other EHEC, but a firm conclusion could not be drawn since only a single strain of other EHEC serotypes were tested (172). Because the focus of AR in E. coli has been limited to O157:H7 strains, the potential of non-0157 EHEC to survive in foods and cause disease has not been fully evaluated. 12 partic to sup or diff. Based EHEC relatec transm EHEC, surviva product DathOgg 1 environ role in f (Ii'seasE human ability 0 theabii Comple. glutamE encOde Purpose. To this point, the consensus view is that E. coli O157:H7 is particularly well suited for adverse environments, but there is little direct evidence to support the hypothesis that bacteria of this serotype have special adaptations or different survival properties from typical E. coli found in the environment. Based on the increasing prevalence of non-0157 EHEC, it is clear that non-0157 EHEC also should be assessed for stress resistance properties, especially those related to survival in food matrices, which are key to both colonization and transmission. By characterizing variation among and between clonal groups of EHEC, the research conducted here is important to accurately determine EHEC survival in acidic foods as well as develop effective means to validate food products and processes that rely on acid to inactivate EHEC and other pathogens. The focus of this dissertation is to characterize EHEC survival in adverse environments, with special emphasis on adaptations related to AR because of its role in foodborne transmission. Survival in food alone is not sufficient to result in disease; bacterial strains must also be able to survive passage through the human stomach. The first part of the research described here investigates the ability of specific EHEC clones to survive under low pH conditions; first observing the ability to utilize specific AR mechanisms and second by quantifying survival in complex acidic environments. One of the most important AR mechanisms is the glutamate decarboxlase system, and the key isozymes of the system are encoded by duplicated genes. Chapter three investigates the evolution of these duplicated genes among distinct evolutionary lineages of pathogenic E. coli. 13 Cha prote the b four a of EH focus. expre: expres such a DUI knr Characterization of EHEC survival in foods and identification of the potential protective effects '(eg., acid habituation) stimulated by various food matrices is the beginning step to understanding survival of EHEC the stomach. Chapter four assesses the effects of storage in an acidic food on the subsequent survival of EHEC in a simulated gastric environment. The final section of this dissertation focuses specifically on E. coli O157:H7 to obtain a global view of gene expression during growth transitions, with specific emphasis on changes in expression of 0157-specific ORFs. Understanding how stressful conditions, such as stationary phase, influence global gene expression further contributes to our knowledge of EHEC ecology. 14 Lar CHAPTER 2 Variation in Acid Resistance Among Shiga Toxin-producing Clones of Pathogenic Escherichia coli Large, T. M., S. T. Walk, and T. S. Whittam. 2005. Appl Environ Microbiol 71:2493-500. 15 Path infec evolv amin< gluco; AR m. toxin-p and 6 . were n individt a highl) CPU/m1 I" SUl'Vi\ better p, ComDare hour. Th less than and SIQnit that this or mechanlSr SUMMARY Pathogenic strains of Escherichia coli, such as E. coli O157:H7, have a low infectious dose and an ability to survive in acidic foods. These bacteria have evolved at least 3 distinct mechanisms of acid resistance (AR) including two amino acid decarboxylase dependent systems (arginine and glutamate) and a glucose-catabolite repressed system. We quantified the survival rates for each AR mechanism separately in clinical isolates representing three groups of Shiga toxin-producing E. coli clones (O157:H7, 026:H11/0111zH8, and 0121:H19), and 6 commensal strains from ECOR group A. Members of the STEC clones were not significantly more acid resistant than the commensal strains using any individual AR mechanism. The glutamate system provided the best protection in a highly acidic environment for all groups of isolates (<0.1 log reduction in CFU/mL per hour at pH 2.0). Under these conditions, there was notable variation in survival rates among the 30 O157:H7 strains. The arginine system provided better protection at pH 2.5, with a range of 0.03 to 0.41 log reduction per hour, compared to the oxidative system, with a range of 0.13 to 0.64 log reduction per hour. The average survival rate for the O157:H7 clonal group was significantly less than that of the other STEC clones in the glutamate and arginine systems, and significantly less than the 026/0111 clone in the oxidative system, indicating that this clonal group is not exceptionally acid resistant using these specific mechanisms. 16 vanet eflen sunfiv pathog infiecfii 277) remsua espech (PH < 3 aCkficf. determj L meChan Denods QIUtama. Suggest asthOSe SYStems which Dla INTRODUCTION Escherichia coli are ecologically versatile bacteria that have adapted to a variety of ecological conditions encountered in both animal hosts and the external environment. These organisms have evolved multiple mechanisms to survive under low pH conditions. E. coli O157:H7, a food and waterborne pathogen, has been considered highly acid resistant in nature because of its low infectious dose (103) and ability to survive in acidic foods (175, 212, 265, 276, 277). In the laboratory, E. coli O157:H7 was shown to be exceptionally acid resistant compared to other enteric bacteria, such as Salmonella enterica (100), especially when exposed to mild acid (pH > 4) prior to exposure to strong acid (pH < 3) (8, 150). Outbreaks of human disease caused by E. coli O157:H7 in acidic foods such as apple juice (65) and salami (47) have stimulated interest in determining the mechanisms behind the acid resistance of this organism. Lin et al. developed assays to separate 3 different acid resistance (AR) mechanisms by which E. coli can survive in low pH environments for extended periods of time: two amino acid decarboxylase dependent systems (arginine and glutamate) and a glucose-catabolite repressed system (154, 155). Their findings suggest that these mechanisms promote survival in low pH environments such as those encountered in the stomach and in acidic food products. The AR systems are governed, in part, by the alternate sigma factor RpoS (42, 201 ), which plays a central role in the regulation and expression of many proteins involved in stationary phase and stress response (108, 144, 227). Mutant RpoS alleles exist in natural populations of enterohemorrhagic E. coli (EHEC) and can 17 lead t lead to acid sensitivity (263). Studies of AR have focused on E. coli O157:H7 because it is one of the most common E. coli types associated with outbreaks and sporadic cases of food and waterborne disease. Diarrheal disease caused by Shiga toxin-producing E. coli (STEC) of serotypes other than O157:H7 (non-0157 STEC) has increased worldwide, with recent outbreaks attributable to serotypes 026:H11, 0111:H8, and O121:H19 (44, 167, 268). These serotypes mark three genetically distinct clonal groups of STEC; EHEC clonal group 1 consists of O157:H7 and its nonmotile relatives, EHEC clonal group 2 consists of serotypes 026:H11, 0111:H8, and strains of serotype 0121:H19 represent a distinct clonal group of STEC (208, 241 ). Non-0157 STEC are common in animal reservoirs, capable of causing the same severe disease as E. coli O157:H7, and are thought to also have a low infectious dose (103); however, these pathogens are associated with disease outbreaks much less frequently than E. coli O157:H7. One reason for this disparity in prevalence may result from differences in the inherent acid resistance of the STEC clones and concomitant survival in acidic foods and low infectious dose. Here we address the question, “Do non-0157 STEC possess the same ability to survive using the defined AR mechanisms as E. coli O157:H7?” To address this question, we assembled a collection of clinical isolates representing three clonal groups of STEC (0121:H19, 026/0111 and O157:H7) that have been screened for RpoS expression and assessed strain-to strain variation in survival at low pH using the three different AR mechanisms. 18 MATERIALS AND METHODS Bacterial strains. A total of 66 pathogenic strains including 30 E. coli O157:H7 strains, 18 026:H11 strains, 4 O111:H8 strains, and 14 0121:H19 strains were used in this study (Table 2.1). The strains were originally isolated by different investigators from sporadic cases of diarrheal illness in the course of surveillance studies of STEC in the USA (2, 85, 94, 124, 139). We also obtained several additional STEC from Dr. James Rudrik of the Michigan Department of Community Health, Dr. Lothar Beutin in Germany, and Dr. Roger Johnson in Canada. The presence of the Shiga toxin 1 and 2 genes (stx1 and stx2) is given in Table 1 as reported in the original publication or to us by the sender. In addition, 6 non-toxigenic, commensal strains from the ECOR (189) group A (ECOR strains 1, 2, 3, 4, 7, 10) were used for comparative purposes. Sequence types and Shiga toxin production. Multilocus sequence analysis of 7 housekeeping genes (aspC, prX, fadD, ich, IysP, mdh, and uidA) was used to characterize the multilocus genotype of each strain. Alleles were identified based on sequence comparisons for each gene and distinct allele combinations were designated as sequence types (STs). The sequencing methods and ST database are available at the STEC website (http://www.shiqatox.net/mlst/index.html). The production of Shiga toxins was determined by enzyme immunoassay (EIA) using the Premier EHEC test (Meridian Diagnostics Inc., Cincinnati, Ohio) with the manufacturer’s instructions. RpoS status. Stock cultures of each strain were screened indirectly for the allelic state of rpoS with the hydrogen peroxidase ll (HPII) assay (26, 275). 19 Fore indivi. Bena transfr incuba by dro indicat (214), in the I- tested f HPll+ C. was dro by the rc pick, Rp Was Use. Ar Were "10. day 0, an and 9row (L8G). LB acid) (LBr, NaNH4Hp‘ For each strain, we picked 96 colonies from an LB agar plate, which were individually inoculated into wells of microtiter plates containing 100 pL of Luria- Bertani (LB) broth. After 6 h of incubation at 37°C, cultures in the 96 wells were transferred to replicate LB agar plates using a 96-pin replicator tool. Plates were incubated for 24 h at 37°C, and were then tested for the expression of catalase by dropping 30% (w/v) hydrogen peroxide on each colony. Vigorous bubbling indicated positive HPII activity, which is strongly correlated to expression of RpoS (214). Colonies were thus classified as RpoS+ if they exhibited vigorous bubbling in the HPII assay. To confirm RpoS status, HPll+ colonies were subsequently tested for glycogen production, which is also controlled by RpoS (109, 144). HPll+ colonies were streaked onto LB agar and grown overnight at 37°C. Iodine was dropped onto individual colonies and glycogen accumulation was detected by the formation of a brown color on the colonies (109). For each strain, a single pick, RpoS+ colony was grown in LB broth, frozen in 10% glycerol at -70°C, and was used as stock for all subsequent AR assays. Acid resistance (AR) mechanism assays. Methods for measuring AR were modified from those of Lin et al. (154, 155) as diagrammed in Fig. 2.1. On day 0, an RpoS+ isolate of each STEC strain was inoculated from freezer stocks and grown overnight in LB medium (pH 7.0) at 37°C (Fig. 2.1). On day 1, 50 pL of overnight culture was used to inoculate 10 mL of LB broth with 0.4% glucose (LBG), LB broth buffered to pH 5.5 with 0.1M MES (morpholineethanesulfonic acid) (LBMES), minimal E medium containing 73 mM K2HPO4, 17 mM NaNH4HP04, 0.8 mM or 12mM MgSO4, 10 mM sodium citrate, and 0.4% glucose 20 meCIlar Freezer stock of —> .' RpoS+ colony . EG + EG + 7‘ ‘ glutamate arginine .‘ pH 2-0 ‘ _. _ pH 2.0 Control Test (2X) Control Test (2X) Control Test (2X) Glutamate Oxidative Arginine system system system Figure 2.1. Diagram of assay conditions to discriminate the three AR mechanisms. 21 (EG). (BHIG these was m used It initial c (EG), and Brain-Heart Infusion broth with a final concentration of 0.6% glucose (BHIG). All media were prepared by filter sterilization. Strains were grown in these media for 22 h at 37°C, with shaking at 120 rpm. Optical density (0Deoo) was measured for each culture after 22 h of growth and this measurement was used to determine the inoculum volume into the test environment to have an initial density of ~ 106 CF U/mL. The ability of the RpoS+ isolates to survive acidic conditions was measured for three AR systems (154, 155): the glutamate-dependent (GLU) system, the glucose-repressed oxidative (OXI) system, and the arginine- dependent (ARG) system (Fig. 2.1). Cultures grown in LBG were tested in the GLU system (EG supplemented with 5.7 mM sodium glutamate at pH 2.0), cultures grown in LBMES were tested in the OXI system (EG at pH 2.5), and cultures grown in BHIG were tested in the ARG system (EG supplemented with 0.6 mM arginine at pH 2.5). Cultures inoculated in EG at pH 2.0 and pH 2.5 without supplements served as controls for the GLU and ARG systems. The control for the OXI system was cells grown overnight in EC pH 7.0 and tested in EC pH 2.5. 4N HCI was used to adjust the pH of the test media, which was then warmed to 37°C before use. All AR tests were conducted at 37°C. Two replicates of each strain were tested per AR system. Samples were taken at 0, 2, and 6 h and plated in duplicate on LB agar using the Autoplate 4000 (Spiral Biotech, Bethesda, MD). Plates were incubated at 37°C for 24 h and then enumerated using the 0 Count software (Spiral Biotech). 22 exan were the a 0.8ml and lo assay. Iogw C experir time pc NC). P differen ComDari conduct To determine if Mg+2 limitation was contributing to cell inactivation (for example, through changing the permeability of the outer membrane), strains were also tested in the EG test environments described as described above, with the addition of 12mM MgSO4 (high Mg”) instead of the originally described 0.8mM MgSO4 (low Mg”). Statistical Analysis. Plate counts were converted to log CFU/mL values, and log decrease per hour for each assay was determined over the period of the assay. We defined the survival rate, AV, as the change in viable cell counts (in Iog1o CFU/mL) per hour and report the mean and standard deviation for each experiment. Analysis of variance was conducted on the log CFU/mL for each time point and the AV values for each assay using SAS (SAS Institute, Cary, NC). Painrvise comparisons were made for each of the clonal groups; significant differences were determined by the Tukey method, adjusted for multiple comparisons. Phylogenetic analysis of the multilocus sequences was conducted with MEGA 3 software (141 ). 23 1251 resoh comrr betwe seroty (Table cluster: Compri: RESULTS Clonal groups. The STEC strains were sequenced at 7 loci (3,753 bp, 1251 codons) and classified into 27 sequence types (STs) based on the alleles resolved (Table 2.1). Most (24/30) of the O157:H7 strains belong to the most common EHEC 1 sequence type (ST-66). The close genetic relationship between the 026:H11 and 0111:H8 strains is reflected by the fact that both serotypes are associated with the common EHEC 2 clone marked by ST-106 (Table 2.1). A neighbor-joining tree places the 27 STEC STs into 3 distinct clusters (Fig. 2.2). The genetic diversity between strains within a cluster comprises only a fraction of the nucleotide divergence between clusters (Table 2.2). The nucleotide diversity within clusters ranged from 0.7% for EHEC 1 to 0.23% for the EHEC 2 group whereas the divergence between clusters ranged from 1.1% to 2.6% (Table 2.2). RpoS status and Stx production. To determine the variability in RpoS expression, we used the HP ll catalase assay to examine 96 colony picks from each of 66 STEC strains (Table 2.1). For 32 of the strains, 100% of the 96 colonies tested were RpoS+. 0f the 30 E. coli O157:H7 strains tested, 15/30 (50%) had more than 90% RpoS+ colonies. In comparison, 19/22 (83%) and 12/14 (86%) strains of E. coli026/0111 and 0121:H19 had more than 90% RpoS+ colonies, indicating significant variation between clonal groups in the proportion of RpoS+ colonies (Gadj = 9.95, p < 0.01 with df = 2). A single RpoS+ colony was selected for each strain and used to create a stock for subsequent experiments. 24 TABLE 2.1. Shiga toxin-producing E. coli strains used in this study. Serotype Sequence Strain Locale (Year) stwenes %Rp_oS+ type Stx EIA Source E. coli O157:H7 TW07591 Mich. (1997) stx2 27 70 + 1 TW07695 Fla. (1997) stx1, stx2 74 60 + 1 TW07700 Calif. (1997) stx1, stx2 91 66 + 1 TWO7702 Ohio (1997) stx2 95 59 + 1 TWO7704 Ohio (1997) stx2 88 66 + 1 TWO7706 Utah (1997) stx1, stx2 88 66 + 1 TW07928 DC. (1998) stx2 99 66 + 1 TW07937 Mass. (1998) stx2 76 66 + 1 1W07938 Mass. (1998) stx1, stx2 100 66 + 1 TW07939 Mass. (1998) stx1, stx2 100 66 + 1 TW07941 Mass. (1998) stx1, stx2 45 66 + 1 TW07943 Mass. (1998) stx1, stx2 96 66 + 1 TW07945 Fla. (1998) stx2 94 66 + 1 TW07949 0.0. (1999) stx1, stx2 8O 66 + 1 TW07950 0.0. (1999) stx1, stx2 51 66 + 1 TW07952 DC (1999) stx1, stx2 82 66 + 1 TW07953 DC. (1999) stx1, stx2 38 66 + 1 1W07956 0.0. (1999) stx1, stx2 94 66 + 1 TW07957 0.0. (1999) stx1, stx2 63 66 + 1 TW07958 DC. (1999) stx1, stx2 53 66 + 1 TWO7961 Ohio (1998) stx1, stx2 96 66 + 1 TWO7962 Ohio (1998) stx1, stx2 82 66 + 1 TW08022 Mont. (2000) stx1, stx2 100 66 + 7 TW08026 Mont. (2000) stx1, stx2 100 71 + 7 TW08027 Mont. (2000) stx1, stx2 100 62 + 7 TW08030 Mont. (2000) stx2 45 72 + 7 TW08080 Mont. (2000) stx2 61 66 - 7 TW08609 Wash. (2000) stx2 100 66 + 7 TW08610 Wash. (2000) stx2 100 66 + 7 TW08612 Wash. (2000) stx1, stx2 100 66 + 7 25 TABLE 2.1, continued 26 Serotype Sequence Strain Locale (Year) stx genes %Rms+ type Stx EIA Source E. coli 026:H11 TW07594 Ariz. (1997) stx1 96 1 10 + 1 TW07595 Neb. (1998) stx1 100 106 + 3 TWO7600 Neb. (1998) stx1 100 106 + 3 TW07622 Mich. (2002) stx1, stx2 4 106 + 5 TW07814 Idaho (1997) stx1, stx2 100 104 + 6 TWO7936 Mass. (1998) stx1 100 10 + 1 TW07948 DC (1999) stx1 97 106 + 1 TW08024 Mont. (2000) stx1 100 1 14 + 7 TW08033 Mont. (2000) stx1 100 102 + 7 TW08038 Mont. (2000) stx1 98 101 + 7 TW08052 Mont. (2000) stx1 76 103 + 7 TW08060 Mont. (2000) stx1 100 1 12 + 7 TW08084 Mont. (2000) stx1 99 99 + 7 TW08569 Germany (1998) stx2 54 1 15 - 2 TW08570 Germany (1999) stx2 94 1 13 + 2 TW08571 Germany (2000) stx1, stx2 100 104 + 2 TW08637 Wash. (2000) stx1 100 106 + 7 TW09184 Mich. (2003) ND 100 106 + 5 E. coli 01 1 1:H8 TW07598 Neb. (1998) stx1, stx2 100 106 + 3 TW07601 Neb. (1998) stx1, stx2 100 106 + 3 TW08642 Wash. (1999) stx1 100 106 + 7 TW08643 Wash. (2001 ) stx1 100 106 + 7 E. coIiO121zH19 TW0761 5 Mich. (2002) stx2 82 182 + 5 TW07927 0.0. (1998) stx1, stx2 100 182 + 1 TW07931 Mass. (1 998) stx2 100 182 + 1 TW08036 Mont. (2000) stx2 1 00 1 82 + 7 TW08040 Mont. (2000) stx2 1 00 183 + 7 TW08042 Mont. (2000) stx2 1 00 1 80 + 7 TW08043 Mont. (2000) stx2 90 182 + 7 TW08055 Mont. (2000) stx2 88 187 + 7 TW08063 Mont. (2000) stx2 99 184 + 7 TABI Seroty, TVV08 TW08r TVV08€ TVVOBE TVVOBE a Shiga t t’Strains I Gennany Rudrik, ht TABLE 2.1, continued Serotype Sequence Strain Locale (Year) stx genes %RpoS+ type Stx EIA Source TW08091 Mont. (2000) stx2 100 179 + 7 TW08646 Wash. (2000) stx2 100 133 - 7 TW08649 Switzerland (2002) stx2 100 185 + 4 TW08650 Switzerland (2002) stx2 100 181 + 4 TW08653 Canada (2002) stx2 100 182 + 4 a Shiga toxin production tested by enzyme-linked immunoassay on single RpoS+ colony. b Strains were obtained from 1, D. Acheson (2, 258); 2, L. Beutin, Robert Koch Institute, Germany; 3, P. Fey (85); 4, R. Johnson, Laboratory of Foodbome Zoonoses, Canada; 5, J. Rudrik, Michigan Department of Community Health; 6, A. O’Brien (95); and 7, P. Tarr (124, 139). 27 Table 2.2. Nucleotide diversity within and between 3 clonal groups of STEC. Distance (%) between groups Clonal Number of Number of Diversity within group isolates STs group (%) EHEC 1 EHEC 2 EHEC 1 30 7 0.07 :I: 0.03 EHEC 2 22 11 0.23 :I: 0.04 2.47 i 0.25 0121:H19 14 9 0.191004 2.601024 1.10:1:0.16 28 FIQUre 2.: parameter ECOR 1, 2, 3,10 ECOR4 103 99 102112 182(6) 184 179 181 135 180 EHEC 2 (026/0111) 0121:H19 0.20% EHEC1 (O157:H7) Figure 2.2. Neighbor joining tree showing the clonal groups of STEC based on multilocus sequence types (STs). Distances were estimated by the Kimura 2- parameter model. Number of isolates of each ST is given in parentheses. 29 The l pl’OdL positir and to. strain STEC had be O157:H cell den (Anders( 0.32. A Variance total vari; hfor the not Show Efl bamerial ‘ StEibilizjng Stability re perceiVed DreSent in Dehheab'e The RpoS+ derivatives of the 66 STEC strains were all tested for Shiga toxin production using the Stx EIA; sixty-three of the 66 RpoS“ isolates were Stx positive (Table 2.1). We further investigated the 3 Stx negative isolates by PCR and found that one had lost the stx2 gene in comparison to the original STEC strain and one had stx2 but did not express it. In the third case the original STEC strain was also negative for the stx2 gene, suggesting that the toxin gene had been lost in a previous transfer. Variation in AR among E. coli O157:H7 strains. Strains of E. coli O157:H7 exhibited variation in survival using each of the AR mechanisms. Final cell densities after 6 h in the GLU system approximated a normal distribution (Anderson-Darling Normality test, p = 0.121) with a mean log1o CFU/mL of 5.43 :I: 0.32. A nested ANOVA indicates that the among-strain component of the variance is highly significant (F = 75.6, p < 0.001) and accounts for 92% of the total variation in cell densities at 6 h. Similar distributions were observed after 6 h for the OXI system (3.09 :1: 0.71) and for the ARG system (4.79 :I: 0.46) (data not shown). Effect of magnesium on AR of E. coli O157:H7. The permeability of the bacterial outer membrane (OM) can be disturbed by chelators that disrupt the stabilizing interactions of Mg”, Ca”, and LPS molecules (255). Decreased OM stability resulting from chelators could be amplifying cell inactivation that is perceived to be a consequence of acid stress alone. For example, the citrate present in the test environments could chelate Mg”, leading to a more permeable OM. 30 (0.8rr of low AVGL; distrib had Sig MgSO. mean A was inc a signifi System, Me”. I and low CI protectioi 0'01 (Star 0157147 Significan. 6 hot” tim faster rate a resUh, th signifiCanU Den'Od. Str To test for the effects of Mg“2 limitation, we challenged strains in both low (0.8mM M9804) and high (12mM MgSO4) Mg” concentration. Under conditions of low Mg”, AVGLU values approximated a normal distribution, with a mean AVGLU of -0.09 1: 0.05 (Fig. 2.3A). AVARG values also approximated a normal distribution, with a mean AVARG of -0.22 :1: 0.08 (Fig. 2.30). The O157:H7 strains had significantly greater survival rates in the GLU and ARG systems when the M9804 levels were increased, and the variation in survival rates decreased. The mean AVGLU with high Mg” was -0.05 :1: 0.02. The mean AVARG when MgSO4 was increased was -0.17 :1: 0.06. In the GLU system, 22/30 (74%) of strains had a significantly greater survival rate with high Mg” than low Mg”. In the ARG system, 14/30 (46%) of strains had a significantly higher survival rate with high Mg”. In contrast, the survival rates were not significantly different between high and low M9” for the OXI system. Comparison of STEC clones. The glutamate system provided the most protection in acidic conditions with average AVGLU values ranging from 0.004 :1: 0.01 (standard error of the mean) for the 026/0111 clone to -0.09 :I: 0.01 for the O157:H7 group (Fig. 2.4A). In the first 2 h of the assay, the 0157 strains had a significantly lower AVGLU compared to the 026/0111 strains. However, in the 2- 6 hour time period, the density of bacteria of the O157:H7 group decreased at a faster rate (AVGLU_t2-t5 = -0.10 :I: 0.01) than strains of the other clonal groups. As a result, the ability of E. coli O157:H7 strains to survive using the GLU system is significantly less (p < 0.001) than the other clonal groups in the 6-hour assay period. Strains of the non-0157 clonal groups were also tested under high Mg” 31 ESQ? 9:505 o5 Eoc— mN 13m mo:_m> >< .0 .Eougm ozfiaxo 9: .8 ON 13m mo:_m> >< .m .Eoygm ofiEmSE 9: .8 oN 13m mo:_m> >< .< .6996 m< some .2 209933 09.52 Be; 25.5 .mcoEccoo «+92 22:55 :9; :_ _m>_>5m 883cm :30 use mcoEocoo «+92 22Ew.8 26. E _m>_>5m 2865 8.86 88.0 .moEEmm 98:3. 92 co cozmsoo Emccfim 2865 23 Lotw Em: 9 to. E0: 520 :_ 00:03 new _m>_>5m among: 2 6026. So: >< 0385 9: 3 03cm. 9m mEmbm ESQ? m< some 052.. 2:93 ~10.va .aoo .m on mcoEm c... .3 £8 o_nm_> 00 Queen 9: 5 gauge n >3 meta _m>_>._:m E cozmtm> .m.~ 230$ 899$ 65:52 and. so: 25$ .. N.2. 2.8... o .. 0nd. 1 0N.0. 1 0rd. 8.0 899m @2636 102%.. I was. 2.56 o m n 09.0. Ema? 99:32.0 «.92 2.5, u N.9). 2.55 o saw... a F. .< mud. 13.0.. 1000. era: rwuns 8.0 32 condi the Cir values acidic 026/O assay, significe clonal g donalg ECOR s Significa System Th system, b Values Far strainS (Fi group ind; 026/0111 arginine s) GLU SYSte variable ac conditions, but there were no significant differences in survival rates for any of the clonal groups with the exception of O157:H7 (data not shown), so all AV values reported are from low Mg” test environments. The oxidative system provided the least amount of protection against acidic environments, with average AVox. values ranging from -0.28 :1: 0.01 for the 026/0111 strains to -0.44 :I: 0.01 for the O157:H7 strains (Fig. 2.43). In this assay, strains in the O157:H7 clonal group decreased in numbers at a significantly (p < 0.01) greater rate (AVox. = -0.44 :t 0.01) than the 026/0111 clonal group. The greatest variability in survival rate among strains of each clonal group was seen in the OXI system (Fig 2.48). The E. coli 0121:H19 and ECOR strains had a similar ability to survive acidic conditions, and were not significantly different from the 026/0111 or O157:H7 groups using the oxidative system as measured by average AVox1 (Fig. 2.4B). The arginine system did not provide as much protection as the glutamate system, but was more effective than the oxidative system at pH 2.5. AVARG values range from -0.11 i 0.01 for 0121 strains to -0.22 1: 0.01 for the O157:H7 strains (Fig. 2.40). Pair wise comparisons of the average AVARG for each clonal group indicate that the O157:H7 strains are significantly different from the 026/0111 and 0121 strains (p < 0.003) but not the ECOR A group in the arginine system (Fig 2.4C). The majority of strains exhibit high survival with the GLU system, whereas the ability to utilize the OXI and ARG systems are more variable across strains. Strain TW08650 (0121:H19) has high survival rates in 33 Survival rate Figure 2 hodzOnt; SerOgrou inmeox System a .0 re A. B. C. a 0.0. é T - . % gs? . . . . . E -o.e ( 0 0157 026/01110121 ECORA 0157 026/01110121 ECORA 0157 026/01110121 ECORA Glutamate system Oxidative system Arginine system Figure 2.4. Box plot of survival rates (AV) for strains in each clonal group. The horizontal bar indicates the mean for each group. A. AV values for each serogroup in the glutamate system at pH 2.0. B. AV values for each serogroup in the oxidative system at pH 2.5. C. AV values for each group in the arginine system at pH 2.0. 34 all thn Strain AVGLU one mi lrzctu Vt donalg 42% of and 52°; 62% is c differenc differenc betIll/cen the total \ the differ? Rp TW07695. were SGIec EaCh strajr l'neigham-Sn exhibits d 31' SUWIVal raIE I all three mechanisms, with AVGLU = -0.01,AVox. = -0.13, and AVARG = -0.12. Strain TWO7962 (O157:H7) exhibits low survival rates in two of the systems, with AVGLU = -0.09,AVox1=-' -0.45, and AVARG = -0.35. Overall, the survival rate using one mechanism is not significantly correlated with ability to survive using another (Ichu 0x1 = 0.42. I'zox1 .. ARG = 0.19. r26... NR. = 0.15). We partitioned the total phenotypic variance in survival rate into within- clonal group and between-clonal group components. For the glutamate system, 42% of the variation is due to differences between strains within a clonal group and 52% is due to differences between clonal groups. For the oxidative system, 62% is due to differences between strains within group, while 33% is due to differences between clonal groups. For the arginine system, 67% is due to differences between strains within a clonal group, and 30% is due to differences between the clonal groups. These results indicate that a substantial fraction of the total variation in acid resistance for a given mechanism is accounted for by the differences between STEC lineages. RpoS+ vs. RpoS' colonies. Eight strains of E. coli O157:H7 (TW07591, TW07695, TWO7702, TWO7704, TW07928, TW07941, TWO7962, and TW08030) were selected and the RpoS+ and RpoS' isolates derived from single colonies for each strain were tested for their ability to survive using the three AR mechanisms. In the glutamate system, the RpoS' isolates (colonies that exhibited slow bubbling in the HP ll assay) all exhibited significantly lower mean survival rates (tloos, so] = 35.97, p < 0.001) than the corresponding RpoS+ isolate 35 at bot. 2.5). l agnMC isolate 2.5). Ir RpoS+ at both levels of Mg”, with the exception of strain TWO7962 in the low Mg” (Fig. 2.5). In the oxidative system, only strains TWO7962 and TW07928 did not exhibit significantly lower (t[o_05, 60] = 24.39, p < 0.001) survival rates for the RpoS- isolates compared to the corresponding RpoS+ isolate at both levels of Mg” (Fig 2.5). In contrast, the RpoS‘ isolates exhibited similar or greater survival than the RpoS+ isolates in the arginine system (Fig. 2.5). 36 Survival rare r5 6 A N 0 a) Figure Mm8m Ullder loxl 0.2 A' +2 +2 8' +2 +2 C' +2 +2 0.8mm Mg 12mm Mg 0.8mm Mg 12mm Mg 0.8mm Mg 12mm Mg 2 0.0 . 1 E . <0 k a E a E .E .0 2 T w ‘ a a co E Q Ei 0.4 ~ - - 0.6 - g B - i RpoS- RpoS+ RpoS- RpoS+ RpoS- RpoS+ RpoS- RpoS+ RpoS- RpoS+ RpoS- RpoS+ Glutamate system Oxidative system Arginine system Figure 2.5. Box plot of survival rates (AV) for paired RpoS+ and RpoS' isolates from 8 strains of E. coli O157:H7 in each AR system. Tests were conducted under low (0.8mM) Mg” and high (12mM) Mg” conditions. 37 and ot survive ability of surv Buchar envnon EHEC s and wha- strains v. and Dati mOGerate 0f acid acl the aCid El Vanabuny 0’5 Colic PIOtect ag,| | s... cap! I rr DISCUSSION E. coli O157:H7 was first thought to have enhanced ability to resist acid and other environmental stresses because of its low infective dose and ability to survive in and be transmitted by acidic foods. Strain-to-strain variation in AR ability has been reported (albeit with a small number of strains), with a continuum of survival abilities classified from highly resistant to sensitive. For example, Buchanan and Edelson observed that the ability to survive in acidic environments varies substantially among isolates within a single serotype; AR of EHEC strains grown to stationary phase was dependent on the type of acidulant and whether the pH-dependent AR system had been induced (35). Benjamin and Datta compared survival of 6 O157:H7 strains in LB at pH 2.5, two of the strains were considered highly acid tolerant, while the other four strains were moderately and slightly acid tolerant (12). Leyer et al. (1995) measured survival of acid adapted and non-adapted cells of 5 strains of E. coli O157:H7, although the acid adapted cells had a better survival rate than the non-adapted, there was variability in survival rates among the strains (150). Lin et al. assayed 11 strains of E. coli O157:H7 for their ability to use three different AR mechanisms to protect against high acidity. Considerable variation existed in the ability of strains to use a particular AR mechanism, indicating that not all O157:H7 strains had the same capability to survive in a low pH environment using a specific AR mechanism (155). Our analysis with 30 O157:H7 isolates collected independently from sporadic cases of disease shows continuous variation in ability to survive at low pH using these AR mechanisms and that survival rates fit 38 a non phen betwe are ad strains examp DUFDOSI differen growth ; (growth. strains \ a single variatior dUe to e acid res large 00 T Stallone eXDless reSDOnS required a normal distribution. In addition, more than half (~50%-70%) of the total phenotypic variation in acid resistance among STEC clones is a result of between-strain variation within a clonal group. This finding suggests that there are additional factors that contribute to variability in acid resistance of STEC strains. Understanding the extent of variation among pathogens of a group, for example of E. coli O157:H7 strains, is of considerable importance, especially for purposes of predictive modeling. Whiting and Golden have identified the differences among strains of the same species as a major source of variation in growth and survival studies; variations determined for four different parameters (growth, survival, thermal inactivation, and toxin production) of E. coli O157:H7 strains were larger than the error calculated from experimental procedures using a single strain or cocktail (269). We also observed that the strain-to-strain variation within a clonal group was larger than variation between replicates or due to experimental error. Thus, drawing conclusions about differences in the acid resistance abilities of a variable group of bacteria will require examination of large numbers of isolates. The alternate sigma factor RpoS governs gene expression during stationary phase and in response to stressful environments. Variation in the expression of this global regulator can contribute to overall variability in stress response. It is clear that expression of the alternate sigma factor rpoS is required for stationary phase induction of acid resistance in E. coli and Shigella 39 (227) static becau to star expres ability c natural Com pie Sensitiv STEC, \- attempt,- OUI' iSOIa selecting Li tOUSethil in reduce resistancf' elth E. C not aQIEe SUWived b COMM-med (227), and Arnold and Kaspar found that as the growth of E. coli O157:H7 neared stationary phase, survival of cells in TSB at pH 2.0 increased dramatically (8). Expression of rpoS can vary in natural populations of E. coli, possibly because rpoS null mutants have a competitive advantage in prolonged exposure to starvation conditions (275). Variation in rpoS, ranging from null to full expression, could account for a component of the variability observed in AR ability of E. coli. Waterman and Small found that mutant rpoS alleles exist in natural populations of STEC during a survey of AR of 58 STEC strains. Complementation with rpoS on a plasmid conferred AR to nine of 13 acid sensitive strains (263). We also observed variation in allelic state of RpoS among STEC, with E. coli O157:H7 strains having the most variability. In this study, we attempted to correct for the potential confounding effects of rpoS null mutants in our isolates by assaying RpoS activity indirectly through the HPII assay (26) and selecting RpoS+ colonies to test AR. Lin et al. determined that functional rpoS is required for a strain to be able to use the oxidative mechanism of acid resistance. A mutant rpoS allele resulted in reduced ability to utilize the arginine and glutamate mechanisms of acid resistance (155). In a comparison of survival using a specific AR mechanism of eight E. coli O157:H7 strains that had rpoS+ and rpoS isolates, our results did not agree with those of Lin et al. In the arginine system, mutant rpoS isolates survived better than the wild type isolates. The allelic state of RpoS was confirmed in these strains by testing colonies for HPII activity with hydrogen peroxide and for glycogen production with iodine. Arginine decarboxylase, 40 encoded by adiA, is positively regulated by CysB (42, 224) and by AdiY, a transcriptional regulator encoded downstream of adiA (235). It is possible that RpoS does not contribute to regulating expression of adiA. However, further studies are needed to determine the role of RpoS and other regulators in governing the arginine AR system. Chelating compounds, as well as organic and inorganic acids, increase the permeability of the outer membrane (3, 107), which has been linked to loss of viability during acid stress in E. coli O157:H7 (127). Bacterial inactivation by acid could be aided by the presence of chelators in the test environment. Citric acid has been shown to be a potent OM permeabilizer in O157:H7, and the perrneabilization effect can be abolished with the addition of 5mM Mg (107). Mg” stabilizes cell membranes (30) by interacting with adjacent LPS molecules (255). Our results indicate that high levels of magnesium in the low pH environment significantly increase the survival rates of most E. coli O157:H7 strains in the glutamate system. The increase in survival is not as dramatic for the arginine system, possibly because this system is tested at a higher pH. Interestingly, bacteria of the other STEC clonal groups, as well as the ECOR group, were not significantly affected by the increased levels of magnesium in these two AR systems. No differences were observed in survival for the 0157 and 026/0111 clones in the OXI system at the two levels of magnesium, but the 0121 group showed a significantly lower survival rate at 12mM Mg in the OXI system. Further work is needed to understand the mechanisms by which Mg” concentration increases survival in acidic environments. 41 and but subs adve thnsp MOSt l seroty ben~er strains Condni the O1 PHise seroty] In vitro studies on AR of E. coli O157:H7 strains, other pathogenic E. coli, and different enteric bacteria have indicated that E. coli O157:H7 can exhibit AR, but whether O157:H7 strains have superior ability to survive in acid has not been substantiated. The consensus seems to be that E. coli O157:H7 is well suited to adverse conditions even though little direct evidence supports the hypothesis that this pathogen is substantially different from other E. coli in AR abilities (168). Most AR studies are based on only a few non-0157 STEC strains of the same serotype, which reduces the statistical power to detect significant differences between groups. For example, Miller and Kaspar observed survival of two strains of E. coli O157:H7 and a single non-pathogenic strain under acidic conditions (175). While a significant difference in survival was reported between the O157:H7 and non-pathogenic strain, their suggestion that resistance to acidic pH is an additional characteristic that distinguishes O157:H7 from other E. coli serotypes (175) is not warranted because other pathogenic E. coli serotypes were not tested. Benjamin and Datta compared survival of 14 EHEC strains in LB at pH 2.5 and 3.0 and found that no correlation could be drawn between acid tolerance and serotype (12), but again only a few non-0157 serotypes were evaluated. McKellar and Knight surveyed a set of 19 EHEC strains for ability to survive in TSB acidified with HCI (pH 2.0) and with acetic acid (pH 4.0). When strains were grouped by source, they found that survival of outbreak strains in acid was significantly greater than strains from human or animal sources. When strains were grouped by serotype, however, strains of E. coli O157:H7 were among the more acid sensitive, but clear conclusions could not be drawn since 42 notha (D261- sunfiv eren 00mpa pathog SUDeno COnditic meChar than be other EHEC serotypes were only represented by a single strain each (172). Berry et al. recently reported a comprehensive comparison of AR between 39 isolates of 0157 and 20 isolates of non-0157 biotype l E. coli, and found no significant differences in the ability of strains from either group to survive in BHI pH 2.5, and concluded that high acid tolerance is not unique to strains of pathogenic E. coli (15). Our results indicate that strains of E. coli O157:H7 do not have superior AR ability compared to strains of E. coli0121zH19 and 026:H11 using the three AR mechanisms. While E. coli O157:H7 clones can survive in low pH using the three described AR mechanisms, there is no evidence that O157:H7 has greater acid resistance in any single system compared to other clones of STEC in these laboratory test environments. In a comparison of three clonal groups of STEC and a clonal group of non- pathogenic ECOR strains, we show that strains of E. coli O157:H7 do not have superior AR ability using three specific AR mechanisms. Under natural conditions, however, E. coli O157:H7 may have the ability to use multiple mechanisms in combination, through some form of epitasis, to achieve higher AR than bacteria of other clonal groups. It is also possible that E. coli O157:H7 has evolved alternative mechanisms, yet to be identified, which contribute to acid resistance in nature. This study suggests that there is no evidence for enhanced AR of O157:H7 by a measurable superiority of one of the three major AR mechanisms of E. coli. There is also compelling evidence for a substantial strain-to-strain component of variation in survival rate; more than 60% of the total variance in survival rates for the oxidative and arginine systems were attributed 43 to be inter- test a STEC under patho to between-strain differences. Our findings demonstrate that because of the high inter-strain variability, a large number of isolates should be examined in order to test and draw conclusions about survival of E. coli O157:H7 and non-0157 STEC. A narrow focus on a few isolates or a single serotype may seriously underestimate the variability that underlies important phenotypic variation in pathogens. 44 used Koch Unfio VVash Ek.Ja resear ACKNOWLEDGEMENTS We would like to thank the following people for kindly providing the strains used in this study: Dr. David W. Acheson, CFSAN; Dr. Lothar Beutin, Robert Koch lnstitut; Dr. Paul D. Fey, University of Nebraska; Dr. Alison D. O’Brien, Uniformed Services University of the Health Sciences; Dr. Phillip l. Tarr, Washington University at St. Louis; Dr. Roger Johnson, University of Guelph; and Dr. James Rudrik from the Michigan Department of Community Health. This research was supported in part by the Michigan Agriculture Experimental Station. 45 CHAPTER 3 Past gene conversions between duplicated glutamate decarboxylase genes (gadAB) in pathogenic Escherichia coli Bergholz, T. M., C. L. Tarr, L. M. Christensen, D. J. Betting, and T. S. Whittam submitted to J. Bacteriology 46 ESch deca invoh vhdeh Enkfl seque strains genes vanab Eflelo. Donut COnve SYHOr allele. Cone SVen1 IESEN SUMMARY Escherichia coli produce two biochemically identical isofon’ns of glutamate decarboxylase, which are encoded by the gadA and gadB genes and are involved in acid resistance. These genes are unusual among the Enterobacteriaceae in that they occur only in E. coli and Shigella. The loci are widely spaced on the E. coli chromosome and the two genes are 97% similar in sequence. Comparison of the nucleotide sequences of the gadA and gadB in 16 strains of pathogenic E. coli revealed 3.8% and 5.0% polymorphism in the two genes, respectively. Alignment of the homologous genes identified a total of 103 variable sites including 21 fixed nucleotide differences between the loci, which are located in the first 82 codons of the genes. Thirty-six variable sites were polymorphic for the same nucleotides in both genes suggesting past gene conversions or intergenic recombination. Phylogenetic analysis based on synonymous substitutions indicated two cases in which specific gadA and gadB alleles were more closely related to one another than to other alleles at the corresponding locus. The results indicate that at least three gene conversion events have occurred after the gad gene duplication in the evolution of this acid resistance mechanism in pathogenic E. coli. 47 15 do ba int: sys 061 E0 net 801 tor INTRODUCTION Escherichia coli can persist in extremely acidic environments (8, 100, 154). This ability to survive at low pH is thought to play a role in the low infective dose of certain foodborne pathogens, such as E. coli O157:H7 (100, 103). Many bacteria have evolved decarboxylation-antiporter systems for maintaining neutral internal pH in acidic environments. The lysine and arginine decarboxlyase systems are common in many enteric bacteria, whereas the glutamate decarboxylase system is present only in E. coli and Shigella among the Enterobacteriaceae. The glutamate decarboxylase system works to maintain neutral cytoplasmic pH by decarboxylation of glutamate and export of y- aminobuytric acid (89). This amino acid-based acid resistance system also acts to repel incoming protons through the inversion of the membrane potential (209). E. coli produce two isozymes of glutamate decarboxylase (GAD) that protect cells in acidic environments. The expression of both isozymes is required for survival at pH 2 (42). The two homologous genes that encode the GAD isozymes are widely spaced on the chromosome (229). The gadA gene is located at 78 minutes (3.67 Mb) on the K-12 chromosome (229), near two transcriptional regulators of the GAD enzymes, gadX and gadW(118, 160), and is part of the acid resistance fitness island (117) (Fig. 3.1). The gadB gene, located at 33 minutes (1.57 Mb) on the K-12 chromosome (229), produces a protein (GadB) that is nearly identical in amino acid sequence to GadA, and is co-transcribed with the adjacent gene, gadC, an antiporter that exports y- aminobutyrate (74, 112, 209).. GadW and GadX have been shown to bind the 48 sam one box, esse forth same 9808 and g found dUplic DI E.( 80d 9‘ same 44-bp promoter sequence, with two binding sites upstream of gadB and one binding site upstream of gadA (248). A 20-bp sequence named the GAD box, located between -52 and -73 bp in the gadA and gadBC promoters, is essential for both acid and stationary phase induction (41 ), and is the binding site for the essential transcriptional activator, GadE (158). The GAD system has the same genomic arrangement in pathogenic E. coli O157:H7 where the gadA and gadBC loci map to homologous locations in the chromosome (197). Both gadA and gadB occur in virtually all E. coli (including Shigella strains) and are not found in other enteric bacteria (102) suggesting that these genes arose by duplication, rather than through horizonatal transfer events, early in the radiation of E. coli and Shigella. Duplicated genes can evolve in complex ways. They can rapidly diverge and evolve new functions; alternatively, they can undergo paralogous gene conversion because of their sequence similarity. This type of intergenic recombination leads to genetic homogenization and has been observed in the multicopy ribosomal genes (152). Here we use sequence comparisons in a well- defined set of pathogenic E. coli strains to address the following questions: To what extent have the duplicated gad structural genes diverged among pathovars? Is there evidence of recent gene conversion between the paralogous gad genes? Has there been divergence, conservation, or homogenization of flanking sequences? To begin to address these questions, we sequenced the gadA and gadB genes in representative strains of enteropathogenic and enterohemorrhagic E. coli (78). These data were used to test hypotheses about 49 the role of mutation, recombination, and natural selection in the molecular evolution of the duplicated genes. 50 1.566 Mb on K12 chromosome glutamate y-aminobutyric acid antiporter gadC <— 1.573 Mb GadE GadW/GadX glutamate decarboxylase l isozyme G AD box gadB pqu <— +— [ITITIIJTIIII—E‘“ v - -"~1—.——{//////////////I//////J gadB R1 gadA R1 V177 //ZT/ZZ ‘/ ”/1717 fl/ '/~"///. gadX <— araC-Iike 3.662 Mb regulator 2050 bp gadB amplicon 2213 F1 2226 bp gadA amplicon 55A F1 ‘« 1w :. . ,: \\XX\\\I gadA I yth +— <— glutamate GAD box decarboxylase 3.666 Mb isozyme GadE GadW/GadX Figure 3.1. Locations of gadA and gadB on the K-12 MG1655 chromosome. The amplicons used for sequencing are marked for each loci. The GAD box is a 20 bp promoter sequence that is the binding site for the essential transcriptional regulator GadE (158). The AraC-like regulatory proteins GadW and GadX bind to two sequences upstream of gadB and a single sequence upstream of gadA (248). 51 MATERIALS AND METHODS Bacterial strains. Twelve strains of E. coli were used as sources of DNA for cloning and sequencing of gadA and gadB (Table 3.1). The strains included laboratory strain K-12 (ATCC 47076) and representative isolates of enteropathogenic E. coli (EPEC) and enterohemorrhagic E. coli (EHEC) that have been classified into clonal groups based on the sequence analysis of conserved housekeeping genes (208). E. coli K-12 was included as a positive control for the cloning and sequencing because the nucleotide sequences of gadA and gadB are known (229) and the entire genome sequence is available (25). For comparative purposes, we also obtained the homologous gad sequences from four completely sequenced genomes including E. coli O157:H7 strains EDL-933 (197) and RIMD 0509952 (referred to here as Sakai) (105), both of which were originally isolated from foodborne outbreaks of hemorrhagic colitis; an 06:H1 uropathogenic E. coli isolate, CFT073 (267), originally recovered from a patient with pyelonephritis (249); and type 2a Shigella flexneri strain 301 (abbreviated Sf301), which was originally isolated from a patient with shigellosis in Beijing, China (126). GAD cloning and sequencing. All strains listed in Table 3.1 were inoculated from stocks stored in Luria-Bertani broth (LB) + 10% glycerol at -80°C and were grown overnight in LB (Becton, Dickinson, Franklin Lakes, NJ) at 37°C. Genomic DNA was isolated from 1 ml of culture using the Purgene DNA isolation kit (Gentra Systems, Minneapolis, MN). The gadA gene was amplified by PCR using locus- 52 Table 3.1. Serotypes and source of 12 E. coli strains sequenced for the glutamate decarboxylase genes, gadA and gadB. Strains were classified into clonal groups based on multilocus sequence comparisons (208). No. Strain segflty'pe Source of original isolation Clonal grog) Rough 1 K-12 (O-:H48) USA (1922) ECOR A 2 E2348/69 O127:H6 UK (1969) EPEC 1 3 DEC 1a 055:H6 USA (1956) EPEC 1 4 DEC 23 055:H6 Congo (1962) EPEC 1 5 93-111 O157:H7 USA (1993) EHEC 1 6 OK-1 O157:H7 Japan (1996) EHEC 1 7 DEC 5d 055:H7 Sri Lanka (1965) EHEC 1 8 DEC 8b O111:H8 USA (1986) EHEC 2 9 2666-74 O26:NM USA (1974) EHEC 2 10 DEC 11a 0128:H2 USA (1975) EPEC 2 11 DEC 12a O111:H2 UK (1950) EPEC 2 12 B170 0111:NM USA (1983) EPEC 2 53 spe TAC 3.1) CCT specific primers (gadAF1 5’-TTAATGCCTCCTCCTTGAGC-3’, gadAR1 5’- TACACGCCGTATATGCAGGA—3’) which produced a 2226 bp amplicon (Fig. 3.1). The gadB gene was amplified separately, (gadBF1 5’- CCTGTCTGTACTGATGTAGCCA -3’, and gadBR1 5’- GACTGAGCAGGAGCAATTGT-3’), and produced a 2050 bp amplicon (Fig. 3.1). PCR reactions contained 1x Amplitaq Gold PCR buffer (Applied Biosystems, Foster City, CA), 0.2mM each dNTP, 2mM magnesium chloride, 1mM each primer, and 3U Amplitaq Gold Taq (Applied Biosystems, Foster City, CA). The following PCR conditions were used: 94°C for 10 min., followed by 35 cycles of 92'C for 1 min., 57°C for 1 min. and 72°C for 2 min. The gadA and gadB amplicons from each strain were purified using the QiaQuick PCR purification kit (Qiagen, Valencia, CA) and then cloned into pCR2.1 using the TA Cloning Kit (lnvitrogen, Carlsbad, CA). Since the known sequences are very similar, the cloning step ensured that only gadA or gadB would be sequenced. Plasmids were extracted using Ultra Clean Mini Plasmid Prep kit (Mobio, Solana Beach, CA) and electrOphoresed on a 1% agarose gel to assure that the insert was present. The amount of plasmid DNA was quantified spectrophotometrically and 90 - 100 ng used per sequencing reaction. Both gadA and gadB were sequenced from the plasmids, using primers that were designed from E. coli K- 12 and EDL-933 gad sequences (GenBank Ascession NO 000913 and NC 002655) (Table 3.2). Cycle sequencing reactions contained 9.0 )ul CEQ DTCS Quick Start premix (Beckman Coulter Inc., Fullerton, CA), 2.0 ,ul of 20 uM primer, approximately 100 ng of 54 Table 3.2. Primer sequences used for gadA and gadB sequencing. Primer Sequence gadB-56F gadA/B204F gadB666F gadB806F gadB1172F gadB284R gadB436R gadA/B89OR gadB1 166R gadA305F gadA635F gadA719F gadA1082F gadA362R gadA281 R gadA1356R 5’ - GAGTCCTTTGCACTI'GCTTAC -3’ 5’ -TTTCTGCCAGACCTGGGAC -3’ 5’ — CTGCACGATGCGCTGGATAA — 3’ 5’ - GATCAGTGCTTCAGGCCACA - 3’ 5’ - CCTGTATGACCTCTCTGAAC — 3’ 5’ -— GGCTGAGATTGCGGATATTC - 3’ 5’ — TCCATACGCTTGCGCCAACG — 3’ 5’ — AACGTTGAACACCAGTTCCT -3’ 5’ — AGAGGTCATACAGGGTGTATCC -3’ 5’ — ATATGGTTGCCGATCTGTGG — 3’ 5’ — CCTACACCGGTAACTATGAG — 3’ 5’ — ACCGGTATCGACATCGACAT -3’ 5’ — GGCCGTATGAGTTCATCTGT — 3’ 5’ - GTGCCAACGGCCTGACCATT - 3’ 5’ — GCGGATTGCGGATA'ITCTTC — 3’ 5’ — CGGGTGATCGCTGAGATATI' — 3’ 55 Pfi‘ 94°— forZ Sep NJ). SUSp Coul EXpo (DNA and a EFsE plasmid containing gadA or gadB and ddH20 to a final volume of 15 ,ul. Prior to addition of DTCS premix and primer, the plasmid was nicked by incubation at 96°C for 903. Amplification utilized an initial denaturing step at 94°C for 1 min, followed by 35 cycles of 96°C for 30 s, 48°C for 30 s, and 60°C for 2 min. Upon completion of cycle sequencing, samples were purified with Sephadex G—50 Fine columns (Amersham Pharrnacia Biotech Inc., Piscataway, NJ), dried under vacuum centrifugation (Savant Instruments Inc., Holbrook, NY), suspended in 40 ul of deionized formamide, and run on a CE02000XL (Beckman Coulter lnc.). Samples were analyzed using the CE02000XL software and then exported for further analysis with the SeqMan module of the Lasergene software (DNASTAR Inc., Madison, WI). Sequences have been submitted to Genbank and are accession numbers gadA (EF547378 —EF547388); gadB (EF551351- EF551361) Statistical analyses. We used two statistical approaches and computer programs to detect recombination between gad sequences. The TOPAL package uses a phylogenetic comparative approach to identify locations of conflicting phylogenetic signal in aligned sequences. It is a graphical method based on sliding a window along a DNA alignment, estimating a tree for the first half of the data in the window using a Least Squares methods, and recording the values of the sum of squares (171 ). The sequence data in the second half of the window are then fit to the same tree topology and the sum of squares is calculated. The difference in the sum of squares (D33) is tabulated for all 56 possible windows, in both fonrvard and reverse directions, and plotted against the position in the sequence. Peaks in the D33 values mark the locations of putative recombination events. We used TOPAL 2.0 (170), which implements a standardization of the distances across windows to reduce the confounding effects of regional rate variation. In our application, a window size of 500 bp was used with a 10-bp increment between windows. Phylogenetic tree construction was based on the Jukes-Cantor distance and the Least Squares method. One hundred randomization tests were run to evaluate the significance of the observed Dss values. To detect paralogous gene conversion between the gad loci, we used GENECONV (Version 1.81) (219), a permutation-based program that assigns probability to matching fragments in aligned DNA sequences. The gadA and gadB sequences from each strain were examined as pairs and calculations were based on silent codon polymorphisms with 10,000 permutations. The number of synonymous substitutions per synonymous site (ds) and the number of nonsynonymous substitutions per nonsynonymous site (dN) were estimated by the modified Nei-Gojobori method using MEGA3 (Version 3.1) (141). Standard errors were estimated by a bootstrap method with 500 replications. Gene phylogenies were inferred by the neighbor-joining algorithm and statistical confidence in the topologies was assessed by bootstrapping using MEGA3. 57 pol. am nuc ami sites non: were nons PrOdi agair distin used RESULTS DNA polymorphism. For the 16 gadA sequences, there were 53 polymorphic nucleotide sites (3.8% of 1401) and 11 (2.4% of 467) polymorphic amino acid positions. The gadB sequences showed a slightly greater level of nucleotide polymorphism (5.0%), which predicted only 9 (1.9%) polymorphic amino acid positions. The rate of synonymous substitution per 100 synonymous sites was 4.13 i 0.64 for gadA and 5.85 i 0.80 for gadB. The rates of nonsynonymous substitution per 100 nonsynonymous sites for gadA and gadB were 0.17 3: 0.06 and 0.14 i 0.05, respectively. The ratios of synonymous to nonsynonymous substitution of 24:1 for gadA and 42:1 for gadB indicate that the products of both genes are highly conserved with purifying selection acting against point mutations that result in amino acid replacements. There are 14 distinct alleles for gadA and 12 distinct alleles for gadB. All sequences were used for subsequent recombination analyses. Putative recombinational breakpoints. As a first step to detecting recombination within and between duplicated gad genes, we used a phylogenetic approach as implemented by the TOPAL software package (170). In our application, a 500-bp sliding window was used to plot the test statistic (Dss) calculated at 10-bp intervals along the aligned 1401 bp of the gad coding regions. At each point a tree was estimated based on the sequence data in the first half of the window, and then Dss was calculated as the difference in the fit (as measured by the sum of squares) between this tree and the data in the first and the second half of the window. This analysis was repeated in both 58 OS 033 FIQUre Statisu, SeqUer C)btain gadA ( 0012 gadA 00090 Dss 0.006)» 0.003” U M j 0 : — . - . : - : - 300 400 500 600 700 800 900 1000 1100 0012 gadB 00091 Dss 0.06 0 0 A 0.003 w 0 c . . 300 400 500 6 W 00 700 800 900 1000 1 100 statistic (difference in the sum of squares) is plotted over the length of the sequence. obtained by parametric bootstrapping. Recombination events were detected for 0.4 gadA and gadB 0.3 0 Dss 0.2 l. 0.1 " rm 0 : : : . - 300 400 500 600 700 800 900 1000 1100 Position in nucleotide sequence Figure 3.2. Putative recombination events as detected by TOPAL. The Dss The horizontal bar indicates the 95% significance level for Dss, gadA (A), gadB (B), and gadA and gadB together (0). 59 directions for the 16 sequences of each locus separately and for the combined 32 sequences. Hypothetical breakpoints of past recombination are identified by significantly large peaks in the running average Dss values. The peaks designate positions where there are discrepancies in the phylogenies based on sequence differences on either side of the peak. For gadA, two potential recombinational breakpoints were identified: one between nucleotides 260-300 and a second between nucleotides 980-1120 (Fig. 3.2A). The secondary Dss peaks near positions 400, 600, and 700 were not found to be significant. Two potential recombinational breakpoints were also identified in gadB and are located at positions 560-590 and 1110-1150 (Fig. 2.2B). The analysis of the combined data for the gadA and gadB alleles reveals three significant Dss peaks in the sliding window analysis (Fig. 3.2C). The peaks centered at positions 575 and 1060 are significant and reflect the primary breakpoints detected for the individual locus comparisons. However, the broad peak at the 5’ ends of the gadA and gadB comparisons is extreme (Dss > 0.35), remaining above the critical value almost to position 500. This large, significant plateau of Dss suggests that the 5' ends of gadA and gadB have a complex history of divergence that cannot be explained by a single recombination event. Paralogous gene conversions. We were primarily interested in evidence of gene conversion or intergenic (paralogous) recombination between gadA and gadB. To detect gene conversion between the loci, we analyzed all 32 gad sequences grouped as gadA and gadB pairs, one group for each of the 16 strains (16 groups with blocksize 2). Eighty-nine polymorphic synonymous sites 60 Ta ger wh are bas Cor Table 3.3. Significant fragments found by Sawyer's method (219) for detecting gene conversion. Comparison 1 is a test for significant outer fragments, whereas comparisons 2-4 are tests for inner fragments. The number differences are the number of polymorphisms within the fragment. Simulated P values are based on 10,000 permutations and are corrected for multiple comparisons. Start End Simulated Comparison position position Differences P value 1. 16 strains 1 111 or 114 9 or 10 0.0041 2. EPEC 2 244 639 24 0.0215 3. EPEC 1 604 1047 26 0.003 4. K-12 700 1176 28 0.008 61 .. .. C .Juaswca..~. ~..:cc:thvvrrh~ w ~ch.uN-.omthWbm.MMMNN~NH Inlnln . . . . . « .(CCCCCCCCCCCOOOOOooocchbhbhccwvwwwwnnmmmnnvvvvvvVVVVVMMMMMMNNNNNNNNNNNNNHHmHHNH HHHHHHHHfiHHddHHHHHHHdHHHHHHHHanfiH ._oo_ 05 comics 9:05 5552.8 050 Le «.39 E 882% EoEmmE 8:5 050...;ch 265m 55 2.200.. <20 095 £929; £989 003% on... .mfibmcm 30053 0523 m B 09098 350385 5:95:58. 02:33 00 80:82 05 EmE 80533 05 06% 39.85 Song 0. no“ E0: one. 9m 09 05 8 BE. 96:53 02620:: on... .N_.-x .208 .m E 88:08 «Sum 05 5ng 55 98:58 00:00.03: EmE Eon of. .mcfibm .Eoo .m or .6 8:00 030 new «Sea 05 E 8% 258.03: 233.9 our 05 0o EoEcm=< and .9". ...‘.U..‘H ............. H .................... AflFOEUv Hfluwo ...‘.U‘.‘.H ............. H .......... U ......... A‘H uuav “luv-no ...:UL.‘H ............. H. .................... A.N unnv @uunno . . .... ...‘.U..‘H ............. a .................... AOW\OVMNuv WInFNHO ...................... ‘............................................‘.............................nU...........-......... aflocflv ‘Nh Eu ................ u ..................... :52 5.58 So ................ u ..................... 32-35 5.53 So ............. 0.0 ..................... :09 5.53 Eu ................ o ..................... 57:. 0.53 .49.... ............ u ..................... a. one 553 l r ............ U ..................... A'F0uwouv Enuflo . l. ----------- O ..................... “no Uflnv an” Hdflo 1 1.»... ........... U --------------------- A‘NH U-nv N“ u HHHO ......... 3 ..................... a: use 0.33 ........... 0 3:53:33 .. «HIM «emu nhvohvmmovun0H0vNHmwovmnnohwomN0hunvnuohfiwndnnn. dn~ Ivan).thn~InIHN000honomhumvrw nhomvomvnufloomhwmmflfldobunmvvnnnNHOQQFFMNNHFUNGv0nnnm>ouvvoorbunudl:. oonnwdddnovvuw NH .FHHHHHHF -- 11111 11111 .HHHflHfi HHHHHAHHHHHHHMHHHHdHHHHHHHHHHnHHHH 62 HHHaHfiAfifiaaHaHaHHHHHH .1<<<:oo 0:00 .60 0000 0002500 .3 00000000 E0E00¢ .006 .:00:_:0.0 0 0:000:00: 05 06000 mm 00:... 0:“ :. 000.0 2:06:28 0P 0:. 00:0,: 0:..50 0.0300 0:... ._00_ 0:0 :0050: 0E0>0 :o.0._0>:00 0:00 60000;: 00000000 0:09:00... :05. 0:005:00 26:0 005 0:200. <20 00.5 E0__:0_: 00x0: 0000:0 0: ._. .0_0>_0:0 26053 050.0 0 >0 00000000 3:69.005 6005:8000: 030:6 00 060000. 05 000E 000:0:000 0:0 0600 0000.005 .03. .000 m :_ 00:0:000 $000 05 :39: 00:0 0:00.000 00.60.02: 0.00.: 060 0: ._. 050.00 .000 .m or .6 00:00 m000 0:0 «000 0:0 :_ 000.0 00.60.00: 0.:0:0> 00.. 0:0 00 EOE:0__< .0.» .0."— ..U.. H. ...... a .......... a Uh.— ‘doocouuoouao ooooo U .8 Amfiosv Hmuwo ..s..ss ...... s .......... s. s. ....... ssssss ...... s ss ... gas. ss.sss - .U. o B ...... a .......... a g ....... @808 ...... U 8 A.“ onav wnnmmo ..s.. s ...... s .......... s..ss ....... ssssss ...... s.. ss .ssssssss .sscs.ssu. ss.ssss s. .................. .s.ssss.sss..s...s.s.s ..... s.s..s s..ss .sssssssss .sssss. as. .2. s ......... s s..s ................ ss.ss s ..ss. s..ss.s ss ................... sssssss. s .s.s.ssssss.sssssssss .s...s. .s..sss ss. s ......... s s..s ................ ss.ss s ..ss. s..ss.s so ................... sssssss. s .s.s..sssss.sssssssss .sss-sss. .s..s.s as... ......... s s..s ................ ss.ss s .ss. s..ss.s ss ................... sssssss. s .s.ss.sssss.ss..sss.s ..-ss. .s..s.s ss..s ......... s..ss s ................ ss.ss s ..ss. s..ss.s ss ................... sssssss. s..s.ss.sssss.sssssssss .sss-ss. .s..sss as... ......... s. s..s ................ ss.ss.s ..ss.. s..ss.s ss ................... sssssss. s..s.s.ssssss.sssssssssw .ss sss. .s.sss . 00.0. dduaaoa.ou¢¢u¢¢¢ Avriuwouv xzuuwo ..oo.o. dduahua.0u¢404¢¢ And can. mundane ..oo.o. ‘duaaua.00¢¢u¢¢¢ AINH unnv NHHHHHO ..oo.o. (409309.00¢¢u<4¢ .oHH can. «unouno ..uo.u (109909.00¢¢u¢¢< AOpan zzunauo .Uo.0 dduaaua.uatdu¢¢¢ warn msssfiw mh¢0hv000vmowfiovNHm00¢0on0§00mN06vawumbflmmammwhoumuomoawnvuowvmowo«m00000hfiwhavocaunaanorvuonoruvamhvmvaN0v0b0momhonvbo 0hamvomvmudoaobwmnNdHOFQmmvvnmmudoomhhnuwdhwN0vvnnaa000vvoahhomuaom00mmmnnvvoomnnoaommhtvnmufiddmmvunadaoobmommnmmuflddfi nmmnnNNNNNNNHHHHHHHdfifiooooocooooooammmmmmm000pbbbbwmowmmw0nmnnnmmvvvvvvvvvvvmnnnmnNNNNNNNNNNNNNHHHHHHH HfiHaHHHHHHHHHHHHHHHHHHHfiHHHHHHHHHH 63 we ex< witl site ma“ EHI inch sign cor. stra mat: 3.4). from nUch 117g that groui t0 ha 3' en QadE estim- SYNC! were identified in the 1,395 aligned bases (the start and stop codons were excluded from the analysis). In the comparison of gadA and gadB sequences within strains, GENECONV (219) detected 13 global outer fragments (runs of sites that are unique to the group) and 10 significant inner fragments (runs of matching sites between loci). The outer fragments were found in all but the five EHEC 1 group strains (EDL-933, OK-1, Sakai, 93-111 and DEC 5d), and included the first 111 or 114 nucleotides in gadA (Table 3.3, Comparison 1). The significant run of synonymous polymorphic sites suggests that a recent conversion event has homogenized the 5' end of the gadA sequences across strains (see Fig. 3.3 and 3.4). The inner fragments include three distinct runs of matching sites between the gad loci within a strain (highlighted in Fig. 3.3 and 3.4). EPEC 2 strains have a significant matching run between gadA and gadB from nucleotide 244 to 629, EPEC 1 strains have a significant run from nucleotide 603 to 1047, and K-12 has a significant run from nucleotide 700 to 1176 (Table 3.3). Each of these runs of matching synonymous sites suggests that past gene conversion between gadA and gadB occurred within a clonal group. It is noteworthy that only the EHEC 1 gadA and gadB sequences appear to have diverged independently with no evidence of past gene conversions in the 3' end of the genes. Gene phylogenies. To assess the level of divergence of the gadA and gadB loci determined by past mutation and gene conversion events, we estimated a gene phylogeny using the neighbor-joining algorithm and the synonymous substitution rate. For this analysis the gad loci were each divided 64 St ()6 tre sen ark: vver cod l Sing) lflasn forml the g Wher grOUp Segue into two regions, based on the first detected significant inner fragments from GENECONV (Table 3.3). The first region analyzed comprised 80 codons (codons 2-81, ATG start codon excluded), which spanned the first 33 polymorphic nucleotide sites through position 243 (Fig. 3.5). The second region comprised the next 385 codons 0' GA stop codon excluded) and included 70 polymorphic sites (Fig. 3.6). The neighbor-joining tree produced by the analysis of synonymous substitutions in the first 80 codons showed a deep split between the gadA and gadB gene clusters (Fig. 3.5). For the 5' region of gadB, the topology is similar to trees constructed from other gene sequences (38). The EPEC 1 cluster of sequences branched off first then the EHEC 1 group splits off from the EPEC 2 and EHEC 2 lineage. The gadB sequences from EPEC 2 and EHEC 2 strains were identical to the K-12 allele in the 5’ region. For the gadA tree, the first 80 codons were nearly identical with only 5 polymorphic sites, three of which are singletons and are not informative. The EPEC 1 strains have substitutions at position 123 and 162 (Fig. 3.3) so that their gadA sequences for the 5' region formed a distinct cluster in the tree (Fig. 3.5). The neighbor-joining tree for codons 82-465 was strikingly different from the gene tree above (Fig. 3.6). In this case, there was no single divergence point between gadA and gadB sequence clusters; instead, there were two cases where the sequences of the two loci were more similar to each other within clonal groups. The first case was for the K-12, EPEC 2 and EHEC 2 groups of sequences. For 65 Codons 2-81 gadA Jr gadB 5.0 __) . O127:H6 (E2348/69) 055:H6 (DEC 13) 69 055:H6 (DEC 2a) 06:H1 (CFTO73) ‘55 K-12 —— 0128sz (DEC 11a) 0111:H2 (DEC 123) D1112NM (B170) 0111:H8 (DEC 8b) 0261NM (2666-74) FZA (Sf301) O157:H7 (93-111) O157:H7 (Sakai) O157:H7 (EDL-933) O157:H7 (OK-1) 8.50 :l: 3.49 T 055:H7 (DEC 5d) 01272H6 (E2348I69) 35-9 7 i 9'15 055:H6 (DEC 1a) 90 055:H6 (DEC 2a) 06:H1 (CFTO73) K-12 01282H2 (DEC 11a) 0111:H2 (DEC 12a) O111:NM (B170) 0111:H8 (DEC 8b) 026:NM (2666-74) F2A (Sf301) O157:H7 (93.111) O157:H7 (Sakai) O157:H7 (EDL-933) O157:H7 (OK-1) 055:H7 (DEC 5d) 4.85 :l: 2.82 Fig. 3.5. Neighbor-joining phylogenies for gadA and gadB based on the the number of synonymous substitutions per 100 synonymous sites (d. x 100). Gene tree for the 5' region of the gad genes spanning codons 2-81. Bootstrap values based on 1000 replicates are given at the internal nodes. Putative gene conversion or duplication events are marked with solid circles. The average ds ( :L- SE) is given for each node of interest. 66 Codons 82-349 — FZA (Sf301) — 0111:NM(B170) 0120112 (DEC 11a) 0111:H2 (DEC 12a) gadA 0111:H8 (DEC 8b) 65 0262NM(2666-74) K-12 99 gadB r K-12 0111:H8 (DEC 8b) T 75» 0262NM (2666-74) 85 0120112 (DEC11a) 1.91;. 0.81 0111:H2 (DEC 12a) O111:NM(B170) O157:H7 (93-111) O157:H7 (EDL-933) 59 10° O157:H7 (Sakai) 7. 7] :t: 1.9/ 0553H7 (DEC 5‘” gadA '— 055:H7 (DEC 5d) O157:H7 (ox-1) 0" O157:H7 (93-111) 01571-17 (Sakai) 5° 0157017 (EDL-933) F2A ($1301) gadB 055:H6 (DEC 2a) O127:H6 (5234869) 055:H6 (DEC 1a) 06:H1 (CFTO73) 06:H1 (CFTO73) 80 D1271H6 (E2348l69) 055:H6(DEC1a) ' 10—4 59 055:H6(DEC 2a) 2.83 :l: 0.97 Fig. 3.6. Neighbor-joining phylogenies for gadA and gadB based on the the number of synonymous substitutions per 100 synonymous Sites (ds x 100). Gene tree for the 3' region of the gad genes including codons 82 to 465 (349 in figure). Bootstrap values based on 1000 replicates are given at the internal nodes. Putative gene conversion or duplication events are marked with solid Circles. The average d5 ( 1: SE) is given for each node of interest. 67 these 6 strains, the gadA sequences were in one cluster and the gadB sequences were in a second, closely related cluster. The gadA and gadB sequences of these two clusters differed, on average, at 1.9% of the synonymous sites (Fig. 3.6). The second case involved the three EPEC 1 strains where again the gadA and gadB fell into two clusters that are more closely related to each other (2.8 % different at synonymous sites) than to the other gadA or gadB sequences. In contrast to these cases, the EHEC 1 gadA and gadB sequences were highly divergent and Share a most recent common ancestral sequence at the deepest part of the phylogeny (Fig. 3.6). Promoter region. Nucleotide sequences were determined for the entire region upstream of the coding sequence for both gadA and gadB (Fig. 3.7). The promoter regions for gadA and gadB have diverged substantially upstream of -72 bp, but were similar to one another closer to the 5’ end of the genes. The proportion of substitutions per site wasthree times higher for the gadB promoter sequences (0.048 1 0.007) compared to the gadA promoter sequences (0.016 :1: 0.005). GadW and GadX have been Shown to bind the same 44-bp promoter sequence, which is centered at 110.5 bp upstream of gadA and 220.5 bp upstream of gadBC (248). Comparisons of the sequences here Showed that the GadW/GadX binding site upstream of gadA was more conserved with only two polymorphic Sites (Fig. 3.7A), whereas the GadW/GadX binding site upstream of gadB (centered at 220.5 bp) contained 7 polymorphic sites (Fig. 3.76), and the sequence identified as a weak binding Site (centered at 110.5 bp) contains 9 polymorphic Sites. The gadA and gadB promoter regions were conserved in the - 68 35 and -10 regions in our strains (Fig. 2.70). A comparison of the K-12 20-bp GAD box sequence, located between -52 and -73 bp in the gadA and gadBC promoters, to those determined here indicated that this region is well-conserved in both promoters, with only two strains that contain a single base difference. 69 $0.0 :. 00.:0._:0E Eu 2:! E00 0500.8 055 0:. 0:0 00:0 0... 0:0 mm- 0: .. .0.0:0.00. 0 .3 0000.0:0 0. 00:00:00 .6: 90.0 0: ._. @000 0:0 «000 503.0: 000:0.0E0 .:0._:0.: 000:00:00 00:.:00:: .00.00..00>:. 0:.0...0 =0 0:0E0 00>:00:00 >_0.0.0E00 00:0 0.8.0:. 00:._ .00....0> .m000 0:0 .0000 .0 00000000 E00000: 00+ 0. 00.100 05:06 :0 .5 00.00.05 00 00:0 0505: X000 0:... .900 .0 00:00:00 E00000: For- 0. man. am 05:06 :0 5.3 00.:0E 0. 0:0 050:... X000 0: .. .0000 .0 00:00:00 E00000: .0..- 0. mm..- 3. 0000.030. :. 0.:0:0> 0.0:.0 ::>> .0:0 0:. 0600 000000.00: 0.0E0._0 0:. >0 00.0:00 0.0 $00 :0 (000 :.::>> 00:0 0.:80E>.0a .900 0:0 «.000 .0 000000000 000:00:00 :0.00. .0.0E0:n. <20 .0... 00+ 0.+ .+ 0.1 001 001 ovi 001 001 0:1 001 001 00.1 no a 0 o a 04. o Do 33 0‘00naugaagdaaaa4ua¢¢moagm009m4yaanaau¢uooa¢¢ ——>- MSS + PBS 8Ah 9h 6.5h pH 2.5 250j.lL — ‘ 006,, = 0.1 i E _. AJ 22°C MSS L370 pH 3.5—’ ""'—‘ pH2.5 P - , 4“ 24h F AJ 4°C MSS freezer stock Figure 4.1. Diagram of assay conditions for measuring survival in the model stomach system (MSS) and apple juice (AJ). To begin an experiment, 0.25 ml of bacterial culture grown in LB broth to an 00500 of 0.1 is transferred to MOPS medium at pH 7.0 (Step 1), and then serially transferred twice to fresh medium. At steps 2 and 3, aliquots of the 6.5 h MOPS culture are added to the M88 (Step 2) or into AJ (Step 3) for a final bacterial concentration of 106 CFU per ml. After a 24h incubation in AJ (Step 4), 15 ml of cultured juice is transferred into the M88 for a 1:10 dilution. The length of incubation time in hours is indicated. 83 the presence of the apple juice in the MSS. Immediately after inoculation and addition of either apple juice or PBS, the contents of the stomacher bags were homogenized in the Stomacher 400C lab blender at 230 rpm for 15s then placed on a rocking platform at 37°C. Contents of the bags were mixed and sampled every hour for three hours. Samples of the SGF-baby food mixture were diluted in PBS and plated on LB agar using an Autoplate 4000 (Spiral Biotech, Norwood, MA). The final samples were taken at 3h post inoculation and were plated on both MacConkey and LB agar to determine the number of injured cells. Plates were incubated at 37°C for 24h before enumerating survivors using the Q Count (Spiral Biotech, NOf'WOOd, MA). The pH of the SGF-baby food mixture was measured after the assay, and remained within 0.03 units of the target pH. Three independent cultures (biological replicates) of each strain were assayed for survival. Survival in the M88 after 24h storage in apple juice. Pasteurized apple juice with no added preservatives was purchased locally. Juice was adjusted to pH 3.5 with the addition of 4N malic acid, sterilized through 0.22pm filters, and stored frozen at -20°C until needed. Thawed juice was mixed vigorously and divided into 100mL aliquots, which were stored at 22° and 4°C one day prior to inoculation. Portions of the same stationary phase cultures in MOPS medium were used to inoculate apple juice at 22°C (AJ 22°C) and apple juice at 4°C (AJ 4°C) with ~106CFU/mL (Fig. 4.1, step 3). Inoculated juice was vortexed briefly and then stored at 22 or 4°C. Samples were taken from each inoculated juice after 24h and plated onto LB agar using an Autoplate 4000. 84 After 24h in apple juice, strains were inoculated into the MSS. prepared as described above. A 15mL aliquot of inoculated juice was transferred to a MSS bag (Fig. 4.1, step 4), homogenized and incubated at 37°C, and sampled as described above. At least three independent biological replicates of each strain were assayed for survival. Plates were incubated at 37°C for 24h before enumerating survivors using the Q Count. Plate counts of each strain after storage in AJ were used as the initial cell density for survival in the M88 after 24h in AJ. Four strains from each clonal group (TW07622, TW07700, TWO7952, TW08022, TW08024, TW08060, TW08264, and TW08642) were tested for survival in the M88 after 24h storage in 4°C apple juice at either pH 3.5 or 7.0. Apple juice was adjusted to pH 7.0 with 4M NaOH and then sterilized using 0.22pm Stericups (Millipore). Strains were inoculated into 100mL of AJ at either pH and stored at 4°C for 24 h. The MSS assay was conducted as described above. Quantitative PCR. TaqMan primer and probe sets targeting gadA, gadB, and mdh were developed by Applied Biosystems Assays-by-Design service based on available genome sequences for these genes (Table 4.2). All TaqMan probes were labeled with FAM fluorescent dye. Primer and probe sets were evaluated for specificity by testing with either gadA or gadB located in the cloning vector pCR 2.1 (Invitrogen). Primer and probe for gadA using gadB template and gadB primer and probe with gadA template were not amplified, as no increase in fluorescence during the Q-PCR run was observed. Six strains from the two 85 Table 4.2. Primers and probes for gadA and gadB detection using quantitative PCR Target Primer pair Probe mdh F - TGGTACAGCAAGTTGCGAAAAC CTGCCCGAAAGCGTG R - GTGGTGTTAACCGGGTI’AGTGATAA gadA F - GGACCAGAAGCTGTTAACGGATTI' CCGCTCAGAACTACTCG R - GCGATAGTAGAAATGGCC‘ITTGC gadB F - GATTCACG'ITI'I'GGTGCGAAGT CTGCGATAGTGGAAATAG R - ACGAGCGTTGCCATCAAGATATAAT 86 clonal groups (0157 strains TW02302, TW07591, TWO7704, TWO7937, TW08022, and TW08264; 026 strains TW07594, TW07595, TW07622, TW08024, TW08033, and TW08642) were selected based on their survival rates in the M88 (two strains that represented the highest, median, and lowest survival rates were selected from each group). Strains were grown in MOPS medium as described above and cultures were sampled 1.5h after entry into stationary phase, the same time point at which cultures were transferred to the M88 in the original assays. Cultures were mixed with 2 volumes RNA Protect Bacterial Reagent (Qiagen) and then centrifuged at 7,500 rpm for 10 min to pellet cells. Cell pellets were stored at —70°C until further use. Total RNA was extracted from the cell pellets using the hot acid phenol- chloroforrn method (165). Two independently grown cultures were sampled for each strain. Genomic DNA was removed from the RNA samples by on-column DNase treatment using the RNeasy kit (Qiagen). Aliquots (2pg) of total RNA were used for cDNA synthesis using the TaqMan Reverse Transcription kit (Applied Biosystems). Reverse transcription reactions contained 1x TaqMan buffer, 5.5 mM magnesium chloride, 500 pM each dNTP, 2.5 uM random hexamers, 4U RNase inhibitor, and 12.5U MultiScribe Reverse Transcriptase and were carried out under the following conditions: 10 min at 25°C, 30 min at 48°C and 5 min at 95°C. A total of 200ng of cDNA was used for quantitative PCR (QPCR) on the 7900HT SDS (Applied Biosystems). QPCR reactions contained 1x TaqMan Buffer A, 5.5 mM magnesium chloride, 200 uM each dNTP, 0.15U AmpErase UNG, 0.3U AmpliTaq Gold, 900 nM forward and reverse 87 primers, and 250 nM TaqMan probe and were carried out under the following conditions: 2 min at 50°C, 10 min at 95°C, followed by 40 cycles of 15s at 95°C and 1 min at 60°C. TaqMan assays were run in triplicate for each culture replicate. Increase in fluorescence over time was monitored and used to quantify expression of gadA, gadB, and mdh in each strain. All QPCR data were analyzed with SDS 2.1 (Applied Biosystems). A standard curve was generated for each primer/probe set using a series of 10-fold dilutions of the cloning vector pCR 2.1 (lnvitrogen) containing either gadA, gadB, or mdh sequence. Relative quantities of each gene in each sample were determined from the standard curve, which was included in each individual QPCR run. The metabolic gene mdh was used as an endogenous control for each sample, which was used to standardize the gadA and gadB expression levels within each sample. To compare expression levels across strains, expression levels of gadA and gadB were normalized to those of E. coli 026:H11 strain TW08033, which had the lowest survival rate in the MSS. Statistical Analysis. Plate counts were converted to log CFU/ml values, and log decrease per hour for each assay was determined over the period of the assay. The survival kinetics were non-linear, so the difference between each pair of time points was averaged to determine the survival rate for each independent replicate of each strain. We defined the survival rate, AV, as the change in viable cell counts (in logic CFU/ml) per hour and report the mean and standard deviation for each experiment. The percentage of injured cells was calculated as the difference between the 3h plate count samples from the MSS 88 (in log1o CFU/mL) on MacConkey agar and LB agar, divided by the total log1o CFU/mL (on LB agar) and multiplied by 100. Analysis of variance was conducted on the log CFU/ml for each time point and the AV values for each assay using SAS 9.0 software (SAS Institute, Cary, NC). Because the survival rates for each clonal group were not normally distributed, painivise comparisons were made between the clonal groups using nonparametric statistical analyses. The Kruskal-Wallis test uses the H test statistic, which approximates the theoretical 38 distribution, and was used to test for significant differences between clonal groups and treatments (193). Distribution of relative expression levels determined by QPCR were found to be non-normal and pair-wise comparisons were made between the clonal groups using the Kruskal-Wallis test to test for significant differences between clonal groups. 89 RESULTS Survival in the model stomach system. To quantify survival in a simulated gastric environment, strains were grown to stationary phase and transferred to the MSS pH 2.5, where the numbers of viable cells were assessed hourly for 3h. AVMSS (log decrease per hour in the MSS) values for each clonal group did not fit a normal distribution (Anderson-Darling test for normality p < 0.005), thus the Kruskal-Wallis rank sum method was used to compare AVMSS between clonal groups. Stationary phase cultures were inoculated directly into MSS pH 2.5 with the addition of either PBS or apple juice (AJ) to determine if the presence of AJ conferred a protective effect in the MSS. The survival rates (AVMSS+sz and AVMSSW) were similar for the 0157 strains (H = 0.40, df = 1, P = 0.52) and the 026/0111 strains (H = 0.38, df = 1, P = 0.53) and thus only AVMSSW was used for further comparisons (Fig. 4.2). E. coli O157:H7 strains had a mean AVMSSW of -0.17 i 0.15 which was significantly greater (H = 6.61, df = 1, P = 0.01) than that of the 026/0111 strains, with a mean AVMSSW of 037 1r 0.18 (Fig. 4.3 columns 1 and 2). Survival in the MSS after incubation in apple juice. The effect of storage in AJ on subsequent survival in gastric acid was tested by using stationary phase cultures to inoculate AJ held at either 22°C or 4°C. After 24h in AJ (Fig. 4.1, step 3), samples were transferred to the MSS where the number of surviving cells were assessed at hourly intervals for 3h. Comparison of AVMSS “4°C (log decrease per hour in the MSS after 24h storage in AJ at 4°C) revealed 90 MSS + PBS 0.0 ~- l .0 N J r Survrval rate (log decrease per hour) 6 j J- 1 * a -0 6 “r o O O 0 e O -0-8 4 m 0157 (n = 14) |:| 026/0111 (n= 12) Figure 4.2. Box plots of survival rates (AV) in the model stomach system (MSS) with the addition of either apple juice (AJ) or phosphate buffered saline (PBS) for strains in each clonal group. The horizontal bar indicates the median for each group. Survival rates were not significantly different between treatments. 91 that O157:H7 strains had a significantly different (H = 12.96, df=1, P = 0.0002) mean survival rate (AVMSS Mac = -0.13 i 0.15) that was more than three times greater than that of the 026/0111 strains (AVMSS AJ4oc = -0.41 i 0.19) (Fig. 4.3, columns 3 and 4). The 0157 strains as a group have a similar survival rate in the MSS before and after storage in AJ at 4°C (Fig. 4.3 columns 1 and 3). When comparing AVMSSW and AVMSS Mac for individual strains within the 0157 group, only strain TWO7702 had significantly higher AVMSS “4°C (P = 0.005). Although 9 of the 14 (64%) 0157 strains had a measurable increase in AVMSS Mac compared to AVMSSW, the mean increase was not significant. The O26/O111 strains also had a similar survival rate in the MSS before and after storage in AJ 4°C (Fig. 4.3 columns 2 and 4). In this case, 7 out of 12 (58%) 026/0111 strains had a measurable decrease in survival rate after storage in AJ 4°C, and two of these strains (TW07595 and TW08570) had a significantly lower AVMSS AJ4oC compared to AVMSS- Storage in AJ held at 22°C decreased survival in the MSS for strains of both clonal groups compared to AVMSSW and AVMSS Mac (Fig. 4.3). E. coli O157:H7 strains had a mean AVMSS “22°C of -0.47 i 0.19, which was significantly greater (H = 5.6, df=1, P = 0.01) than that of the O26/O111 strains, with a mean AVMSS AJ22°C of -0.68 i 0.28 (Fig 4.3, columns 5 and 6). By comparing survival within clonal groups, we found that both 0157 and O26/O111 strains have significantly lower AVMSS Amoc compared to either AVMSS or AVMSS AJ4°C- 92 MSS after MSS after 0 2 q MSS AJ 4°C AJ 22°C 0.0 . g .- O .8 -0.2 -- I 3 ~93 “ Ti 9 8- -o.4 -~ 4 a 0 . i > (O O ' ~ ‘3 C E 9 -0.6 " o a o 0 . (0.8 e — o) -0.8 fl . g 0 -10 .. [:j O157(n=14) [j 026/0111 (n =12) -1.2 -» I Figure 4.3. Box plots of survival rates (AV) in the model stomach system (MSS) before and after incubation in either apple juice at 4°C (AJ 4) or apple juice at 22°C (AJ 22) for strains in each clonal group. The horizontal bar indicates the median for each group. 0157 strains have a significantly higher (H = 12.96, df = 1, P = 0.0003) survival rate compared to the 026 strains in the MSS and MSS-I-AJ 4. 93 Quantifying injured cells in the MSS. Injured cells were quantified by plating on LB and MacConkey agar after 3h in the MSS, and calculating the log difference in CFU per ml after 24h on the two media. The mean log difference between the injured and uninjured cells after 3h in the MSS was 0.49 i 0.31 for the 0157 strains, which was significantly less than 1.37 i 0.88 for the 026/0111 strains (Table 4.3, H = 5.12, df = 1, P = 0.02). The log difference between the El ‘ —‘__5L.». injured and uninjured population after 3h in the MSS was similar for the 0157 strains before and after storage in AJ 22°C, but there was significantly lower injury (0.32 i 0.27, H = 10.58, df = 1, P = 0.001) in the MSS after AJ 4°C (Table 4.3). The comparison of the log difference injury for the 026/0111 strains in the MSS indicates a significant difference between MSS and MSS AJ 4°C (H = 9.55, df=1, P = 0.002), and the log difference of injured and uninjured cells is significantly lower for the 026 strains (H = 14.5, df = 1, P = 0.001) in MSS AJ at 22°C compared to the MSS without pre-exposure to AJ (Table 4.3). Survival in the MSS after storage in AJ 4°C at pH 3.5 and 7.0. After observing that STEC strains, regardless of clonal group, survived better after cold storage (i.e., AVMss N400 > AVMSS mm), we tested whether low temperature or increased acidity led to the protective effect on subsequent survival in the MSS. Four strains from each clonal group were selected and incubated in AJ 4°C at either pH 3.5 or pH 7.0. After 24h, part of the inoculated juice was transferred to the MSS to assay survival. Incubation in AJ 4°C pH 3.5 did not 94 Table 4.3. Log difference of noninjured and injured cells (mean :I: sd) measured after 3h in the model stomach system (MSS) following incubation in apple juice (AJ) at 4°C or 22°C. Log difference of noninjured and injured cells MSS after MSS after Clonal group No. strains MSS alone AJ 4°C AJ 22°C O157:H7 14 0.49 :l: 0.31 0.32 :l: 0.27 0.63 i: 0.34 026/0111 12 1.37 i 0.88 0.77 :l: 0.48 0.58 :t 0.41 95 affect survival rates for the O157:H7 or 026/0111 strains compared to survival in the MSS before incubation in AJ, as seen previously (Fig. 4.4). The O157:H7 strains had a mean AVMSS AJ4°C pH 7,0 of -0.71 i 0.34, which was similar compared to the 026/0111 strains with a mean AVMSS Mac pH 7,0 of -0.85 i 0.34. Storage in AJ 4°C pH 7.0 significantly (1‘ = -7.29, df = 57, P < 0.0001) decreased survival in the MSS for both the 0157 and 026/0111 strains compared to their original 1 i AVMss, which indicates that the acidity of the apple juice contributed more to l subsequent acid challenge and survival than cold temperature, although the malate ion concentration would be different in the AJ pH3.5 compared to the AJ pH 7.0. Transcription of glutamate decarboxylase in stationary phase. The glutamate decarboxylase system has been shown to provide the greatest acid resistance in E. coli (77, 145, 155) and is critical to survival of O157:H7 during passage through the gastrointestinal tract of calves (202). We hypothesized that the difference in MSS survival between 0157 and 026/0111 clonal groups results from differences in the level of transcription of glutamate decarboxylase genes, gadA and gadB. To test this, TaqMan assays were developed targeting gadA and gadB, which encode the glutamate decarboxylase isozymes. Primer and probe sets were found to be specific to gadA or gadB when each set was tested with either gadA or gadB as a template (see Materials and Methods). Transcript levels of gadA and gadB were measured for six strains of each group that represented the range of survival rates for the group (two strains with the highest, median, and lowest survival rates). 96 MSS after MSS after MSS AJ 4°C pH 3.5 AJ 22°C pH 7 0.0 'i' ' -0.2 1~ -o.4 t l; I -0.6 -— I ~' . Sun/[val rate (log decrease per hour) -0.8 -- .4 :' Iii...“ —‘ -1.0 -- 0157 (n = 4) 42 'r [J 026/0111 (n= 4) J I Figure 4.4. Box plots of survival rates (AV) in the model stomach system (MSS) before and after incubation in either apple juice pH 3.5 at 4°C (AJ 4 pH 3.5) or apple juice at pH 7.0 at 4°C (AJ 4 pH 7.0) for 4 strains from each clonal group. The horizontal bar indicates the median for each group. 97 0157 strains had significantly greater relative levels of transcription of both gadA (H = 10.46, df = 1, P = 0.0012) and gadB (H = 15.42, df = 1, p <0.001) compared to the 026 strains (Fig. 4.5). Relative transcription levels of gadA and gadB were highly correlated (:2 = 0.89). Transcription of gadA was more variable than gadB among strains in both clonal groups. Relative transcription levels of either gadA or gadB were not positively correlated with survival rates in the MSS (r2980... = 0.21, 1293,13 = 0.15), indicating that other factors play a role in enduring the MSS environment. 98 gadA gadB O 10 -- __ S '17) 8 a- 8 a 6 4_ 7"1‘ 0 e 35 3:3 4 .- 16 Q) , e m 2 -- J.- é + 0 -I- E 0157 (n = 6) [j 026/0111 (n = 6) Figure 4.5. Box plots of relative transcription values for gadA and gadB for 6 strains of 0157 and 6 strains of 026. Correlating the increase in fluorescence over time to a standard curve for each target gene and then normalizing across strains determined the relative transcription for each strain. The horizontal bar indicates the median for each group. The 0157 strains have significantly higher transcription of gadA (P = 0.0012) and gadB (p <0.0001) compared to the 026 strains. 99 DISCUSSION Gastric acidity has long been seen as one of the first lines of defense against foodborne pathogens, and the ability to survive gastric acidity is thought to be associated with the infectious dose of a pathogen. Epidemiological data from outbreaks suggest that E. coli O157:H7 has a low infectious dose, estimated to be approximately 100 cells (59). Human volunteer studies have been utilized in the past to assess the infectious dose of an organism; this approach is not feasible for E. coli O157:H7 or other STEC because of the potential of severe disease complications resulting from the effects of Shiga toxin. Instead, inferences about infectious dose have been drawn from the investigation of the survival of STEC strains in experiments designed to simulate the conditions in the human digestive tract. Previous studies that have compared survival of pathogenic E. coli in simulated gastric environments have focused mainly on E. coli O157:H7. Arnold and Kaspar, for example, measured survival of strains of several pathotypes of E. coli in SGF pH 1.5 and found that EHEC O157:H7 strains persisted longer compared to strains of enteroinvasive E. coli, enteropathogenic E. coli, and Shigella dysenten'ae (8). Roering et al. compared survival among strains of Salmonella enterica serovar Typhimurium DT104, Listeria monocytogenes, and E. coli O157:H7 and found that O157:H7 strains endured for 3h in SGF pH 1.5 compared to survival of 20 min or less for Salmonella and Listeria strains (213). In addition, Tamplin observed that 01 57:H7 strains had a 4-fold higher survival rate when mixed with cooked ground beef and inoculated into SGF compared to Shigella strains (239). In his 100 study, the average survival for 10 O157:H7 strains in SGF plus cooked ground beef (—0.08 i- 0.07) was greater than the average survival rate observed here in the MSS (—0.17 i 0.15), which comprised SGF plus homogenized turkey. The MSS is a valuable tool for evaluating survival of bacterial cells in a gastric environment after ingestion of a meal and was devised to address the criticism that survival studies in SGF alone do not accurately the complexity of the gastric environment after eating (128). Food remains the predominant transmission route for E. coli O157:H7 (206), so creating a gastric environment that accounts for the impact and quality of food is important. The gastric environment is dynamic, and pH fluctuates considerably after ingestion of food (80). De Jonge and colleagues recently modeled survival of E. coli O157:H7 under dynamic pH profiles designed to mimic those of the human gastric environment after eating, and found that acid adaptation can occur during passage through the gastric barrier (75). While the MSS utilized in this study does not account for variable conditions, such as changes in pH and rate of gastric emptying, it does contain food particles and digestive enzymes, and yielded highly reproducible results with STEC strains. Environmental stresses may damage cellular components and impair the ability of microorganisms to survive subsequent stress. Many studies examining survival in gastric environments have not quantified the percentage of cells injured. In an extensive survey of AR of 0157 and non-0157 commensal bovine isolates, Berry et al. found that O157:H7 strains had a significantly higher percentage of injured cells compared to the non-0157 commensal isolates when 101 exposed to Brain Heart Infusion broth at pH 2.5 for 6h (15). In these cases, AR was tested in a variety of media at different lengths of time, and isolated from different sources (bovine vs. human), and 0157 isolates were compared to commensal isolates, which could account for differences observed in the percentage of injured cells. Murinda and colleagues found that O157:H7 strains were more acid resistant compared to non-0157 STEC when testing at pH 2.5 for 30s (177), although cells were recovered from acid stress on selective media, which also indicates the effect of injury on non-0157 STEC. This observation is consistent with our finding that the amount of injured cells after 3h in the MSS was significantly greater for the O26/O111 strains compared to the 0157 strains. These results indicate that differences among strains in the AR phenotype may also be evident in the injury status of cells. Prior exposure to mild acidity can increase the subsequent AR of E. coli upon exposure to extreme acidity. As E. coli O157:H7 has been transmitted in acidic foods, it has the potential to become more acid resistant and thus have improved survival in passage through the stomach. However, severe stress imposed by food matrices could also lead to increased cell injury and decreased bacterial survival. Relatively few studies have explored the impact of incubation in particular food matrices on the ability of bacterial cells to survive subsequent acid stress. The results of the experiments presented here indicate that there is no significant increase in survival rates of the 0157 or 026/0111 populations in the MSS after storage in AJ for 24h. We found that the storage temperature of inoculated AJ significantly influenced subsequent survival of STEC in the MSS, 102 with incubation at 22°C leading to decreased survival in the MSS. A similar temperature effect was found for E. coli O157:H7 inoculated into apple juice, with survival in SGF at pH 2.5 greater after cold storage (at 4°C) versus storage at room temperature (21°C) (254). Naim and colleagues (179) found a strong protective effect of sausage on the ability of E. coli O157:H7 to survive gastric acid. They inoculated three O157:H7 strains into sausage batter and measured survival in an in vitro digestive challenge after fermentation and drying, and found that digestive juices offered no real challenge to the bacterial cells ingested within dry fermented sausage (179). Three strategies have been identified for acid tolerance in bacteria — those involved with changes in membrane composition, the use of enzymatic or physiological mechanisms to maintain internal pH, and systems to repair and prevent damage to essential cellular components by pH (63, 117). It is clear that E. coli O157:H7 alters its membrane fatty acid profiles (273, 274), decreases membrane permeability in response to acidic conditions (127), and induces several mechanisms that maintain internal pH (155, 209). In addition, expression of the DNA binding protein, Dps, protects DNA from damage resulting from acid stress (63, 180). Differences in any of these systems could account for variation in AR within and between STEC clonal groups. Our experiments indicate that there is a significant between-group component of variation resulting from differences in the transcription of glutamate decarboxylase genes, gadA and gadB. Glutamate decarboxylase has been shown to be critical to survival of O157:H7 during passage through calves (202) and important to survival under 103 extremely acidic conditions such as the stomach. Here we observed difference among clonal groups in the transcription levels of gadA and gadB, however, these levels were not highly correlated with survival rates among strains within groups, suggesting that other factors are also contributing to the AR and survival rates in the model stomach. In a comparison of two clonal groups of STEC, we have shown that E. coli O157:H7 strains have a superior ability to survive in a simulated gastric environment compared to the non-0157 STEC serotypes 026:H11 and O111:H8 strains. This difference, however, is alleviated when cultures are held at stationary phase for longer periods of time. The findings suggest that E. coli O157:H7 cells rapidly achieve an enhanced state of AR in early stationary phase, an ability that may underlie the low infectious dose of this pathogen. 104 ACKNOWLEDGEMENTS We thank the following people for kindly providing the strains used in this study: David W. Acheson, CFSAN; Lothar Beutin, Robert Koch Institute; Peter Feng, CFSAN; Paul D. Fey, University of Nebraska; James Rudrik, Michigan Department of Community Health; Alison D. O’Brien, Uniformed Services University of Health Sciences; Philip I. Tarr, Washington University at St. Louis; and Tetsuya Hayashi. Genome Information Research Center, Osaka University. We also thank Shannon Manning for helpful discussion. We thank Megan Parsons for assistance with the model stomach assays and Annette Thelen for excellent technical assistance with the Q-PCR assays. This work was supported by USDA NRI Competitive Grant number 2005- 35201-16362 and in part by the Michigan Agricultural Experiment Station. TMB was supported by a DHS graduate fellowship from The National Center for Food Protection and Defense. 105 CHAPTER 5 Global transcriptional response of Escherichia coli 01 57:H7 to growth transitions in glucose minimal medium 106 SUMMARY Global patterns of gene expression of Escherichia coli K-12 during growth transitions have been extensively investigated, however, comparable studies of E. coli O157:H7 have not been explored, particularly with respect to factors regulating virulence genes and genomic islands specific to this pathogen. To examine the impact of growth phase on the dynamics of the transcriptome, O157:H7 Sakai strain was cultured in MOPS minimal media (0.1% glucose), RNA harvested at 10 time points from early exponential to full stationary phase, and relative gene expression was measured by co-hybridization on high—density DNA microarrays. Analysis of variance (R/MAANOVA, Fs test) identified 442 (36%) of 1239 0157-specific ORFs and 2110 (59%) of 3647 backbone ORFs that changed in expression significantly over time. QT cluster analysis placed 2468 of the 2552 significant ORFs into 12 groups; each group representing a distinct expression pattern. ORFs from the largest cluster (n=1078) decreased in expression from late exponential to early stationary phase: most of these ORFs are involved in functions associated with steady state growth. Also represented in this cluster are ORFs of the TAI island, encoding tellurite resistance and urease activity, which decreased ~4-fold. Most ORFs of the LEE island, with the exception of espZ, decreased ~2-fold by early stationary phase. ORFs encoding proteins secreted via the LEE encoded type III secretion system, such as tccP and espJ, also decreased in expression from exponential to stationary phase. Three of the clusters (n=154) comprised genes that are transiently upregulated at the transition into stationary phase and included genes involved in nutrient 107 scavenging. Upregulated genes with an increase in mRNA levels from late exponential to early stationary phase belonged to one cluster (n=923) which includes genes involved in stress responses (eg. gadAB, osmBC, and dps). These transcript levels remained relatively high for >3h in stationary phase. The Shiga toxin genes (stx1 AB and stx/28) were significantly induced after transition into stationary phase. Expression of 307 0157-specific ORFs was modulated in a growth dependent manner. These results provide a baseline transcriptional profile that can be compared to patterns of gene expression of this important foodborne pathogen under adverse environmental conditions. 108 INTRODUCTION Enterohemorrhagic Escherichia coli (EHEC), a food and water- borne pathogen of zoonotic origin, are an important cause of acute gastroenteritis in humans. O157:H7 is the predominant serotype of enterohemorrhagic E. coli causing illness in the United States (173). One of the hallmarks of EHEC pathogenesis is the formation of attaching and effacing lesions on intestinal epithelial cells (90, 256) , resulting in intimate adherence of the bacterial cell to the intestinal mucosa. The ability to form AlE lesions is encoded by the Locus of Enterocyte Effacement (LEE), a 35 kb pathogenicity island (169). Another virulence characteristic of EHEC is the production of Shiga toxin, a two component toxin that inhibits protein synthesis in eukaryotic cells (194). The genes for production of Shiga toxin are located in intact or partial genomes of lambda prophages that are inserted into the chromosome (110, 220). Together, the LEE and Shiga toxins are considered to be two of the principal virulence determinants that mediate development of hemorrhagic colitis and the life- threatening hemolytic uremic syndrome (133). Regulation of these virulence factors is influenced by global regulators as well as regulators specific to the virulence factor. Many of the virulence factors in EHEC are carried on ‘foreign’ DNA, which is not present in E. coli K-12, and their expression is typically regulated by transcriptional regulators also carried on these mobile elements. EHEC can utilize global regulators that are common to all E. coli as well as 0157-specific regulators to control expression of virulence factors (133, 230). For example, the expression of LEE is dependent on growth 109 phase and responds to nutrient downshifts via ppGpp signaling through DksA to Ler (_I=EE encoded regulator) and PchA (181). LEE expression can also be upregulated by activation of Ler in the presence of bicarbonate ions (1 ), and also through quorum sensing, where autoinducer 3 activates LEE expression through Ler (232). Shiga toxin expression is linked to expression of the phage lysis genes (259, 260) and can be controlled by the iron responsive regulator Fur (36). In addition to the known virulence determinants, E. coli O157:H7 Sakai and EDL- 933 contain ~ 1.3 Mb of sequence not found in E. coli K-12; these sequences occur as part of the genetic material of 18 Iysogenic phages and 6 phage-like elements that have integrated into the Sakai genome (105) and as 177 0157 strain-specific islands (OI number) in the EDL-933 strain (197). Recently it was identified that many effector proteins that are secreted via the LEE-encoded type three secretion system are specified by genes distributed throughout the genome on Iambdoid phages (247). E. coli is metabolically versatile and can adapt to growth under a wide range of conditions, adaptations essential to a bimodal lifestyle either in the primary habitat within animal hosts or in secondary habitats as free living cells in the natural environment (101). Slowing metabolic activity during stationary phase reflects a survival mechanism in nutrient poor environments in which bacteria undergo a variety of morphological and physiological changes. In stationary phase, E. coli are more resistant to a number of stresses, including pH stress (100, 227) and osmotic stress (125), both of which occur during transit through the colonization of the host intestinal tract. Important to the transmission 110 route of E. coli O157:H7 is the ability to persist in adverse environments until entering a new host. Hence studying growth transitions and how they relate to the lifestyle of E. coli O157:H7 is essential to understanding the persistence and spread of this pathogen from the bovine reservoir to foods and to humans. Microarrays have been utilized to observe changes induced in the transcriptome of E. coli during growth transitions and in response to different environmental stresses. Global transcription profiling of E. coli K-12 indicates a coordinated response to growth arrest (57). The sigma factor RpoS plays a crucial role in transcription of stationary phase and stress response genes (142, 195, 266). Microarray studies have been conducted to understand how E. coli K- 12 responds to growth at low pH (106, 252), on rich and minimal media (240), during anaerobic growth (66, 132, 215), and growth transitions in rich medium (10). Transcriptional responses of O157:H7 during growth on plasma membranes (72) and upon exposure to norfloxacin (111), have been investigated, and microarrays have been used to identify genes regulated by quS in E. coli O157:H7 (233), but a full picture of global gene expression during growth transitions, while studied in E. coli K-12, has not been investigated in the food borne pathogen E. coli O157:H7. A key component to understanding the adaptations and ecology of E. coli O157:H7 is to determine how the 0157-specific genomic elements respond to one of the critical stages of growth, stationary phase. It has been well- documented that some virulence factors are regulated by environmental cues that signal the presence of the host environment, but what is less understood is 111 how these virulence factors are expressed in a non-host environment. Here we address the question of how O157:H7 responds to growth transitions by determining the changes in global gene expression patterns in O157:H7 during the transition from exponential to stationary phase in minimal medium, with particular interest in expression patterns of 0157-specific genes. The transcriptome of E. coli O157:H7 Sakai was examined by comparative microarray hybridizations at 10 different time points during exponential and stationary phase and compared to obtain a temporal pattern of gene expression. 112 MATERIALS AND METHODS Growth conditions. E. coli O157:H7 RIMD0509952 (Sakai) involved in a radish sprout outbreak (174) was stored at —70°C in LB broth and 10% glycerol. The Sakai strain was inoculated into 10mL of LB broth from freezer stocks and grown to ODsoo = 0.1. This ~4h period of growth in LB allowed cells to recover before transfer to MOPS minimal media. MOPS minimal media (10x) was prepared as described in Neidhardt et al. (187) with the addition of 20mL of micronutrient solution instead of 10mL. MOPS minimal media contained 100mL of 10x MOPS minimal media, 10mL of 0.132M K2HPO4, and 5mL of 20% D—glucose per liter. Cultures were inoculated into 50mL of MOPS minimal media with 0.1% glucose at a ratio of 1:200 and grown to stationary phase at 37°C with shaking. Cultures in MOPS were transferred to 100mL MOPS in 250mL plastic culture flasks (Nalgene, Rochester, NY.) at a ratio of 1:75 and grown for 9h before transfer again to 100mL MOPS in 250mL plastic culture flasks at a ratio of 1:30 and this culture was used to sample RNA during growth (Fig. 5.1). Cultures were sampled at 3h (ODeoo :02), 4h (ODeoo z0.45), 4.33h (ODeoo z0.73), 4.66h (ODeoo zO.93), 5h (00600 z1), 5.5h, 6h, 7h, 7.5h, and 8h. After 5h, the culture density remained at ODeoo z1. Residual dissolved oxygen tension was monitored using a dOz probe and Consort C535 multimeter (Topac Instrumentation, Hingham, Mass.) for three independent cultures in MOPS minimal medium. RNA isolation. At each time point, 4mL of culture was mixed with 8mL of RNAProtect (Qiagen, Valencia, Calif), vortexed, and centrifuged at 4°C, 113 ODeoo Exp. iL.Exp§Trn.§ E. St. Stationary .1).-- -l- 0.1 -u- l l 3 4 5 6 7 8 Time (h) Figure 5.1. Average growth of E. coli O157:H7 Sakai in MOPS minimal medium plotted as the increase in cell density, measured at ODsoo, over time. Error bars represent the standard deviation of four culture replicates. Samples were taken from the culture at the time points plotted and RNA was extracted. The growth phases are marked; Exp.= exponential growth, L. Exp.= late exponential growth, Trn = transition to stationary phase, E. St = early stationary phase, and Stationary = stationary phase. 114 7500 rpm for 10 min to pellet cells. The supernatant was removed and cell pellets stored at —70°C for 1 week or less before RNA extraction. Cell pellets were suspended in 700pL of 95°C lysis buffer (20mM sodium acetate pH 5.2, 2mM EDTA pH 8.0, and 0.5% SDS), held at 95°C for 303 then mixed with 700uL of 65°C Acid-Phenol: Chloroform, pH 4.5 (with IAA, 125:2521) (Ambion, Austin, Tex.). Samples were held at 65°C with periodic shaking for at least 6 minutes before centrifuging at 12000 rpm for 10 min. Supernatant was extracted again with acid-phenol:chloroform and then with chloroformzisoamyl alcohol (24.1 ). RNA was precipitated for at least 1 hour at —20°C in 2.5V 100% ethanol and 1/10V 3M sodium acetate pH 5.2. RNA was suspended in 1mM sodium citrate pH 6.5 and stored at —70°C. RNA samples were purified and treated with Dnase using the Rneasy kit (Qiagen). RNA quality was assessed by observing the 23s and16s rRNA bands by electrophoresis on a formaldehyde-agarose gel. RNA was checked for complete DNA digestion by PCR. cDNA synthesis and hybridizations. Reverse transcription reactions contained 6pg RNA, 2ug random primers (lnvitrogen, Carlsbad, Calif.), 1x first strand buffer (lnvitrogen), 10mM D'I'I', 400U Superscript II (lnvitrogen), 0.5mM each dATP, dCTP, and dGTP, 0.3mM dTTP, and 0.2mM amino-allyl dUTP. 30pL reactions were incubated overnight at 42°C. cDNA was purified using PCR cleanup columns (Qiagen) with Phosphate wash buffer (5mM K2HPO4, pH 8.0, 80%EtOH) and phosphate elution buffer (4mM K2HPO4, pH 8.5). Amino-allyl labeled cDNA was dried and resuspended in 0.1 M sodium carbonate pH 9.3 and coupled with either Cy3 or Cy5 (Amersham Biosciences, Piscataway, NJ.) 115 Uncoupled dye was removed by another purfication using the PCR cleanup kit. Concentration of cDNA and amount of incoporated dye was measured for each sample using a Nanodrop spectrophotometer (Ambion). Oligo microarray. The E. coli oligo set version 1.0.1 (Operon) was printed onto Corning UltraGaps (Corning Incorporated, Acton, Mass.) coated slides at the Research Technology Support Facility at Michigan State University. The Qiagen oligo set contained 5,978 probes specific for three E. coli strains,K—12 (MG1655) (25), O157:H7 Sakai (105), and O157:H7 EDL-933 (197), in which 5,943 probes were 70-mer oligonucleotides, and 35 probes had lengths that ranged from 41- 69 bp. The oligo set also contained 12 randomized negative control 70-mer oligonucleotides. There are a common set of 3807 probes that target 3807 ORFs present based on the genome sequences of E. coli K-12 MG1655 , E. coli O157:H7 EDL-933 , and O157:H7 Sakai , which we refer to as the backbone ORFs. There are also 1741 probes targeting ORFs that are specific to one or two of the genomes, typically to O157:H7 EDL-933 and Sakai. These probes are referred to as the 0157-specific ORFs. Probes designed to target K-12-specific or EDL-933 specific ORFs were not considered after the data analysis stage because these genes are not present in the Sakai genome. We used the Qiagen annotation for the probe set, and analyzed only the 4886 probes that were assigned a ECs number. Hybridization conditions. Arrays were cross-linked by exposure to 600 mJ UV before blocking in 1% SDS, 5x SSC, and 1mg/mL BSA at 42°C for 1 hour. After blocking, arrays were washed 2x 5 min in 0.1x SSC and 2x 303 in H20. Dried 116 arrays were placed into hybridization cassettes (TeIeChem lntemational, Sunnyvale, Calif.) and the cDNA samples were suspended in 10 mM EDTA, denatured at 95°C for 5 min and then mixed with 40uL of SlideHyb buffer 1 (Ambion) and loaded under a coverslip onto the array. Hybridizations were canted out at 47°C for 16-18h. After hybridization, arrays were washed in 2x SSC, 0.5% SDS 37C for 5 min, followed by 2x 5 min in 0.1x SSC, 0.1% SDS 37°C, and then 2x 2.5min in room temperature 0.1x SSC. Arrays were scanned using an Axon 4000b scanner (Molecular Devices, Sunnyvale, Calif) and images were analyzed using GenePix 6.0 (Molecular Devices). The 3h sample served as a common reference and was hybridized with all subsequent samples. Array data have been deposited at the NCBI Gene Expression Omnibus (Accession xxxx). Data analysis. Raw intensity values for all probes on each array were normalized using pin-tip LOWESS (204) in R v.2.2.1 (245) with the MAANOVA (v. 0.98-8) package (205). Signals from two replicate probes on each array were averaged and '092 transformation applied. Pairwise differences in transcription levels over time were determined using a mixed model ANOVA in R/MAANOVA. The ANOVA modeling allows for consideration of appropriate error structures for experiments with multiple sources of variation in microarray measurements (136). The random effects of the model were biological replicate and array effects, whereas the fixed effects were time point and dye effects (70). Significant changes in expression over time were determined using the Fs test (71) with the False Discovery Rate (FDR) correction applied and a cutoff p-value 117 of 0.00001. Pair wise contrasts of time points were estimated by the t—test in MAANOVA. Contrast p-values were corrected for multiple testing by using the FDR correction. Probes with significant changes in expression over time were grouped by the time points at which the significant change occurred, as determined by the contrast analysis (FDR corrected p value < 0.05). QT clustering of significant ORFs was conducted in MeV v. 3.1 (www.tigr.org) with a diameter of 0.4 and a minimum cluster size of 15. 118 RESULTS Analysis of expression ratios. The results are based on O157:H7 Sakai cultures collected from 10 time points of growth in minimal media with 4 replicates at each point for a total of 40 samples. The samples collected from the mid-exponential phase 3-hour time point (Fig. 5.1) were used as a common reference for hybridization and analysis for all subsequent time points in a blocked reference design (234). This design allowed for indirect comparison between all time points via the 3-hour samples (64, 271 ). We incorporated the dye-swaps data among the four biological replicates which can confound dye effects and biological replicate effects. However, if there is significant variation in the rate of dye incorporation from one labeling reaction to another, this would result in large dye-effects compared to biological replicate effect (137), an effect not found in our experiments. In general, the variance estimates for the biological replicate effects were small (median 0.0008), compared to the variance estimates (<0.5%) for the array effects (median 0.16). Most ORFs (4362/5886, 90%) had similar signal intensities for the mid- and late- exponential phase samples, indicating that there were few overall differences in expression between these time points (Fig. 5.2A and 5.2B). In contrast, there were many differences in expression found between mid-exponential phase samples and samples from the transition point (Fig. 5.20) and early stationary phase (Fig. 5.2D). 119 Average signal intensity for 4h samples Average signal intensity for 4.6h samples 00° 0 swearing: ,gois rifled: ii- _; O 0) l J Average signal intensity for 6h samples 3: Average signal intensity for 5h samples I I I I I l I I I I I I I 102 1o3 104 1o5 102 104 105 Average signal intensity for mid-exponential phase samples (3h) Average signal intensity for mid-exponential phase samples (3h) Figure 5.2. Average signal intensity plots for four of the time points sampled. The normalized signals of 4,886 ORFs for E. coli O157:H7 Sakai were averaged from four hybridizations, representing four biological replicates for the 3h vs 4h comparison (A.), the 3h vs 4.6h comparison (3.), the 3h vs 5h comparison (C.), and the 3h vs. 6h comparison (0.). The Lowess line is plotted for each graph in white. 120 Summary of significant changes in gene expression over time. A total of 4,886 O157:H7 Sakai ORFs were targeted on the microarray, and a significant (p < 0.00001) change in expression for at least one time point was detected for 3,050 ORFs. The false discovery rate (FDR) correction for multiple testing applies a step-wise correction to the p values, and this was used to narrow the dataset by increasing the stringency. The 2,552 ORFs with the lowest p values after correction (p-value < 1 x 10”) were considered for further analysis. Trends in gene expression identified by significant contrasts between subsequent time points and QT clustering. Expression profiles of the 2,552 ORFs with significant differences were clustered into 12 groups using QT clustering (Fig. 5.3), which groups similar expression profiles based on jackknife correlations (113). The majority of ORFs were classified into 12 groups; only 84 were not placed into a cluster. ORFs were also classified based on the time intervals at which a significant change in gene expression occurred. A subset of 524 ORFs had differences in expression from mid- to late-exponential phase (3h to 4h) (Fig. 5.4). The majority of ORFs, 1,668 of 4,886 (34%), had significantly changed in transcript level at the stationary phase transition point (4.6 to 5h). Additionally, 1,045 of 4,886 (20%) ORFs significantly changed during the transition point to early stationary phase (5 to 5.5h). Significant changes in exponential phase — response to limiting oxygen. There are 524 ORFs with significant differences in transcript level during 121 3 3 3 QTC 1 - 1078 genes QTC 2 - 923 genes QTC 3 - 114 genes 2 -i- 2 -l- 2 -- i 1 a 1 -~ 1 +- 0 o «r o -- o -- 3.1 -- .1 - .11 2 -~ -2 -- 2 -- -3 i i i l t s a t i t i i i e : i 1 i : s 3 3 3 QTC 4 - 96 genes QTC 5 - 54 genes QTC 6 - 47 genes 2 e- 2 -- 2 -- .5 1 "" 1 - 1 -r- is? go -' 0 -u- 0 —n— N 3-1 “F .1 -I— -1 -1— -2 -- -2 ~- -2 -— 3 : i i I i 4. .3 i i i I i . -3 + r i i 1 1 3 QTC 7 - 43 genes 3 QTC 8 - 32 genes 3 QTC 9 - 22 genes 2 " 2 -- 2 .— c 1 —L 1-- 1 -~ .9 g o a o -— 0 -~ , i N -1 -- 1 -- 1-r 3 -2 -1l— .2 -i- 2 di— -3 i t 1‘ t 4. t -3 ‘ ‘r t i i i 3 I t i I l t 3 QTC 10 - 21 genes 3 QTC 11 - 19 genes 3 QTC 12 - 17 genes 2 —- 2 -- 2 ._ g 1 -- 1 -- 1 -- a 5 o ~~ o -- o -- N 3-1 -- 1 -*- 1 ~1- 2 A— 2 + 2 ->— -3#ll#ti-3.tiiI%-3trtttr 3 4 5 6 7 8 3 4 5 6 7 8 3 4 5 6 7 8 Tune (h) Time (h) Time (h) Figure 5.3. QT clusters of the 2,552 significant ORFs. Average expression profiles were determined for the ORFs in each cluster and plotted with the standard deviation for each time point. 122 3-4h 4-4.3h 4.3-4.6h 4.6-5h 5-5.5h 5.5-6h Percent of ORFs with significant increase in expression Percent of ORFs with significant decrease in expression a 20- — backbone genes —0157-specific genes 25 Figure 5.4. Proportion of backbone genes (those common to E. coli K12 and O157:H7) and O157:H7 specific genes that are significantly differentially expressed, either increasing or decreasing, at specific time intervals. The ORFs with significant changes in expression between two time points were determined from the contrast analysis. 123 Table 5.1. Significant ORFs (p value < 1 x 107) with greater than 4-fold significant change between 3h and 4h. I092 Ecs change in OT number Genea function expression clusterb ECsO236 0157 Unknown function -2.89 10 ECsO482 cyoE protoheme IX farnesyltransferase -2.08 10 ECsO483 cyoD cytochrome c ubiquinol oxidase subunit IV -3.18 10 ECsO484 cyoC cytochrome c ubiquinol oxidase subunit lIl -3.32 10 ECsO485 cyoB cytochrome c ubiquinol oxidase subunit l -3.87 10 ECsO486 cyoA cytochrome c ubiquinol oxidase subunit II -3.89 10 ECsOGGO dcuC transport of dicarboxylates 2.20 6 ECsO746 ' M succinate dehydrogenase, cytochrome b556 -3.15 10 ECsO747 sth succinate dehydrogenase, hydrophobic subunit -2.66 10 ECsO748 MA succinate dehydrogenase -3.06 10 ECsO749 s_dfi§ succinate dehydrogenase -2.79 10 ECsO750 - orf, hypothetical protein -2.24 10 ECs0751 sucA 2-oxoglutarate dehydrogenase -2.20 10 ECsO768 gm cytochrome d terminal oxidase 2.56 4 ECsO769 9mg cytochrome d terminal oxidase 2.58 4 ECsO916 yliH putative receptor 3.84 2 ECsOQ79 dmsA anaerobic dimethyl sulfoxide reductase 3.71 4 ECsOQ81 dmsC anaerobic dimethyl sulfoxide reductase 2.08 4 ECs0986 pflB formate acetyltransferase 1 2.04 6 ECs1572 pepT putative peptidase T 2.64 2 ECs1728 narK nitrite extrusion protein 2.98 6 ECs1729 narG nitrate reductase 1, alpha subunit 2.25 4 ECs1741 adhE CoA-linked acetaldehyde dehydrogenase 2.50 6 ECs1756 ych putative outer membrane protein 4.15 4 ECs1875 ych probable amidotransferase subunit -2.28 2 ECs2027 ydcl putative transcriptional regulator LYSR-type -2.41 2 ECsZO78 fdnG formate dehydrogenase-N, nitrate-inducible 2.30 6 ECsZ150 yde orf, hypothetical protein 4.70 4 ECsZZ93 ynfE putative oxidoreductase, major subunit 2.83 6 ECsZ457 yde orf, hypothetical protein 2.03 6 ECs2463 ynjE putative thiosulfate sulfur transferase 3.47 1 E032614 yecH orf, hypothetical protein 3.09 6 ECs3061 fruB PTS system, fructose-specific Ilepr component 2.02 1 ECs3084 dst disulfide oxidoreductase 2.04 1 ECs3088 ccmC heme exporter protein C 2.19 1 ECs3090 ccmA ATP binding protein of heme exporter A 2.64 1 EC33091 napC cytochrome c-type protein 3.55 1 ECs3092 napB cytochrome c-type protein 3.30 1 ECs3093 napH ferredoxin-type protein: electron transfer 3.04 un ECs3180 ackA acetate kinase 2.28 6 ECs3226 ych orf, hypothetical protein 3.02 6 ECs3445 yflD putative formate acetyltransferase 4.48 6 ECs3460 yfiA putative yhbH sigma 54 modulator 2.12 2 ECs3582 hypA pleiotrophic effects on 3 hydrogenase isozymes 2.92 6 124 Table 5.1, continued. log2 Ecs change in QT number Genea function expression clusterb ECs3583 hypB guanine-nucleotide binding protein 2.10 6 plays structural role in maturation of all 3 EC53586 hypE hydrogenases 2.10 6 ECs3799 O157 orf; hypothetical protein 2.92 6 ECs3800 O157 orf; hypothetical protein 2.33 6 ECs3802 O157 putative ATP-binding protein of ABC transport system 2.40 6 ECs3833 ansB periplasmic L-asparaginase II 3.41 4 probable cytochrome NilFe component of ECs3880 hbe hydrogenase-2 2.05 4 ECs4040 yth orf, hypothetical protein 3.11 1 ECs4216 nirB nitrite reductase (NAD(P)H) subunit 3.55 1 ECs4343 £154 periplasmic binding protein for nickel 3.46 1 ECs4344 yi_kB_ transport of nickel, membrane protein 2.10 6 ECs4347 LlrE ATP-binding protein of nickel transport system 2.85 1 ECs4398 yth putative cytochrome C peroxidase 2.86 6 ECs4456 yial orf, hypothetical protein 2.40 4 ECs4750 yigl orf, hypothetical protein -2.55 10 ECs4834 sodA superoxide dismutase, manganese -2.40 7 ECs4874 gldA glycerol dehydrogenase, (NAD) 2.86 6 ECS5052 nrfA periplasmic cytochrome c(552) 3.08 un ECsSO53 nrfB formate-dependent nitrite reductase 2.86 6 ECs5054 nrfC formate-dependent nitrite reductase; Fe-S centers 2.96 un fumarate reductase, anaerobic, iron-sulfur protein ECsS134 frdB subunit 2.50 8 EC55135 frdA fumarate reductase, anaerobic, flavoprotein subunit 2.11 8 ECS5214 nrdG anaerobic ribonucleotide reductase activating protein 2.02 6 ECs5215 nrdD anaerobic ribonucleoside-triphosphate reductase 3.54 6 ECs5298 yjiM orf, hypothetical protein 2.76 8 “ Genes in bold have been identified as being under the control of the transcriptional regulator FNR (66, 132). Underlined genes have been identified as being controlled by the transcriptional regulator ArcA (156). Genes with “0157" indicate that these are 0157-specific ORFs. b QT cluster un. means unassigned 125 exponential phase - a total of 113 genes changed only in exponential phase, whereas the remainder had altered levels of expression in exponential as well as stationary phase. ORFs that decreased in expression included those encoding enzymes of the TCA cycle, such as sucA, and sdhABCD (Table 5.1, Fig. 5.3, QTC 10). Many of the ORFs (94/524, 18%) that changed in transcript level are regulated by the transcription factor FNR (66, 132), whereas a smaller set of ORFs (29/524, 5%) are known to be regulated by ArcA (156). Genes that increased included those encoding anaerobic electron acceptors, fumarate reductase, frdAB, and DMSO reductase, dmsA and dmsC, as well as a DMSO reductase paralog, ynfE (Table 5.1, Fig. 5.3, QTC 6). Genes encoding the cytochrome c biogenesis system (ccmABCDEFG) and the cytochrome c-like protein (napB and napC) also had a significant increase from 3h to 4h. The operon encoding the low affinity cytochrome oxidase bo (cyoABCD) decreased 9 to15- fold in expression from 3h to 4h (Table 5.1). Concomitantly, expression of the high affinity cytochrome oxidase bd, cydAB, increased 4-5 fold. Expression of the high and low affinity cytochrome oxidases is directly influenced by the level of available oxygen in the media (4, 250). The altered expression of the cytochrome oxidases as well as the increase in F NR-controlled, anaerobic associated genes and decrease in TCA cycle enzymes is consistent with the hypothesis that oxygen became limiting under the culture conditions that were used. Subsequent measurement of the residual dissolved oxygen tension in the culture during growth indicates that residual dissolved 02 levels decreased almost to zero (95% decrease in 02) during exponential phase (Fig. 5.5). As 126 cultures entered stationary phase, the O2 levels began to increase, as the demand for 02 by the culture decreased. Significant change in 0157-specific genes during exponential phase. In contrast to the backbone genes, only a small proportion (6%) of the 0157- specific ORFs changed significantly in expression measured during exponential phase (Fig. 5.4). While statistically significant, most of the differences in expression level for these 61 genes were not substantial ranging from 1.2- to 2- fold. Many of these ORFs are of unknown function, but of the 45 0157-specific ORFs that increased from 3-4 h, 9 of them were from a single Sakai phage (Sp) 15 region which encodes the Shiga toxin 1 phage, and contains 5 ORFs from Sp 5 which encodes the Shiga toxin 2 phage. Interestingly, ECs2381 a locus whose expression has been shown to contribute to colonization in the bovine gastrointestinal tract (82), was also significantly induced 1.8 fold from 3 to 4 h. There were 4 additional 0157-specific ORFs that had a greater than 4-fold change in expression from 3-4 h (Table 5.1). These 4 ORFs were not associated with any of the Sakai phages, and have unknown or putative functions. Genes with significant changes in expression at the stationary phase transition point. The majority of significant changes in transcript level occurred at the stationary phase transition point (Fig. 5.4). There were 176 ORFs with a 4- 127 1O 8.. Q 6' U) a 5 4- 9 2- 0— Time (h) Figure 5.5. Residual dissolved oxygen in the MOPS minimal medium during growth of E. coli O157:H7 Sakai. Mean rDOT from three independent cultures is plotted with the error bars indicating the standard deviation. 128 Table 5.2. Significant ORFs (p value < 1 x 107) with greater than 4-fold significant changes in expression at the stationary phase transition point (4.6 to 5h). logZ , logZ change in OT change in QT Ecs no. Gene8 expression cluster Ecs no. Genea expression cluster ECsOO34 dapB -2.63 1 ECs1286 0157 -2.09 1 ECsOO35 fllfi -4.23 1 ECs1426 mdoG -2.46 5 EC30068 araC 2.57 2 ECs1440 gy_r_C -2.65 1 ECs0072 tpr -2.25 1 ECs1468 plsX -2.16 1 ECsOO83 fruL -2.06 1 ECs1479 ptsG -3.09 1 EC50124 speD -2.74 1 ECs1492 mfd -2.33 1 ECsO149 dksA 2.12 2 ECs1604 ych -2.19 1 E0501 50 still 2.83 2 ECs1662 0157 -2.53 1 ECsO191 yan 2.42 2 ECs1684 dadA 2.34 2 ECsOZO7 dniR -2.11 1 ECs1710 ychH 4.61 2 ECs0334 0157 -2.07 1 ECs1712 m; -2.68 1 ECsO374 0157 2.42 2 ECs1859 mb -2.70 1 ECsO401 mhpR 4.33 3 ECs1954 0157 -2.61 1 ECsO416 0157 3.55 2 ECs1955 0157 -3.22 1 ECsO468 rin -2.27 1 ECs2011 yan 2.34 2 ECsO483 cyoD 2.06 10 EC32021 all/3 3.60 2 ECsO485 cyoB 2.01 1 0 E052027 ydcl 3.35 2 ECsO489 LOLA 2.10 2 ECs2031 ych 2.99 2 EC30584 purK -2.27 1 ECsZO37 ych 2.71 2 ECsOS85 purE -2.66 1 ECs2083 sch -2.18 1 EC50610 cusC -2.31 1 ECle82 0157 3.21 2 ECsO61 1 cusF -2.82 1 EC52293 ynfE -2.97 6 ECsO61 2 cusD -2.62 1 ECs2296 yan —2.03 6 ECsO61 3 yde -2.70 1 ECsZ362 lhr -2.16 1 ECs0670 dacA -2.32 1 ECs2367 purR -2.51 1 ECsOB81 mg 2.78 2 EC32431 yniA 3.52 2 ECsO773 tolR -2.29 1 ECsZ433 yniC 2.44 3 ECsO782 aroG -3.09 1 ECs2445 osmE 2.33 2 ECsO800 yth -2.13 1 ECs2510 yeaV 2.06 3 ECs0814 O1 57 -2.1 5 1 ECs2692 yodD 2.96 2 ECsO869 yth 3.14 2 ECs2714 espJ -2.01 1 ECs0890 gas 2.64 2 ECsZ737 pchC 3.43 2 ECs0943 artJ -2.81 1 ECs2814 yeeD -2.09 1 ECsO966 M 2.75 2 ECsZ816 yeeF -2.55 1 EC30968 clpA 2.89 3 ECs2839 0157 -2.23 1 ECs0986 pflB -2.61 6 ECs2840 wde -2.14 1 ECs1008 MB 2.15 2 ECs2847 0157 -2.25 1 ECs1029 Mg -2.88 1 ECs3038 yeiT 2.14 3 ECs1037 rmf 2 .24 2 ECs3060 fruK -2.1 3 1 ECs1072 0157 -2.13 5 ECs3061 fruB -2.38 1 ECs1091 0157 3.24 2 ECs3136 yfaX -2.04 1 E031 137 yccC -2.27 1 ECs3149 menC -2.06 1 ECs1138 ych -2.78 1 ECs3180 ackA -2.33 6 E051 139 ych -2.74 1 ECs3194 argT 4.08 2 129 Table 5.2. continued. Iog2 logZ change in OT change in QT Ecs no. Gene“ expression cluster Ecs no. Gene“ expression cluster ECs3196 purF -2.94 1 E034188 hopD 2.38 2 E033197 cva -3.35 1 E034192 M -2.57 1 E03321 2 mepA -2.29 1 E034193 [ps_L -2.67 1 E033287 ptsH -2.66 1 E034204 yheT -2.00 1 E033296 cysP -2.17 1 E034249 yth -2.53 1 E033361 purM -4.24 1 ECs4294 M 2.48 2 E033387 sseA 2.18 2 ECs4343 nikA -3.22 1 E033403 hcaR 2.50 3 ECs4347 nikE -2.38 1 ECs3423 purL -2.45 1 E034366 )thg 4.26 2 ECs3448 tGC 2.24 2 E034481 lIdP 2.39 2 ECs3463 tyrA 3.67 3 E034490 gpml -3.35 1 ECs3464 aroF 3.30 3 E034512 [mn_B -2.38 1 ECs3471 yle 358 1 E034553 0157 -2.10 1 E033546 eer -3.47 1 E034575 9300 -2.00 1 E033559 sr1A_1 2.08 3 E034632 yidA -2.02 1 E033595 rpoS 2.80 2 E034638 rpmH -2.47 1 E033606 M -2.59 1 ECs4639 mpA -2.92 1 E033640 pyrG -2.98 1 E034679 gtLE -2.33 1 E033659 fucO 2.17 3 ECs4716 rho -2.56 5 E033675 m -4.09 1 E034720 wecC -2.34 1 E033677 r908 21 7 1 E034741 xerC -2.08 1 E033701 yqu 2.50 3 E034756 yigL -2.18 1 E033743 ygeW 2.08 2 ECs4759 metE -2.76 1 E033749 yqu 2.45 2 E034791 glnL -2.80 1 ECs3799 O1 57 -2.69 6 ECs4870 m_etE -3.65 1 E033800 O1 57 -2.43 6 ECs4885 gag -3.62 1 ECs3802 O1 57 -2.43 6 ECs4887 argC -3.99 1 E03381 1 yggG 2.28 2 ECs4888 argB -3.62 1 E03381 8 metK -3.03 1 E034893 O1 57 2.09 2 ECs3840 nqu 2.44 2 E03490? M -3.02 1 E033931 g_lg§ 2.94 2 5034908 LU -2.64 1 E034052 argG -2.68 1 ECs4909 M -2.60 1 E034103 rpsl -2.28 1 ECs4929 purH -4.14 1 E0341 32 yth -2.54 1 E034931 LnflA; -3.78 1 E0341 50 3mg 2.24 2 E034932 aceB 5.35 2 ECs4156 mscL 2.01 2 E035004 0157 2.14 2 E034161 M -2.17 1 E035046 ych -3.81 1 ECs4162 IpsK -2.36 1 E035051 acs 5.55 2 E034164 rpmJ -2.03 1 E035067 O1 57 2.21 2 E034165 prlA -2.53 1 E035153 M -2.06 1 130 Table 5.2, continued. logz logZ change in OT change in OT Ecs no. Gene“ expression cluster Ecs no. Gene‘3 expression cluster E035164 yij 4.54 2 E035240 ngR 2.24 2 E035192 cysQ 2.01 2 E035259 O1 57 -2.09 1 E035231 argl -3.29 1 E035354 rob 2.15 2 E035235 vaIS -2.58 1 ECs5360 yti 2.03 2 a Gene names in bold have been found to be regulated by RpoS (142, 195, 266) and those underlined are genes that increase or decrease in stationary phase in E. coli K-12 (57). Genes with “0157” indicate that these are 0157-specific ORFs. 131 fold or greater change in expression at the stationary phase transition point (Table 5.2). 154 ORFs that significantly increased or decreased during the transition to stationary phase (4.6-5h) also increased or decreased in early stationary phase (5-5.5h) (Table 5.3). A decrease in expression of ribosomal genes and genes involved in nucleotide and amino acid synthesis was observed from 4.6-5h and 5-5.5h (Tables 5.2 and 5.3, Fig. 5.3 QTC 1). Many of the genes that were significantly down-regulated upon entry into stationary phase have been also observed in E. coli K-12 (57) (Tables 4.2 and 4.3). Expression of rpoS, encoding the stationary phase sigma factor, increased 7 fold during the transition point and the rpoS transcriptional activator, dksA, increased 4.3 fold at the same interval (Table 4.2). Increased expression of genes known to be regulated by RpoS, such as bolA, dps, osmE, ath and ung occured at the stationary phase transition point (Tables 5.2 and 5.3, Fig. 5.3 QTC 2). The greatest increase in transcriptional level for an upregulated gene (45-fold) during the stationary phase transition was acs, encoding acetyl CoA synthetase (Table 4.2). Acs expression is controlled in part by RpoS (226). Also induced at a very high level during the transition point into stationary phase were 8098 (40 fold) and 309A (32 fold), which encode components of the glyoxalate shunt. Increased expression of acetyl CoA synthetase as well as genes of the glyoxalate shunt is consistent with the hypothesis that cells are scavenging acetate produced during rapid growth on glucose in exponential phase. 132 Table 5.3. Significant ORFs (p value < 1 x 107) with greater than 4-fold significant changes in expression from late exponential to early stationary phase (4.6 to 5.5h) Log2 Log2 change in OT change in OT Ecs no. Gene3 expression cluster Ecs no. Gene8 expression cluster E030026 rpsT -4.14 1 E032450 ydjS 3.94 2 E0301 18 806E -3.27 1 E032451 - 4.1 1 2 5030123 yacL 3.1 7 2 E032452 - 5.88 2 E030172 {if -4.58 1 E032453 - 4.62 2 E030173 p yrH -3.15 1 E032454 a_rgQ 6.97 2 E030199 yaeC -3.14 1 E032463 - -4.04 1 E030201 abc -3.32 1 E032467 gdhA -5.15 1 E030248 fadE 4.57 2 E032492 mg 4.08 2 E030383 1ah_0 4.53 2 E032527 manX -2.95 1 E030384 prpR 3.63 2 E032528 manY -3.21 1 E03041 5 aqu 5.21 2 E032546 - 4.09 2 E030502 deA -3.80 1 E032609 araF 3.61 2 E030540 0157 3.52 2 E032613 ltn -3.88 1 E030541 0157 4.1 1 2 E032888 yegP 4.47 2 E030990 serC -3.29 1 E032892 yegS 3.38 2 E030991 aroA -4.24 1 E033022 yohc 6.12 2 ECs1050 yccV 3.88 2 ECs3029 0157 4.04 2 ECs1266 phoH 7.69 2 E033030 0157 3.38 2 ECs1438 yceP 4.78 2 E033077 M -3.64 1 E031466 yceD -2.86 1 E0331 14 gyrA -3.38 1 E031603 QLLB -3.58 1 E033181 pta -3.50 1 ECs1683 y_cg§ 5.36 2 E033224 fadJ 4.26 2 ECs1831 yciG 2.91 2 ECs3225 fadl 4.50 2 ECs1849 aan 4.07 2 E033227 [Q 5.39 2 E031874 ych 4.73 2 ECs3259 - 3.57 2 ECs1875 ych 5.14 2 E033271 - 3.06 2 E031997 ynaF 3.42 2 ECs3286 gals -3.68 1 E032028 ych 3.64 2 E033338 purC -5.04 1 E032036 ych 3.78 2 E033401 csiE 3.77 2 E032044 ydcs 5.22 2 ECs3445 yfiD -6.36 6 E032045 ydcT 5.23 2 ECs3465 yflL 3.81 2 E032082 ath 3.62 2 E033469 lplS -4.28 1 E032092 - 3.70 2 E033470 gln_D -5.46 1 E0321 1 3 O1 57 -3.24 1 E033472 rpsP -5.57 1 E032120 - 4.17 2 E033520 ygaT 4.88 2 E032145 ydel 3.74 2 E033523 gabT 3.14 2 E032150 - -5.10 4 E033619 913;! -4.11 1 E032295 - -2.81 6 ECs3742 ygeV 3.64 2 E032346 tyrS -2.97 1 E033750 191.1 4.14 2 E032385 ynhG 3.80 2 E033771 - 3.52 2 E032389 ynhD 3.58 2 ECs3784 serA -3.66 1 133 Table 5.3, continued L092 L092 change in OT change in QT Ecs no. Gene3 expression cluster Ecs no. Genea expression cluster E033810 tktA -4.16 1 E034433 yij 3.88 2 E033896 thE 2.86 2 E034464 aldB 3.02 2 E033948 rpsU -3.14 1 ECs4473 yin 3.78 2 E033955 yng 5.55 2 E034475 mtlA 2.84 2 E033963 ygjL 4.06 2 E03461 1 iIvN 4.84 2 ECs4034 yth 3.67 2 ECs4612 ilvB 4.41 2 ECs4040 yth -3.40 1 E034633 yidB 3.85 2 E034091 gig -3.19 1 ECs4675 QILG -3.77 1 E034104 M! -4.05 1 ECs4676 _a_tLA -4.02 1 E0341 12 yth 3.56 2 E034677 gal-l -4.50 1 E034128 accC —3.51 1 ECs4678 QILF -4.08 1 E034133 173 -3.76 1 E034680 '21th -3.27 1 E0341 59 m -3.99 1 E034686 asnA -3.68 1 ECs4160 [2% -4.1 1 1 ECs4758 metR -3.26 1 E034166 rpIO -3.63 1 ECs4774 fadB 6.58 2 E034167 rme -3 .70 1 E034792 glnA -5.34 1 E034168 rpsE -3.75 1 ECs4884 yijP -3.62 1 E034171 rpsH -4.00 1 E034889 argH -4.04 1 E034173 LIE -4.26 1 E034906 rle -4.80 1 E034174 M -4.14 1 E034928 purD -4.58 1 ECs4175 m_N -4.46 1 E034933 aceA 4.98 2 E0341 76 4:30 -4.30 1 E034934 aceK 2.90 2 ECs4177 rme -4.12 1 E035007 lysC -3.43 1 E034179 4330 -5.07 1 ECs5049 actP 4.17 2 ECs4180 [M_\/ -4.74 1 E035050 yjc_H 3.78 2 E034181 113§ -4.14 1 E035089 ghnB 4.22 2 E034182 M -4.83 1 E035109 yde 3.07 2 E034183 mm -5.75 1 ECs5123 groE§ -3.66 . 1 E034184 rpID -6.50 1 E035176 IpsF -4.84 1 E034185 rplC -4.69 1 E035177 mg -5.00 1 E034186 rsz -5.93 1 E035178 M -4.46 1 E034189 0157 3.24 2 E035179 _rb_ll -3.90 1 E034191 rug -3.27 1 E035194 yth 4.59 2 E034212 fic 4.44 2 E035205 m 6.29 2 E034213 yth 4.1 3 2 E035222 MB. -2.95 1 E034299 uggB 6.46 2 E035298 yjiM -1 .43 8 a Gene names in bold have been found to be regulated by RpoS (142, 195, 266) and those underlined are genes that increase or decrease in stationary phase in E. coli K-12 (57). Genes with “0157” indicate that these are 0157-specific ORFs. 134 0157-specific genes that decreased in the transition to stationary phase include those encoding the 0157 LPS antigen (E032839, E032840, and E032847) with a greater than 4 fold decrease in expression (Table 5.3, Fig. 5.3 QTC 1) as well as E032835, ECs2836, E032838, E032841, E032844, and E032845 (1.6 to 3 fold decrease) from that genomic region. E032113, part of the F9 fimbrial operon (157), decreased in expression from exponential to stationary phase (Table 5.3). Other ORFs in this operon, E032107-2112, also decreased significantly (~1.5 fold) upon entry into stationary phase. 0157-specific ORFs that increased in stationary phase included E032182, E032737, and E035067, which encode putative transcriptional regulators. E032737 (pchC) encodes a perC-Iike homolog reported to regulate LEE transcription (122). E030415, encoding a perimplasmic ferric iron transporter, had the greatest increase in expression of the 0157-specific ORFs, with a 37-fold increase during the transition point (Table 4.2). Genes with transient expression at the stationary phase transition point. A subset of 120 ORFs were expressed transiently from late exponential to early stationary phase. These ORFs changed significantly during the 4.6 to 5h time interval, and then again during the 5 to 5.5h interval, but in the opposite direction (Fig. 5.3 QTC 3, 11, and 12). Genes with this pattern of transcription included many involved in nutrient scavenging and turnover. Genes encoding both the anaerobic (glpABC) and aerobic (glpD) sn-glycerol 3 phosphate dehydrogenase increased >4 fold during the transition into stationary phase, and then decreased 135 significantly in early stationary phase (Table 5.4). This pattern is consistent with the understanding that the glp metabolic system serves as a salvage pathway for glycerol derived from degradation of phospholipids and triacylglycerol. The murPQ operon was expressed transiently, with a 5.3 fold increase followed by a 2 fold decrease in early stationary phase. The products of murPQ are involved in peptidoglycan turnover (253) suggesting that membrane components are being recycled and utilized for energy. A dicarboxylate transporter, dctA, was induced 24-fold upon entry into stationary phase, then decreased 4-fold in early stationary phase (Table 5.4). Only 12 O1 57-specifc ORFs were expressed transiently upon entry into stationary phase. ECs1772, a putative intestinal colonization factor, decreased upon entry into stationary phase, then significantly increased in expression (>4 fold increase from 5 to 5.5h) (Table 5.4). Genes with a significant change in expression in early stationary phase. The interval with the second largest number of significantly modulated ORFs was in early stationary phase, during the 5 to 5.5h interval (Fig. 5.4). A number of these ORFs that had significant increases in expression levels during this interval are transcriptional regulators, such as the AraC-like regulators gadW and gadX, involved in regulation of the glutamate decarboxylase system of acid resistance (Table 4.5). Genes encoding DNA binding proteins, such as hha and cpr, also increased in transcription in early stationary phase. Components of the glycerol metabolism pathway were activated in early stationary phase, including ung 136 Table 5.4. Significant ORFs (p value < 1 x 107) with transient expression at the stationary phase transition point with expression changes greater than 4-fold. log2 logz change in change in expression expression QT Ecs no. Genea 4.6-5h 5-5.5h cluster E030057 deA -2.43 0.66 1 E030287 0157 -1.09 2.91 2 E030358 betB 3.07 -1 .25 3 E030359 betl 2.99 -1 .48 3 E030418 0157 2.19 -0.64 3 ECsO486 cyoA 2.77 -1 .08 10 ECsO746 sdhC 2.93 -1 .74 10 E030747 sth 2.55 -1 .59 10 ECsO748 sdhA 2.76 -1 .38 10 E030778 nadA -2.57 0.76 1 E030946 artl -2.66 0.40 5 E030987 focA -3.38 1 .1 1 6 E030994 rpsA -3.78 0.66 1 E031261 putP 2.80 -0.92 10 E031741 adhE -4.01 1.17 6 ECs1744 oppB -2.05 1 .36 7 E031772 0157 -0.68 2.22 2 E032383 pku -2.42 1 .49 un E032387 sufS 0.71 2.17 2 E032481 yde -2 .44 0.84 6 E032482 ydjl -2.18 0.81 6 E032486 yeaC 3.00 -0.57 2 E032487 yeaA 3.08 -0.60 3 ECs3042 mng 3.96 -1 .13 3 ECs3043 galS 2.14 -0.87 3 E033117 nrdA 1.83 -3.64 1 1 E033125 ngT 1 .65 -2.22 4 E033126 ngA 3.61 -2.78 3 E03312? glpB 2.59 -2.15 3 ECs3128 glpC 2.92 -2.94 un ECs3142 ytbF -2.54 1.59 12 ECs3143 mm -2.50 1 .03 5 E033144 yfbH -2.66 0.98 5 ECs3299 yer 2.39 -0.89 3 E033300 yfeV 2.28 -0.99 3 137 Table 5.4 continued. log2 logZ change in change in expression expression QT Ecs no. Genea 4.6-5h 5—5.5h cluster E033305 ypeA 2.05 -0.81 3 E033393 hscB -1 .37 2.51 7 E033748 yqu 2.23 -0.78 3 ECs4109 mdh 2.05 -1.40 un ECs4269 glpD 4.17 -1 .82 3 ECs4408 dctA 4.59 -1 .93 3 ECs4498 rfaF -2.10 0.85 12 ECs4705 ilvD -2.03 1 .03 12 ECs4726 yifM 3.51 -3.62 un ECs4750 yigl 2.77 -1.14 10 E034840 fieF 3.76 -3.29 3 E034851 ngK 3.54 -3.73 4 E034852 glpF 5.22 -4.1 0 3 E034918 yjaE 4.21 -1.58 3 E035298 yjiM 0.71 -2.14 8 E035313 yti 1.38 -3.28 un 138 3 Genes with “0157” indicate that these are 0157-specific ORFs. Table 5.5. Significant ORFs (p value < 1 x 107) with greater than 4-fold expression changes in early stationary phase. logz log2 changein changein expression QT expression QT Ecs no. Jene 5-5.5h cluster Ecs no. gene 5-6h cluster E030342 ykgC 2.30 2 E030568 90! 4.04 2 E030513 hha 2.55 2 E030569 gip 3.56 2 E030514 ybaJ 2.16 2 E030570 ybe 4.89 2 E030732 ybgA 2.27 2 E030571 ybe 3.27 2 E030733 phrB 2.15 2 E032095 yddV 4.27 2 ECsO769 cydB -2.03 4 E032122 ydeZ 4.88 2 E030957 poxB 2.47 2 E032123 yneA 5.14 2 E030973 tpr -2.25 1 E032124 yneB 4.46 2 ECs1155 cpr 2.26 2 E032705 yedU 4.54 2 E031236 IomW 2.21 2 E031 756 yciD -3.21 4 E032047 ych 3.59 2 E032048 ych 2.35 2 E032386 ynhA 2.75 2 E032388 ynhC 2.89 2 E032493 yeaH 4.26 2 E032529 manZ -2.65 1 E033395 ich 2.61 2 ECs3396 yfllO 2.04 2 ECs4046 rpsO -2.20 1 E034295 ugpQ 2.20 2 ECs4296 ung 2.60 2 E034297 ung 2.12 2 ECs4305 livK -2.23 1 ECs4395 gadW 2.10 2 ECs4396 gadX 2.76 2 E034534 intL 2.50 2 E034673 atpC -2.24 1 E035108 yjdl 2.20 2 E035269 yth 3.09 2 139 and ung, components of the glycerol 3 phosphate transporter, and ugpQ, the cytosolic glycerophosphoryl diester phosphodiesterase (Table 5.5). Expression of LEE island genes upon entry to stationary phase. Of the 41 LEE island ORFs, 23 had a significant (p < 0.00001) differences in expression in growth into stationary phase under minimal media (Fig. 5.6). Most of these ORFs were downregulated in transition to stationary phase (between 4.6 and 5h). Three of these ORFs, espB, espD, and espZ, also had a significant change in expression in early stationary phase (5h to 5.5h). The greatest reduction in expression was seen in espB and espD, which both decreased 5.5- fold from late log to early stationary phase. Expression of espA also decreased > 5- fold over this interval, but the error variance for the espA probe was large and thus was not deemed significant. While the overall trend for the LEE island genes was a decrease in expression during the transition to stationary phase, the exception was espZ which had a 3.2 fold increase in expression from late log to early stationary phase, indicating that this ORF may be activated by stationary phase associated regulators. Expression changes of LEE effectors. More than 60 putative effectors that are translocated through the LEE-encoded TTSS were recently identified (247) and of these TTSS effectors, 33 were found here to have significant (p < 0.00001) changes in expression over time (Fig. 5.7). A total of 7 TTSS effector genes 140 LEE3 Int: : - Figure 5.6. Heatmap of gene expression over time for the ORFs of the LEE island. An asterisk next to an ORF name denotes that expression levels for at least one time point was statistically significantly different from expression levels for the other time points for that ORF. 141 UEE‘J‘. : ,- 1092 1.8 0 1092 -1.8 Figure 5.7. Expression heatmap of TTSS effector ORFs that had a significant change in expression over time. 142 Table 5.6. Significant changes in transcript levels of type III secretion system effectors. E03 number Gene name Logz change in expression from 4.6h to 5.5h E030061 espY1 3.03 E030073 esp Y2 1 .24 ECs1825 espM1 1 .47 E031994 nleGZ-Z 1 .41 E0321 56 nleGZ-3 1 .63 E032229 nleGZ-4 -1 .1 5 E032714 espJ -2.18 E032715 tccP -1 .39 ECs3488 nle66-3 1 .1 1 ECSs4554 espB -2.41 E034561 tir -1.72 E034571 espZ 1 .69 E034643 espL3 1 .77 143 (nleA, nleG, espM1, espM2, nleGZ-2, eSpZ, and nleGZ-3) increased in expression between mid- and late- exponential phase (3h to 4h). At the stationary phase transition point, 11 ORFs increased in expression and 7 ORFs had a significant decrease in expression (espJ, tccP, tir, espF, espB, espM2, and nleD) (Fig. 5.7). In early stationary phase, eleven ORFs were upregulated, with the exception of espB (2.5 fold decrease). From late exponential to early stationary phase (4.6h to 5.5h), 8 ORFs had an increase in expression > 2- fold, with espY1 increasing 9 fold. Over this same time interval, 5 'I'I'SS effector ORFs had a 2 fold or greater decrease in expression, with espB decreasing 5.3 fold (Table 5.6). Significant increase of genes from the acid fitness region in stationary phase. A cluster of 12 genes located in the E. coli K-12 genome at position 3652706 to 3665603 bp has recently been termed the acid fitness island (AFI) (117). In E. coli O157:H7 Sakai, the homologous region contains a 9 kb insertion between yhiF (ECs4378) and yhiD (E034388). This AFI genomic region contains multiple transcriptional regulators that control expression of the glutamate dependent acid resistance (GDAR) system, such as GadE, GadX, and GadW, that are also known to influence LEE expression (225, 243). The microarray comparisons demonstrate that the expression of the AFI genes increased significantly from exponential to stationary phase, and included some of the greatest increases in transcript level among the whole genome (Fig. 5.8). 144 3h 4h 4 3h ECs4377 slp ECs4378 yhiF ECs4388 yhiD E034389 hdeB ECs4390 hdeA I " I ECs4391 hdeD E034392 gadE m E034393 yhiU ECs4394 yhiV E034395 gadW ECs4396 gadX ECs4397 gadA u IE; 1;- .- |092 3 0 km 4 Figure 5.8. Expression heat map of 12 genes from the acid fitness region (AFI). Expression values determined from the ANOVA analysis are represented colorimetrically, with dark red representing expression = 3 and dark green representing expression = -4 on a log2 scale. Over half of the AFI genes (7/12) had a significant increase in expression from mid- to late- exponential phase 145 Transcript levels of gadA increased 111 fold and levels of gadE increased 85 fold from mid-exponential (3h) to stationary phase (6h). Over half of the AFI genes (7/12) had a significant increase in expression from mid- to late- exponential phase (Fig. 5.8). Transcript levels of gadA, gadE, and slp also increased significantly from 4.3 to 4.6h, which is different from the expression profiles of the other known stationary phase activated genes that increase from 4.6-5h. Increased expression of Shiga toxin genes upon entry into stationary phase. Over 50% of Sp 15 ORFs and 60% of Sp 5 ORFs had a significant (p < 0.00001) change in expression over time. The genes encoding Shiga toxin 1, stx1A and stx1B, as well as those encoding Shiga toxin 2, stx2A and stXZB, had a significant difference in expression between mid- and late- log phase (3 to 4h, Fig.5.9). Stx1 genes decreased ~ 1.8 fold over this interval, while Sb<2 genes increased ~ 1.7 fold. Stx1 genes increased significantly in early stationary phase, with a ~ 4- fold increase from 5 to 6h. A significant 25 fold increase for stsz occurred at the transition point into stationary phase, but a concomitant increase was not observed for the stx2A ORF on the microarray. The ORFs encoding the phage replication factors 0 and P, as well as the CH regulatory protein and the Q antiterminator protein, increased in expression upon entry into stationary phase, but it is difficult to attribute these ORFs to a specific phage, as there are multiple copies of these ORFs throughout the genome. 146 2.0 0.5 a- \\ s : .9 to e g r a 0.0 -- x a) (g, -0.5 ~- 9. -1.0 ~- + stx1A . + stx1B -1,5 -- —I— stx2A —0— stx28 -2.0 I I I I I I 3 4 5 6 7 8 Time (h) Figure 5.9. Expression profiles of stx1 and stx2 over time. 147 A. 2.0 + ureD 1 5 _ -'v— ureA ' -I- ureC —<>— ureE c _ —. 2 1 0 2 g 0.5 - X e g: 0.0 — ...l -0.5 — -1.0 - .105 I I I I I 3 4 6 7 8 Time (h) B. 2.0 + terZ c .2 m to 2 G x o N D) O .l Time (h) Figure 5.10. Expression profiles of the urease genes (A.) and the genes encoding tellurite resistance (B.). 148 x! Expression changes in tellurite resistance genes and urease genes. Proteins involved in tellurite resistance are encoded on a pathogenicity island (TAI) along with genes encoding urease and are present in a single copy in E. coli O157:H7 Sakai (105). Significant changes in expression over time were detected only for terZ, terD, and terE (Fig. 4.10A). Probes targeting terB and terC are not present on the array. Expression of these three genes increased significantly between mid- and late- exponential phase (~ 1.5 fold) and then decreased 5.9- to 7.4- fold during the transition into stationary phase and in early stationary phase (4.6h to 5.5h interval). These three ORFs were grouped into QT cluster 1. The ureD, ureA, ureC, ureE, and ureG ORFs had a significant decrease in expression from late exponential to early stationary phase (1.5 to 3 fold) and were grouped into QT cluster 1 (Fig. 5108). Also located on this genomic island is the adherence factor iha (242), but there is no significant change in iha mRNA over time. 149 DISCUSSION By monitoring global transcript levels over time, we have measured the numerous alterations in gene expression patterns of E. coli O157:H7 when growing exponentially and then entering stationary phase in minimal medium. E. coli O157:H7 responded to low 02 by increasing expression of genes encoding anaerobic electron acceptors and decreasing expression of genes encoding the TCA cycle. 0157-specific ORFs, including nleA, nleG, and espZ, which encode TTS effectors, were upregulated during this interval. As cells entered stationary phase, ~ 50% of the ORFs in the genome responded with significant changes in transcript level. The global trends include downregulation some of virulence factors, such as the components of the TTSS and some of the TTSS effectors, in stationary phase. Other TTS effectors and the genes encoding Shiga toxin increased in transcript level upon entry to stationary phase. Entry into stationary phase was characterized by the increase in expression of known stress and survival related genes, such as the osmotic stress proteins encoded by osmB and osmC, and the glutamate dependent acid resistance system encoded by gadA and gadBC. As E. coli O157:H7 cells grew in the minimal medium, 02 levels in the medium decreased and were not replenished from the atmosphere (Fig. 5.5). The decrease in available oxygen was reflected in the transcriptome, with significant increases in genes encoding anaerobic electron acceptors, decrease in transcript levels of TCA cycle genes, and increase in expression of pyruvate formate Iyase and concomitant decrease in expression of the pyruvate 150 dehydrogenase complex. 5 % (73/1239) of 0157-specific ORFs also had significant changes in expression over the same time interval, which leads one to speculate that these 0157-specific ORFs may be regulated by transcription factors that influence expression of respiratory genes, such as FNR and ArcA. Virulence factor expression in enteric pathogens, such as Salmonella Typhimurium and Vibrio cholera, can be controlled by the redox state regulators FNR and ArcA (86, 140, 222). A search for the FNR and ArcA binding sequences 500bp upstream of all ORFs in the Sakai genome (using Pattern Search on coliBASE (61 )) did not identify any of the ORFs that were significantly expressed or repressed in exponential phase. It may be that these 0157-specific ORFs are regulated indirectly through FNR or ArcA, or are modulated in a completely different manner. Recently Ando et al. (6) demonstrated that anaerobic growth in the presence of the electron acceptors nitrate or TMAO accelerates maturation of the TTSS, independently of new protein synthesis . They found that the narGHlJ operon must be present for the maturation of the 'I‘I'SS. Our findings indicate that the expression of this operon increases significantly in late exponential phase, which is consistent with the narGHIJ operon contribution to the maturation of the TTSS in this environment. Much is known about regulation of LEE expression; multiple circuits of activation and repression have been identified. LEE expression can be induced in response to a number of environmental stimuli, and while common transcriptional regulators and signaling systems can activate LEE expression, they all typically act on Ler or the recently described GrlA and GrlR regulators 151 located on the LEE island (76), which then alter expression of other LEE genes. LEE expression is known to be maximal in late exponential phase (232, 272). Nakanishi et al. determined that LEE can be stimulated during nutrient downshifts through activation of ppGpp and DksA (181). The pattern of espB expression observed during induction of LEE with ppGpp is similar to our observation, in that espB expression increases from mid- to late-exponential phase, and then decreases in stationary phase. LEE transcription can also be stimulated by quorum sensing, via the auto-inducer 3 signaling system that activates ler (232). The quorum sensing activated regulator QseA which activates LEE transcription (231 ), had a decrease in transcript levels from exponential to stationary phase (Fig. 5.3 QTC 1). SdiA is a longer-term stationary phase- related quorum sensing system and represses expression of virulence factors, including LEE (131). Many of the common E. coli regulators that have been shown to repress LEE expression, such as Hha (223), SdiA, GadE (243), YhiF (243), and IHF (91, 151), all increased in mRNA levels from exponential to stationary phase (Table 4.2, Fig. 5.3 QTC 2). Other LEE activators, such as PchC (122), had a significant increase in mRNA in stationary phase. The decrease in expression of LEE genes from exponential to stationary phase that we observed is similar to the expression changes of LEE genes when O157:H7 is attached to red blood cells (72). In comparing our dataset to the genes that were found to be significantly up- or down-regulated in O157:H7 attached to red blood cells, we found that of the 299 ORFs reported to be 152 downregulated in attached cells, 185 had significant changes in expression over time, and 73% of these ORFs were classified in OT cluster 1 (Fig. 5.3), which includes the LEE genes. Of the 105 ORFs reported to be upregulated in attached cells, 89 had significant changes in expression over time, and 56% of these significant ORFs were placed in OT cluster 2 (Fig. 5.3), suggesting that many of the ORFs expressed or repressed during attachment to plasma membranes are also regulated by growth phase. EspZ is known to be translocated through the TTSS, but its function is unknown’ (130). One of the main differences among the LEE genes is the increase in espZ mRNA upon entry into stationary phase, which at first glance, appears to be an anomaly, but is part of a larger pattern when examined in the context of the LEE and non-LEE encoded type III secretion system effectors. Not as much known about regulation of non-LEE encoded TTSS effector genes. The presence of some of the 'I'I'SS effectors in the genome is variable among 0157 strains (190). The TTSS effectors tccP and espJ are located on Sakai phage 14 (EDL-933 OI#79); tccP (espFu) is necessary for actin recruitment and NE lesion formation in E. coli O157:H7, while espJ influences colonization dynamics (37, 73, 97). Reading et al. recently determined that the Al-3, epinephrine, and norepinephrine response regulator QseF (yflrA) is an indirect transcriptional activator of tccP (207). Our data indicate that as Sakai 0157:H7 entered stationary phase, qseF mRNA levels decrease significantly (2 fold) (QTC 1) and tccP-espJ levels decreased significantly as well. This decrease in tccP and espJ transcript levels in stationary phase is in contrast to a 153 previous study, where the activity of the tccP-espJ promoter was found to be the same in exponential vs. stationary phase cultures of EDL-933 in Dulbecco’s Modified Eagle Medium as monitored by a gfp fusion (96). Some of the TTSS effectors appear to respond to low 02 conditions, as 7 effectors increased significantly from mid- to late-exponential phase when oxygen levels decreased. Of the TTSS effectors with significant changes in expression over time, ~50% have decreased expression from exponential to stationary phase, while the other half have increased expression in stationary phase. Some of these TTSS effectors are predicted to be pseudogenes; one example is espL3, an ORF that was found to have a significant increase in expression at the stationary phase transition point. Many of the TTSS effectors with increased expression levels in stationary phase are members of the NleG family (Table 5.7), one of the largest families of ITS effectors that contain many duplications (247). The utility of the increased expression of some of the TTS effectors in stationary phase is unclear and warrants further investigation. The ability to resist low pH is a crucial adaptation and critical component of the low infectious dose of E. coli O157:H7. The GDAR system provides the greatest protection against low pH, allowing survival at levels of gastric acidify (e.g. pH 2.0) as long as glutamate is present in the environment (155). Genes encoding the glutamate decarboxylase isozymes and the glutamate-gamma- aminobutyric acid antiporter are under tight control and can be activated via a number of environmental cues and transcriptional regulators (88), and many components of the system are located in the AF I region (117). Transcription 154 factors located on the AF I, such as GadE and GadX, regulate a large number (>40) of ORFs outside of the GAD system in E. coli K-12 (117, 159), and it is known that these factors serve as repressors of LEE transcription (225, 243), but it is unknown what the effect of these regulators are on a genome-wide scale in E. coli O157:H7. Important questions remain about the factors regulating the Shiga toxins which are encoded by lambda-like bacteriophages (220). These phages can be transferred between many E. coli strains in nature and were the principal virulence factors acquired as a key step in the evolution of E. coli O157:H7. (208). How they become influenced in expression and regulated when acquired intoanew genome is not fully understood. Early work described two types of Shiga toxin (188) encoded by Stx1 and Stx2, both located within the late operons of the Stx-encoding phages (185, 199). Stx1 expression is regulated in part by the iron-dependent transcriptional repressor, Fur (36). Exposure of E. coli O157:H7 to low iron conditions result in an increase of stx1 transcription (259). Our growth experiments indicate that the stx1 genes decrease in expression from mid- to late-exponential phase, at the same interval when 02 levels decrease (Fig. 4.9). Iron becomes more soluble as 02 concentration decreases, which could lead to the decreased stx1 expression. Stx1 expression can be induced at 2 other promoters in addition to the Fur binding site, those that require the phage anti-terminator Q (186) and the phage anti-terminator N (259). Stx2 transcription is initiated at the late phage promoter and is dependent on the anti-repressor Q protein (260). The production of Stx2 is part of the lytic cycle of the 155 . - .1. bacteriophage. Prophages can be induced by antimicrobials, and antimicrobials have been shown to induce stx2 expression (9, 138). Herold et al. found that phage genes were induced upon exposure of E. coli O157:H7 EDL-933 to norfloxacin; phage late genes and stx2 were induced 150-fold (111). Levels of stx2 mRNA are higher in starved O157:H7 compared to exponentially growing cells (148), indicating an overall trend of stx2 expression induction by stress. Here we observed that stx1 and stx2 expression increased in stationary phase. ‘ Herold et al. also reported that stx2A is transcribed more efficiently than stXZB, and may be regulated at the posttranscriptional level (111). This difference in transcription efficiency between stx2A and stXZB was also reflected in our data, as stXZB levels did not appear to increase as the stx2A levels did in stationary phase (Fig. 4.9). The genomic island termed the Tellurite and Adherence Island (TAI) contains genes encoding tellurite resistance, urease, and an adherence factor, iha (242). The urease operon and the tellurite genes are thought to be ubiquitously distributed in EHEC strains (183, 244) but is not present in O157:NM or 055:H7 strains (92). Urease activity has been detected in relatively few EHEC strains. In O157:H7 Sakai, the urease genes have been detected at the transcript level, but a functional protein is not produced due to a stop codon present in ureD (182). Saridakis et al. surveyed 25 O157:H7 strains that were isolated from humans, cattle, and pigs for ureC and the ability to resist low pH using the urease enzyme. UreC was detected in all of the isolates but there was no evidence for urease-mediated acid resistance (218). Here we see that the genes encoding 156 urease are expressed during exponential growth and transcript levels decrease significantly as cells enter stationary phase. The genes encoding tellurite resistance are functional in EHEC, and can be separated into two classes - those that are expressed constitutively during growth (terDEZ) and those that are expressed only during growth in the presence of tellurite (terBCF) (244). Our data show this as well, with terD, E, and Z all expressed during exponential growth, with a significant decrease in transcript levels upon entry to stationary phase, and no significant changes in transcript levels for terA and F. Transition to stationary phase is characterized by complex physiological changes to the bacterial cell. Many of the changes in expression patterns observed for O157:H7 in this study have also been observed for E. coli K-12 entering stationary phase (57). The sigma factor RpoS plays a critical role in transcribing genes associated with stationary phase and stress response (142, 195, 266), many of these genes, such as bolA, dksA, osmC, and osmB, have been identified in this study as well. We still do not know about most of the 0157-specific ORFs, and what regulators control their expression. Here we have classified some of the well-known O157 genes, as well as 0157 genes of unknown function, as being modulated during growth transitions. Global transcriptional profiling of isogenic mutants for a specific regulator have been conducted for a number of regulators in E. coli K-12. As we know more about the regulation of virulence loci, such as the LEE, by global regulators, it would be very useful to perform these profiling experiments in E. coli O157:H7 and other pathogenic E. coli. 157 t 374'" Here we provide the first description of global transcriptome profiling of O157:H7 during growth transitions. Many 0157-specific ORFs responded to the growth transition with significant changes in gene expression. As a large proportion of 0157-specific ORFs are of unknown function, these data are important for comparison to other growth conditions and elucidating the function and factors regulating these unknown ORFs. This provides the opportunity to identify ORFs of interest, based on expression profiles, and compare expression within 0157 population or with other EHEC populations. It is useful to understand what transcriptional changes occur from exponential to stationary phase, as these transitions are important to the lifestyle of O157:H7. 158 ACKNOWLEDGEMENTS The authors thank Katherine Schaeffer for assistance with the dog measurements and Galeb Abu-Ali, Peter Bergholz, and Sivapriya Kailasan Vanaja for helpful scientific discussions and critical review of the manuscript. This work was supported by Food Safety NRI #2005-35201-16362 from the United States Department of Agriculture. This work will be submitted to BMC Microbiology. 159 Chapter 6 Summary and Synthesis 160 .2'. A. _. .1" .F'. The purpose of the work presented here was to characterize the phenotypic and genetic basis of survival that contributes to the transmission of EHEC. One advantage of these studies is the use of multiple strains representing specific clonal groups of E. coli for comparison. This work has contributed to the greater body of knowledge about EHEC acid resistance by showing that E. coli O157:H7 can be more acid resistant compared to non-0157 strains (i.e. members of the EHEC 2 group) in complex acidic environments, and this may be due, in part, to differences in expression of the glutamate decarboxylase acid resistance system. For comparing survival between O157:H7 and non-0157 EHEC, multi- locus sequence typing was used to define groups of EHEC based on genetic similarity. Under these growth conditions and test environments, 0157 strains were no more AR than non-0157 EHEC. Results indicate that the ability to utilize these mechanisms when induced both by acid and stationary phase are similar among all EHEC strains. Under natural conditions in complex acidic environments, however, E. coli O157:H7 may have the ability to use multiple mechanisms in combination, through some form of epitasis, to achieve higher AR than E. coli of other clonal groups. The glutamate decarboxylase acid resistance system was found to provide the greatest protection against low pH in the comparison of AR mechanisms. Of the specific amino acid dependent AR mechanisms, the lysine and arginine decarboxlyase systems are common in many enteric bacteria, whereas the GAD system is present only in E. coli and Shigella among the 161 Enterobacteriaceae. A similar GAD system is present in some Gram-positive bacteria, including the food borne pathogens Listeria monocytogenes and Clostridium perfringens (228). The GAD system in Listeria is capable of protecting L. monocytogenes in low pH environments, such as gastric fluid (67) and acidic foods (68), and also is involved in growth at mildly acidic pH (69). While the GAD system is not involved in survival of E. coli O157:H7 in low pH foods (202), it is essential for survival of O157:H7 Sakai in the MSS (129) and is expressed in E. coli K-12 during growth at pH 5.5 (252). Variations in levels of GAD activity between different strains of L. monocytogenes was found to correlate significantly with levels of tolerance to gastric fluid (67). In the MSS study reported here, we found that transcript levels of gadA and gadB were significantly higher in the 0157 strains compared to the 026 strains, but this was not significantly correlated with survival rate in the MSS. The glutamate decarboxylase enzyme is encoded by two duplicated genes, gadA and gadB, and strains were selected from diverse groups of pathogenic E. coli, based on genetic similarity, to determine evolutionary history of the glutamate decarboxylase enzymes. A gene conversion event had occurred in all pathogenic lineages studied, with the exception of EHEC 1. Recent evidence of differential expression of gadA and gadB between EHEC 0157 and EHEC 2 strains (14), and between EHEC 0157 and Shigella flexneri (22) suggest that regulation of GAD expression is distinct in EHEC 0157 compared to other E. coli strains. 162 Survival in a complex acidic environment was assessed for the same populations of EHEC. O157 strains survived better than 026 strains in this simulated gastric environment and the 026 strains had a higher proportion of injured cells after 3h in the gastric environment. Previous studies that have compared survival of pathogenic E. coli in simulated gastric environments have focused mainly on E. coli O157:H7. In SGF pH 1.5, EHEC O157:H7 strains persisted longer compared to strains of enteroinvasive E. coli, enteropathogenic E. coli, and Shigella dysenteriae (8). O157:H7 strains endured for 3h in SGF pH 1.5 compared to survival of 20 min or less for Salmonella and Listeria strains (213). In addition, O157:H7 strains had a 4-fold higher survival rate when mixed with cooked ground beef and inoculated into SGF compared to Shigella strains (239). In his study, the average survival for 10 O157:H7 strains in SGF plus cooked ground beef (—0.08 i 0.07) was greater than the average survival rate observed here in the MSS (—0.17 i 0.15), which comprised SGF plus homogenized turkey. Incubation in particular food matrices can influence the ability of bacterial cells to survive subsequent acid stress. For E. coli O157:H7 and O26:H11, storage in apple juice at 4°C did not influence subsequent survival for both groups of strains, and storage in apple juice at 22°C resulted in decreased survival for both groups. A similar temperature effect was found for E. coli O157:H7 inoculated into apple juice, with survival in SGF at pH 2.5 greater after cold storage (at 4°C) versus storage at room temperature (21°C) (254). For foodborne pathogens that can multiply at 4°C, such as L. monocytogenes, 163 storage of an inoculated food product at refrigeration temperatures results in increased acid resistance as the bacteria can multiply during storage (87, 236). Bacterial growth conditions and the intrinsic and extrinsic properties of the food product clearly play a role in survival of food borne pathogens. Determining how foodborne pathogens respond and adapt to these stressful conditions will lead to development of new methods to decrease and eliminate survival of these pathogens in foods. Global patterns of gene expression. This work provided the first description of global transcriptome profiling of O157:H7 during growth transitions. Many 0157-specific ORFs responded to the growth transition with significant changes in gene expression. As a large proportion of 0157-specific ORFs are of unknown function, these data are important for comparison to other growth conditions and elucidating the function and factors regulating these unknown ORFs. This provides the opportunity to identify ORFs of interest, based on expression profiles, and compare expression within 0157 population or with other EHEC populations. It is useful to understand what transcriptional changes occur from exponential to stationary phase, as these transitions are important to the lifestyle of O157:H7. Global transcriptional profiling has been utilized to characterize stress responses and pathogenesis in other foodborne pathogens, such as Salmonella enterica Typhimurium, L. monocytogenes, and Vibrio cholera. These studies have identified genes involved in bile (203) and antimicrobial (238) resistance as well as the role of IHF in stationary phase gene expression (162) in Salmonella 164 . Typhimurium. Transcription profiling of L. monocytogenes indicates that 17% of the genome is activated during intracellular growth (60) and that virulence genes as well as stress response genes are activated by the sigma factor SigB (134). Gene expression duing biofilm development has been assessed for Vibrio cholera (221 ). As more data are collected from transcription profiling of foodborne pathogens, this will allow researchers to highlight common and unique strategies used by these organisms to survive under stresses that are representative of those found in foods, processing environments, and the human host Future investigations. Numerous questions still remain about survival of EHEC 1 and EHEC 2 in adverse environments. The GAD sequencing results indicated that a gene conversion event has occurred in all clonal groups tested, with the exception of EHEC 1. A natural extension of that work would be to then investigate how the conversion event may influence expression of these genes. Many of the strains that were sequenced were RpoS deficient isolates, and would not serve as an appropriate comparison of GAD expression, as RpoS influences GAD expression. A new set of strains has been selected to represent EPEC 1 and 2, EHEC 1 and 2, and include Shigella strains (129) and only RpoS positive isolates will be used to elucidate differences in GAD expression across these strains. Global transcriptional response of O157:H7 exposed to apple juice is currently being investigated (13). The main transcriptional responses to the combined osmotic and acidic stress of the apple juice were increased production 165 of genes involved in membrane stress and protein mis-folding in the periplasm. Inactivation of the membrane stress transcriptional activator cpxP significantly decreased in survival of O157:H7 in apple juice. By comparing genes induced by apple juice in O157:H7 to those induced in 055:H7, an 0157-specific response has been identified. These data lead to hypotheses about the evolution of regulation of these stress response genes in O157:H7 from 055:H7, and remain to be tested. The studies described here have only used EHEC strains isolated from cases of human disease. Recent work suggests that the population of EHEC in the bovine reservoir is more diverse than the population isolated from humans (17). It is unclear why only a sub-population of bovine EHEC cause disease in humans, but there are many hypotheses that could be tested, such as that the bovine EHEC are not as resistant to stresses that may be encountered during food processing and storage compared to the human EHEC. 166 10. 11. REFERENCES Abe, H., I. Tatsuno, T. Tobe, A. Okutani, and C. Sasakawa. 2002. Bicarbonate ion stimulates the expression of locus of enterocyte effacement-encoded genes in enterohemorrhagic Escherichia coli O157:H7. Infect Immun 70:3500-9. Acheson, D. W., K. Frankson, and D. Willis. 1998. Multicenter prevalence study of Shiga toxin-producing Escherichia coli. In Abstracts of the 98th General Meeting of the American Society for Microbiology, American Society for Microbiology, Washington DC. abstr. 0-205. Alakomi, H. L., E. Skytta, M. Saarela, T. Mattila-Sandholm, K. Latva-Kala, and l. M. Helander. 2000. Lactic acid perrneabilizes gram-negative bacteria by disrupting the outer membrane. Appl Environ Microbiol 66:2001-5. Alexeeva, S., K. J. Hellingwerf, and M. J. Teixelra de Mattos. 2002. Quantitative assessment of oxygen availability: perceived aerobiosis and its effect on flux distribution in the respiratory chain of Escherichia coli. J Bacteriol 184:1402-6. Allerberger, F., A W. Friedrich, K. Grif, M. P. Dierich, H. J. Dornbusch, C. J. Mache, E. Nachbaur, M. Freillnger, P. Rleck, M. Wagner, A. Caprloll, H. Karch, and L. B. Zimmerhackl. 2003. Hemolytic-uremic syndrome associated with enterohemorrhagic Escherichia coli 026:H infection and consumption of unpasteurized cow's milk. Int J Infect Dis 7:42-5. Ando, H., H. Abe, N. Sugimoto, and T. Tobe. 2007. Maturation of functional type III secretion machinery by activation of anaerobic respiration in enterohaemorrhagic Escherichia coli. Microbiology 153:464-73. Arnold, C. N., J. McElhanon, A. Lee, R. Leonhart, and D. A. Siegele. 2001. Global analysis of Escherichia coli gene expression during the acetate- induced acid tolerance response. J Bacteriol 183:2178-86. Arnold, K. W., and C. W. Kaspar. 1995. Starvation- and stationary-phase- induced acid tolerance in Escherichia coli O157:H7. Appl Environ Microbiol 61 :2037-9. Azizoglu, R. 0., and M. Drake. 2007. Impact of antibiotic stress on acid and heat tolerance and virulence factor expression of Escherichia coli O157:H7. J Food Prot 70:194-9. Baev, M. V., D. Baev, A. J. Radek, and J. W. Campbell. 2006. Growth of Escherichia coli MG1655 on LB medium: determining metabolic strategy with transcriptional microarrays. Appl Microbiol Biotechnol 71:323-8. Barkocy-Gallagher, G. A., T. M. Arthur, M. Rivera-Betancourt, X. Nou, S. D. Shackelford, T. L. Wheeler, and M. Koohmaraie. 2003. Seasonal prevalence of Shiga toxin-producing Escherichia coli, including O157:H7 and non-0157 serotypes, and Salmonella in commercial beef processing plants. J Food Prot 66:1978-86. 167 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. Benjamin, M. M., and A. R. Data. 1995. Acid tolerance of enterohemorrhagic Escherichia coli. Appl Environ Microbiol 61 :1669-72. Bergholz, T. M. unpublished data. Bergholz, T. M., and T. S. Whittam. 2007. Variation in acid resistance among enterohaemorrhagic Escherichia coli in a simulated gastric environment. J Appl Microbiol 102:352-62. Berry, E. D., G. A. Barkocy-Gallagher, and G. R. Siragusa. 2004. Stationary- phase acid resistance and injury of recent bovine Escherichia coli 0157 and non- 0157 biotype l Escherichia coli isolates. J Food Prot 67:583-90. Besser, R. E., S. M. Lett, 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. E., N. Shaikh, N. J. Holt, P. l. Tarr, M. E. 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 73:671-9. Bettelheim, K. A. 2003. Non-0157 Verotoxin-Producing Escherichia coli: A Problem, Paradox, and Paradigm. Exp Biol Med (Maywood) 228:333-44. Bettelheim, K. A. 2000. Role of non-0157 VTEC. Symp Ser Soc Appl Microbiolz38S-508. Beuchat, L. R. 1999. Survival of enterohemorrhagic Escherichia coli O157:H7 in bovine feces applied to lettuce and the effectiveness of chlorinated water as a disinfectant. J Food Prot 62:845-9. Beumer, R. R., J. de Vries, and F. M. Rombouts. 1992. Campylobacterjejuni non-culturable coccoid cells. Int J Food Microbiol 15:153-63. Bhagwat, A. A., and M. Bhagwat. 2004. Comparative analysis of transcriptional regulatory elements of glutamate-dependent acid-resistance systems of Shigella flexneri and Escherichia coli O157:H7. FEMS Microbiol Lett 234:139-47. Blanco, J., M. Blanco, J. E. Blanco, A. Mora, E. A. Gonzalez, M. l. Bernardez, M. P. Alonso, A. Coira, A. Rodriguez, J. Rey, J. M. Alonso, and M. A. Usera. 2003. Verotoxin-producing Escherichia coli in Spain: prevalence, serotypes, and virulence genes of O157:H7 and non-0157 VTEC in ruminants, raw beef products, and humans. Exp Biol Med (Maywood) 228:345-51. ‘ Blaser, M. J., and L. S. Newman. 1982. A review of human salmonellosis: l. lnfective dose. Rev Infect Dis 4:1096-106. 168 25. 26. 27. 28. 29. 30. 31. 32. 33. Blattner, F. R., G. Plunkett, 3rd, 0. A. Bloch, N. T. Perna, 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 277:1453-74. Bohannon, D. E., N. Connell, J. Keener, A. Tormo, M. Espinosa-Urgel, M. M. Zambrano, and R. Kolter. 1991. Stationary-phase-inducible "gearbox" promoters: differential effects of katF mutations and role of sigma 70. J Bacteriol 1 73:4482-92. Bopp, D. J., B. D. Sauders, A. L. Waring, J. Ackelsberg, N. Dumas, E. Braun- Howland, D. Dziewulski, B. J. Wallace, M. Kelly, T. Halse, K. A. Musser, P. F. Smith, D. L. Morse, and R. J. leberger. 2003. Detection, isolation, and molecular subtyping of Escherichia coli O157:H7 and Campylobacterjejuni associated with a large waterborne outbreak. J Clin Microbiol 41:174-80. Breuer, T., D. H. Benkel, R. L. Shapiro, W. N. Hall, M. M. Winnett, M. J. Linn, J. Neimann, T. J. Barrett, 8. Dietrich, F. P. Downes, D. M. Toney, J. L. Pearson, H. Rolka, L. Slutsker, and P. M. Griffin. 2001. A multistate outbreak of Escherichia coli O157:H7 infections linked to alfalfa sprouts grown from contaminated seeds. Emerg Infect Dis 7:977-82. Brewster, D. H., M. I. Brown, D. Robertson, G. L. Houghton, J. Blmson, and J. C. Sharp. 1994. An outbreak of Escherichia coli O157 associated with a children's paddling pool. Epidemiol lnfect112:441-7. Brock, T. D. 1962. Effects of magnesium ion deficiency on Escherichia coli and possible relation to the mode of action of novobiocin. J Bacteriol 84:679-82. Brooks, J. T., D. Bergmire-Sweat, M. Kennedy, K. Hendricks, M. Garcia, L. Marengo, J. Wells, M. Ying, W. Bibb, P. M. Griffin, R. M. Hoekstra, and C. R. Friedman. 2004. Outbreak of Shiga toxin-producing Escherichia coliO111zH8 infections among attendees of a high school cheerleading camp. Clin Infect Dis 38:190-8. Brooks, J. T., E. G. Sowers, J. G. Wells, K. D. Greene, P. M. Griffin, R. M. Hoekstra, and N. A. Strockbine. 2005. Non-0157 Shiga toxin-producing Escherichia coli infections in the United States, 1983-2002. J Infect Dis 192:1422-9. Brown, J. L., T. Ross, T. A. McMeekin, and P. D. Nichols. 1997. Acid habituation of Escherichia coli and the potential role of cyclopropane fatty acids in low pH tolerance. Int J Food Microbiol 37:163-73. Buchanan, R. L., and S. G. Edelson. 1996. Culturing enterohemorrhagic Escherichia coli in the presence and absence of glucose as a simple means of evaluating the acid tolerance of stationary-phase cells. Appl Environ Microbiol 62:4009-13. 169 35. 36. 37. 38. 39. 40. 41. 42. 43. 45. 46. Buchanan, R. L., and S. G. Edelson. 1999. pH-dependent stationary-phase acid resistance response of enterohemorrhagic Escherichia coli in the presence of various acidulants. J Food Prot 62:211-8. Calderwood, S. B., and J. J. Mekalanos. 1987. Iron regulation of Shiga-like toxin expression in Escherichia coli is mediated by the fur locus. J Bacteriol 169:4759-64. Campellone, K. G., D. Robbins, and J. M. Leong. 2004. EspFU is a translocated EHEC effector that interacts with Tir and N-WASP and promotes ch-independent actin assembly. Dev Cell 7:217-28. Capitanl, G., D. De Biase, C. Aurizl, H. Gut, F. Bossa, and M. G. Grutter. 2003. Crystal structure and functional analysis of Escherichia coli glutamate decarboxylase. Embo J 22:4027-37. .FEM7-‘I Caprioll, A., l. Luzzl, F. Rosmini, C. Restl, A. Edefonti, F. Perfumo, C. Farlna, A. Gogllo, A. Gianvltl, and G. Rizzoni. 1994. Community-wide outbreak of hemolytic-uremic syndrome associated with non-0157 verocytotoxin-producing Escherichia coli. J Infect Dis 169:208-11. Carter, A. 0., A. A. Borczyk, J. A. Carlson, B. Harvey, J. C. Hockln, M. A. Karmall, C. Krishnan, D. A. Korn, and H. Llor. 1987. A severe outbreak of Escherichia coli O157:H7—associated hemorrhagic colitis in a nursing home. N Engl J Med 317:1496-500. CastanIe-Cornet, 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 147:709-15. Castanie-Cornet, 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 :3525-35. CDC. 1995. Community outbreak of hemolytic uremic syndrome attributable to Escherichia coliO111:NM-South Australia 1995. MMWR Morb Mortal Wkly Rep 44:550-1, 557-8. CDC. 2000. Escherichia coliO111zH8 outbreak among teenage campers-Texas, 1999. MMWR Morb Mortal Wkly Rep 49:321-4. CDC. 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. CDC. 1997. Escherichia coli 0157:H7 infections associated with eating a nationally distributed commercial brand of frozen ground beef patties and burgers— Colorado, 1997. MMWR Morb Mortal Wkly Rep 46:777-8. 170 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. CDC. 1995. Escherichia coli O157:H7 outbreak linked to commercially distributed dry-cured salami--Washington and California, 1994. MMWR Morb Mortal Wkly Rep 44:157-60. CDC. 1991. Foodbome outbreak of gastroenteritis caused by Escherichia coli O157:H7--North Dakota, 1990. MMWR Morb Mortal Wkly Rep 40:265-7. CDC. 2002. Multistate outbreak of Escherichia coli O157:H7 infections associated with eating ground beef—United States, June-July 2002. MMWR Morb Mortal Wkly Rep 51:637-9. CDC. 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. CDC. 1995. Outbreak of acute gastroenteritis attributable to Escherichia coli serotype O1042H21-Helena, Montana, 1994. MMWR Morb Mortal Wkly Rep 44:501-3. CDC. 1999. Outbreak of Escherichia coli O157:H7 and Campylobacter among attendees of the Washington County Fair-New York, 1999. MMWR Morb Mortal Wkly Rep 48:803-5. CDC. 2000. Outbreak of Escherichia coli O157:H7 infection associated with eating fresh cheese curds--Wisconsin, June 1998. MMWR Morb Mortal Wkly Rep 49:911-3. CDC. 1997. Outbreaks of Escherichia coli O157:H7 infection and cryptosporidiosis associated with drinking unpasteurized apple cider- Connecticut and New York, October 1996. MMWR Morb Mortal Wkly Rep 46:4-8. CDC. 2001. Outbreaks of Escherichia coli O157:H7 infections among children associated with farm visits-Pennsylvania and Washington, 2000. MMWR Morb Mortal Wkly Rep 50:293-7. CDC. 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. Chang, D. E., D. J. Smalley, and T. Conway. 2002. Gene expression profiling of Escherichia coli growth transitions: an expanded stringent response model. Mol Microbiol 45:289-306. Chang, Y. Y., and J. E. Cronan, Jr. 1999. Membrane cyclopropane fatty acid content is a major factor in acid resistance of Escherichia coli. Mol Microbiol 33:249-59. Chart, H. 2000. VTEC enteropathogenicity. Symp Ser Soc Appl Microbiol:12S- 238. 171 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. Chatterjee, S. S., H. Hossain, S. Otten, C. Kuenne, K. Kuchmina, S. Machata, E. Domann, T. Chakraborty, and T. Haln. 2006. Intracellular gene expression profile of Listeria monocytogenes. Infect Immun 74:1323-38. Chaudhuri, R. R., A. M. Khan, and M. J. Pallen. 2004. coliBASE: an online database for Escherichia coli, Shigella and Salmonella comparative genomics. Nucleic Acids Res 32:D296-9. Cheville, A. M., K. W. Arnold, C. Buchrleser, C. M. Chang, and C. W. Kaspar. 1996. rpoS regulation of acid, heat, and salt tolerance in Escherichia coli 0157:H7. Appl Environ Microbiol 62:1822-4. Chat, 8. 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 66:3911-6. Churchill, G. A. 2002. Fundamentals of experimental design for cDNA microarrays. Nat Genet 32 Suppl:490—5. 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 O157:H7 infection from unpasteurized commercial apple juice. Ann lntem Med 130:202-9. Constantinidou, C., J. L. Hobman, L. Griffiths, M. D. Patel, 0. W. Penn, J. A. Cole, and T. W. Overton. 2006. A reassessment of the FNR regulon and transcriptomic analysis of the effects of nitrate, nitrite, NarXL, and NarQP as Escherichia coli K12 adapts from aerobic to anaerobic growth. J Biol Chem 281 :4802-15. Cotter, P. D., C. G. Gahan, and 0. Hill. 2001. A glutamate decarboxylase system protects Listeria monocytogenes in gastric fluid. Mol Microbiol 40:465-75. Cotter, P. D., K. O'Rellly, and 0. Hill. 2001. Role of the glutamate decarboxylase acid resistance system in the survival of Listeria monocytogenes L028 in low pH foods. J Food Prot 64:1362-8. Cotter, P. D., S. Ryan, C. G. Gahan, and 0. Hill. 2005. Presence of GadDI glutamate decarboxylase in selected Listeria monocytogenes strains is associated with an ability to grow at low pH. Appl Environ Microbiol 71 :2832-9. Cul, X., and G. A. Churchill. 2003. Statistical tests for differential expression in cDNA microarray experiments. Genome Biol 4:210. Cul, 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. Dahan, 8., S. Knutton, R. K. Shaw, V. F. Crepin, G. Dougan, and G. Frankel. 2004. Transcriptome of enterohemorrhagic Escherichia coli O157 adhering to eukaryotic plasma membranes. Infect Immun 72:5452-9. 172 73. 74. 75. 76. 77. 78. 79. 80. 81. 82. 83. Dahan, S., S. Wiles, R. M. La Ragione, A. Best, M. J. Woodward, M. P. Stevens, R. K. Shaw, Y. Chang, S. Knutton, A. Philllps, and G. Frankel. 2005. EspJ is a prophage-carried type III effector protein of attaching and effacing pathogens that modulates infection dynamics. Infect Immun 73:679-86. De Blase, D., A. Tramontl, F. Bossa, and P. Visca. 1999. The response to stationary-phase stress conditions in Escherichia coli: role and regulation of the glutamic acid decarboxylase system. Mol Microbiol 32:1198-211. de Jonge, R., K. Takuml, W. S. Ritmeester, and F. M. van Leusden. 2003. The adaptive response of Escherichia coli 0157 in an environment with changing pH. Journal of Applied Microbiology 94:555-560. Deng, W., J. L. Puente, S. Gruenheid, Y. Li, B. A. Vallance, A. Vazquez, J. Barba, J. A. Ibarra, P. O'Donnell, P. Metalnlkov, K. Ashman, S. Lee, D. Goode, T. Pawson, and B. B. Finlay. 2004. Dissecting virulence: systematic and functional analyses of a pathogenicity island. Proc Natl Acad Sci U S A 101:3597-602. Diez-Gonzalez, F., and Y. Karaibrahimoglu. 2004. Comparison of the glutamate-, arginine- and lysine-dependent acid resistance systems in Escherichia coli O157:H7. J Appl Microbiol 96:1237-44. Donnenberg, M. S., and T. S. Whittam. 2001. Pathogenesis and evolution of virulence in enteropathogenic and enterohemorrhagic Escherichia coli. J Clin Invest 107:539-48. Doyle, M. P., T. Zhao, J. Meng, and S. Zhao. 2001. Escherichia coli O157:H7, p. 171-191. In L. R. B. Michael P. Doyle, and Thomas J. Montville (ed.), Food Microbiology: Fundamentals and Frontiers, 2nd ed. ASM Press, Washington DC. Dressman, J. B., R. R. Berardi, L. C. Dermentzoglou, T. L. Russell, 8. P. Schmaltz, J. L. Barnett, and K. M. Jarvenpaa. 1990. Upper gastrointestinal (GI) pH in young, healthy men and women. Pharrn Res 7:756-61. Dutyshev, D. l., E. L. Darli, N. P. Fomenkova, I. V. Pechik, K. M. Polyakov, S. V. Nikonov, N. S. Andreeva, and B. S. Sukhareva. 2005. Structure of Escherichia coli glutamate decarboxylase (GADalpha) in complex with glutarate at 2.05 angstroms resolution. Acta Crystallogr D Biol Crystallogr 61:230—5. Dzlva, F., P. M. van Dlemen, M. P. Stevens, A. J. Smith, and T. S. Wallis. 2004. Identification of Escherichia coli O157:H7 genes influencing colonization of the bovine gastrointestinal tract using signature-tagged mutagenesis. Microbiology 150:3631-45. Elder, R. 0., J. E. Keen, G. R. Siragusa. G. A. Barkocy-Gallagher, M. Koohmaraie, and W. W. Laegreld. 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. 173 85. 86. 87. 88. 89. 90. 91. 92. 93. 94. 95. Feldman, K. A., J. C. Mohle-Boetani, J. Ward, K. Furst, S. L. Abbott, D. V. Ferraro, A. Olsen, and S. B. Werner. 2002. A cluster of Escherichia coli O157: nonmotile infections associated with recreational exposure to lake water. Public Health Rep 117:380-5. Fey, P. D., R. S. Wickert, M. E. Rupp, T. J. Safranek, and S. H. Hlnrichs. 2000. Prevalence of non-O157:H7 shiga toxin-producing Escherichia coli in diarrheal stool samples from Nebraska. Emerg Infect Dis 6:530—3. Fink, R. C., M. R. Evans, S. Porwollik, A. Vazquez-Torres, J. Jones-Carson, B. Troxell, S. J. Libby, M. McClelland, and H. M. Hassan. 2007. FNR Is a Global Regulator of Virulence and Anaerobic Metabolism in Salmonella enterica Serovar Typhimurium (ATCC 140283). J Bacteriol 189:2262-73. Formato, G., I. Geornaras, l. M. Barmpalla, P. N. Skandamls, K. E. Belk, J. A. Scanga, P. A. Kendall, G. C. Smith, and J. N. Sofos. 2007. Effect of acid adaptation on growth during storage at 10 degrees 0 and resistance to simulated gastric fluid of Listeria monocytogenes inoculated onto bologna formulated with or without antimicrobials. J Food Prot 70:65—9. Foster, J. W. 2004. Escherichia coli acid resistance: tales of an amateur acidophile. Nat Rev Microbiol 2:898-907. Foster, J. W., and M. Moreno. 1999. Inducible acid tolerance mechanisms in enteric bacteria. Novartis Found Symp 221 :55-69. Frankel, G., A. D. Phillips, l. Rosenshine, G. Dougan, J. B. Kaper, and S. Knutton. 1998. Enteropathogenic and enterohaemorrhagic Escherichia coli: more subversive elements. Mol Microbiol 30:911-21. Friedberg, D., T. Umanski, Y. Fang, and l. Rosenshine. 1999. Hierarchy in the expression of the locus of enterocyte effacement genes of enteropathogenic Escherichia coli. Mol Microbiol 34:941-52. Friedrich, A. W., R. Kock, M. Bielaszewska, W. Zhang, H. Karch, and W. Mathys. 2005. Distribution of the urease gene cluster among and urease activities of enterohemorrhagic Escherichia coli O157 isolates from humans. J Clin Microbiol 43:546-50. Gajiwala, K. S., and S. K. Burley. 2000. HDEA, a periplasmic protein that supports acid resistance in pathogenic enteric bacteria. J Mol Biol 295:605-12. Gansheroff, L. J., and A. D. O'Brien. 2000. Escherichia coli O157:H7 in beef cattle presented for slaughter in the US: higher prevalence rates than previously estimated. Proc Natl Acad Sci U S A 97:2959-61. Gansheroff, L. J., M. R. Wachtel, and A. D. O'Brien. 1999. Decreased adherence of enterohemorrhagic Escherichia coli to HEp-2 cells in the presence of antibodies that recognize the C-terrninal region of intimin. Infect Immun 67:6409-17. 174 96. 97. 98. 99. 100. 101. 102. 103. 104. 105. 106. Garmendia, J., and G. Frankel. 2005. Operon structure and gene expression of the espJ—tccP locus of enterohaemorrhagic Escherichia coli O157:H7. FEMS Microbiol Lett 247: 1 37-45. Garmendia, J., A. D. Phillips, M. F. Carlier, Y. Chang, S. Schuller, O. Marches, S. Dahan, E. Oswald, R. K. Shaw, S. Knutton, and G. Frankel. 2004. chP is an enterohaemorrhagic Escherichia coli O157:H7 type III effector protein that couples Tir to the actin-cytoskeleton. Cell Microbiol 6:1167-83. Goh, S., C. Newman, M. Knowles, F. J. Bolton, V. Hollyoak, S. Rlchards, P. Daley, D. Counter, H. R. Smith, and N. Keppie. 2002. E. coli O157 phage type 21/28 outbreak in North Cumbria associated with pasteurized milk. Epidemiol Infect 129:451-7. Goodson, M., and R. J. Rowbury. 1989. Habituation to Normally Lethal Acidity by Prior Growth of Escherichia coli at a Sub-Lethal Acid Ph Value. Letters in Applied Microbiology 8:77-79. Gorden, J., and P. L. Small. 1993. Acid resistance in enteric bacteria. Infect Immun 61:364-7. Gordon, D. M., S. Bauer, and J. R. Johnson. 2002. The genetic structure of Escherichia coli populations in primary and secondary habitats. Microbiology 148:1 51 3-22. Grant, M. A., S. D. Weagant, and P. Feng. 2001. Glutamate decarboxylase genes as a prescreening marker for detection of pathogenic Escherichia coli groups. Appl Environ Microbiol 67:31 10-4. Griffin, P. M., and R. V. Tauxe. 1991. The epidemiology of infections caused by Escherichia coli O157:H7, other enterohemorrhagic E. coli, and the associated hemolytic uremic syndrome. Epidemiol Rev 13:60-98. Hansen, A. M., Y. Qiu, N. Yeh, F. R. Blattner, T. Durfee, and D. J. Jln. 2005. SspA is required for acid resistance in stationary phase by downregulation of H- N8 in Escherichia coli. Mol Microbiol 56:719-34. Hayashi, T., K. Makino, M. Ohnishi, K. Kurokawa, K. lshii, K. Yokoyama, C. G. Han, E. Ohtsubo, K. Nakayama, T. Murata, M. Tanaka, T. Tobe, T. llda, H. Takaml, T. Honda, 0. Sasakawa, N. Ogasawara, T. Yasunaga, S. Kuhara, T. Shlba, M. Hattori, and H. Shinagawa. 2001. Complete genome sequence of enterohemorrhagic Escherichia coli O157:H7 and genomic comparison with a laboratory strain K-12. DNA Res 8:11-22. Hayes, E. T., J. C. Wilks, P. Sanfilippo, E. Yohannes, D. P. Tate, B. D. Jones, M. D. Radmacher, S. S. BonDurant, and J. L. Slonczewski. 2006. Oxygen limitation modulates pH regulation of catabolism and hydrogenases, multidrug transporters, and envelope composition in Escherichia coli K-12. BMC Microbiol 6:89. 175 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. 118. Helander, I. M., and T. Mattila-Sandholm. 2000. Fluorometric assessment of gram-negative bacterial permeabilization. J Appl Microbiol 88:213—9. Hengge-Aronls, R. 2002. Signal transduction and regulatory mechanisms involved in control of the sigma(S) (RpoS) subunit of RNA polymerase. Microbiol Mol Biol Rev 66:373-95, table of contents. Hengge-Aronis, R., and D. Fischer. 1992. Identification and molecular analysis of gigs, a novel growth-phase- regulated and rpoS-dependent gene involved in glycogen synthesis in Escherichia coli. Mol Microbiol 6:1877-86. Herold, S., H. Karch, and H. Schmidt. 2004. Shiga toxin-encoding bacteriophages—genomes in motion. Int J Med Microbiol 294:115-21. 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. Hersh, B. M., F. T. Farooq, D. N. Barstad, D. L. Blankenhorn, and J. L. Slonczewski. 1996. A glutamate-dependent acid resistance gene in Escherichia coli. J Bacteriol 178:3978-81. Heyer, L. J., S. Kruglyak, and S. Yooseph. 1999. Exploring expression data: identification and analysis of coexpressed genes. Genome Res 9:1106-15. Hilborn, E. D., J. H. Mermln, 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 lntem Med 159:1758-64. Hilborn, E. D., P. A. Mshar, T. R. Florentine, 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 lnfect124z31-6. Hiruta, N., T. Murase, and N. Okamura. 2001. An outbreak of diarrhoea due to multiple antimicrobial-resistant Shiga toxin-producing Escherichia coli 026:H11 in a nursery. Epidemiol lnfect127:221-7. Hommais, F., E. Krin, J. Y. Coppee, C. Lacrolx, E. Yeramian, 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., E. Krin, C. Laurent-Winter, O. Soutourlna, 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. Mol Microbiol 40:20-36. 176 119. 120. 121. 122. 123. 124. 125. 126. 127. 128. 129. 130. Hrudey, S. E., P. Payment, P. M. Huck, R. W. Gillham, and E. J. Hrudey. 2003. A fatal waterborne disease epidemic in Walkerton, Ontario: comparison with other waterborne outbreaks in the developed world. Water Sci Technol 47:7- 14. Hussein, H. S., and T. Sakuma. 2005. Prevalence of shiga toxin-producing Escherichia coli in dairy cattle and their products. J Dairy Sci 88:450-65. Hussein, H. S., B. H. Thran, M. R. Hall, W. G. Kvasnicka, and R. C. Torell. 2003. Verotoxin-producing Escherichia coli in culled beef cows grazing rangeland forages. Exp Biol Med (Maywood) 228:352-7. 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 150:2357-571. Jay, M. T., V. Garrett, J. C. Mohle-Boetani, M. Barros, J. A. Farrar, R. Rlos, 8. Abbott, R. Sowadsky, K. Komatsu, R. Mandrell, J. Sobel, and S. B. Werner. 2004. A multistate outbreak of Escherichia coli O157:H7 infection linked to consumption of beef tacos at a fast-food restaurant chain. Clin Infect Dis 39:1-7. Jelacic, J. K., T. Damrow, G. S. Chen, S. Jelacic, M. Blelaszewska, M. Ciol, H. M. Carvalho, A. R. Melton-Celsa, A. D. O'Brien, and P. l. Tarr. 2003. Shiga Toxin—Producing Escherichia coli in Montana: Bacterial Genotypes and Clinical Profiles. J Infect Dis 188:719-29. Jenkins, D. E., S. A. Chalsson, and A. Matin. 1990. Starvation-induced cross protection against osmotic challenge in Escherichia coli. J Bacteriol 172:2779-81. Jin, Q., Z. Yuan, J. Xu, Y. Wang, Y. Shen, W. Lu, J. Wang, H. Liu, J. Yang, F. Yang, X. Zhang, J. Zhang, G. Yang, H. Wu, D. Qu, J. Dong, L. Sun, Y. Xue, A. Zhao, Y. Gao, J. Zhu, B. Kan, K. Dlng, S. Chen, H. Chang, Z. Yao, B. He, R. Chen, D. Ma, 8. Qiang, Y. Wen, Y. Hou, and J. Yu. 2002. Genome sequence of Shigella flexneri 2a: insights into pathogenicity through comparison with genomes of Escherichia coli K12 and 0157. Nucleic Acids Res 30:4432-41. Jordan, K. N., L. Oxford, and C. P. O'Byrne. 1999. Survival of low-pH stress by Escherichia coli O157:H7: correlation between alterations in the cell envelope and increased acid tolerance. Appl Environ Microbiol 65:3048-55. 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. unpublished data. Kanack, K. J., J. A. Crawford, I. Tatsuno, M. A. Karmall, and J. B. Kaper. 2005. SepZ/EspZ is secreted and translocated into HeLa cells by the enteropathogenic Escherichia coli type III secretion system. Infect Immun 73:4327-37. 177 131. 132. 133. 134. 135. 136. 137. 138. 139. 140. 141. 142. 143. Kanamaru, K., I. Tatsuno, T. Tobe, and C. Sasakawa. 2000. SdiA, an Escherichia coli homologue of quorum-sensing regulators, controls the expression of virulence factors in enterohaemorrhagic Escherichia coli O157:H7. Mol Microbiol 38:805—16. Kang, Y., K. D. Weber, Y. Qlu, P. J. Kiley, and F. R. Blattner. 2005. Genome- wide expression analysis indicates that FNR of Escherichia coli K-12 regulates a large number of genes of unknown function. J Bacteriol 187:1135-60. Kaper, J. B., J. P. Nataro, and H. L. Mobley. 2004. Pathogenic Escherichia coli. Nat Rev Microbiol 2:123-40. Kazmlerczak, M. J., S. C. Mlthoe, K. J. Beer, and M. Wiedmann. 2003. Listeria monocytogenes sigma B regulates stress response and virulence functions. J Bacteriol 185:5722-34. Keene, W. E., 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. Kerr, M. K. 2003. Linear models for microarray data analysis: hidden similarities and differences. J Comput Biol 10:891-901. Kerr, M. K., M. Martin, and G. A. Churchill. 2000. Analysis of variance for gene expression microarray data. J Comput Biol 7:819—37. Klmmitt, P. T., C. R. Hanlvood, and M. R. Barer. 2000. Toxin gene expression by shiga toxin-producing Escherichia coli: the role of antibiotics and the bacterial SOS response. Emerg Infect Dis 6:458-65. Klein, E. J., J. R. Stapp, C. R. Clausen, D. R. Boster, J. G. Wells, X. Qin, D. L. Swerdlow, and P. l. Tarr. 2002. Shiga toxin-producing Escherichia coli in children with diarrhea: a prospective point-of-care study. J Pediatr 141:172-7. Krishnan, H. H., A. Ghosh, K. Paul, and R. Chowdhury. 2004. Effect of anaerobiosis on expression of virulence factors in Vibrio cholerae. Infect Immun 72:3961-7. Kumar, S., K. Tamura, and M. Nei. 2004. MEGA3: Integrated software for Molecular Evolutionary Genetics Analysis and sequence alignment. Brief Bioinforrn 5:150-63. Lacour, 8., and P. Landlni. 2004. SigmaS-dependent gene expression at the onset of stationary phase in Escherichia coli: function of sigmaS—dependent genes and identification of their promoter sequences. J Bacteriol 186:7186-95. Lampel, K. A., and A. T. Maurelli. 2001. Shigella species, p. 247-261. In M. P. Doyle, L. R. Beuchat, and T. J. Montville (ed.), Food Microbiology: Fundamentals and Frontiers, 2nd ed. ASM Press, Washington DC. 178 144. 145. 146. 147. 148. 149. 150. 151. 152. 153. 154. 155. 156. Lange, R., and R. Hengge-Aronis. 1991. Identification of a central regulator of stationary-phase gene expression in Escherichia coli. Mol Microbiol 5:49-59. 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 :2493-500. Lee, S. J., and J. D. Gralla. 2001. Sigma38 (rpoS) RNA polymerase promoter engagement via -10 region nucleotides. J Biol Chem 276:30064-71. Lee, S. M., and J. Chen. 2004. Survival of Escherichia coli O157:H7 in set yogurt as influenced by the production of an exopolysaccharide, colanic acid. J Food Prot 67:252-255. Leenanon, B., D. Elhanafi, and M. A. Drake. 2003. Acid adaptation and starvation effects on Shiga toxin production by Escherichia coli O157:H7. J Food Prot 66:970-7. Levine, M. M., J. G. Xu, J. B. Kaper, H. Lior, V. Prado, B. Tall, J. Nataro, H. Karch, and K. Wachsmuth. 1987. A DNA probe to identify enterohemorrhagic Escherichia coli of O157:H7 and other serotypes that cause hemorrhagic colitis and hemolytic uremic syndrome. J Infect Dis 156:175-82. Leyer, G. J., L. L. Wang, and E. A. Johnson. 1995. Acid adaptation of Escherichia coli O157:H7 increases survival in acidic foods. Appl Environ Microbiol 61 :3752—5. Li, M., l. Rosenshine, S. L. Tung, X. H. Wang, D. Friedberg, C. L. Hew, and K. Y. Leung. 2004. Comparative proteomic analysis of extracellular proteins of enterohemorrhagic and enteropathogenic Escherichia coli strains and their ihf and ler mutants. Appl Environ Microbiol 70:5274-82. Liao, D. 2000. Gene conversion drives within genic sequences: concerted evolution of ribosomal RNA genes in bacteria and archaea. J Mol Evol 51:305- 17. Licence, K., K. R. Oates, B. A. Synge, and T. M. Reid. 2001. An outbreak of E. coli O157 infection with evidence of spread from animals to man through contamination of a private water supply. Epidemiol Infect 126:135-8. Lin, J., l. 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. P. Smith, K. C. Chapin, H. S. Balk, G. N. Bennett, and J. W. Foster. 1996. Mechanisms of acid resistance in enterohemorrhagic Escherichia coli. Appl Environ Microbiol 62:3094-100. Liu, X., and P. De Wulf. 2004. Probing the ArcA-P modulon of Escherichia coli by whole genome transcriptional analysis and sequence recognition profiling. J Biol Chem 279:12588-97. 179 157. 158. 159. 160. 161. 162. 163. 164. 165. 166. 167. 168. Low, A. S., F. Dzlva, A. G. Torres, J. L. Martinez, T. Rosser, S. Naylor, K. Spears, N. Holden, A. Mahajan, J. Findlay, J. Sales, D. G. Smith, J. 0. Low, M. P. Stevens, and D. L. Gally. 2006. Cloning, expression, and characterization of fimbrial operon F9 from enterohemorrhagic Escherichia coli O157:H7. Infect Immun 74:2233-44. Ma, 2., S. Gong, H. Richard, D. L. Tucker, T. Conway, and J. W. Foster. 2003. GadE (YhiE) activates glutamate decarboxylase-dependent acid resistance in Escherichia coli K-12. Mol Microbiol 49:1309-20. Ma, 2., N. Masuda, and J. W. Foster. 2004. Characterization of EngS-YdeO- GadE branched regulatory circuit governing glutamate-dependent acid resistance in Escherichia coli. J Bacteriol 186:7378-89. Ma, 2., H. Richard, D. L. Tucker, T. Conway, and J. W. Foster. 2002. Collaborative regulation of Escherichia coli glutamate-dependent acid resistance by two AraC-Iike regulators, GadX and GadW (thW). J Bacteriol 184:7001-12. MacDonald, D. M., M. Fyfe, A. Paccagnella, A. Trinidad, K. Louie, and D. Patrick. 2004. Escherichia coli O157:H7 outbreak linked to salami, British Columbia, Canada, 1999. Epidemiol Infect 132:283-9. Mangan, M. W., S. Lucchlni, V. Danlno, T. O. Crolnln, J. C. Hinton, and C. J. Dorman. 2006. The integration host factor (IHF) integrates stationary-phase and virulence gene expression in Salmonella enterica serovar Typhimurium. Mol Microbiol 59: 1 831 -47. Mac, Y., M. P. Doyle, and J. Chen. 2001. Insertion mutagenesis of woe reduces acid and heat tolerance of enterohemorrhagic Escherichia coli O157:H7. J Bacteriol 183:381 1-5. Mao, Y., M. P. Doyle, and J. Chen. 2006. Role of colanic acid exopolysaccharide in the survival of enterohaemorrhagic Escherichia coli O157:H7 in simulated gastrointestinal fluids. Lett Appl Microbiol 42:642-7. Masuda, N., and G. M. Church. 2002. Escherichia coli gene expression responsive to levels of the response regulator Eng. J Bacteriol 184:6225-34. Masuda, N., and G. M. Church. 2003. Regulatory network of acid resistance genes in Escherichia coli. Mol Microbiol 48:699-712. McCarthy, T. A., N. L. Barrett, J. L. Hadler, B. Salsbury, R. T. Howard, D. W. Dingman, C. D. Brinkman, W. F. Bibb, and M. L. Cartter. 2001. Hemolytic- Uremic Syndrome and Escherichia coli 0121 at a Lake in Connecticut, 1999. Pediatrics 108:E59. McClure, P. J., and S. Hall. 2000. Survival of Escherichia coli in foods. Symp Ser Soc Appl Microbiol:61S-7OS. 180 169. 170. 171. 172. 173. 174. 175. 176. 177. 178. 179. 180. 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. McGuire, G., and F. Wright. 2000. TOPAL 2.0: improved detection of mosaic sequences within multiple alignments. Bioinforrnatics 16:130-4. McGuire, G., F. Wright, and M. J. Prentice. 1997. A graphical method for detecting recombination in phylogenetic data sets. Mol Biol Evol 14:1 125-31. McKellar, R. 0., and K. P. Knight. 1999. Growth and survival of various strains of enterohemorrhagic Escherichia coli in hydrochloric and acetic acid. J Food Prot 62:1466-9. Mead, P. S., L. Slutsker, V. Dietz, L. F. McCaig, J. S. Bresee, C. Shapiro, P. M. Griffin, and R. V. Tauxe. 1999. Food-related illness and death in the United States. Emerg Infect Dis 5:607-25. Michino, H., K. Araki, S. Mlnami, S. Takaya, N. Sakai, M. Miyazakl, A. One, 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. Miller, L. G., and C. W. Kaspar. 1994. Escherichia coli O157:H7 acid tolerance and survival in apple cider. J Food Prot 57:460-464. Morgan, D., C. P. Newman, D. N. Hutchinson, A. M. Walker, B. Rowe, and F. Majid. 1993. Verotoxin producing Escherichia coli 0 157 infections associated with the consumption of yoghurt. Epidemiol lnfect111:181-7. Murinda, S. E., L. T. Nguyen, T. L. Landers, F. A. Draughon, A. G. Mathew, J. S. Hogan, K. L. Smith, D. D. Hancock, and S. P. Oliver. 2004. Comparison of Escherichia coli Isolates from humans, food, and farm and companion animals for presence of Shiga toxin-producing E. coli virulence markers. Foodbome Pathog Dis 1:178-84. Mutaku, l., W. Erku, and M. Ashenafi. 2005. Growth and survival of Escherichia coli O157:H7 in fresh tropical fruit juices at ambient and cold temperatures. Int J Food Sci Nutr 56:133-9. Naim, F., S. Messier, L. Saucier, and G. Plette. 2004. Postprocessing in vitro digestion challenge to evaluate survival of Escherichia coli O157:H7 in fermented dry sausages. Appl Environ Microbiol 70:6637-42. Nair, S., and S. E. Flnkel. 2004. Dps protects cells against multiple stresses during stationary phase. J Bacteriol 186:4192—8. 181 181. 182. 183. 184. 185. 186. 187. 188. 189. 190. 191. 192. 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. Mol Microbiol 61 :194-205. Nakano, M., T. lida, and T. Honda. 2004. Urease activity of enterohaemorrhagic Escherichia coli depends on a specific one-base substitution in ureD. Microbiology 1 50:3483-9. Nakano, M., T. lida, M. Ohnishi, K. Kurokawa, A. Takahashl, T. Tsukamoto, T. Yasunaga, T. Hayashi, and T. Honda. 2001. Association of the urease gene with enterohemorrhagic Escherichia coli strains irrespective of their serogroups. J Clin Microbiol 39:4541-3. Naylor, S. W., D. L. Gally, and J. 0. Low. 2005. Enterohaemorrhagic E. coli in veterinary medicine. Int J Med Microbiol 295:419-41. Neely, M. N., and D. l. Friedman. 1998. Arrangement and functional identification of genes in the regulatory region of Iambdoid phage H-19B, a carrier of a Shiga-like toxin. Gene 223:105-13. Neely, M. N., and D. I. Friedman. 1998. Functional and genetic analysis of regulatory regions of coliphage H-1QB: location of shiga-like toxin and lysis genes suggest a role for phage functions in toxin release. Mol Microbiol 28:1255-67. Neidhardt, F. C., P. L. Bloch, and D. F. Smith. 1974. Culture medium for enterobacteria. J Bacteriol 119:736-47. O'Brien, A. D., J. W. Newland, S. F. Miller, R. K. Holmes, H. W. Smith, and S. B. Formal. 1984. Shiga-like toxin-converting phages from Escherichia coli strains that cause hemorrhagic colitis or infantile diarrhea. Science 226:694-6. Ochman, H., and R. K. Selander. 1984. Standard reference strains of Escherichia coli from natural populations. J Bacteriol 157:690-3. Ogura, Y., K. Kurokawa, T. Ooka, K. Tashiro, T. Tobe, M. Ohnishi, K. Nakayama, T. Morimoto, J. Terajima, H. Watanabe. S. Kuhara, and T. Hayashi. 2006. Complexity of the genomic diversity in enterohemorrhagic Escherichia coli 0157 revealed by the combinational use of the 0157 Sakai OligoDNA microarray and the Whole Genome PCR scanning. DNA Res 13:3-14. Oliver, J. D., and J. B. Kaper. 2001. Vibrio species, p. 263-300. In M. P. Doyle, L. R. Beuchat, and T. J. Montville (ed.), Food Microbiology: Fundamentals and Frontiers, 2nd ed. ASM Press, Washington DC. Olsen, S. J., G. Miller, M. Kennedy, 0. Higgins, J. Walford, G. McKee, K. Fox, W. Bibb, and P. Mead. 2002. A waterborne outbreak of Escherichia coli O157:H7 infections and hemolytic uremic syndrome: implications for rural water systems. Emerg Infect Dis 8:370-5. 182 193. 194. 195. 196. 197. 198. 199. 200. 201. 202. 203. 204. Ott, R. L., and M. Longnecker. 2001. An Introduction to Statistical Methods and Data Analysis, 5th ed. Duxbury, Pacific Grove, California. Paton, J. C., and A. W. Paton. 1998. Pathogenesis and diagnosis of Shiga toxin-producing Escherichia coli infections. Clin Microbiol Rev 11 :450-79. Patten, C. L., M. G. Klrchhof, M. R. Schertzberg, R. A. Morton, and H. E. Schellhorn. 2004. Microarray analysis of RpoS-mediated gene expression in Escherichia coli K-12. Mol Genet Genomics 272:580-91. Paunio, M., R. Pebody, M. Keskimakl, M. Kokkl, P. Ruutu, S. Olnonen, V. Vuotarl, A. Slitonen, E. Lahtl, and P. Lelnikki. 1999. Swimming-associated outbreak of Escherichia coli O157:H7. Epidemiol Infect 122:1-5. Perna, 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. Boutln, Y. Shao, L. Miller, E. J. Grotbeck, N. W. Davis, A. Lim, E. T. Dimalanta, K. D. Potamousls, J. Apodaca, T. S. Anantharaman, J. Lin, C. Yen, D. C. Schwartz, R. A. Welch, and F. R. Blattner. 2001. Genome sequence of enterohaemorrhagic Escherichia coli O157:H7. Nature 409:529-33. 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 Plunkett, 6., 3rd, D. J. Rose, T. J. Durfee, and F. R. Blattner. 1999. Sequence of Shiga toxin 2 phage 933W from Escherichia coli O157:H7: Shiga toxin as a phage late-gene product. J Bacteriol 181:1767—78. Pradel, N., V. Livrelli, C. De Champs, J. B. Palcoux, A. Reynaud, F. Scheutz, J. Sirot, B. Joly, and C. Forestier. 2000. Prevalence and characterization of Shiga toxin-producing Escherichia coli isolated from cattle, food, and children during a one-year prospective study in France. J Clin Microbiol 38:1023-31. Price, S. B., C. M. Chang, 0. W. Kaspar, J. C. Wright, F. J. DeGraves, T. A. Penfound, M. P. Castanie-Cornet, and J. W. Foster. 2000. Role of rpoS in acid resistance and fecal shedding of Escherichia coli O157:H7. Appl Environ Microbiol 66:632—7. Price, 8. B., J. C. Wright, F. J. DeGraves, M. P. CastanIe-Cornet, 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 70:4792-9. Prouty, A. M., I. E. Brodsky, J. Manos, R. Belas, S. Falkow, and J. S. Gunn. 2004. Transcriptional regulation of Salmonella enterica serovar Typhimurium genes by bile. FEMS Immunol Med Microbiol 41:177-85. Quackenbush, J. 2002. Microarray data normalization and transformation. Nat Genet 32 Suppl:496—501. 183 205. 206. 207. 208. 209. 210. 211. 212. 213. 214. 215. 216. 217. RIMAANOVA, http://www.jax.org/staff/churchilI/labsite/softwarelanova/nnaanova 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. Reading, N. 0., A. G. Torres, M. M. Kendall, D. T. Hughes, K. Yamamoto, and V. Sperandlo. 2007. A Novel Two-Component Signaling System That Activates Transcription of an Enterohemorrhagic Escherichia coli Effector Involved in Remodeling of Host Actin. J Bacteriol 189:2468-76. Reid, S. D., C. J. Herbelln, A. C. Bumbaugh, R. K. Selander, and T. S. Whittam. 2000. Parallel evolution of virulence In pathogenic Escherichia coli. Nature 406:64-7. Richard, H., and J. W. Foster. 2004. Escherichia coli glutamate- and arginine- dependent acid resistance systems increase internal pH and reverse transmembrane potential. J Bacteriol 186:6032—41. Richard, H. T., and J. W. Foster. 2003. Acid resistance in Escherichia coli. Adv Appl Microbiol 52:167-86. Riley, L. W., R. S. Remls, S. D. Helgerson, H. B. McGee, J. G. Wells, B. R. Davis, R. J. Hebert, E. S. Olcott, L. M. Johnson, N. T. Hargrett, P. A. Blake, and M. L. Cohen. 1983. Hemorrhagic colitis associated with a rare Escherichia coli serotype. N Engl J Med 308:681-5. Riordan, D. C., G. Duffy, J. J. Sheridan, R. C. Whiting, l. 8. Blair, and D. A. McDowell. 2000. Effects of acid adaptation, product pH, and heating on survival of Escherichia coli O157:H7 in pepperoni. Appl Environ Microbiol 66:1726-9. Roering, A. M., J. B. Luchansky, A. M. lhnot, S. E. Ansay, C. W. Kaspar, and S. C. lngham. 1999. Comparative survival of Salmonella typhimurium DT 104, Listeria monocytogenes, and Escherichia coli O157:H7 in preservative-free apple cider and simulated gastric fluid. Int J Food Microbiol 46:263-9. Sak, B. D., A. Eisenstark, and D. Touatl. 1989. Exonuclease Ill and the catalase hydroperoxidase II in Escherichia coli are both regulated by the katF gene product. Proc Natl Acad Sci U S A 86:3271-5. Salmon, K. A., S. P. Hung, N. R. Steffen, R. Krupp, P. Baldl, G. W. Hatfield, and R. P. Gunsalus. 2005. Global gene expression profiling in Escherichia coli K12: effects of oxygen availability and ArcA. J Biol Chem 280:15084-96. Santoyo, G., and D. Romero. 2005. Gene conversion and concerted evolution in bacterial genomes. FEMS Microbiol Rev 29:169-83. Sargeant, J. M., M. W. Sanderson, R. A. Smith, and D. D. Griffin. 2003. Escherichia coli 0157 in feedlot cattle feces and water in four major feeder-cattle states in the USA. Prev Vet Med 61 :127—35. 184 218. 219. 220. 221. 222. 223. 224. 225. 226. 227. 228. 229. 230. Saridakis, C. E., R. P. Johnson, A. Benson, K. Ziebell, and C. L. Gyles. 2004. Influence of animal origin and lineage on survival of Escherichia coli O157:H7 strains in strong and weak acid challenges. J Food Prot 67:1591-6. Sawyer, S. 1989. Statistical tests for detecting gene conversion. Mol Biol Evol 6:526-38. Schmidt, H. 2001. Shiga-toxin-converting bacteriophages. Res Microbiol 152:687-95. Schoolnik, G. K., M. I. Voskull, D. Schnappinger, F. H. Yildiz, K. Meibom, N. A. Dolganov, M. A. Wilson, and K. H. Chang. 2001. Whole genome DNA microarray expression analysis of biofilm development by Vibrio cholerae 01 E1 Tor. Methods Enzymol 336:3—18. Sengupta, N., K. Paul, and R. Chowdhury. 2003. The global regulator ArcA modulates expression of virulence factors in Vibrio cholerae. Infect Immun 71 :5583-9. Sharma, V. K., and R. L. Zuerner. 2004. Role of hha and ler in transcriptional regulation of the esp operon of enterohemorrhagic Escherichia coli O157:H7. J Bacteriol 186:7290—301 . Shl, X., and G. N. Bennett. 1994. Effects of rpoA and cysB mutations on acid induction of biodegradative arginine decarboxylase in Escherichia coli. J Bacteriol 176:7017—23. ' Shin, 8., M. P. Castanie-Cornet, 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. Mol Microbiol 41:1133-1150. Shin, 8., S. G. Song, D. 8. Lee, J. 6. Pan, and 0. Park. 1997. Involvement of iclR and rpoS in the induction of acs, the gene for acetyl coenzyme A synthetase of Escherichia coli K-12. FEMS Microbiol Lett 146:103-8. Small, P., D. Blankenhorn, D. Welty, E. Zinser, and J. L. Slonczewski. 1994. Acid and base resistance in Escherichia coli and Shigella flexneri: role of rpoS and growth pH. J Bacteriol 176:1729—37. Small, P. L., and S. R. Waterman. 1998. Acid stress, anaerobiosis and gadCB: lessons from Lactococcus lactis and Escherichia coli. Trends Microbiol 6:214—6. 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—6. Spears, K. J., A. J. Roe, and D. L. Gally. 2006. A comparison of enteropathogenic and enterohaemorrhagic Escherichia coli pathogenesis. FEMS Microbiol Lett 255: 1 87-202. 185 231. 232. 233. 234. 235. 236. 237. 238. 239. 240. 241. Sperandio, V., C. C. Li, and J. B. Kaper. 2002. Quorum-sensing Escherichia coli regulator A: a regulator of the LysR family involved in the regulation of the locus of enterocyte effacement pathogenicity island in enterohemorrhagic E. coli. Infect Immun 70:3085—93. Sperandio, V., J. L. Mellles, W. Nguyen, 8. Shin, and J. B. Kaper. 1999. Quorum sensing controls expression of the type III secretion gene transcription and protein secretion in enterohemorrhagic and enteropathogenic Escherichia coli. Proc Natl Acad Sci U S A 96:15196-201. Sperandio, V., A. G. Torres, J. A. Giron, and J. B. Kaper. 2001. Quorum sensing is a global regulatory mechanism in enterohemorrhagic Escherichia coli O157:H7. J Bacteriol 183:5187-97. Stelbel, J. P., and G. J. Rosa. 2005. On reference designs for microarray experiments. Stat Appl Genet Mol Biol 4:Article36. Stim-Herndon, 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 142:131 1-20. Stopforth, J. D., Y. Yoon, l. M. Barrnpalia, J. Samelis, P. N. Skandamls, and J. N. Sofos. 2005. Reduction of Listeria monocytogenes populations during exposure to a simulated gastric fluid following storage of inoculated frankfurters formulated and treated with preservatives. Int J Food Microbiol 99:309-19. Swerdlow, D. L., B. A. Woodruff, R. 0. Brady, P. M. Griffin, 8. Tippen, H. D. Donnell, Jr., E. Geldreich, B. J. Payne, A. Meyer, Jr., J. G. Wells, and et al. 1992. A waterborne outbreak in Missouri of Escherichia coli O157:H7 associated with bloody diarrhea and death. Ann lntem Med 117:812-9. Tamayo, R., A. M. Prouty, and J. S. Gunn. 2005. Identification and functional analysis of Salmonella enterica serovar Typhimurium PmrA-regulated genes. FEMS Immunol Med Microbiol 43:249-58. Tamplin, M. L. 2005. Inactivation of Escherichia coli O157:H7 in Simulated Human Gastric Fluid. Appl Environ Microbiol 71:320—5. Tao, H., C. Bausch, 0. Richmond, F. R. Blattner. and T. Conway. 1999. Functional genomics: expression analysis of Escherichia coli growing on minimal and rich media. J Bacteriol 181:6425-40. Tarr, C. L., T. M. Large, 0. L. Moeller, D. W. Lacher, P. I. Tarr, D. W. Acheson, and T. S. Whittam. 2002. Molecular characterization of a serotype 0121:H19 clone, a distinct Shiga toxin-producing clone of pathogenic Escherichia coli. Infect Immun 70:6853-9. 186 242. 243. 244. 245. 246. 247. 248. 249. 250. 251. 252. Tarr, P. l., S. S. Bilge, J. C. Vary, Jr., 8. Jelacic, R. L. Habeeb, T. R. Ward, M. R. Baylor, and T. E. Besser. 2000. Iha: a novel Escherichia coli O157:H7 adherence-conferring molecule encoded on a recently acquired chromosomal island of conserved structure. Infect Immun 68:1400-7. Tatsuno, l., K. Nagano, K. Taguchl, L. Rong, H. Mori, and C. Sasakawa. 2003. Increased adherence to Caco-2 cells caused by disruption of the thE and yhiF genes in enterohemorrhagic Escherichia coli O157:H7. Infect Immun 71 :2598-606. Taylor, D. E., M. Rocker, M. Keelan, L. K. Ng, I. Martin, N. T. Perna, N. T. Burland, and F. R. Blattner. 2002. Genomic variability of O islands encoding tellurite resistance in enterohemorrhagic Escherichia coli O157:H7 isolates. J Bacteriol 184:4690-8. R Core Development Team. 2005. R: A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. Teunis, P., K. Takuml, and K. Shinagawa. 2004. Dose response for infection by Escherichia coli O157:H7 from outbreak data. Risk Anal 24:401-7. Tobe, T., S. A. Beatson, H. Tanlguchi, H. Abe, C. M. Bailey, A. Flvlan, R. Younis, 8. 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. Proc Natl Acad Sci U S A 103:14941-6. Tramontl, A., M. De Canio, l. Delany, V. Scarlato, and D. De Blase. 2006. Mechanisms of transcription activation exerted by GadX and GadW at the gadA and gadBC gene promoters of the glutamate-based acid resistance system in Escherichia coli. J Bacteriol 188:8118—27. Trifillis, A. L., M. S. Donnenberg, X. Cui, R. 6. Russell, 8. J. Utsalo, H. L. Mobley, and J. W. Warren. 1994. Binding to and killing of human renal epithelial cells by hemolytic P-fimbriated E. coli. Kidney Int 46:1083-91. Tseng, C. P., J. Albrecht, and R. P. Gunsalus. 1996. Effect of microaerophilic cell growth conditions on expression of the aerobic (cyoABCDE and cydAB) and anaerobic (narGHJl, frdABCD, and dmsABC) respiratory pathway genes in Escherichia coli. J Bacteriol 178:1094-8. Tsuji, H., K. Hamada, S. Kawanishi, A. Nakayama, and H. Nakajima. 2002. An outbreak of enterohemorrhagic Escherichia coli 0157 caused by ingestion of contaminated beef at grilled meat-restaurant chain stores in the Kinki District in Japan: epidemiological analysis by pulsed-field gel electrophoresis. Jpn J Infect Dis 55:91-2. Tucker, D. L., N. Tucker, and T. Conway. 2002. Gene expression profiling of the pH response in Escherichia coli. J Bacteriol 184:6551-8. 187 253. 254. 255. 256. 257. 258. 259. 260. 261. 262. 263. 264. Uehara, T., K. Suefujl, T. Jaeger, C. Mayer, and J. T. Park. 2006. MurQ Etherase is required by Escherichia coli in order to metabolize anhydro-N- acetylmuramic acid obtained either from the environment or from its own cell wall. J Bacteriol 188:1660—2. Uljas, H. E., and 8. C. lngham. 1998. Survival of Escherichia coli O157:H7 in synthetic gastric fluid after cold and acid habituation in apple juice or trypticase soy broth acidified with hydrochloric acid or organic acids. J Food Prot 61:939- 47. Vaara, M. 1992. Agents that increase the permeability of the outer membrane. Microbiol Rev 56:395-411. Vallance, B. A., 0. Chan, M. L. Robertson, and B. B. Finlay. 2002. Enteropathogenic and enterohemorrhagic Escherichia coli infections: emerging themes in pathogenesis and prevention. Can J Gastroenterol 16:771-8. Varma, J. K., K. D. Greene, M. E. Reller, S. M. DeLong, J. Trottler, S. F. Nowicki, M. DiOrio, E. M. Koch, T. L. Bannerman, S. T. York, M. A. Lambert- Fair, J. G. Wells, and P. S. Mead. 2003. An outbreak of Escherichia coli O157 infection following exposure to a contaminated building. Jama 290:2709-12. Wagner, P. L., D. W. Acheson, and M. K. Waldor. 1999. lsogenic Iysogens of diverse shiga toxin 2-encoding bacteriophages produce markedly different amounts of shiga toxin. Infect Immun 67:6710-4. Wagner, P. L., J. leny, M. N. Neely, D. W. Acheson, D. I. Friedman, and M. K. Waldor. 2002. Bacteriophage control of Shiga toxin 1 production and release by Escherichia coli. Mol Microbiol 44:957-70. Wagner, P. L., M. N. Neely, X. Zhang, D. W. Acheson, M. K. Waldor, and D. I. Friedman. 2001. Role for a phage promoter in Shiga toxin 2 expression from a pathogenic Escherichia coli strain. J Bacteriol 183:2081-5. Watanabe. Y., K. Ozasa, J. H. Mermin, P. M. Griffin, K. Masuda, S. lmashuku, and T. Sawada. 1999. Factory outbreak of Escherichia coli O157:H7 infection in Japan. Emerg Infect Dis 5:424-8. Waterman, S. R., and P. L. Small. 1998. Acid-sensitive enteric pathogens are protected from killing under extremely acidic conditions of pH 2.5 when they are inoculated onto certain solid food sources. Appl Environ Microbiol 64:3882-6. Waterman, S. R., and P. L. Small. 1996. Characterization of the acid resistance phenotype and rpoS alleles of Shiga-like toxin-producing Escherichia coli. Infect Immun 64:2808-11. Waterman, S. R., and P. L. Small. 1996. Identification of sigma S-dependent genes associated with the stationary-phase acid-resistance phenotype of Shigella flexneri. Mol Microbiol 21:925-40. 188 265. 266. 267. 268. 269. 270. 271. 272. 273. 274. 275. 276. Weagant, S. D., J. L. Bryant, and D. H. Bark. 1994. Survival of Escherichia coli O157:H7 in mayonnaise and mayonnaise-based sauces at room and refrigerated temperatures. J Food Prot 57:629-631. Weber, H., T. Polen, J. Heuvellng, V. F. Wendlsch, and R. Hengge. 2005. Genome-Wide Analysis of the General Stress Response Network in Escherichia coli: {sigma}S-Dependent Genes, Promoters, and Sigma Factor Selectivity. J Bacteriol 187:1591-603. Welch, R. A., V. Burland, G. Plunkett, 3rd, P. Redford, P. Roesch, D. Rasko, E. L. Buckles, S. R. Liou, A. Boutin, J. Hackett, D. Stroud, G. F. Mayhew, D. J. Rose, 8. Zhou, D. C. Schwartz, N. T. Perna, H. L. Mobley, M. S. Donnenberg, and F. R. Blattner. 2002. Extensive mosaic structure revealed by the complete genome sequence of uropathogenic Escherichia coli. Proc Natl Acad Sci U S A 99:17020-4. Werber, D., A. Fruth, A. Liesegang, M. Littmann, U. Buchholz, R. Prager, H. Karch, T. Breuer, H. Tschape, and A. Ammon. 2002. A multistate outbreak of Shiga toxin-producing Escherichia coli 026:H11 infections in Germany, detected by molecular subtyping surveillance. J Infect Dis 186:419-22. Whiting, R. C., and M. H. Golden. 2002. Variation among Escherichia coli O157:H7 strains relative to their growth, survival, thermal inactivation, and toxin production in broth. Int J Food Microbiol 75:127-33. Whittam, T. 8., and T. M. Bergholz. 2007. Molecular subtyping, source tracking, and food safety, p. 93-136. In Microbial Source Tracking, J.W. SantoDomingo and M. Sadowsky, (ed). ASM Press, Washington D. C. Yang, Y. H., and T. Speed. 2002. Design issues for cDNA microarray experiments. Nat Rev Genet 3:579—88. Yona-Nadler, C., T. Umanski, S. Aizawa, D. Friedberg, and I. Rosenshine. 2003. Integration host factor (IHF) mediates repression of flagella in enteropathogenic and enterohaemorrhagic Escherichia coli. Microbiology 149:877-84. Yuk, H. G., and D. L. Marshall. 2004. Adaptation of Escherichia coli O157:H7 to pH Alters Membrane Lipid Composition, Verotoxin Secretion, and Resistance to Simulated Gastric Fluid Acid. Appl Environ Microbiol 70:3500-5. Yuk, H. G., and D. L. Marshall. 2005. Influence of acetic, citric, and lactic acids on Escherichia coli O157:H7 membrane lipid composition, verotoxin secretion, and acid resistance in simulated gastric fluid. J Food Prot 68:673-9. Zambrano, M. M., D. A. Siegele, M. Almiron, A. Tormo, and R. Kolter. 1993. Microbial competition: Escherichia coli mutants that take over stationary phase cultures. Science 259: 1 757-60. Zhao, T., and M. P. Doyle. 1994. Fate of enterohemorrhagic Escherichia coli O157:H7 in commercial mayonnaise. J Food Prot 57:780-783. 189 277. Zhao, T., M. P. Doyle, and R. E. Besser. 1993. Fate of enterohemorrhagic Escherichia coli O157:H7 in apple cider with and without preservatives. Appl Environ Microbiol 59:2526-30. 190 lI)1111111111111)Iii)