IN VITRO AND IN VIVO COLD SHOCK RESPONSE IN VIBRIO VULNIFICUS Except where reference is made to the work of others, the work described in this dissertation is my own or was done in collaboration with my advisory committee. This dissertation does not include proprietary or classified information Suttinee Limthammahisorn Certificate of Approval: Yolanda J. Brady, Co-Chair Covadonga R. Arias, Co-Chair Associate Professor Associate Professor Fisheries and Allied Aquacultures Fisheries and Allied Aquacultures Richard K. Wallace David LaDon Swann Professor Associate Research Professor Fisheries and Allied Aquacultures Fisheries and Allied Aquacultures Yifen Wang Joe F. Pittman Assistant Professor Interim Dean Biosystems Engineering Graduate School IN VITRO AND IN VIVO COLD SHOCK RESPONSE IN VIBRIO VULNIFICUS Suttinee Limthammahisorn A Dissertation Submitted to the Graduate Faculty of Auburn University in Partial Fulfillment of the Requirement for the Degree of Doctor of Philosophy Auburn, Alabama August 4, 2007 iii IN VITRO AND IN VIVO COLD SHOCK RESPONSE IN VIBRIO VULNIFICUS Suttinee Limthammahisorn Permission is granted to Auburn University to make copies of this dissertation at its discretion, upon request of individuals or institutions and at their expense. The author reserves all publication rights. Signature of Author Date of Graduation iv VITA Suttinee Limthammahisorn, daughter of Pravit and Pranee Sukaphatana, was born on December 19, 1964, in Chonburi, Thailand. She graduated from Assumption College Sriracha High School, Chonburi, Thailand in 1983. She attended Kasetsart University, Thailand and graduated with a Bachelor of Science degree majoring in Fisheries in 1988. She obtained a Master of Science degree in Environmental Technology Management from Mahidol University, Thailand in 1997. She entered Graduate school in the Department of Fisheries and Allied Aquacultures at Auburn University in August 2003. She is a senior fishery biologist in Department of Fisheries, Thailand. She married Somkiat Limthammahisorn on December 31, 1990. They have one son, Meen (Max) Limthammahisorn. v DISSERTATION ABSTRACT IN VITRO AND IN VIVO COLD SHOCK RESPONSE IN VIBRIO VULNIFICUS Suttinee Limthammahisorn Doctor of Philosophy, August 4, 2007 (M.Sc., Mahidol University, 1997) (B.Sc., Kasetsart University, 1988) 167 Typed Pages Directed by Yolanda J. Brady and Covadonga R. Arias Vibrio vulnificus is a serious food-borne pathogen associated with the consumption of raw oysters. The Interstate Shellfish Sanitation Conference (ISSC) recommends refrigeration of Gulf Coast oysters after harvesting to minimize pathogen risks. Cooling temperature control of harvested oysters could potentially induce cold shock response in V. vulnificus. In order to determine the potential for cold shock mediated survival of V. vulnificus, my research investigated in vitro and in vivo cold shock response in V. vulnificus regarding its growth and the molecular mechanisms underlying this phenomenon. The effect of cold shock on V. vulnificus cultures was investigated under ISSC guidelines by the drop plate method. V. vulnificus cultures were incubated at 35, 25, 20, and 15?C for 10 h, 12 h, 14 h, and 36 h, respectively, and then abruptly shifted to 7.2?C and 4?C for 168 h. The mean counts of V. vulnificus after 7 day refrigeration at 7.2?C vi was significantly greater (P<0.05) than at 4?C in all treatments. Cold adaptation response enhanced cell culturability and was observed when cells were adapted to 15?C prior to cold shock at 7.2?C. When V. vulnificus was artificially inoculated into oysters, its growth during cold shock was investigated using colony hybridization. After oysters were kept at 35?C and 25?C for 10 h and 12 h, respectively, and abruptly shifted to 4?C, cells exhibited a moderate decline in culturability. Oysters kept at 15?C prior to a 4?C shift for 168h, showed a slight decline in V. vulnificus cell numbers. These findings suggested that incubation of V. vulnificus at 15?C could induce cold adaptation and should be avoided during post-harvest conditions. Four cold shock genes (csp s ) encoding small cold shock proteins in V. vulnificus were cloned. Analysis of in vitro and in vivo cold shock gene expression was displayed by reverse transcriptase polymerase chain reaction (RT-PCR). All csp s tested in this study can not be considered cold shock inducible genes under in vitro condition, while some csp s showed weak induction after cold shock under in vivo experiments. Besides, the oxidative related genes (oxyR and katG) were induced during in vitro study, while these transcripts were not detected during in vivo experiment. The transcript levels of 16SrRNA and rpoS were constant throughout the study. Overall, these results demonstrated that gene expression of V. vulnificus was different between in vitro and in vivo conditions during cold shock. vii ACKNOWLEDGEMENTS The author extremely grateful to Dr. David B. Rouse, Department Head of Fisheries and Allied Aquaculture and my major advisor Dr. Yolanda J. Brady for providing me an opportunity to pursue research and obtain a doctorate from this country. I express my gratitude and obligation to my co-advisor, Dr. Covadonga R. Arias for her guidance and inspiring comments and suggestions throughout my research. I am grateful for time and expertise offered by my committee members, Dr. Richard K. Wallace, Dr. LaDon Swann, and Dr. Yifen Wang. I thank Karl J. Hayden and Briana Keaffer for their technical support in the Fish Health Laboratory. I greatly appreciate the assistance and support from Scott Rikards and the staff of AU Shellfish Lab at Dauphin Island during the depuration of oysters. I would like to thank the administration staff in the Department of Fisheries and Allied Aquacultures for their help and support. The friendship and encouragement extended to me by the graduate students in the department is also gratefully acknowledged. Most of all, I acknowledge the love of my son. He has been the motivating force pushing me forward to obtain the degree. My special thanks and deep appreciation go to my husband for his patience, unwavering support, encouragement and sacrifice during these years. viii Style manual or journal used: Applied and Environmental Microbiology Computer software used: Window XP, Microsoft Word 2003, Microsoft Excel 2003 Adobe Photoshop 7.0, Vector NTI Advance 9.0 ix TABLE OF CONTENTS LIST OF TABLES??????????????????????????.....xi LIST OF FIGURES??????????????????????????...xii I. INTRODUCTION???????????????????????.1 II. LITERATURE REVIEW???????????????????......5 III. COLD SHOCK EFFECT ON GROWTH RATE AND SURVIVAL OF VIBRIO VULNIFICUS IN CULTURE BROTH UNDER SHELLSTOCK TEMPERATURE CONTROL CONDITIONS ?????28 Abstract Introduction Materials and methods Results Discussion IV. EFFECT OF COLDSHOCKING ON SURVIVAL OF VIBRIO VULNIFICUS IN SHELLSTOCK OYSTERS??????..........48 Abstract Introduction Materials and methods Results Discussion V. IN VITRO COLD SHOCK GENE EXPRESSION OF VIBRIO VULNIFICUS?????????????????????.65 Abstract Introduction Materials and methods Results Discussion VI. IN VIVO COLD SHOCK GENE EXPRESSION OF VIBRIO VULNIFICUS?????????????????????.94 Abstract Introduction Materials and methods Results Discussion x VII. CONCLUSIONS??????????????????????...113 CUMULATIVE BIBLIOGRAPHY??????????????????..117 APPENDICES???????????????????????...............143 xi LIST OF TABLES 1. Computed theoretical pI and M w values of cold shock proteins encoded by putative cold shock genes of V. vulnificus??????????.86 2. Primers design for cold shock genes of V. vulnificus? ?????????...87 3. Primers used for RT-PCR of cold shock genes and stress-related genes for the in vitro study?????????????????................88 4. Primers used for RT-PCR of cold shock genes and stress-related genes for the in vivo study????????????????????...108 xii LIST OF FIGURES 1. Growth of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 at 35?C????????????????????????????...41 2. Growth of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 at 25?C????????????????????????????...41 3. Growth of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 at 20?C????????????????????????????...42 4. Growth of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 at 15?C????????????????????????????...42 5. Growth of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 at 7.2?C??????????????????????????..........43 6. Growth of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 at 4?C?????????????????????????????.43 7. Cold shock-survival response of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 in MB when cells were incubated at 35?C for 10 h and cold shocked at 7.2?C for 168 h??????????????44 8. Cold shock-survival response of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 in MB when cells were incubated at 25?C for 12 h and cold shocked at 7.2?C for 168 h??????????????44 9. Cold shock-survival response of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 in MB when cells were incubated at 20?C for 14 h and cold shocked at 7.2?C for 168 h??????????????45 10. Cold shock-survival response of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 in MB when cells were incubated at 15?C for 36 h and cold shocked at 7.2?C for 168 h??????????????45 xiii 11. Cold shock-survival response of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 in MB when cells were incubated at 35?C for 10 h and cold shocked at 4?C for 168 h??????????????...46 12. Cold shock-survival response of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 in MB when cells were incubated at 25?C for 12 h and cold shocked at 4?C for 168 h??????????????...46 13. Cold shock-survival response of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 in MB when cells were incubated at 20?C for 14 h and cold shocked at 4?C for 168 h??????????????...47 14. Cold shock-survival response of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 in MB when cells were incubated at 15?C for 36 h and cold shocked at 4?C for 168 h??????????????...47 15. Total heterotrophic bacteria, total Vibrio, and V. vulnificus (CFU/g oyster meat) during depuration of eastern oysters (Crassostrea virginica)??????????????????????..62 16. pH of oyster meat during cold shock experiment????????????..62 17. Cold shock-survival response of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 in oysters when oysters were incubated at 35?C for 10 h and cold shocked at 4?C for 168 h???????????..63 18. Cold shock-survival response of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 in oysters when oysters were incubated at 25?C for 12 h and cold shocked at 4?C for 168 h???????????..63 19. Cold shock-survival response of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 in oysters when oysters were incubated at 15?C for 36 h and cold shocked at 4?C for 168 h???????????..64 20. Vibrio vulnificus gene expression profiles: internal control: 16SrRNA, and cold shock genes: csp1, csp3, csp4, and csp5 during cells were cold shocked from 35?C 10 h to 4?C 168 h ????????????????..89 21. Vibrio vulnificus gene expression profiles: stress related genes: rpoS, oxyR, and katG during cells were cold shocked from 35?C 10 h to 4?C 168 h ?????..90 xiv 22. Vibrio vulnificus gene expression profiles: internal control: 16SrRNA, and cold shock genes: csp1 and csp3 during cells were cold shocked from 35?C 10 h to 15?C 36 h and from 15?C 36 h to 4?C 168 h ??????????.91 23. Vibrio vulnificus gene expression profiles: cold shock genes: csp4 and csp5 during cells were cold shocked from 35?C 10 h to 15?C 36 h and from 15?C 36 h to 4?C 168 h ??????????????????????.........92 24. Vibrio vulnificus gene expression profiles: stress-related genes: rpoS, oxyR, and katG during cells were cold shocked from 35?C 10 h to 15?C 36 h and from 15?C 36 h to 4?C 168 h ??????????????????..........93 25. Vibrio vulnificus gene expression profiles: internal control: 16SrRNA, and cold shock genes: csp1, csp3, csp4, and csp5 during oysters were cold shocked from 35?C 10 h to 4?C 168 h ??????????????109 26. Vibrio vulnificus gene expression profiles: stress related gene: oxyR, katG, and rpoS during oysters were cold shocked from 35?C 10 h to 4?C 168 h ????.110 27. Vibrio vulnificus gene expression profiles: internal control: 16SrRNA, and cold shock genes: csp1, csp3, csp4, and csp5 during oysters were cold shocked from 25?C overnight to 15?C 36 h and from 15?C 36 h to 4?C 168 h? ????.111 28. Vibrio vulnificus gene expression profiles: stress-related genes: oxyR, katG, and rpoS during oysters were cold shocked from 25?C overnight to 15?C 36 h and from 15?C 36 h to 4?C 168 h ???????????????..112 1 I. INTRODUCTION 2 Vibrio vulnificus is a Gram-negative, ubiquitous, estuarine bacterium that is abundant in water, sediments, mollusks, crustaceans, and finfish along the U.S. Gulf Coast (38, 170). Vibrio vulnificus has been blamed for the reduced prices and consumer demand for oysters (Crassostrea virginica), especially Gulf Coast oysters, because of negative publicity associated with health risks. Oysters have served as vectors in the transfer of this indigenous estuarine pathogen to humans (38). Every year millions of Americans consume raw oysters. For some people, ingestion of raw or undercooked oysters containing this bacterium can cause serious illness and even death. This pathogen may cause severe wound infection and primary septicemia (99). Nearly all fatal primary septicemia cases are associated with the consumption of raw oysters. The United States Food and Drug Administration (USFDA), the Centers for Disease Control and Prevention (CDC), the Interstate Shellfish Sanitation Conference (ISSC), and the oyster industry faced with a major challenge in managing public health risk of V. vulnificus. In 2001, the ISSC proposed the ?Vibrio vulnificus Risk Management Plan?, and set a goal for reduction of illnesses caused by the consumption of commercially harvested raw or undercooked oysters by at least 60% by the year 2007 and 2008 (44). The ISSC recommended that post-harvest treated consumable oysters must contain less than 30 colony forming units of V. vulnificus per gram of oyster meat (45). Typically, the concentration of V. vulnificus at consumption is influenced by ambient air temperature during harvest, the time from harvest until the oysters are placed under refrigeration, the time it takes the oyster to cool down under refrigeration, and the length of refrigeration time until consumption (183). Potential infection by this bacterium could be reduced by limiting time from harvest to refrigeration in at-risk-harvesting areas and by using proper 3 refrigeration temperatures. In one effort to achieve the ISSC goal, restrictive shellstock temperature control measures have been recommended by FDA/ISSC and implemented by the industry. This control limits the time from harvesting of oysters to refrigeration. The length of time depends upon the average monthly maximum water temperature at the time of harvest. Once placed under temperature control, shellstock shall be maintained at 10?C or below throughout processing, distribution and storage (45). However, many bacteria, especially food-borne pathogens have been reported to manifest their tolerance to cold temperatures by altering their cellular protein profiles and continuing to grow (6, 36, 55, 75, 90). This physiological response of living cells to temperature downshift is referred to as cold shock response (52). The cooling strategy recommended by ISSC may potentially cause a cold shock response in V. vulnificus inhabitating oysters. To date, cold shock response in V. vulnificus has not been fully elucidated although this pathogen greatly impacts oyster industry and public health. The impact of cold shock response on the physiology of this pathogenic microorganism in oysters under typical industry practices from harvest to consumption, particularly in regard to their survival, needs to be further investigated. Although a wide array of experimental data addressed the role of cold shock response in many bacteria, no clear indication on the molecular mechanisms underlying this phenomenon has been proposed in V. vulnificus so far. Only two reports have investigated a V. vulnificus response mechanism to cold temperature. In 1995, McGovern and Oliver (110) showed that when V. vulnificus is subjected to a moderate temperature downshift from 23 to 13?C, protein synthesis declines but new cold-induced proteins are expressed. In 1999, Bryan et al. (17) reported that a brief cold adaptation 4 period for 3 h at 15?C enhanced survival of V. vulnificus at 6?C. They concluded that cold adaptive protective proteins might enhance survival and tolerance at cold temperature. Moreover, cold shock has been shown to alter gene expression in bacterial cells in order to compensate for this cold stress. Little is known about V. vulnificus gene expression in response to cold shock. While several reports have investigated in vitro and in situ gene expression of V. vulnificus during temperature downshift in the ?viable but non-culturable? (VBNC) state (46, 160, 161), in vivo gene expression of this pathogen has not yet been reported in response to the cold or any other situation. Since oysters can act as reservoirs for V. vulnificus, it is crucial to study the cold shock gene expression while this pathogen is embedded in oysters. This scientific knowledge from in vitro and in vivo studies of cold shock response of V. vulnificus under USFDA/ISSC shellstock temperature control guidelines will assist in the development of an effective risk reduction strategy for V. vulnificus septicemia associated with the consumption of raw oysters. 5 II. LITERATURE REVIEW 6 Overview Raw oysters have the potential to cause human disease because they serve as vectors for viral, bacterial and parasitic pathogens (156). The greatest numbers of seafood-associated illnesses are from the consumption of contaminated raw shellfish (such as mussels, oysters, clams) as a result of their filter feeding habit that concentrate particles present in the surrounding waters. Two general groups of pathogenic bacteria may be present in coastal sea water and may be entrapped by shellfish: the first group consists of nonindigenous bacterial pathogens that enter the water from animals and humans, the latter includes indigenous bacteria in the marine environment, mainly members of the family Vibrionaceae (148). Typically, bacterial pathogens associated with fecal contamination have represented only 4% of the shellfish associated outbreaks, while naturally-occurring bacteria accounted for 20% of shellfish-associated illnesses and 99% of the deaths. Most of these indigenous bacteria fall into the family Vibrionaceae which includes the genus Vibrio (101). The Vibrio genus comprises more than 30 species, some of them being pathogenic to human and had been associated with food- borne related diseases (22). 7 1. Vibrio vulnificus Vibrio vulnificus is recognized as the most invasive and rapidly lethal Vibrio and is the leading cause of reported seafood-related deaths in the United States (11). It accounts for 95% of seafood related deaths in the U.S. It is an opportunistic human pathogen that can cause gastroenteritis, septicemia, and wound infections in humans. Typically, V. vulnificus infections are associated with the consumption of raw oysters and can produce rapidly fatal septicemia in individuals (151). In humans, V. vulnificus is associated with two disease patterns, primary septicemia and wound infection. The major pattern of infection is a rapidly progressing septicemia. Septicemia caused by this pathogen occur in individuals with chronic liver disease shortly after eating raw oyster to several days of infection, with a fatality rate of over 60%. In immunocompromised hosts, V. vulnificus infection can cause high fever, nausea, myalgia, and abdominal cramps 24 to 48 hours after eating contaminated food. Because the organism can cross the intestinal mucosa rapidly, sepsis and cutaneous bullae can occur within 36 hours of the initial onset of symptoms. Cases are most commonly reported during warm weather months (April- November) and often are associated with eating raw oysters. Progression of the illness can be very rapid from asymptomatic to death within 24 hours. About 75% of patients with Vibrio septicemia have pre-existing hepatic diseases. Wound infections with V. vulnificus usually develop after trauma and exposure to the marine environment by handling seafood products and progress rapidly. Wound infections most commonly present cellulites, and the fatality rate can be up to 25%. About one-third of patients with wound infections have some underlying disease. 8 However, infections that occur through wounds are not limited to immunocompromised individuals. The etiological role of V. vulnificus in gastroenteritis has not been proven. An indole-negative biogroup of this Vibrio, which is pathogenic to eels, has also been associated with human infections (99). Characteristics of Vibrio vulnificus Vibrio vulnificus is a Gram-negative, oxidase-positive, obligately halophilic, motile, curved bacillus with a single polar flagellum. It belongs to the family Vibrionaceae, and is distinguished from other members of the Vibrio genus by its ability to ferment lactose. Vibrio vulnificus can grow in NaCl with concentration as high as 5% and does not require Mg +2 or K + for its growth. Consequently, it can easily be cultured in a variety of media. Vibrio vulnificus was first isolated in 1964 by the U.S. Centers for Disease control (CDC), although at that time it was mis-identified as a virulent strain of V. parahaemolyticus (165). It was finally recognized as a distinct species in 1976 after researchers realized that many clinical cases of foodborne septicemias and wound infections were caused by a pathogen with unique characteristics from other Vibrio species. The name ? vibrio? meaning ?to quiver?, and the word ?vulnificus? is derived from Latin word ?vulnus? meaning ?wound? and was given its official taxonomic status in 1980 (116). Vibrio vulnificus has been classified based on biotypes, lipopolysaccharide (LPS) antigens, capsule, and most recently genetic sequences. There are three biotypes of V. vulnificus according to their different biochemical and biological properties. Biotype 1 9 is ubiquitous in estuarine environments, is an opportunistic human pathogen, produces indole and ornithine decarboxylase, and exhibits several immunologically distinct lipopolysaccharide (LPS) types (66, 100). Biotype 2 is primarily related with eels but can be pathogenic to humans, is characterized by a homogeneous lipopolysaccharide (LPS), and is negative for indole production and ornithine decarboxylase (3, 10). A third group, biotype 3, was recently identified and causes wound infections and bacteremia in humans handling St. Peter?s fish (Tilapia spp.) in Israel (12). Geographical range distribution and environmental conditions The presence of V. vulnificus is not associated with pollution. It is normally found in estuarine environments, flourishes in warm temperatures. It is part of the bacterial flora of the marine environment, and is distributed worldwide (175). Vibrio vulnificus has been isolated from waters of widely varying temperatures and salinities (187). Several studies along the Gulf of Mexico and the east and west coasts of the United States have shown strong relationships between salinity, water temperature, and the isolation of V. vulnificus (38, 125, 133, 134, 169). Prevalence of V. vulnificus in the marine ecosystem is correlated with seasonality since water temperature plays a significant role in controlling the densities of this bacterium. Vibrio vulnificus has been recovered from waters with a range of temperatures from 9?C to 31?C, but has seldom been recovered from cold waters and has frequently been found in high concentrations in shellfish and surface water during the summer, April to September, when water temperatures were above 25?C. In fact, V. vulnificus optimal temperature growth is 10 35?C, and like all human pathogens can also grow at 37?C. However, this bacterium has never been responsible for human infections in that area where summer temperatures are optimal for its growth (117). Under suboptimal temperatures the number of V. vulnificus increases during 6 days incubation at 13?C to 22?C, but V .vulnificus survives poorly at less than 8.5?C (85). Pfeffer et al. (145) reported that water temperature accounted for most of the variability in the concentration of V .vulnificus, and that low water temperatures (<10?C) have a negative effect for its isolation. There have been several theories to explain the seasonality observed in V. vulnificus isolation. One possible reason that might explain why V. vulnificus is difficult to culture during cold-water months is that this bacterium can enter a viable but nonculturable state known as VBNC. It has been proven that V. vulnificus enters a VBNC state at temperatures below 10?C under low nutrient conditions (126). Typically, U.S. Gulf Coast oysters harbor about 10 3 to 10 4 V. vulnificus cells g -1 during the warmer months from April through October but typically less than 10 cells g -1 during colder months (32, 171). Illnesses caused by V. vulnificus have been reported every year, primarily along the Gulf of Mexico in the U.S. during the summer months. In general, the relationship between salinity and V. vulnificus presence is not as evident as between temperature and prevalence. It has been isolated from waters with a wide range of salinities from 1 to 34 parts per thousand (ppt) although estuarine waters with salinities between 6 and 16 ppt are considered optimal for V. vulnificus populations. Salinities greater than 25 ppt have been shown to adversely affect survival of this bacterium under laboratory conditions. However, V. vulnificus has been isolated from Mediterranean waters and shellfish with salinities around 35 ppt. 11 Isolation and identification methods for Vibrio vulnificus Several isolation and identification techniques have been developed to distinguish V. vulnificus from other vibrios. Thiosulfate citrate bile salts sucrose (TCBS) agar was originally used for isolation of pathogenic vibrios and became widely used for the isolation of Vibrio spp. from environmental samples. This medium differentiates V. cholerae from V. parahaemolyticus and V. vulnificus on the basis of sucrose fermentation (85, 153). Sodium dodoecyl sulfate-polymyxin B-sucrose agar (SPS) was used to separate V. vulnificus from V. parahaemolyticus, which are both negative for sucrose fermentation and had similar morphology on TCBS (92). Many selective and differential media have been developed for V. vulnificus isolation including V. vulnificus (VV) agar (15), cellobiose-polymyxin B colistin (CPC) agar (107), modified CPC (mCPC) agar (171), V. vulnificus enumeration (VVE) agar (112), V. vulnificus agar (VVA) (188), cellobiose-colistin (CC) agar (65), V. vulnificus medium (VVM) (20), modified VVM (VVMc) agar (21). Confirmation assays for V. vulnificus include biochemical or physiological tests, immunological assays, DNA probes, and PCR analysis (60). For V. vulnificus enumeration in oysters, USFDA recommended two approaches. One uses a most probable number (MPN) enrichment step coupled with biochemical or molecular confirmation. The second performs a direct isolation on selective media followed by colony hybridization with a DNA probe (VVAP) (82). The VVAP probe method has been widely used for molecular confirmation of V. vulnificus (21, 65). A quantitative PCR or real-time PCR assay is now available for V. vulnificus detection (18, 138). However, this method does not discriminate live from dead cells. 12 Reverse transcription PCR is an alternative method that can be used to detect RNA from viable but nonculturable (VBNC) cells of V. vulnificus since only live cells yield a positive result (46). Methods by reducing Vibrio vulnificus in oysters In order to reduce the number of unwanted V. vulnificus cells to acceptable levels for human consumption, several studies have been investigating methods that will eliminate this pathogen. One of the most common methods in controlling pathogen proliferation in oysters is cold temperatures (30, 31, 129, 170). However these studies have shown that cold storage cannot eliminate this bacterium to safe levels for consumption. Thoroughly cooking the oysters then storing them under either hot (60?C or higher) or cold (4?C or lower) conditions completely eliminates the bacteria, but this methodology is unfit for the half-shell market (64). Ionized or gamma irradiation (69, 73) is effective at eliminating Vibrio but this process requires an expensive facility and is not yet federally approved. The use of additives such as organic acids (Generally Recognized as Safe, GRAS) (166, 167), butylated hydroxyanisole (BHA), or diacetyl (11), although effective in reducing Vibrio contamination, produce unacceptable flavor changes. The use of hot sauce (168) was not successful in reducing numbers of this bacterium. Depuration has been proved to be effective in removing indicator bacteria but it is ineffective in reducing the numbers of this pathogen. In addition, some bacteria may persist and even multiply during depuration (86, 170). Exposing oysters to high-salinity (30 to 34 ppt) environments seemed an effective method to eliminate V. vulnificus in 13 oysters; however, this method requires a period of time of more than one month (117). Rapid chilling by ice immersion at 7.2?C (149) has been shown to reduce V. vulnificus, but not to recommended limits. Recently, electrolyzed oxidizing water (150) was found to be effective in eliminating this bacteria in oysters, but this method directly increased the mortality of oysters. To date, most efforts to reduce Vibrio contamination in live oysters have not been successful. Post-harvest treatments have been recommended in order to reduce V. vulnificus numbers in oysters. Currently, several effective post-harvest treatments have been used commercially and approved by the Food and Drug Administration. These methods include high hydrostatic pressure (9), low-temperature pasteurization (5), and individual quick freezing (63). These treatments; however, required the use of specialized equipment, which increases the cost of the post harvest process. 2. The United States Oyster industry Oyster industry profile Two major species of oysters are produced in the U.S.: Crassostrea virginica (Atlantic oyster or Eastern oyster) and the non-native Crassostrea gigas (Pacific oyster). The amount of eastern oysters accounts for roughly 75 percent of total U.S. harvests. The U.S. annual yields of eastern oysters were 10,112 metric tons in year 2005 (121). Oysters are typically produced in three different ways: natural, managed and cultivated. Natural oysters grow and reproduce without human intervention and are available for harvest by 14 anyone with the appropriate licenses and permits. In contrast, managed natural oysters are supervised by harvesters who scrape the oyster bed periodically to reduce clustering. In the case of cultivated oysters, immature oysters are transported to man-made beds where they are allowed to mature. In the Northwest and Northeast of the U.S., a significant amount of shellstock oysters are produced on cultivated beds, whereas in the Gulf waters historically, they have been primarily harvested from wild reefs (119). All oyster dealers in the U.S. must be certified under the National Shellfish Sanitation Program (NSSP) in order to market oyster products in interstate commerce. Processing plants that ship oysters are certified as interstate or intrastate shippers. Interstate certified shippers may ship oysters across state lines, whereas intrastate certified shippers may ship oysters only within their state borders. Interstate shippers are inspected and certified by individual states, which later provide the list to the USFDA. The USFDA uses this information to publish the Interstate Certified Shellfish Shippers List. Intrastate shippers are also inspected and certified by individual states, but those states manage their own lists. State agencies involved in certification include the departments of health, marine resources, agriculture, national resources, and wildlife and fisheries (119). The FDA carefully monitors oysters from their growing waters through processing plants, and finally to retail outlets. The NSSP maintains a Manual of Operations, which is routinely revised by the Interstate Shellfish Sanitation Conference (ISSC), a voluntary group composed of members from FDA, state health regulatory agencies and industry (105). Generally, there are four main sectors in the oyster industry; harvesters, wholesalers, processors, and retailers. Oyster harvesting includes wild and cultivated 15 oysters. Harvested oysters are generally delivered to wholesalers and processors, and in some cases, they are directly delivered to restaurants or other retail outlets. They are generally sold in dozens, by the bushel bag, or in bushel-fraction boxes (105). Processors normally sell their products as fresh, raw, shucked, processed halfshell, smoked, cooked, canned, and breaded oysters. Oysters generally reach consumers as live in the shell, fresh, frozen, or as canned products. Shucked oyster meats are graded and sold according to size in 8 ounce (226 g) and 12 ounce (342 g) cup, or in pint (0.55L), quart (1.1 L), or gallon (3.78 L) containers (144). Oyster harvesting is done by handpicking, tonging, and dredging from oyster boats. Oysters are harvested throughout the year; but the meat yield differs between seasons (104). United States per capita oyster consumption was 0.11 kg in the year 2003. Consumers judge the quality of the raw half-shell oysters based on organoleptic characteristics including appearance (size, shape, color), odor, flavor (sweetness and saltiness), and texture (firmness). Consumers prefer cup-shaped oysters where the meat fills the shell. Color is less important to consumers, but unpleasant odors are regarded as an indication of spoilage. The public prefers fresh oysters that are mild and salty in flavor, have no off-flavor, and are tender (not mushy) in texture (26, 105). Individuals generally consume cooked oysters at home and both raw and cooked oysters in restaurants. Demand for in-shell oysters is assumed to be high during the summer travel and vacation season; whereas demand for shucked oysters peaks during the winter holiday season (4). The U.S. imported approximately 10.48 million kilograms of farmed and wild oysters during 2004. The value of imports during that period was $ 47 million. South 16 Korea is the major exporter of oysters to the U.S., accounting for 53.6% of total U.S. oyster imports in 2006. The U.S. imports oysters in live, fresh, frozen, dried, salted, brine, canned smoked and canned forms. On the other hand, in 2004, the U.S. exported approximately 3.4 million kilograms of oysters overseas. France was the major importer of oysters accounting for 24 percent of total U.S. exports (105, 122). Current regulation strategies for post-harvest oysters In the U.S., the Interstate Shellfish Sanitation Conference (ISSC), a voluntary, cooperative association of states, U.S. Food and Drug Administration (FDA), National Marine Fisheries Service (NMFS), Environmental Protection Agency (EPA), and shellfish industry, adopted guidance to be incorporated in the National Shellfish Sanitation Program (NSSP) that requires states, which produce shellfish reported to harbor V. vulnificus, to develop ?Vibrio vulnificus Risk Management Plans?. The goal of this strategy is to reduce the rate of etiologically confirmed shellfish-borne V. vulnificus septicemia illnesses from the consumption of commercially harvested raw or undercooked oysters by at least 60 percent at the end of year 2008 (44). The core reporting states include Florida, Texas, Louisiana, and California (the biggest oyster importer). Additionally, in order to minimize the risk from this pathogen, ISSC adopted a ? NSSP Guide for the Control of Molluscan Shellfish 2005?. Under the section of ?Control of shellfish harvesting? the time from harvesting to refrigeration is limited. The length of time depends upon the average monthly maximum water temperature (AMMWT) at the time of harvest. These controls set the maximum time interval from 17 harvest to refrigeration at 36 h when AMMWT in the harvest area is less than 18?C, 14 h when AMMWT is 18?C to 23?C, 12 h when AMMWT is 23?C to 28?C, and 10 h when AMMWT exceeds 28?C. Once placed under temperature control, shellstock shall be continuously maintained at 10?C or below until final sale to the consumer. This guideline suggests that consumable oysters should not contain more than 30 MPN of V. vulnificus per gram of oyster meat (45). Currently, the USFDA and ISSC have approved the use of post-harvest treatments which was implemented by 2006 (Interstate Shellfish Sanitation Commission Model Ordinance 03-121) (60). 3. Cold shock response in bacteria The cold shock response is a physiological response of living cells to temperature downshift. In terms of thermal stress, cold shock stress (temperature down shift) is poorly understood compared with heat shock. Most of the cold shock response studies have been performed with Escherichia coli and Bacillus subtilis as model system (51). Microbial response to low temperature involves changes in cellular physiology that have been documented as survival strategy in other species. Several cold shock proteins have been identified not only in E. coli but also in many other bacteria, including mesophilic, psychrophilic, and psychrotrophic bacteria such as Bacillus subtilis (55), Enterococcus faecalis (139), and Lactococcus lactis (90, 140), Lactococcus lactis subs. lactis, Lactobacillus helveticus, Streptococcus thermophilus, Pediococcus pentosaceus (89), Pseudomonas fragi (61, 113, 114), Arthrobacter globiformis (7, 8), Bacillus 18 psychrophilus (184), Listeria monocytogenes (6, 147), Pseudomonas putida (59), Salmonella typhimurium (35), Salmonella enteritidis (75), Aeromonas hydrophila (71), Vibrio vulnificus (110), Vibrio cholerae (36), Sinorhizobium meliloti (123, 124), Rhizobium (28), and Aquaspirillum articum (152). Because temperature has wide-ranging effects on growth and survival, bacteria have developed responses that allow them to adapt to change in temperature. The evolution of cold shock genes is one those adaptive mechanisms. Many species of bacteria have been shown to alter gene expression in response to a shift to low temperatures (6, 7, 55, 80, 109, 140). Since cold has a rate-depressing effect on almost all aspects of physiological performance, an increase in mRNA transcripts may not necessarily correspond to increased levels of the translated protein. Increases in mRNA may simply reflect changes in mRNA turnover parameters with cooling. These changes should not be associated with cold-shock gene induction. In general, bacterial cells must cope with several recognized problems that arise following cold shock: a) membrane fluidity is too low, b) superhelical density of the DNA is too high for opening out the double helix, c) enzyme activities decrease profoundly, but probably to different extents, therefore protein levels must be adjusted, d) protein folding may be too low or inefficient, e) ribosomes must be adapted to function property at low temperatures, and f) secondary structures in RNA might affect initiation of translation (54, 78, 140). A decrease in membrane fluidity can be overcome by increasing an unsaturated fatty acid and di-unsaturated phospholipids in membrane. Cold shock proteins are transiently induced to overcome the translational block (72). Lowering the temperature, especially down to 1-8?C, brings up alterations in translational 19 processes, including the dissociation of polysomes and the accumulation of whole ribosomes and their subunits. It should be noted that cold shock suppresses protein synthesis just at the stage of initiation of RNA translation. Once the synthesis of a polypeptide chain has started, its elongation will continue irrespective of the cold shock conditions (54, 78). Cold shock proteins have been classified based on level of increased expression after a temperature downshift. Class I Csps, which are at low level of expression at 37?C, increased synthesis in more than 10-fold in response to cold shock, whereas class II Csps are present at 37?C and increase modestly upon cold shock. However, the majority of class I and II Csps that have been described are chaperones involved in enabling ribosomes to translate mRNAs at low temperatures (173). Cold shock synthesis of Csps seems to be regulated mainly at the post-transcriptional level. Individual mRNA for each Csp seems to play a main role in cold shock response. It is believed that the cold shock response serves as an adaptive mechanism, and it has been suggested that proteins of the CspA family, together with other Csps, play a role in protecting cells from damage due to the cold. Most of the free living bacteria possess at least one cold shock-inducible CspA homologue, which likely functions as an RNA chaperone (77). In E. coli, when culture temperature of exponentially growing cells is shifted from 37 to 10?C, a lag phase is observed before growth is reinitiated (80). Similar to the heat shock response, to compensate for this stress, E. coli reacts to temperature downshifts by inducing a specific pattern of gene expression called ?cold shock response?, which includes induction of a set of proteins defined as cold shock proteins (78, 80). These cold 20 shock proteins are typically classified into two groups based on their expression patterns (173). Class I cold shock proteins are expressed at an extremely low level at 37?C, but are dramatically induced upon cold shock. Class I includes CspA, CspB, CspG, CspI, CsdA, RbfA, NusA, and PNP. Class II cold shock proteins are expressed at a certain level at 37?C, and are induced moderately upon cold shock. Class II included IF-2, H-NS, RecA, ? subunit of DNA gyrase, Hsc66, HscB, trigger factor, dihydrolipoamide acetyltransferase, and pyruvate dehydrogenase (lipoamide). Three of the cold shock proteins so far identified in E. coli appear to be associated with translation; RbfA, a ribosome binding factor, CsdA, an RNA unwinding protein, and IF2, an initiation factor (173). These proteins are thought to be involved in the preservation of ribosome function at colder temperature. In E. coli, where the induction of cold shock genes has been most extensively studied (173), nine genes encoding CspA-like proteins, cspA to cspI have been identified (190). Among them, cspA (50), cspB (97), cspG (120), and cspI (180) have shown to be cold shock inducible. Among cold shock inducible genes, cspA has been extensively investigated for its cold shock induction (190). Low temperature regulation of the E. coli cspA gene is involved in both transcriptional and post transcriptional mechanisms. In V. vulnificus, when cells responded to a moderate temperature downshift, from 23?C to 13?C, a decline in protein synthesis was observed, but cells continued to grow at a lower rate (110). In contrast, when V. vulnificus was exposed to more dramatic temperature shifts, cell growth was halted and protein synthesis stopped (127). When V. vulnificus culture was shifted from 35?C to 6?C, it underwent a transition to the VBNC state, but when cells were adapted to 15?C prior to change to 6?C they remained viable 21 and culturable. This result suggested that the exposure to an intermediate cold temperature (15?C) caused a cold adaptive response helping V. vulnificus remain culturable at a much colder temperature. The survival and tolerance at cold temperatures could be due to the expression of cold adaptive proteins other than the major cold shock protein as found in E. coli and other microorganisms (17). 4. Genes involved in general stress responses rpoS The rpoS gene encodes a sigma factor, ? S , RpoS, which is required for expression of a large number of genes in response to various stresses, including nutrient limitation, osmotic challenge, and during growth into stationary phase (103). The rpoS gene has been found in a variety of Gram-negative bacteria, and its function and regulation have been studied extensively in the enteric species E. coli and Salmonella enterica serovar typhimurium. RpoS can act as a virulence factor since it activates some of the genes necessary for pathogenesis in several bacteria, including S. enterica, E. coli, and Shigella flexneri (42, 135, 159). In fact, rpoS expression is induced when these bacteria enter mammalian cells (24). The rpoS levels tend to be very low in rapidly growing cells not exposed to any stress. However, the rpoS levels increase in response to a variety of environmental stresses, which included starvation for limiting nutrients, stationary phase, high osmolarity, and high or low temperature. RpoS is a highly unstable protein under non-stress conditions, but exposure to various stresses differentially affects rpoS 22 transcription and translation, as well as the rate of proteolysis. Research on the regulation of rpoS gene expression has mainly focused on the control of RpoS translation (96, 103) and RpoS proteolysis (96, 155) oxyR OxyR is a peroxide sensor and a transcriptional activator of the oxidase stress response. OxyR has been proved to be a positive regulator of the expression of nine hydrogen-peroxide-inducible genes, including those encoding catalase (katG), glutathione reductase (gor), glutaredoxin (grxA), alkyl hydroperoxide reductase (ahpCF), oxyS (a regulatory RNA), fur (ferric uptake regulation), and dps (a nonspecific DNA- binding gene) (27, 115). Catalase is part of a global system regulated by OxyR (27). OxyR-regulated enzymes are involved in the defense against reduced-oxygen compounds. OxyR is required for adaptation to H 2 O 2 (74). An oxyR gene has been identified in 31 bacterial genomes (178). The position of oxyR is variable in the different bacterial genomes, but it is often next to a gene involved in oxidative stress protection and regulated by OxyR. In E. coli it is known that the expression of peroxide defense genes is activated by OxyR in the exponential phase. In the stationary phase, expression of these genes is controlled by the starvation-induced sigma factor RpoS. OxyR mediates a response to oxidative stress in E. coli. The change in oxyR expression during aerobic growth is completely abolished in the absence of oxygen (52). 23 katG There are two types of structural catalases in bacteria: a bifuntionally catalase- peroxidase (HPI) and a monofunctional catalase (HPII). Most bacteria appear to express one or more catalases in response to peroxidase stress. Vibrio vulnificus has been shown to have only one catalase gene, katG, encoding the catalase-hydroperoxidase (94). KatG is induced during the bacterial exponential phase (174). The expression of the katG is controlled by the positive transcriptional activator OxyR, and it is induced in response to exogenous H 2 O 2 (115), which directly sense oxidative stress (88, 193). Bacteria produce H 2 O 2 under aerobic conditions to be toxic, and katG is primarily responsible for its detoxification (137). 5. Molecular diagnosis techniques DNA probe hybridization The ISSC established the methodology for enumerating V. vulnificus in shellfish under Chapter 9 of the FDA Bacteriological Analytical Manual, 9 th edition, 2004 (176). Two analytical schemes for isolating and enumerating V. vulnificus are described. The first method utilizes the most probable number (MPN) analysis coupled with identification of suspect isolates using biochemical profiles, DNA probe colony hybridization, or Polymerase Chain Reaction (PCR). The second uses two direct plating methods followed by hybridization with DNA probes for colony identification. This 24 technique has been used to enumerate V. vulnificus in oysters in several studies (187, 188). This method provides a more rapid and accurate evaluation of post harvest treatment of oysters (60). The sensitivity of DNA probe hybridization assays for detection of vibrios in oyster homogenates approaches 10 CFU/g oyster meat. The identification of bacteria by DNA probe hybridization is based on the presence and absence of particular genes. The physical basis for gene probe tests stem from the structure of DNA molecules themselves (84). DNA hybridization tests may be performed in many ways; one technique is the colony hybridization assay (58, 106). An aliquot of a homogenized food is spread on the appropriate agar plate medium. After incubation, the colony pattern is transferred to a solid support (usually membrane or paper filter). Then, the cells are lysed by a combination of high pH and temperature, which also denature and fix the DNA to the support. The solid supports with the attached target DNA are incubated with an enzyme-labeled specific probe. The labeled probe failing to attach to the target is removed by stringent washing (177). The radioactive probe DNA that is bound to the target on the support is often detected by a colorimetric reaction. A chromogenic substrate then is added to the membrane. Where the probe-associated enzyme is present, a color spot will develop. Each spot represents a bacterial colony that has arisen from a single cell. The number of cells harboring the target gene in the original sample can be calculated by multiplying the number of spots by the dilution factor (177). 25 Reverse transcription polymerase chain reaction (RT-PCR) Reverse transcription polymerase chain reaction (RT-PCR) is an approach for converting and amplifying a single-stranded RNA template to yield abundant double- stranded DNA products. This molecular approach was first described in 1987 (179) and its use in the analysis of steady-state messenger RNA levels has become increasingly widespread in recent years. Detection of low abundance mRNAs by RT-PCR has become a standard technique to determine gene expression in tissues and culture cells. RT-PCR is most useful to determine if a specific mRNA is present in sufficient numbers to be amplified to a detectable level and is a more sensitive alternative to Northern Blot Analysis (172). RT-PCR technique involves the amplification of target RNA molecules by sequentially combining reverse transcription (RT) and polymerase chain reaction (PCR). It is a technique similar to conventional PCR, except that the starting material is RNA rather than DNA. Because of this copy, a DNA must first be made from RNA, utilizing an enzyme known as reverse transcriptase. Once the copy of the DNA is made, the PCR proceeds as usual (136). The greatest advantage of the RT-PCR in the analysis of mRNA includes its sensitivity, versatility, and rapid turn-around time (29). The major shortcoming of RT-PCR is reliability and accuracy because the results are based solely on the quality and quantity of starting RNA sample (189). Since RT-PCR was first described for the detection of low abundance mRNAs, the method has been extended to quantitative measurements of relative and absolute levels of single and multiple mRNAs. Methods for obtaining RNA for RT-PCR have also been extended to include analysis of mRNA 26 levels in clinical samples and even in single cells. Measurement of specific mRNA species is of major importance for studies on gene expression. RT-PCR is one of the most applied techniques in biomedical research for disease pathogenesis. The use of RT-PCR is also increasing in the clinical research and diagnostic fields where it is becoming a popular method for the identification of infectious agents, as well as for detecting markers indicatives of certain types of cancers and genetic disorders (16). Several studies have investigated the potential of mRNA detection using reverse transcription-PCR (RT-PCR) as a viability marker (13, 46, 93, 158). Detection of mRNA is thought to be a good viability marker, due to its central role in cell metabolism and its very short half life. RT-PCR was used to study gene expression both in situ and in vitro for VBNC state of V. vulnificus (160, 161). Statement of research objectives: In summary bacteria adapted to a period of low temperature growth could promote survival, which is referred as cold shock response. Vibrio vulnificus, a food-borne pathogen could potentially develop this cold shock phenomenon during exposure to low temperature. Oysters can serve as reservoirs for V. vulnificus and transmit this bacterium to humans through the ingestion of raw oysters. If the potential for cold shock-mediated survival of V. vulnificus exists, the cooling strategy recommended by USFDA/ISSC may potentially induce this response in V. vulnificus and result in its persistence in oyster. Not only could cold shock response affect microbial survival, but it may also alter gene expression. It was critical to examine the impact of 27 cold shock response of V. vulnificus in oysters regarding its survival and physiology. The expected outcome of this study is to provide scientific data for the development of control strategies of V. vulnificus risk associated with raw oyster consumption. The overall benefit would be to enhance quality and safety of shellfish to ensure consumer confidence in U.S. Gulf Coast oysters. This research focuses on the cold shock response of V. vulnificus both in culture (in vitro) and in live oysters (in vivo). The objectives of this study are: 1. to investigate the effect of cold shock on growth of V. vulnificus under mimicked controlled shellstock temperature conditions as proposed by USFDA/ISSC in vitro. 2. to examine the effect of cold shock on V. vulnificus growth in shellstock oysters using colony blot hybridization as a tool to enumerate V. vulnificus persisting in oysters. 3. to determine which genes of V. vulnificus might be altered during cold shock 4. and to use RT-PCR technology to analyze V. vulnificus gene expression during cold shock both in vitro and in vivo. 28 III. COLD SHOCK EFFECT ON GROWTH RATE AND SURVIVAL OF VIBRIO VULNIFICUS IN CULTURE BROTH UNDER SHELLSTOCK TEMPERATURE CONTROL CONDITIONS 29 ABSTRACT The growth rate of V. vulnificus was investigated at different temperatures. The drop plate method was used to determine the culturability of the cells. All the four strains tested displayed a similar growth rate when incubated at the same temperature (P>0.05), except at 20?C (P<0.05). At 35 and 25?C V. vulnificus cultures showed a mean growth rate of 1.23 and 1.169 hour -1 , respectively. At 20?C cells multiplied with a mean growth rate that ranged from 0.752 hour -1 in clinical strains to 0.982 hour -1 in environmental strains. In contrast, at 15?C V. vulnificus showed little growth with a mean growth rate of 0.004 hour -1 . At 4?C and 7.2?C cell numbers decreased over time until non-detectable levels were recorded at 72 h. The response of V. vulnificus to cold shock was investigated under ISSC guidelines. Vibrio vulnificus was grown at 35, 25, 20, and 15?C, and then subjected to subsequent low temperatures at 7.2 or 4?C for one week. Vibrio vulnificus cells rapidly multiplied at elevated temperatures (35, 25, and 20?C). After cold shock at 7.2?C cell numbers decreased by 1 to 2 log units. However, the highest reduction in culturability was observed after cold shock at 4?C. When cells were adapted to 15?C prior to cold shock at 7.2?C, a slight increase in cell counts were observed. Results suggested that a cold adaptive mediated survival mechanism in V. vulnificus existed when cells were adapted to 15?C prior to shock at 7.2?C. The effective refrigeration temperature in reducing the multiplication of V. vulnificus is at 4?C. However, optimal and sub-optimal growth temperatures (35, 25, 20?C) allowed V. vulnificus cell counts to reach high number that could not be reduced to safety levels by refrigeration conditions even at 4?C. 30 INTRODUCTION Vibrio vulnificus is a Gram-negative bacterium naturally inhabitating marine and estuarine environments. V. vulnificus is a mesophilic bacterium with maximum growth occurring at 35?C, pH of 7.5 to 8.0, and with 2% NaCl present in the medium (68). In the U.S., V. vulnificus is a major concern for the public health authorities since 95% of all deaths associated to seafood consumption are caused by this bacterium (132). V. vulnificus infections can cause a primary septicemia with fatalities up to 60% (132), especially in patients with underlying diseases. The majority of those cases are associated to the consumption of raw oysters. High V. vulnificus counts in oysters are correlated with elevated water temperatures in summer months which explains the increase incidence of the disease during warmer months (157). These epidemiological findings suggest that infections caused by V. vulnificus are dose dependent. Although the infectious dose for V. vulnificus has not been established for humans, the invasiveness and high fatality rates that characterize V. vulnificus infections makes the reduction of V. vulnificus counts in oysters to non-detectable levels a primary goal for the Interstate Shellfish Sanitation Conference (ISSC), as well as for the Gulf Coast oyster industry. Under optimal conditions, V. vulnificus exhibits one of the fastest growth rates of all bacteria but cell culturability rapidly declines when this bacterium is subjected to refrigeration temperatures (30, 33, 110, 128, 170). In order to minimize consumer risks, the ISSC has established specific guidelines for the control of molluscan shellfish at harvest, transportation, shucking, packing and shipping (44). Shellstock temperature control is a critical point within that set of guidelines. Shellstock temperature control is 31 defined as the management of the environmental temperature of shellstock by means of ice, mechanical refrigeration or other approved methods which are capable of lowering the temperature of the shellstock and which will maintain it at 10?C or less. A specific time-temperature matrix has been defined for molluscan shellfish collected from waters where problems associated to V. vulnificus have been reported. The warmer the water temperatures are at time of harvest, the shorter the time required from harvest to temperature control conditions (45). Not only shellstock temperature control but also various actions have been adopted in order to minimize the risk of oyster containing V. vulnificus. ISSC has adopted industry-implemented post-harvest controls to reduce V. vulnificus loads in oyster shellstock, which may include time-temperature, post harvest treatment (i.e. hydrostatic pressure, cool pasteurization, IQF, and irradiation - pending approval), and rapid chilling. The post-harvest treatment technologies currently account for less than 10% of all domestic raw oyster sales in the U.S. (98). Moreover, these post-harvest treatments are effective for eliminating V. vulnificus, but each has some drawbacks. Cool pasteurization (5) may kill bacteria but can change the oyster flavor and texture. Hydrostatic pressure processing (9) can eliminate this bacterium but it cause the meat to detach from the shell. All of these processes kill the harvested oyster and increase product costs. The use of accurate shellstock time-temperature controls offers several advantages over post-harvest treatments since the procedure does not require specialized technology. The temperature control strategies for minimizing V. vulnificus risks by the ISSC were proposed following reports by Cook and Ruple (33). These authors stated that storage of summer harvested oyster shellstock from the U.S. Gulf Coast at 10?C prevented the multiplication of 32 V. vulnificus, however, recent studies have been shown that V. vulnificus can adapt to cold temperatures below 7.2?C, following exposure to intermediate temperatures (?15?C) (17, 110). Previous literature has reported that 4?C seems to be more effective than 7?C in reducing V. vulnificus numbers to non-detectable levels (162). Understanding the relationship between V. vulnificus numbers and temperature in culture broth may help predict the growth pattern of V. vulnificus in oysters. To better characterize the effects of refrigeration conditions in V. vulnificus growth, the objectives of this study were (a) to establish the growth rate of different V. vulnificus strains under different temperatures, and (b) to determine V. vulnificus growth under ISSC recommended shellstock temperature control conditions. MATERIALS AND METHODS Bacterial strains, media, and growth conditions Two clinical strains of V. vulnificus, ATCC 27562 (type strain), and C7184Tr (avirulent, translucent morphotype provided by J.D. Oliver, University of North Carolina, Charlotte, NC) were used in this study. Two environmental strains (VV3, VV4) isolated from eastern oyster (Crassostrea virginica) samples collected from Mobile Bay, Alabama were also included. Identification and classification of both environmental isolates were performed by API 20E (bioMerieux, Inc., Hazelwood, MO) and confirmed by PCR analysis with V. vulnificus specific primers (43). Cells were maintained and routinely grown on marine agar (MA) (Difco, Detroit, MI.) at 35?C. 33 Growth rate of V. vulnificus Growth curves for V. vulnificus were determined by inoculating one single colony into 25 ml of marine broth (MB) (Difco) and grown in a shaker incubator at 35?C. At various time intervals, 5 ml aliquots of culture were taken and the optical density (OD) at 450 nm was determined. Simultaneously and in order to correlate OD with colony forming units (CFU), the drop plate method described by Herigstad et al. (62) was used. One hundred ?l of culture broth was serially diluted and plated onto MA and incubated overnight at 35?C and colonies counted after 24 h incubation (Appendix. 1). Growth curves at different temperatures (35, 25, 20, 15, 7.2, and 4?C), were determined following the above mentioned protocol with the following modifications. Four V. vulnificus strains were compared at the same time for each experiment. One hundred and fifty ml of MB was inoculated with cells at OD 0.025 which equals a cell count of approximately 10 4 CFU/ml. Flasks were then incubated at 35, 25, 20, 15, 7.2, and 4?C. Samples were taken at 0, 1, 2, 3, 4, 5, 6, 8, 10, and 12 h of incubation, viable plate counts were done in triplicate on MA plates by the drop plate method. For experiments carried out at 7.2?C and 4?C, cells were cultured and sampled for culturability until no CFU will be detected. For each temperature experiment, the cell growth rate was calculated by simple linear regression of ln (CFU/ml) over on time (hours) for five or six points in the linear portion of the growth curve. 34 Growth patterns under cold shock conditions The effect of temperature downshift (cold shock) on V. vulnificus growth was studied by imitating the time-of-harvest to time-of-refrigeration conditions that are currently recommended by the ISSC (44). The culturability of the cells after cold shock was monitored and compared at two different refrigerated temperatures (7.2?C and 4?C). Four different growth temperatures and times prior to cold shock were compared: 35?C for 10 h, 25?C for 12 h, 20?C for 14 h, and 15?C for 36 h. After those incubation periods, cells were transferred to either 7.2?C or 4?C and incubated for 168 h. At various time intervals from initial growth to 168 h, the culturability of the cells was monitored by the drop plate method. Statistical analysis All bacterial samples were analyzed in triplicate. The standard error was calculated for all of the treatment replications. To determine if difference between data set existed, the Randomized Complete Block Design (RCBD), F-test was performed using one-way Analysis of Variance (ANOVA) analysis procedure in the Statistical Analysis System, SAS 9.1.3 (SAS Institute, Cary, NC). 35 RESULTS Growth rate of V. vulnificus All four strains exhibited a very similar growth rate (Fig. 1-6) when incubated at the same temperature (P>0.05), with the exception of temperature at 20?C (P<0.05). When cells were cultured at 35?C, V. vulnificus viable counts increased by 5 log over a period of 7 to 10 h, reaching 10 9 CFU/ml during the stationary phase. The growth rate at 35?C was between 1.218 to 1.250 hour -1 (Fig.1). A similar result was obtained at 25?C where growth rates ranged from 1.149 to 1.174 hour -1 (Fig.2). At 20?C only the two environmental strains were able to reach a viable count of 10 9 CFU/ml after a 12 h period. Both clinical strains showed a decline in cell growth by one to two log reductions with the growth rate ranged from 0.655 to 0.982 hour -1 (Fig.3). Small positive growth was observed at 15?C with the growth rate ranged from 0.003 to 0.005 hour -1 (Fig.4). When cells were incubated under refrigeration temperatures at 4?C, total viable counts started to decrease 12 h post-inoculation. Cells incubated at 7.2?C started to lose culturability 24 h post-inoculation. After 72 h of incubation, no viable cells were observed at either 7.2 or 4 ?C. No difference between strains was observed in this experiment (Fig. 5-6). 36 Growth patterns under cold shock conditions Figures 7 to 14 depict cold shock survival response of V. vulnificus, when cells were subjected to different temperatures and times growing conditions. Under ISSC guidelines conditions, V. vulnificus reached 10 9 CFU/ml when cells were incubated at 35 and 25?C for 10 and 12 h, respectively (Fig. 7-8). At 20?C, only environmental strains were able to reach those numbers after 14 h of incubation, while clinical strains plateaued at 10 8 CFU/ml (Fig. 9). At 15?C no growth was recorded prior to cold shock (Fig. 10). After cold shock at 7.2?C, a steady decrease in V. vulnificus cells numbers was observed at 35, 25, and 20?C. After 168 h of incubation at 7.2?C, an overall cell count reduction by one log was observed in all cases with the exception of VV3 and VV4 grown at 25?C (Fig. 7-9). After cold shock, these two strains experienced a reduction of two log units in cell count numbers. Nevertheless, when cells were allowed to grow up to 10 8 CFU/ml or higher levels, they maintained a cell culturability of 10 7 CFU/ml or higher during 7 days under refrigeration conditions at 7.2?C. Conversely, when cells were incubated at 15?C for 36 h and then shifted to 7.2?C, total cell counts slightly increased during the length of the study (Fig. 10). This increment in cell counts represented only a 0.2 log in magnitude but showed an adaptation response to 15?C. A similar pattern was found when V. vulnificus cells were incubated at 35, 25, 20?C, and then shifted to 4?C (Figure 11-13). Cell counts started to decrease a few hours after cold shock. Cells grown at 35?C required up to 50 h of incubation at 4?C before starting to lose culturability. Cells grown at 25?C and 20?C showed a reduction in total cell counts in less than 12 h at 4?C. A similar trend was observed when cells were 37 subjected at 7.2?C (Fig. 7-9). When cells were incubated at 15?C for 36 h and then cold shocked at 4?C a decrease in total cell counts was observed starting 72 h after cold shock (Fig. 14). This result was different when cells were cold shocked at 7.2?C. After 168 h of incubation at 7.2?C, cell counts averaged 10 4 CFU/ml while after 168 h of incubation at 4?C remaining viable cells were down to 10 2 CFU/ml. Statistical analyses showed that the mean of V. vulnificus counts after 7 day refrigeration at 7.2?C were significantly greater (P<0.05) than at 4?C in all treatments. DISCUSSION The response of living cells to temperature downshift has been termed cold shock (72). Several factors have been reported to affect bacterial growth during cold shock including temperature (67), incubation time at low temperature (164), chill rate (111), phase of inoculums (111, 143), cell concentration (49, 111, 163), plating medium (49), and the opportunity for the cells to repair the damage caused by the cold shock (37). Several studies have reported cold shock resulted in the decline in cell culturability of V. vulnificus (30, 33, 81, 83, 118, 129). There has also been evidence that cold shock adaptation promoted survival of V. vulnificus (17). Results of this study indicate there was no difference in the growth pattern between the clinical and environmental strains of V. vulnificus tested. The optimal temperature condition assayed was 35?C while incubation at 15?C or lower prevented V. vulnificus growth. The current data are in agreement with the maximum (37?C) and minimum (13?C) temperatures reported for V. vulnificus growth (81). This study 38 demonstrated that clinical strains ATCC 27562 and C7814Tr had a worse performance rate at 20?C than environmental strains, and suggests that environmental strains might be better adapted to suboptimal temperatures than clinical strains. When V. vulnificus growing patterns were investigated under the time-of-harvest to time-of-refrigeration conditions recommended by the ISSC, the results revealed that all strains tested had similar growing patterns. V. vulnificus cells rapidly multiplied in culture broth at elevated temperature (35, 25, 20?C) while no positive growth was detected at 15?C and loss of culturability occurred under refrigeration conditions. Growth of V. vulnificus in vitro can be prevented when exposed to 7.2?C and 4?C, however, V. vulnificus cells, although unable to multiply at low temperatures, remain culturable for up to 7 days under refrigeration conditions. If V. vulnificus cell numbers are allowed to reach high numbers under favorable conditions, neither 7.2?C nor 4?C will reduce cell culturability to non-detectable levels (at least in a 7 day period). Data presented here show that cold shock adaptation enhances survival of V. vulnificus, when cells are cultured in vitro. A cold adaptation response was observed after subjecting V. vulnificus cells to 15?C for a few hours before cold shock. In fact, cells adapted at 15?C were able to slightly increase their numbers at 7.2?C. This finding is in agreement with Bryan et al (17) who verified that a cold adaptation period at 15?C enhanced V. vulnificus survival at 6?C. This cold adaptive response in V. vulnificus induced at 15?C could prevent the cells from entering the VBNC state when exposed to a much colder temperature. McGovern and Oliver (110) reported the induction of cold response proteins in V. vulnificus after temperature downshift from 23?C to 12?C. This result suggested that cellular protein expression changed during or after temperature 39 shifting from high temperature to low temperature. The adaptation of V. vulnificus to cold temperature after priming at 15?C may be due to the induction of cold-adaptive protective proteins. Cold temperature adaptation effect was lost when cells were transferred to colder refrigeration conditions (4?C instead of 7.2?C). Not only were 15?C-acclimated V. vulnificus cells unable to grow at 4?C, but their numbers drastically declined over time. This striking difference between close temperature conditions has been pointed out in other studies. Stanford reported that 4?C and 7?C differently affected the rate of entry into dormancy (162). These stronger negative effects caused by 4?C versus 7?C might warrant the use of stringent refrigeration conditions by the shellfish industry. The temperature and time of cold shock treatment clearly affects the cold shock response of V. vulnificus, which could enhance survival during cold stress conditions. The current findings show that the shellstock temperature control guidelines suggested by ISSC might not be effective in reducing V. vulnificus numbers. When V. vulnificus cells are exposed to optimal or suboptimal temperatures, cell numbers rapidly reached 10 8 to 10 9 CFU/ml, and which can be reduced by refrigeration only in a few orders of magnitude. Minimizing V. vulnificus growth by limiting the time from harvest to refrigeration should be strictly enforced. Further studies analyzing V. vulnificus response to cold shock in vivo should be carried out in order to provide the industry with safer practices. In conclusion, this study showed that exposure of V. vulnificus to optimal or suboptimal temperatures for a brief period increases the number of cells to such high 40 levels that they can not be safely reduced by refrigeration afterwards. These data also suggest that 4?C is the appropriate refrigeration temperature. 41 At 35?C 4 5 6 7 8 9 10 0 2 4 6 8 10 12 14 Hour l o g CF U/ m l ATCC 27562 C7184Tr VV3 VV4 FIG. 1. Growth of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 at 35?C At 25?C 4 5 6 7 8 9 10 02468101214 Hour l og C F U / m l ATCC 27562 C7184Tr VV3 VV4 FIG. 2. Growth of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 at 25?C 42 At 20?C 4 5 6 7 8 9 10 02468101214 Hour l o g CF U/ m l ATCC 27562 C7184Tr VV3 VV4 FIG. 3. Growth of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 at 20?C At 15?C 4 4.1 4.2 4.3 4.4 4.5 4.6 4.7 4.8 4.9 5 02468101214 Hour l og C F U / m l ATCC 27562 C7814Tr VV3 VV4 FIG. 4. Growth of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 at 15?C 43 At 7.2?C 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 0 1020304050607080 Hour l o g CF U/ m l ATCC 27562 C7184Tr VV3 VV4 FIG. 5. Growth of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 at 7.2?C At 4?C 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 0 1020304050607080 Hour l o g CF U/ m l ATCC 27562 C7184Tr VV3 VV4 FIG. 6. Growth of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 at 4?C 44 At 35?C for 10h and cold shocked at 7.2?C for 168h 4 5 6 7 8 9 10 0 20 40 60 80 100 120 140 160 180 200 Hour l o g CF U/ m l ATCC 27562 C7184Tr VV3 VV4 FIG. 7. Cold shock-survival response of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 in MB when cells were incubated at 35?C for 10 h and cold shocked at 7.2?C for 168 h, vertical arrow indicated cold shock at 7.2?C. At 25?C for 12 h and cold shocked at 7.2?C for 168h 4 5 6 7 8 9 10 0 20 40 60 80 100 120 140 160 180 200 Hour lo g C F U / m l ATCC 27562 C7184Tr VV3 VV4 FIG. 8. Cold shock-survival response of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 in MB when cells were incubated at 25?C for 12 h and cold shocked at 7.2?C for 168 h, vertical arrow indicated cold shock at 7.2?C. 45 At 20?C for 14h and cold shocked at 7.2?C for 168h 4 5 6 7 8 9 10 0 20 40 60 80 100 120 140 160 180 200 Hour l o g CF U/ m l ATCC 27562 C7184Tr VV3 VV4 FIG. 9. Cold shock-survival response of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 in MB when cells were incubated at 20?C for 14 h and cold shocked at 7.2?C for 168 h, vertical arrow indicated cold shock at 7.2?C. At 15?C for 36h and cold shocked at 7.2?C for 168h 4 4.1 4.2 4.3 4.4 4.5 4.6 4.7 4.8 4.9 5 0 20 40 60 80 100 120 140 160 180 200 220 Hour l o g CF U/ m l ATCC 27562 C7184Tr VV3 VV4 FIG. 10. Cold shock-survival response of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 in MB when cells were incubated at 15?C for 36 h and cold shocked at 7.2?C for 168 h, vertical arrow indicated cold shock at 7.2?C. 46 At 35?C for 10h and cold shocked at 4?C for 168h 4 5 6 7 8 9 10 0 20 40 60 80 100 120 140 160 180 200 Hour l o g CF U/ m l ATCC 27562 C7184Tr VV3 VV4 FIG. 11. Cold shock-survival response of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 in MB when cells were incubated at 35?C for 10 h and cold shocked at 4?C for 168 h, vertical arrow indicated cold shock at 4?C. At 25?C for 12h and cold shocked at 4?C for 168h 4 5 6 7 8 9 10 0 20 40 60 80 100 120 140 160 180 200 Hour l o g CF U/ m l ATCC 27562 C7184Tr VV3 VV4 FIG. 12 Cold shock-survival response of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 in MB when cells were incubated at 25?C for 12 h and cold shocked at 4?C for 168 h, vertical arrow indicated cold shock at 4?C. 47 At 20?C for 14h and cold shocked at 4?C for 168h 4 5 6 7 8 9 10 0 20 40 60 80 100 120 140 160 180 200 Hour l og C F U / m l ATCC 27562 C7184Tr VV3 VV4 FIG. 13. Cold shock-survival response of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 in MB when cells were incubated at 20?C for 14 h and cold shocked at 4?C for 168 h, vertical arrow indicated cold shock at 4?C. At 15?C for 36h and cold shocked at 4?C for 168h 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 20 40 60 80 100 120 140 160 180 200 220 Hour l og C F U / m l ATCC 27562 C7184Tr VV3 VV4 FIG. 14. Cold shock-survival response of V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4 in MB when cells were incubated at 15?C for 36 h and cold shocked at 4?C for 168 h, vertical arrow indicated cold shock at 4?C. 48 IV. EFFECT OF COLDSHOCKING ON SURVIVAL OF VIBRIO VULNIFICUS IN SHELLSTOCK OYSTERS 49 ABSTRACT The response of V. vulnificus cells in oysters to cold shock was evaluated. Depurated oysters were inoculated with three strains of V. vulnificus at 10 6 CFU/g oyster meat. Both clinical and environmental strains of V. vulnificus presented similar growth numbers in oysters under the same treatment (p>0.05). Three cold shock conditions were examined. When oysters were transferred to 35 and 25?C, V. vulnificus numbers multiplied by a half log after 10 h and 12 h incubation, respectively. When oysters were shifted to 4?C for one week, V. vulnificus exhibited a moderate decline in cells viability. No growth of V. vulnificus was observed after oysters were adapted to 15?C for 36 h, but cell viability decreased slightly after they were cold shocked at 4?C for one week. This decline in cell culturability varied from experiment to experiment (p<0.05) but was generally on the order of a half to two log reduction after cold shock for one week. This study indicates that cells adapted to 15?C prior to change to 4?C enhanced cell culturability of V. vulnificus while it is embedded in oysters. Conditions that favor adaptation to cold temperatures should be avoided in V. vulnificus control strategies. 50 INTRODUCTION Raw or under cooked oysters pose a human health risk, as these oysters tend to concentrate microorganisms from the surrounding waters via their filter feeding. In the U.S., Vibrio vulnificus is an autochthonous microorganisms in the Gulf of Mexico waters where its growth is favored by relatively high water temperatures (85). Gulf Coast oysters (Crassostrea virginica) have been reported to be the primary source for V. vulnificus related illnesses and deaths. From 1988-1996, Center for Disease Control and Prevention (CDC) received reports of 302 cases of V. vulnificus infections from the Gulf Coast States, of these, 141 cases were associated with seafood ingestion, and 96% of these cases were due to raw oyster consumption (19). V. vulnificus is the most serious food-borne pathogen with more deaths reported annually from this pathogen than any other seafood-related disease (11). This pathogen is known to cause gastroenteritis, severe wound infection and primary septicemia (99). Nearly all fatal primary septicemia cases were associated with the consumption of raw oysters. The case fatality rate for V. vulnificus septicemia exceeded 60% (19). High case-fatality rate by V. vulnificus was reported in persons with preexisting liver disease or immunocompromised conditions (165). This bacterium is one of the most costly in terms of economic loss. FDA estimated that V. vulnificus-related illnesses cost over $120 million annually, or about 60% of the total cost of shellfish-related illnesses each year (44). Legal harvesting of oysters is limited to areas free of feacal contamination, however oysters harvested from approved sites may contain V. vulnificus. During summer, V. vulnificus levels in the Gulf Coast oysters can ranged from 10 3 to 10 5 CFU/g 51 oyster and appears to replicate within the oyster tissue (170). This bacterium may be presented in oysters in relative low numbers at harvest but may multiply to more hazardous levels if they are exposed to time or temperature abuse. Storage times and refrigerated temperature are known to effect the survival of V. vulnificus. Cook et al. (30) reported that V. vulnificus could multiply in the oyster shellstock if they are held at temperatures above 13?C. It has been proven that the shorter the time oyster shellstock remained outside refrigeration after harvest, the lower the number of V. vulnificus were detected in them (31). Conversely, refrigeration temperature is known to be detrimental to the survival of V. vulnificus. The majority of studies that focused on the effect of low temperatures to V. vulnificus survival pointed up the need to refrigerate oyster shellstock quickly after harvesting (30, 31, 81, 129). In order to limit the multiplication of V. vulnificus in post-harvest oysters, restrictive shellstock temperature control measures that requires refrigeration of Gulf coast oysters after harvesting have been recommended by the shellfish authorities, the Interstate Shellfish Sanitation Conference (ISSC) and implemented by the industry (44). However, fatalities associated to V. vulnificus infectious associated with the consumption of raw oysters have not decreased in recent years. The cold shock phenomenon, a physiological response for survival strategy of living cells to temperature downshift has been reported in several bacteria (51). Bryan et al. (17) found that the effectiveness of rapid chilling or freezing of oysters to eliminate V. vulnificus levels in shellfish may be compromised by product handling procedures that permitted cold adaptation. When V. vulnificus culture was abruptly shifted from 35?C to 6?C conditions, a non-culturable state was induced, whereas cells 52 adapted to 15?C prior to change to 6?C condition, remained viable and culturable. Recently, Quevedo et al. (149) reported that chilling of oysters on ice at 7.2?C resulted in a small decline of endogenous V. vulnificus and concluded that this technique was not sufficient to eliminate this pathogen in postharvest oysters. Cooling regimes in control measures may potentially induce cold shock adaptation that could enhance V. vulnificus survival. It is critical to determine whether the potential for cold shock-mediated survival of V. vulnificus exists while it presents in oyster shellstock. The effect of cold adaptation of this pathogenic microorganism in oysters under industry practices from harvest to consumption, particularly in regard to their survival need to be further investigated. The objective of this study was to examine the effect of cold shock to V. vulnificus growth in oyster shellstock. MATERIALS AND METHODS Bacterial strains and growth conditions Two clinical strains of V. vulnificus, ATCC 27562 (type strain), and C7184Tr (avirulent, translucent morphotype) provided by J.D. Oliver, University of North Carolina, Charlotte, NC and one environmental strain (VV3) isolated from oyster samples collected from Mobile Bay, Alabama were used in these studies. Cells were grown in marine broth with shaking and maintained in marine agar (MA) (Difco, Detroit, MI) at 35?C. 53 Depuration of oyster samples Twelve hundred live shellstock eastern oysters, one year old, 50 to 80 mm in diameter, were provided by the Auburn University Shellfish Laboratory (AUSL), Department of Fisheries and Allied Aquacultures, Dauphin Island, AL. Oysters were acclimated in 19 mm mesh size Durethane oyster bags, (ADPI Enterprise Inc., Philadelphia, PA) for 2 weeks in filtered water flow-through system before depuration. Oysters were scrubbed with a sterile scrub-brush under running tap water to remove extraneous material. Cleaned oysters were treated under UV filtered seawater system (UV flow 3000 ?w-sec/cm 2 ), using a water flow-through system (180 l/h) for 7 days at the AUSL. During depuration, oysters were fed twice a day with marine microalgae concentrate (Shellfish diet 1800, Reedmariculture, Campbell, CA). Salinity, temperature, and dissolved oxygen were determined everyday. The mortality of oysters was checked every morning before feeding. Bacteriological analysis Ten oysters were randomly selected before and during depuration once a day for bacteriological analysis as follows: total heterotrophic aerobic plate count, total Vibrio counts, and V. vulnificus counts. Oysters were washed and scrubbed in tap water with sterile brushes, shucked and weighed in sterile equipment to avoid cross-contamination. Oyster meat sample and the shell liquid from ten oysters were pooled together and homogenized using a MCI mini Waring? blender for 90s. The homogenate was then 54 diluted 1:1 (wt/wt) in phosphate-buffered saline (PBS) and homogenized for another 30s. Ten-fold dilutions from the homogenate were made with PBS. The sample (100?l) was spread with a sterile glass rod on MA in triplicate for total aerobic plate count. Another 100 ?l was spread on Thiosulfate Citrate Bile-Salt and Sucrose (TCBS) agar (Difco) in triplicate for total Vibrio plate count. One hundred ?l of homogenated oyster was spread on T1N1 agar (1%NaCl, 1% tryptone, 1.5% agar) in triplicate for V. vulnificus enumeration. All plates were dried at room temperature until homogenate was fully diffused in the agar. The plates were inverted and incubated 24 h at 35?C. Vibrio vulnificus counts in oysters were enumerated using a species-specific oligonucleotide probe (vvhA) and colony blot hybridization procedure previously described by Quevedo et al (149). After 24 h incubation, colony lifts from T1N1 spread plates were overlaid with Whatman #541 filters papers. Filters with transferred colonies were heated to dryness in a microwave oven for 2 min on the highest setting in lysis solution (0.5 M NaOH, 1.5M NaCl), followed by neutralization in ammonium acetate buffer and washed in standard saline citrate buffer. Filters were treated with proteinase K (20?g/ml) and hybridized with the alkaline phosphatase-labeled vvhA probes (DNA Technologies, Aarhus, Denmark) under stringent conditions (56?C) for 1 h. Filters were washed in standard saline citrate buffer with 1% sodium dodoecyl sulfate at hybridization temperature, followed by additional washing with standard saline citrate buffer at room temperature to overcome nonspecific hybridizations signal. NBT/BCIP (nitrobluetetrazolium/5-bromo-4-chloro-3-indolylphosphate toluidinium) substrate (Roche Diagnosis, Indianapolis, Ind.) was used for detection of hybridization positive colony. Control filters were developed concurrently for all the time points. 55 Artificial inoculation After one week, live depurated oysters were cooled on ice and immediately transported to the Bacterial Genomics Laboratory, Department of Fisheries and Allied Aquacultures, Auburn University, AL for cold shock experiments. Live depurated oysters were inoculated with V. vulnificus by self-inoculation filtration. Four 10 gallon (38 L) aquaria containing 27 L of autoclaved seawater was prepared. One hundred fifty oysters were placed into each aquarium at 25?C and allowed to acclimate up to 24 h. A fresh culture of each strain of V. vulnificus was grown in MB until exponential growth phase (10 8 CFU/ml, OD 450 =0.15) was reached. Three liters of V. vulnificus culture were introduced into each aquarium at approximately 10 6 CFU per gram of oyster. Control aquaria contained oysters without introduced V. vulnificus and were acclimated as described for test aquaria. Cold shock experiments After V. vulnificus inoculations, oysters remained at 25?C overnight prior to temperature shift. To examine cold shock effect on V. vulnificus survival in oysters, three different cold shock conditions were compared. The first experiment, oysters were transferred to 35?C for 10 h, then chilled by ice-water immersion until the internal temperature in oysters reached 4?C. Oysters were then transferred to a 4?C incubator for 168 h. In the second experiment, oysters were transferred to 25?C for 12 h then chilled to 4?C and incubated for 168 h. For the last experiment, oysters were transferred to 15?C 56 for 36 h, and chilled to 4?C as described above and incubated at 4?C for 168 h. The internal temperature of the oysters was monitored using a digital thermometer inserted through a hole drilled in the shell to the meat of oysters throughout the study (5). The external ambient air temperature was also recorded and the pH of oyster meat measured periodically. V. vulnificus enumeration in oysters At various time intervals from t = 0 to t = 168 h, V. vulnificus counts in oysters were enumerated using a species-specific oligonucleotide probe (vvhA) and colony blot hybridization procedure previously described by Quevedo et al (149). Statistical analysis All bacterial samples were analyzed in triplicate. The standard error was calculated for all of the treatment replications. To determine if difference between data set existed, the Randomized Complete Block Design (RCBD) using F-test was performed using one-way ANOVA analysis procedure in the Statistical Analysis System, SAS 9.1.3 for Windows (SAS Institute, Cary, NC). 57 RESULTS Total bacteria, total Vibrio, and V. vulnificus during depuration Before oysters were depurated by UV filtered seawater system, the total bacteria count was 10 4 CFU/g oyster meat. After depuration for 7 days, CFU/g oyster declined about two log units. Similar decline in CFU/g oyster of total Vibrio count from TCBS plate was observed. Numbers of V. vulnificus in oysters was around 10 2 CFU/g oyster before depuration. After five days of depuration, no CFU/g oyster of V. vulnificus was detected (Fig. 15). pH of oyster meat The pH of oysters after artificially inoculation with V. vulnificus at 25?C was 6.6. After oysters were transferred to 35?C for 10 h the pH slightly increased, but then decreased to 6.2 after oysters were kept at 4?C for one week. When oysters were kept at 25?C for 12 h, pH remained constant. After oysters were shifted at 4?C for one week, pH of oyster meat decreased from 6.6 to 6.2. When oysters were transferred to 15?C for 36 h, the pH was constant. After oysters were shifted to 4?C, the pH decreased from 6.6 to 6.3 (Fig. 16). The pH range across the study (pH 6.2 to 6.6) was lower than the optimal pH described for V. vulnificus (pH 7.5 to 8.0). 58 V. vulnificus growth in shellstock oysters during cold shock Colony hybridization was used to enumerate V. vulnificus in oysters during cold shock. No V. vulnificus was detected in control oysters at anytime during any treatment. The average initial V. vulnificus numbers in artificially inoculated oysters for each experiment was 1?10 6 CFU/g oyster. When V. vulnificus-contaminated oysters were incubated at 35?C for 10 h, cells increased by one half log unit. After oysters were placed at 4?C cell numbers declined moderately by two log units, with the final counts of 3?10 4 CFU/g oyster meat (Fig 17). Similarly, when V. vulnificus was incubated at 25?C for 12 h, cell numbers slightly increased but after cold shock at 4?C numbers decreased by 1.5 log units, with the final counts of 6.5? 10 4 CFU/g oyster meat (Fig 18). When V. vulnificus contaminated oysters were acclimated at 15?C for 36 h V. vulnificus numbers remained constant, and slightly decreased only by a half log unit, with final numbers around 5.3?10 5 CFU/g oyster meat after downshift to 4?C (Fig 19). Overall, V. vulnificus numbers were not significantly different among strains (ATCC 27562, C7184Tr, VV3) after cold shock for one week under the same temperature treatment (p>0.05). The decline in cell culturability varied from experiment to experiment (p<0.05) but was generally on the order of a half to two log reduction after cold shock for one week. 59 DISCUSSION The present study successfully demonstrated a consistent method to artificially inoculate V. vulnificus into live oysters. Numbers of V. vulnificus found in oyster shellstock were very low (10 2 CFU/g oyster) before depuration treatment. After 5 days of depuration, the cells were not detected by colony hybridization. Previous literature reported that the sensitivity of DNA probe colony hybridization assays for detection of V. vulnificus approached 10 CFU/g oyster. The FDA currently recommends that postharvest-treatment should reduce V. vulnificus to non-detectable level (<30 CFU/g) (45). This result revealed that the depuration treatment used in this study was successful in eliminating V. vulnificus to non-detectable levels. When V. vulnificus-contaminated oysters were kept in a warm environment (at 35 and 25?C), even for 10 to 14 h, the numbers of V. vulnificus in oysters increased by a half log. This finding supports previous reports of how elevated temperature abuse in oysters contributes to an increase in endogenous V. vulnificus levels in post-harvest shellstock oysters (30, 31, 81). Recently, Gooch (53) reported that V. vulnificus multiplied in live oysters held at 26?C with an increase of 1.2 log CFU/g 10 h post harvest. There was evidence that V. vulnificus growing in vitro had a faster growth rate than in vivo (14). The difference in V. vulnificus numbers in oysters may be attributed to oyster defense mechanisms, specifically in hemocytes which posses strong phagocytic properties and play a role in the survival of V. vulnificus in oyster tissues (47). Vibrio specific Bacteriophages are another factor that may effect populations of V. vulnificus within oysters (39). 60 The decrease in the persistence of viable V. vulnificus cells by one to two orders of magnitude was observed when oysters were transferred to cold temperature (4?C). This result is in agreement with Kasper and Tamplin (81), who reported one log reduction of natural V. vulnificus numbers during cold storage at 4?C. The same authors also concluded that this temperature is critical to control V. vulnificus growth in oyster shellstock. In this study, however, the reduction observed at this level would still leave large concentrations of V. vulnificus remaining in oysters. These data show that the endogenous population of V. vulnificus might behave similarly to the laboratory grown culture when it responded to the cold during embedded in oysters. Vibrio vulnificus remained culturable but no growth was observed when oysters were transferred to 15?C. After a temperature down shift from 15?C to 4?C for one week, V. vulnificus numbers in oysters decreased by less than a half log unit. This result showed that V. vulnificus final counts at this temperature condition (15?C to 4?C) was one log higher in cell survival when compared with 35 and 25?C to 4?C. Data presented here documented cell adaptation to cold temperatures since 15?C enhanced cell culturability. This finding is in agreement with Datta and Bhadra (36) who reported that priming of cells at 15?C allow bacteria to adapt and grow at 10?C. Exposure to an intermediate temperature (15?C) may cause a cold adaptive response helping V. vulnificus remain in the culturable state when exposed to a much colder temperature (17). It is well documented that low temperatures can cause V. vulnificus cells to enter the viable but non-culturable (VBNC) state (182, 185). The decrease in cell culturability that was observed under refrigeration conditions during the study could be attributed to the VBNC state (131). The methodology used only accounted for culturable cells and the 61 effect, if any, of the VBNC was not measured. If the reduction in number of cells was by a reversion into the VBNC state, no net reduction may have occurred. Further investigation is necessary to evaluate the VBNC state during cold shock under shellstock temperature control recommended by ISSC. Moreover, the pH of the oysters slightly decreased throughout the cold shock study but probably did not effect V. vulnificus growth and survival in the current study. Optimum pH for V. vulnificus growth is around pH 7 and minimum growth pH is pH 5.0 (95). Goatcher (49) found pH not to be a factor in vibrio cell death in oysters following cold shock. It was clear in this in vivo study that the capability for this pathogen to survive in oysters at cold temperature for a considerable period of time exists. Based on this result high temperatures and cold adaptation conditions should be avoided to reduce V. vulnificus numbers in oysters. This study suggests that cooling oysters at 4?C after harvesting could limit the multiplication of V. vulnificus while embedded in oyster shellstock, but might not be effective in order to reduce V. vulnificus numbers to the recommended limit by FDA (<30 CFU/g oyster). The guideline from ISSC should be evaluated due to the potential for cold adaptation-mediated survival of V. vulnificus. 62 0 1 2 3 4 5 6 0 20 40 60 80 100 120 140 160 180 Hour l o g C F U / g oyster m e at Total bacteria Vibrio count Vibrio vulnificus Fig. 15. Total heterotrophic bacteria, total Vibrio, and V. vulnificus (CFU/g oyster meat) during depuration of eastern oysters (Crassostrea virginica). pH of oysters meat during cold shock 6 6.1 6.2 6.3 6.4 6.5 6.6 6.7 6.8 6.9 7 01234567 sampling point p H 15 degree C 25 degree C 35 degree C Fig. 16. pH of oyster meat during cold shock experiment, vertical arrow indicated cold shock point. 63 At 35?C for 10h and cold shocked at 4?C for 168h 4 4.5 5 5.5 6 6.5 7 0 20 40 60 80 100 120 140 160 180 200 Hour l og C F U / g oyst er ATCC 27562 C7184Tr VV3 FIG. 17. Cold shock-survival response of V. vulnificus strain ATCC 27562, C7184Tr, and VV3 in oysters when oysters were incubated at 35?C for 10 h and cold shocked at 4?C for 168 h, vertical arrow indicated cold shock at 4?C. At 25?C for 12h and cold shocked at 4?C for 168h 4 4.5 5 5.5 6 6.5 7 0 20 40 60 80 100 120 140 160 180 200 Hour l og C F U / g oy s t er ATCC 27562 C7184Tr VV3 FIG. 18. Cold shock-survival response of V. vulnificus strain ATCC 27562, C7184Tr, and VV3 in oysters when oysters were incubated at 25?C for 12 h and cold shocked at 4?C for 168 h, vertical arrow indicated cold shock at 4?C. 64 At 15?C for 36h and cold shocked at 4?C for 168h 4 4.5 5 5.5 6 6.5 7 0 20 40 60 80 100 120 140 160 180 200 220 Hour l o g C F U / g oys t e r ATCC 27562 C7184Tr VV3 FIG. 19. Cold shock-survival response of V. vulnificus strain ATCC 27562, C7184Tr, and VV3 in oysters when oysters were incubated at 15?C for 36 h and cold shocked at 4?C for 168 h, vertical arrow indicated cold shock at 4?C. 65 V. IN VITRO COLD SHOCK GENE EXPRESSION OF VIBRIO VULNIFICUS 66 ABSTRACT Four cold shock genes (csp s ) encoding small cold shock proteins of V. vulnificus were cloned. In vitro gene expression of csp s and stress-related genes in four different strains of V. vulnificus was analyzed by reverse transcriptase PCR. Two different temperature profiles were compared. First, V. vulnificus cultures were incubated at 35?C for 10 h and then cold shocked at 4?C for 168 h. The transcript levels of csp5, and a stress regulator gene, rpoS, were found to be constant in both clinical and environmental strains throughout the study. The synthesis of katG, a catalase gene, was detected at different time points after cold shock, while the transcript of oxyR, a transcriptional activator of oxidative stress response, was expressed at all time points. The induction of csp1 and csp3 varied among strains, while csp4 transcript was down-regulated after cold shock in all strains. In the second experiment, after overnight incubation at 35?C, V. vulnificus cultures were transferred to 15?C for 36 h and then cold shocked at 4?C for 168 h. A constitutive expression of csp1 and rpoS was observed, while transcription of csp3 and csp5 was discontinued after cold shock at 4?C. The synthesis of oxyR mRNA was induced during the acclimation step at 15?C. KatG gene expression was induced after shift to 4?C in some strains. The expression of csp4 was repressed after cold shock at 4?C. The constitutive expression of 16SrRNA indicated that V. vulnificus was viable during cold shock treatment. These results revealed that gene expression in V. vulnificus varied during cold shock in vitro. It can be concluded that the cold shock genes of V. vulnificus tested in this study are not cold inducible. 67 INTRODUCTION Temperature has wide-ranging effects on microbial growth and survival, and therefore, bacteria have developed responses that allowed them to adapt to change in temperature. Bacteria can exhibit a cold shock response by altering gene expression in response to a shift to low temperature (6, 7, 55, 80, 109, 140). Cold shock response inhibits the synthesis of most cellular proteins resulting in a growth lag period, called acclimation phase (72). Cold shock proteins are synthesized to overcome the deleterious effects of cold shock (146). Cold shock proteins (Csps) are small proteins (around 7 kDa) and the Csp family has been found in nearly all bacteria (41). Several functions have been postulated for Csps. These proteins could act as a transcription activators, RNA chaperones and could help to minimize secondary folding (55, 56). Moreover, Csps has been reported to play a role in protection against freezing (186). Csps are not expressed only at low temperatures; they are also induced by other environmental stresses. Most of the cold shock response studies have been performed with Escherichia coli and Bacillus subtilis as model systems (51). Out of the nine Csps found in E. coli, five were not induced at low temperatures. CspA is considered essential for cold adaptation and has been found to be expressed during the stationary growth phase (50, 146). CspD was induced when under nutritional stress during the stationary growth phase (190). Bacillus subtilis possesses three Csps: CspB, CspC, and CspD that are induced upon cold shock (57). Recently, Datta and Bhadra (36) investigated the major cold shock proteins in V. cholerae and reported that when V. cholerae cells were shifted from 37?C to 15?C, CspV and CspA were produced at low temperature and concluded 68 that cspV was a cold shock inducible gene and was tightly regulated at the transcription level. Vibrio vulnificus is a gram negative naturally occurring bacterium that causes septicemia, wound infections, and gastroenteritis. Transmission typically occurs through ingestion of contaminated raw or undercooked seafood, especially raw oysters harvested from the Gulf of Mexico during warm weather months (151). Vibrio vulnificus infections can develop into a fatal septicemia, especially in patients who suffer from immune disorders, liver disease, or hemochromatosis with death occurring within 48 h after onset of symptoms (48). The incidence of outbreak of V. vulnificus due to oysters is a significant concern of the oyster industry and public health agencies. Numerous studies have investigated the response of V. vulnificus to low temperatures (17, 110, 118, 128, 130, 149, 181, 185) although most of these studies focused on the viable but non- culturable (VBNC) state that can be induced in V. vulnificus when subjected to low temperatures. This VBNC state might explain the difficulty of culturing this bacterium from marine samples during winter months (for a recent review on the VBNC state see Oliver, 2005 (131)). However, the VBNC state is not the only response to cold temperatures observed in V. vulnificus. In 1995, McGovern and Oliver (110) showed that when V. vulnificus is subjected to a temperature downshift from 23?C to 13?C, protein synthesis is reduced but new cold-induced proteins are expressed. In addition, an adaptive response to cold temperatures in V. vulnificus has been described by Bryan et al.(17). Their study showed that V. vulnificus cells shifted from 35?C to 4?C entered the VBNC state in few days. However, when cells were acclimated from 35?C to 15?C and then shifted to 6?C, they remained viable for the duration of the study. Their 69 results also suggested that cold-induced proteins played a role in this cold adaptive response. However, no evidence of a cold shock inducible gene was reported and no gene involved in the cold shock response in V. vulnificus has yet been identified. To date, the molecular mechanisms by which V. vulnificus adapt to low temperature are not well understood. To better explain the physiological mechanisms response to cold shock at the molecular level, the small cold shock family genes (csp s ) obtained from two clinical and two environmental strains of V. vulnificus were cloned. Analysis of V. vulnificus cold shock genes encoding cold shock proteins (Csps) that exhibited significant increased or decreased in mRNA abundance caused by the temperature downshift were performed. Moreover, rpoS, the stress regulator which is used by bacteria in the adaptive response through various environmental stresses was examined. KatG, a periplasmic catalase gene (HPI) and oxyR, a positive regulator of catalase gene (both involved in oxidative stress) were also analyzed in response to exposure to cold temperature. Reverse transcription PCR was used to analyze gene expression of all these genes in all strains used in this study. 70 MATERIALS AND METHODS 1) Cloning of the small cold shock family genes (csp s ) of V. vulnificus Bacterial strains and growth conditions Two clinical strains of V. vulnificus, ATCC 27562 (type strain), and C7184Tr (avirulent, translucent morphotype, provided by J.D. Oliver, University of North Carolina, Charlotte, NC) were used in this study. Two environmental strains (VV3, VV4) isolated from oyster samples collected from Mobile Bay, AL were also included. Cells were grown in marine broth (Difco, Detroit, Mich.) at 35?C. Primer design Putative cold shock genes were identified in V. vulnificus based on amino acid- amino acid similarities with V. cholerae cold-shock proteins (Table 1). Open reading frame (ORF) sequences from five putative cold shock genes (csp1-5), (see Appendix 2 to 6), were obtained from the whole V. vulnificus genome sequence available at the National Center for Biotechnology Information (NCBI) (GenBank accession no. NC0044459, NC004460 for V. vulnificus CMCP6, and GenBank accession no. NC005139, NC005140 for V. vulnificus YJ016) (23, 91) (Table 2). 71 DNA extraction DNA was extracted from the four strains of V. vulnificus (ATTC 27562, C7184Tr, VV3, and VV4) using DNeasy kit following the manufacturer?s instructions (QIAGEN, Valencia, CA). Briefly, 1.5 ml of V. vulnificus cells were placed in a 2 ml microcentrifuge tube centrifuged for 10 min at 3,000?g. The supernate was extracted and resulting pellet was resuspended in 180 ?l of buffer ATL, plus 20 ?l of proteinase K (10 mg/ml). The sample was incubated at 55 o C for 3 h until cells were completely lysed or appeared viscous. Samples were vortexed for 15 s, with 200 ?l of lysis buffer, followed by incubation at 70?C for 10 min. The resulting DNA precipitated with 200 ?l ethanol 95% (vol/vol). Washing buffers (AW1 and AW2) were added sequentially to the DNeasy Mini spin column and centrifuged at 3,500?g. The supernatants were discarded, and the DNA was resuspended in distilled water. DNA was quantified by Gene Quant spectrophotometer (Amersham Pharmacia Biotech, Sweden) and stored at -20?C for amplification. PCR condition and DNA amplification Genes were PCR amplified in a total volume of 50 ?l. Unless otherwise stated, all PCR reagents were purchased from Promega (Promega, Madison, WI). One hundred ng of DNA were used as template in a reaction containing; 0.2 U of Taq polymerase, 0.2 mM of dNTP mix, 1X PCR buffer, 2.5 mM MgCl 2 and 0.2 mM of each primer. Primers were designed in order to amplify the entire ORFs (Table 2). The reaction mix 72 was cycled through the following temperature profile: 95?C for 5 min; 95?C for 30 sec, 50?C for 30 sec, 72?C for 45 sec for a total of 35 times. The PCR reaction was ended at 72?C for 10 min and thereafter cooled down to 4?C. PCR products were analyzed through a 1% agarose gel. Gels were run for about 60 min at 102 V in TAE 1X buffer (0.04 M Tris/acetate, EDTA 1mM) mixed with ethidium bromide (0.5?g/ml). A 100-bp ladder was used as a standard marker. Following electrophoresis, the PCR products were visualized under a UV (245 nm) transilluminator. DNA purification The selected DNA bands were purified from the gels using the GENECLEAN III kit (Q-BIO gene, Carlsbad, CA) followed the manufacturer?s instruction. The DNA band from the agarose gel was excised and transferred to polypropylene tube. Six hundred ?l of NaI solution was added to cover the gel and incubated at 50 o C in water bath for 5 min. After the agarose block was melted, the NaI/DNA solution was resuspended with 5 ?l of EZ-GLASSMILK, mixed by vortexing 1 min, and the NaI/DNA/EZ-GLASSMILK solution was incubated at room temperature for 5 min (vortexed every two min). The solution was centrifuged at 11,000?g for 5 sec, the supernatant was discarded, and the pellet containing the EZ-GLASSMILK with bound DNA was saved. The washing step was done twice by adding 500 ?l of provided washing solution, then the solution was resuspended and centrifuged at 10,000?g for 5 sec, and the supernatant was discarded. The resulting pellet was air dried and resuspended in 5 ?l of ddH 2 O. After centrifugation 73 at 11,000?g for 1 min the supernatant of DNA was removed and transferred to a new tube. Cloning procedure Ligation of purified DNA to linearized plasmid The ligation mixture was prepared by using PGEM-T-easy vector system (Promega). The mixture of 5 ?l of Genecleaned DNA from purification step included 1 ?l of pGEM-T easy vector (50ng/?l), 5 ?l of 2X rapid ligation buffer, and 1 ?l of T4 DNA ligase (3U/?l) was incubated at room temperature for 3 h. Transformation The ligation mixture was mixed with 50 ?l of JM 109 competent cells (Promega) and incubated on ice for 20 min. The bacterial-plasmid mixture was heat shocked by placing the mixture in a 42?C water bath for 45 sec without shaking. Tubes were removed and incubated on ice for another 5 min. The transformed bacterial mixture was pooled in a tube containing 900 ?l Luria-Betani Medium (LB), incubated in a water bath at 37?C for 1 h, centrifuged for 1 min at 500?g, and resuspended in 75 ?l of LB. 74 Plating of recombinant E. coli cells Transformed E. coli cells were plated onto LB + ampicillin + IPTG/X-GAL agar and incubated at 37?C for 16-18 h for blue/white screening. White colonies were picked using a toothpick. Each colony was transferred to 5 ml LB broth/ 5?l ampicillin (50mg/ml), and incubated at 37?C for 16-18 h. Two ml of cell cultures were kept in 10% sterile glycerol at -80?C, while 3 ml of the cultures were used for plasmid isolation. Plasmid isolation Aurum TM plasmid mini kit (Bio-Rad laboratories, Hercules, CA) was used for plasmid isolation. Briefly, 1.5 ml of cell cultures was pelleted by centrifugation at 11,000?g for 5 min at room temperature, and the supernatant was removed by pipetting. Bacterial cell pellets were resuspended in 250 ?l of resuspension solution by vortexing. The suspensions were mixed with 250 ?l of lysis solution until the solution became viscous and slightly clear. The solutions were mixed with 350 ?l of neutralization solution until a visible precipitate was formed. Then the neutralized lysate was centrifuged for 5 min. The supernatant (plasmid DNA) was washed twice with washing solution via a plasmid mini column then transferred to a new 1.5 ml eppendorf tube. Fifty ?l of sterile deionized water was added to the column, incubated for 1 min, and centrifuged to elute the plasmid. The mini column was discarded and the eluted plasmid DNA stored at 4?C. Plasmid DNA samples were quantified using a GeneQuant Spectrophotometer (Amersham Pharmacia Biotech). Plasmid DNA was stored at -20?C. 75 Clone characterization/testing In order to check the presence of insert, the EcoRI restriction digestion analysis was performed. The plasmid vector has two EcoRI restriction sites on its sequence. DNA sequences were inserted between these two EcoRI recognition sites. The EcoRI restriction digestion was used to digest the ligated DNA insert. Briefly, a mixture of 5?l of plasmid DNA, 1 ?l of EcoRI enzyme (12U/?l), 2 ?l of 10X buffer and 12 ?l of deionized water was incubated at 37?C for 4 h. The restriction digestion products were resolved on 1 % agarose gels at 102 V in TAE 1X buffer (0.04 M Tris/acetate, EDTA 1mM) stained with ethidium bromide (0.5?l/ml) for 30 min. A 100-bp ladder was used as a standard marker. Following electrophoresis, the gel was visualized under a UV (245 nm) transilluminator. Sequencing and alignment Plasmid concentrations were adjusted to 200 ng/?l and sequenced at the Auburn University Genomics & Sequencing Lab (AU-GSL) using the T7 promoter and the ABI 3100 Genetic Analyzer. The sequences were assembled and aligned with the Vector NTI Advance 9.0 software (Informax, Frederick, MD). The nucleotide sequences of the csp genes of V. vulnificus were compared with that of other bacterial species using the BLAST program (Basic Local Alignment Search Tool) (1). 76 In vitro gene expression of cold shock genes and stress-related genes of V. vulnificus during cold shock Bacterial collection during cold shock treatment Expressions of four putative cold shock genes (csp s ) and additional genes involved in stress response; rpoS, oxyR, and katG of V. vulnificus were investigated under cold shock conditions. In order to mimic the conditions V. vulnificus might experience under the ISSC guidelines, the experiment used the following temperature profiles: a) cells (10 4 CFU/ml) were incubated at 35?C for 10 h and b) at 15?C for 36 h, and then transferred to an incubator shaker at 4?C for up to 168 h. Samples were taken at different time points and kept in RNAlater RNA stabilization solution (Ambion, Austin, TX.) and stored at 4?C until RNA isolation analysis. All four strains were analyzed simultaneously. Total RNA isolation and DNA elimination RNA Plus mini kit (QIAGEN, Valencia, CA) was used to extract total RNA from all the bacterial samples (according to the manufacturer?s protocol). Each bacterial sample was homogenized using QIAshredder homogenizers, and the lysate transferred to a gDNA Eliminator spin column. The column was centrifuged at 5,600?g, then discarded. The filtrate was mixed with 70% ethanol, and 750 ?l of the sample was transferred to an RNeasy spin column, and centrifuged for 15 sec at 5,600?g. The filtrate 77 was discarded. Seven hundred ?l of buffer RW1 was added to the column, and centrifuged at 5,600?g for another 15 sec to wash the spin column membrane. The filtrate was discarded. Then 500 ?l of Buffer RPE was added to wash the column and centrifuged at 5,600?g. This wash step was repeated once. Then the RNA was eluted with 50 ?l RNase-free water by centrifugation. RNA concentrations were quantified spectrophotometerically and stored at -80?C. In order to eliminate carryover DNA in the RNA extract, 1 ?g of RNA was treated for 30 min at 37?C with 150 U of DNase (Ambion). The DNase was then inactivated by heating at 65?C for 10 min. Reverse transcriptase PCR The presence and expression of csp1, csp3, csp4, csp5, rpoS, oxyR, and katG was tested using reverse transcriptase-PCR (RT-PCR) amplification. cDNA was synthesized using the cDNA synthesis kit (Amersham Biosciences, Piscataway, NJ). According to the manufacturer?s instructions, random hexamers were used to synthesize the first strand cDNA. Twenty ?l of RNA solution (5?g total RNA plus RNase-free water) was heated to 65?C for 10 min, and then was chilled on ice. The cDNA synthesis reaction mixture consisted of 20 ?l total RNA (3 ?g/?l), 1 ?l of primer, 1 ?l of DDT solution, and 11 ?l of Bulk first-strand reaction mix. The cDNA synthesis reaction was incubated at 37?C for 1h, then the completed first-strand reaction was heated to 90?C for 5 min to denature the RNA-cDNA duplex, and the reverse transcriptase was inactivated and chilled on ice. PCR amplification was performed as followed: 2 ?l of the cDNA was used as template. Primers targeting the eight genes were used for specific PCR (Table 3). The 78 absence of DNA contamination in RNA samples was verified by PCR using RNA as a template and internal control primers UFUL and URUL. A PCR negative control (no nucleic acid template) was also carried out. V. vulnificus DNA was used as a positive control for PCR. The expression level of the 16SrRNA gene was used as constitutive expression control (internal control) in the RT-PCR. All PCRs and RT-PCR were conducted using the appropriate annealing temperature according to the primer set used. PCR amplification followed this profile: initial heating step at 95?C for 5 min followed by 35 cycles of denaturation (95?C for 30 sec), annealing (temperature varying according to the primers used (Table 3) for 30 sec), and extension (72?C for 45 sec). The final extension ended at 72?C for 10 min and thereafter cooled to 4?C. The amplification product was separated by horizontal gel electrophoresis in a 1 % agarose gels at 102 V in TAE 1X buffer (0.04 M Tris/acetate, EDTA 1mM) stained with ethidium bromide (0.5?l/ml) for 30 min. A 100-bp ladder was used as a standard marker. Following electrophoresis, the gel was visualized under a UV light (245 nm) transilluminator. RESULTS Cloning of csp s genes from V. vulnificus By comparing the genome sequence of V. vulnificus strain YJ016 and CMCP6 (23, 91) with V. cholerae, several putative cold shock genes were identified. Using sequence data for the whole genome of V. vulnificus strain YJ016 and CMCP6 (23, 91), four putative cold shock genes in V. vulnificus YJ016 and five putative cold shock genes 79 in V. vulnificus CMCP6 were recognized (Table 2, nucleotide sequence see appendix 2-6). The four putative cold shock genes obtained from strain YJ016 shared 100% nucleotide similarity with those of strain CMCP6. These five predicted cold shock genes encoding for cold shock proteins were designated as csp1, csp2, csp3, csp4, and csp5. The entire sequences csp1 to 5 of V. vulnificus were amplified from clinical (ATCC 27562, C7184Tr) and environmental strains (VV3, VV4). All V. vulnificus isolates produced PCR products of the expected size, except csp2. This gene identified in CMCP6 strains was not found in any of the four V. vulnificus strains analyzed, and therefore was not included in the study. PCR products of these four cold shock genes were amplified from the clinical and the environmental strains of V. vulnificus and were cloned. The isolates were sequenced and the amplified DNA sequences were compared for homologous amino acid sequences using BLAST. A BLAST-X analysis of predicted Csp sequences indicated that Csp1 obtained from V. vulnificus ATCC 27562, C7184Tr, VV3 and VV4 strains was 100% identical to cold shock protein of V. vulnificus CMCP6 and YJ016 (GenBank accession number NP761574 and NP 934297), and 81.4% identity to cold shock transcriptional regulator CspA from V. cholerae (TIGR accession number VCA0166). Csp3 showed 98% similarity with the sequence of cold shock-like protein, CspD of V. vulnificus CMCP6 and YJ016 (GenBank Accession number NP760979 and NP935117), and 81.6% identity to the cold shock-like protein from V. cholerae CspD (TIGR accession number VC1142). Csp4 protein sequence revealed 98% matching with the protein sequence of V. vulnificus cold shock protein from GenBank accession number NP762479 and 937123, and was 88.4% identical to a cold shock DNA-binding domain protein of V. cholerae (TIGR accession number VCA0184). The deduced sequence of 80 Csp5 was found to be 98% identical to cold shock protein of V. vulnificus CMCP6 and YJ016 (GenBank accession number NP762464 and NP937108), and was 85.7% identity to cold shock protein CspV of V. cholerae (TIGR accession number VCA0933) (Table 1). The deduced amino acid sequences for csp1, csp3, csp4, and csp5 resulted in a 7 to 8 kDa proteins. The csp1 and csp3 genes are presented in the large chromosome (chromosome I), whereas the csp4 and csp5 are presented in the small chromosome (chromosome II). The M w and isoelectric point (pI) were also calculated and found to be similar to the values described for V. cholerae CSPs (Table 1) (36). In vitro expression of cold shock genes and stress-related genes of V. vulnificus during cold shock Gene expression analysis of cold shock genes and stress-related genes were investigated in pure culture using RT-PCR technology at two different temperature profiles. In the first experiment, V. vulnificus was incubated at 35?C for 10 h (T 0 ) and cells transferred to 4?C for one week. The mRNA transcripts of the 16SrRNA (house keeping gene), rpoS, a stress regulator gene, and csp5, a cold shock gene were detected and remained constant through the temperature downshift. Expression levels of the transcriptional activator of oxidative stress response, oxyR, were also detected at all time points although some intensity variations were observed. The expression of csp1 and csp3, encoding cold shock proteins of V. vulnificus varied between strains. Csp1 and csp3 mRNA synthesis was induced 24 h after cold shock only in C7184Tr strain. These genes appeared to be constitutively expressed in ATCC 27562 and in both environmental 81 strains. In addition, induction of katG (a catalase gene) was observed in all strains of V. vulnificus but at different time points after cold shock from 35?C to 4?C. In contrast, cold shock appeared to repress csp4 expression in all V. vulnificus strains used (Figure 20-21). The second experiment investigated gene expression when aliquots of V. vulnificus cells grown overnight at 35?C (T 0A ) were transferred to 15?C for 36 h (T 0B ), and then cold shocked at 4?C for one week. Continuous expression of 16SrRNA, rpoS, and csp1 was observed at 35?C, 15?C, and 4?C in clinical and environmental strains across the duration of the study. Additionally, csp3 and csp5 transcripts were constantly detected from 35?C through the acclimation period at 15?C and then at 4?C although some variations in expression levels at 4?C were observed. Csp4 expression was detected during the first cold shock (35?C to 15?C) and was down-regulated after transferred to refrigerate conditions. OxyR and katG showed different patterns of gene regulation. While oxyR mRNA was induced during the acclimation phase at 15?C, katG mRNA was induced after downshift to 4?C in some strains. Once expression of oxyR occurred, gene expression remained fairly constant in all strains. Interestingly, katG showed an induction pattern 48h after cells were transferred to 4?C only in the environmental strain VV3. On the contrary, VV4 expressed katG immediately after cold shock at 4?C but transcripts levels steady decreased overtime. Expression of katG in clinical strains was seldom detected (Fig.22-24). 82 DISCUSSION Several food-borne pathogens have been reported to manifest their tolerance to cold temperatures by altering their cellular protein profiles and continuing to grow (6, 36, 55, 75, 90). The cold shock response is a physiological response of living cells to temperature downshift which involves de novo protein synthesis (52, 54, 123). From this current study, RT-PCR analysis revealed that csp s transcripts were constitutively expressed before and after cold shock with a few exceptions. Expression of csp1, csp3, and csp5 was detected at all sampling times in strains ATCC 27562, VV3 and VV4 with or without the 15?C cold adaptation period. Interestingly, csp1 and csp3 were not expressed in the avirulent mutant C7184tr at 35?C but was induced 24 h post cold shock. However, csp1 and csp3 expression was detected in C7184Tr during the 15?C adaptation period. The induction of csp1 transcript (homologous to cspA) in C7184Tr strain which has a translucent morphotype after cold shock may be essential for cell protection during cold adaptation. It is believed that the cold shock response serves as an adaptive function and it has been suggested that cspA, together with other csp s , played a role in protecting cells from damage due to cold stress. The increased levels of cspA after cold shock have been reported to be necessary for compensating for higher stability of secondary structures in RNA at low temperature. The expression of V. vulnificus csp1 and csp5 (homologous genes to cspA and cspV, respectively) remained constant despite cold shock. This current result contrast the data reported from V. cholerae response to cold shock. Datta and Bhadra (36) reported a strong induction of cspA and cspV after cold shock at 15?C. The current result seems to indicate V. vulnificus cold shock response is 83 different from that trigger in V. cholerae. These present findings showed that all V. vulnificus cold shock genes tested in this study are not cold inducible. Csp4 showed a different expression pattern in all strains. This gene was expressed at 35?C and 15?C but its expression faded away 24 to 48 h post cold shock at 4?C. Despite sequence similarity with other cold-inducible genes, csp1, csp3, csp4, and csp5 cannot be considered class I csp genes (173), since no upregulation by cold shock was observed. Similarly, no differences in induction levels were detected (although some genes seemed to be repressed overtime at 4?C) and therefore did not follow the expression pattern of csp class II genes either. It can conclude that although V. vulnificus presented some of the csp genes described in other species, they are not upregulated during cold shock. In addition to csp s gene expression this study investigated other genes that have been implicated in response to environmental stressors in V. vulnificus (70, 94, 142, 154). As expected rpoS, a general stress regulator, was constitutively expressed thorough the study (160, 161). Generally, rpoS was expressed when response to various stress, including nutrient limitation, osmotic challenge, and during stationary growth phase (103). This finding suggests that the presence of rpoS transcript may be significant for cell survival during cold stress. OxyR regulated the expression of the periplasmic catalase (katG) in V. vulnificus and both genes have been shown to be repressed by low temperatures (94, 160). The current data show that levels of oxyR transcripts increased after 4?C cold shock while katG expression was slightly induced at 4?C between 60 min and 72 h post cold shock depending on strain. During the adaptation period at 15?C, oxyR was induced and reach high expression levels at 36 h. Expression levels decreased at 4?C overtime, except in strain VV3 which displayed high levels of oxyR throughout the 84 experiment. The expression of oxyR, a peroxide sensor in this research indicated V. vulnificus may experience a H 2 O 2 toxicity under aerobic condition during growing in culture broth. Expression of katG during the 15?C period was weak but after cold shock at 4?C main differences were observed between strains. While the type strain ATCC 27562 did not expressed this gene, a weak expression was detected in C7184Tr after cold shock. Interestingly, VV3 and VV4 presented opposite expression patterns. KatG expression in VV3 was gradually induced after 6 h post cold shock. On the contrary, VV4 exhibited high levels of katG immediately after cold shock but expression faded out until non-detectable levels at 168 h post cold shock. Smith and Oliver (160) showed that katG expression could be detected in both clinical and environmental strains of V. vulnificus up to 240 h post cold shock at 4?C. However, these authors were inducing the VBNC state and cells were maintained in low nutrient medium. In this present study use of rich medium might explain the differences in katG and oxyR expression since reactive oxygen species (ROS) between high and low nutrient media are likely to be different. Additionally, results show that the synthesis of the 16SrRNA housekeeping gene was never repressed following an abrupt temperature downshift. This result is in agreement with previous studies where housekeeping genes were similarly expressed at optimal and low temperatures (72). Detection of ribosomal mRNA is thought to be a good viability marker in many bacteria (13, 46, 93, 158). The constitutive expression of 16SrRNA mRNA indicated the viability of V. vulnificus occurred throughout the cold shock experiment. 85 In conclusion, no cold shock gene induction was observed in all four genes tested, although the differences in oxyR and katG expressions observed during cold shock suggest that V. vulnificus cells encountered not only cold stress but also oxidative stress. Further investigation of the role of these cold shock genes and stress-related genes and their regulation in V. vulnificus will help scientists to better understand the survival of this pathogen during cold shock. 86 TABLE 1. Computed theoretical pI and M w values of cold shock proteins encoded by putative cold shock genes of V. vulnificus Gene Chromosome pI M w (KDa) GenBank Accession no. Homologous gene in V. cholerae TIGR Accession no. V. cholerae * (36) % identity of amino acid with V. cholerae csp1 chromosome I 8.09 7.599 NP761574 cspA VCA0166 81.4 csp3 chromosome I 6.03 8.398 NP760979 cspD VC1142 81.6 csp4 chromosome II 6.54 7.521 NP762479 DNA-binding domain VCA0184 88.4 csp5 chromosome II 8.12 7.554 NP762464 cspV VCA0933 85.7 * TIGR, The Institute for Genome Research, from Datta and Bhadra (36) 86 87 TABLE 2. Primers design for cold shock genes of V. vulnificus Bacteria strain GenBank Accession Length bp Tm (?C) Primer sequences V. vulnificus CMCP6 (Chromosome I) V. vulnificus YJ016 (Chromosome I) (a) VV12757 VV1504 213 50 ATGTCTGGTAAAATGACTGG TTAGATAACTGTAACGTTAGAA V. vulnificus CMCP6 (Chromosome I) (b) VV12532 468 50 ATGAAAGGGAAGATTCTTCG CTACCAGCCGCCACATAA V. vulnificus CMCP6 (Chromosome I) V. vulnificus YJ016 (Chromosome I) (c) VV12119 VV2324 231 50 ATGTATAGCATGGCTACAG TTACTTCGTTGCTTGTACTT V. vulnificus CMCP6 (Chromosome II) V. vulnificus YJ016 (Chromosome II) (d) VV20519 VVA1067 210 50 ATGTCTAACACAGTAACCG TTAAAGCAGTGTGATCTCTT V. vulnificus CMCP6 (Chromosome II) V. vulnificus YJ016 (Chromosome II) (e) VV20503 VVA1052 213 50 ATGTCTAATAAAGTAACTGGTT TTATAGCGCAGTTACGTTTG (a): designated csp1, (b): designated csp2, (c): designated csp3, (d): designated csp4, and (e): designated csp5 88 TABLE 3. Primers used for RT-PCR of cold shock genes and stress-related genes for the in vitro study Gene GenBank Accession Number Characteristics Size of amplified fragment (bp) Tm (?C) used in RT-PCR Primer sequence References csp1 VV12757 VV1504 Cold shock protein, 213 50 ATGTCTGGTAAAATGACTGG TTAGATAACTGTAACGTTAGAA This study csp3 VV12532 Cold shock-like protein (CspD) 231 50 ATGTATAGCATGGCTACAG TTACTTCGTTGCTTGTACTT This study csp4 VV12119 VV2324 Cold shock protein 210 50 ATGTCTAACACAGTAACCG TTAAAGCAGTGTGATCTCTT This study csp5 VV20519 VVA1067 Cold shock protein 213 50 ATGTCTAATAAAGTAACTGGTT TTATAGCGCAGTTACGTTTG This study oxyR AY102627 Control katG expression, mediated catalase activity. 496 57 CGAATACCTAGTGGCTTTGGC GCATGGCACAGCAACACTTA This study rpoS AY187681 Stress regulator 391 50 GTGCATTACGAGGCGATGAA ATTTTGCCCACATCGTCAAC This study katG VV12755 Catalase gene 504 57 GTTGGTACCTCAACTACCGC GGATGTCTAGATTGAGGGCC (142) 16SrRNA X76333 X76334 Housekeeping gene 492 57 GCCTAACACATGCAAGTCGA CGTATTACCGCGGCTGCTGG (25) 88 89 16SrRNA M + - T 0 15m 30m 60m 6h 12h 24h 48h 72h 96h 120h 144h 168h ATCC 27562 C7184Tr VV3 VV4 csp1 M + - T 0 15m 30m 60m 6h 12h 24h 48h 72h 96h 120h 144h 168h ATCC 27562 C7184Tr VV3 VV4 csp3 M + - T 0 15m 30m 60m 6h 12h 24h 48h 72h 96h 120h 144h 168h ATCC 27562 C7184Tr VV3 VV4 csp4 M + - T 0 15m 30m 60m 6h 12h 24h 48h 72h 96h 120h 144h 168h ATCC 27562 C7184Tr VV3 VV4 csp5 M + - T 0 15m 30m 60m 6h 12h 24h 48h 72h 96h 120h 144h 168h ATCC 27562 C7184Tr VV3 VV4 FIG. 20. Vibrio vulnificus gene expression profiles: internal control: 16SrRNA, and cold shock genes: csp1, csp3, csp4, and csp5 during cells were cold shocked from 35?C 10 h to 4?C 168 h (T 0 : after 10 h incubation at 35?C, vertical arrow indicated cold shock at 4?C). 90 rpoS M + - T 0 15m 30m 60m 6h 12h 24h 48h 72h 96h 120h 144h 168h ATCC 27562 C7184Tr VV3 VV4 oxyR M + - T 0 15m 30m 60m 6h 12h 24h 48h 72h 96h 120h 144h 168h ATCC 27562 C7184Tr VV3 VV4 katG M + - T 0 15m 30m 60m 6h 12h 24h 48h 72h 96h 120h 144h 168h ATCC 27562 C7184Tr VV3 VV4 FIG. 21. Vibrio vulnificus gene expression profiles: stress-related genes: rpoS, oxyR, and katG during cells were cold shocked from 35?C 10 h to 4?C 168 h (T 0 : after 10 h incubation at 35?C, vertical arrow indicated cold shock at 4?C). 91 FIG. 22. Vibrio vulnificus gene expression profiles: internal control: 16RrRNA, and cold shock genes: csp1 and csp3 during cells were cold shocked from 35?C 10 h to 15?C 36 h , and from 15?C 36 h to 4?C 168 h (T0 A : after overnight incubation at 35?C, T0 B : after 36 h incubation at 15?C, the 1 st vertical arrow indicated cold shock at 15?C, the 2 nd vertical arrow indicated cold shock at 4?C). 91 92 FIG. 23. Vibrio vulnificus gene expression profiles: cold shock genes: csp4, and csp5 during cells were cold shocked from 35?C 10 h to 15?C 36 h , and from 15?C 36 h to 4?C 168 h (T0 A : after overnight incubation at 35?C, T0 B : after 36 h incubation at 15?C, the 1 st vertical arrow indicated cold shock at 15?C, the 2 nd vertical arrow indicated cold shock at 4?C). 92 93 FIG. 24. Vibrio vulnificus gene expression profiles: stress-related genes: rpoS, oxyR, and katG during cells were cold shocked from 35?C 10 h to 15?C 36 h , and from 15?C 36 h to 4?C 168 h (T0 A : after overnight incubation at 35?C, T0 B : after 36 h incubation at 15?C, the 1 st vertical arrow indicated cold shock at 15?C, the 2 nd vertical arrow indicated cold shock at 4?C). 93 94 VI. IN VIVO COLD SHOCK GENE EXPRESSION OF VIBRIO VULNIFICUS 95 ABSTRACT Analysis of V. vulnificus gene expression in oysters (in vivo) during cold shock was performed at two different temperature conditions. RT-PCR was used to analyze the presence and expression of cold shock genes (csp s ) and stress-related genes of V. vulnificus in oysters. During the first experiment, oysters were contaminated with V. vulnificus overnight at 25?C, transferred to 35?C for 10 h, and then cold shocked at 4?C for 168 h. Expression of 16SrRNA, a house keeping gene, and rpoS, a stress regulator was detected at all time points. The induction of csp5, a cold shock gene occurred after cold shock at 4?C, while the expression of csp4 varied between strains. The expression of csp1, csp3, oxyR (which mediated catalase activity) and katG, a catalase gene were not detected throughout the study period in neither clinical nor environmental strains. In the second experiment, oysters were contaminated with V. vulnificus and after overnight at 25?C, were cold adapted at 15?C for 36 h prior change to 4?C for 168 h. Constitutive gene expression pattern of 16SrRNA was observed whereas rpoS exhibited discontinuous gene expression throughout the study period. The transcript of csp5 was not continuous expressed at all time points. The synthesis of csp3 and csp4 mRNA was weakly expressed and varied among strains. The expression of oxyR and katG mRNA was down-regulated during 35?C, 15?C, and in the cold. These result demonstrated that V. vulnificus displayed variations of in vivo cold shock genes expression. This is the first report on V. vulnificus gene expression under in vivo conditions. 96 INTRODUCTION Vibrio vulnificus, a mesophilic shellfish-borne pathogen, is present as natural microbiota in warm coastal waters throughout most parts of the world (22, 85, 133, 169, 187). V. vulnificus thrives in warm seawater and is frequently isolated from shellfish from the Gulf of Mexico (85). Infection caused by this pathogen is a serious concern for the oyster industry and public health in the United States. Most cases of illnesses or fatality occurred from consumption of raw oysters harvested from the Gulf region (Alabama, Florida, Louisiana, Mississippi, and Texas) and a few cases from oysters harvested from the Northeast, and the Northwest in the U.S. (19). Primary septicemia, the most severe disease, was observed in individuals who had consumed raw oysters and was limited almost exclusive to persons with underlying liver disease (22). The occurrence and prevalence of V. vulnificus is seasonal with numbers peaking in summer month, with a range from May to October (32, 85). However, during the winter months, the decline in numbers of detectable V. vulnificus in coastal water and shellfish has been attributed to its entry into the viable but non-culturable (VBNC) state (30, 126). Low temperature has been reported to be one of the most important factors in limiting V. vulnificus survival. V. vulnificus was not capable of growing when incubated at temperature below 10?C, however low temperatures are not lethal to V. vulnificus (33) since cells can entry to the VBNC state when temperature is decreased below 5?C (185). Bryan et al. (17) reported that a brief cold adaptation period for 3 h at 15?C enhanced survival of V. vulnificus at 6?C. They concluded that cold-adaptive protective proteins might enhance survival and tolerance at cold temperature. Moreover, a downshift to low 97 temperature has been previously considered to change protein expression in V. vulnificus (110). The physiological response for survival strategy of living cells to temperature downshift have been reported in several bacteria (51). Microbes exposed to abrupt temperature downshift underwent severe physiological disturbance, such as changes in supercoiled DNA, reduction in membrane fluidity, and the formation of stable secondary structures in DNA and RNA that impaired replication, transcription and protein synthesis. To overcome the injurious effects of cold shock and ensured that cellular activity will resume or be maintained at low temperature, bacteria have developed a transient adaptive response, the cold shock response (78, 141). During that period, the expression of a subset of specific proteins termed cold shock proteins (Csps) is induced (146, 173). Csps are widespread among bacteria (41) and are not only required during the cold shock adaptive response but also play a general role in cellular function adaptation to various growth conditions (40). Cold shock has been shown to alter genes expression in many species of bacteria (6, 7, 55, 80, 109, 140). However, little is known about V. vulnificus gene expression in response to cold shock. Cold shock response in V. vulnificus has not been fully elucidated although this pathogen greatly impacts the oyster industry and public health. The preventive measures suggested by USFDA/ISSC, required refrigeration of oysters after harvest, may potentially induce a cold shock phenomenon in V. vulnificus. The molecular mechanisms of V. vulnificus response to cold shock need to be unraveled. While several reports have been investigated in vitro and in situ gene expression of V. vulnificus during the VBNC state (46, 160, 161), in vivo gene expression of this pathogen has not yet been attempted under any conditions. Since oysters can act as 98 reservoirs and transmission vectors for V. vulnificus, it is crucial to study the cold shock gene expression while this pathogen is embedded in oysters. Specifically, the investigation of cold shock response of V. vulnificus should be based on the result of the examination in oysters as results may differ from those obtained through artificial inoculation. In an effort to elucidate the in vivo gene expression of V. vulnificus during cold shock, reverse transcriptase PCR (RT-PCR) technology was used to monitor the presence and the expression of cold shock genes (csp s ), encoding cold shock protein, and stress related genes in two clinical and one environmental strains. In this study, I examined the transcription of small putative cold shock family genes of V. vulnificus (csp1, csp3, csp4, and csp5), encoded for cold shock proteins. In addition, rpoS, the stress regulator which is used by cells in adaptive response in various environmental stresses was monitored. KatG, a periplasmic catalase gene (HPI) and oxyR, a positive regulator of catalase gene which both involved in oxidative stress were analyzed in response to exposure to cold temperature. MATERIALS AND METHODS Bacterial strains and growth conditions Two clinical strains of V. vulnificus, ATCC 27562 (type strain), and C7184Tr (avirulent, translucent morphotype) provided by J.D. Oliver, University of North Carolina, Charlotte, NC and one environmental strain (VV3) isolated from Mobile Bay, 99 AL were used in these study. Cells were grown in marine broth (MB) with shaking and maintained on marine agar (MA) (Difco, Detroit, MI) at 35?C Oyster samples One-year-old live shellstock eastern oysters (Crassostrea virginica) (50-75 mm diameter) were depurated using a UV filtered (UV flow 3000 ?w-sec/cm 2 ), with a water flow-through system (180 l/h) for 7 days at Auburn University Shellfish Laboratory (AUSL), Department of Fisheries and Allied Aquacultures, Dauphin Island, AL and transported immediately to the Bacterial Genomics Laboratory, Department of Fisheries and Allied Aquacultures, Auburn University for processing. The oyster samples for each experiment were transported and stored at 4?C. Artificial inoculation Live depurated oysters were artificially-inoculated with V. vulnificus by self- inoculation filtration. Four aquaria (38 L) containing 27 L of autoclaved seawater were prepared. One hundred fifty oysters were placed into each aquarium at 25?C and were allowed to acclimate for 24 h. A fresh culture of each strain of V. vulnificus was grown in MB until exponential growth phase (10 8 CFU/ml, OD 450 =0.15) was reached. In each aquarium, 3 L of V. vulnificus cultures were used to inoculate oysters at approximately 10 6 CFU per gram of oyster. Oysters with no V. vulnificus artificially inoculation (control) was prepared and acclimated at the same condition. 100 Cold shock experiment After V. vulnificus inoculation, oysters remained at 25?C overnight prior to temperature downshift. Two different cold shock conditions were compared. During the first experiment oysters were transferred to 35?C for 10 h, and then chilled by ice-water immersion until the internal temperature in oysters reach to 4?C. Oysters were placed in a low temperature incubator at 4?C and incubated for 168 h. For the second experiment, inoculated oysters were transferred to 15?C for 36 h, chilled by ice-water immersion until the internal temperature reached 4?C. Oysters were incubated at 4?C for an additional 168 h. Internal oyster temperature and external ambient air temperature were monitored throughout the study. Reverse transcription-PCR (RT-PCR) detection (i) Sampling. The gastro-intestinal tract of oysters containing with V. vulnificus was sampled from these two different cold shock experiments. Samples were collected at various time intervals from time 0 h to 168 h. Three oyster gastro-intestinal tracts were pooled from each time point and kept in RNalater RNA stabilization reagent (Ambion, Austin, TX) and stored at 4?C for further RNA extraction analysis. 101 (ii) RNA isolation and contaminated DNA elimination Thirty milligrams of mixed tissue samples were used for RNA isolation. RNA Plus mini kit (QIAGEN, Inc., Valencia, CA) was used to extract total RNA from all oyster samples (according to the manufacturer?s protocol, as described in Chapter IV). RNA concentrations were quantified in Gene Quant spectrophotometer and stored at -80?C until being processed. In order to eliminate carryover DNA in the RNA extract, 1 ?g of RNA was treated for 30 min at 37?C with 150 U of DNase (Ambion, Austin, TX). Treated RNA was heated at 65?C for 10 min to inactivate DNase. (iii) Reverse transcription or cDNA synthesis cDNA synthesized using the cDNA synthesis kit from Amersham Biosciences. The presence and expression of seven genes in V. vulnificus was investigated using RT- PCR. Primers targeting four cold shock genes plus three stress-related genes were used (Table 4). The absence of DNA contamination in RNA samples was verified by PCR, using RNA as a template with primer target 16SrRNA universal primers (UFUL and URUL). Negative controls with no template nucleic acid were carried out. V. vulnificus DNA was used as a positive control for the reaction. Primers against V. vulnificus specific target sequences in the 16SrRNA gene were used to detect V. vulnificus constitutive gene expression. All PCRs were conducted using the annealing temperature varying according to the primer set used. PCR amplification consisted of the following temperature profile: initial heating step at 95?C for 5 min followed by 35 cycles of 102 denaturation (95?C for 30 sec), annealing (temperature varying according to the primers used (Table 1) for 30 sec), and extension (72?C for 45 sec). The final extension was ended at 72?C for 10 min and thereafter cooled to 4?C. The amplification product was separated by horizontal gel electrophoresis in a 1% agarose gel and PCR products were visualized by ethidium bromide staining and UV translumination. RESULTS The initial mRNA analysis by RT-PCR was 18 h (overnight incubation) after oysters kept at 25?C were exposed V. vulnificus. After treatment oysters were transferred to either 35?C for 10 h and 15?C for 36 h and downshifted to 4?C for one week. V. vulnificus displayed variations of in vivo cold shock genes expression after a temperature downshift. During the first experiment, when oysters were transferred from 25?C to 35?C for 10 h, and abruptly shifted to 4?C (Fig 23, 24), the expression of 16SrRNA and rpoS were detected or constitutively expressed at all time points in both clinical (ATCC 27562 and C7184Tr) and environmental strain (VV3). Both ATCC 27562 and C7184Tr strains exhibited weak induction of csp4 transcript at 25?C whereas only C7184Tr transcript was induced at 35?C. After a change to cold temperature, this gene was early induced (after 60 min at 4?C) in both ATCC 27562 and C7184Tr, while this transcript exhibited late expression after 168 h incubation at 4?C in VV3. The csp5 synthesis was early expressed after 30 min cold shock at 4?C in ATCC 27562, but detected later after 6 h cold incubation in both C7184Tr and VV3. The transcript of csp1 and csp3 were undetectable or down-regulated throughout the study period in both 103 clinical and environmental strains. The same expression pattern was shared by the two oxidative stress related genes, oxyR and katG which exclusively down-regulated during at 35 or 15?C and in the cold. When oysters were transferred from 25?C to 15?C for 36 h and downshifted at 4?C (Fig 25, 26), the synthesis of 16SrRNA was constantly expressed throughout the study period. RpoS was also detected but showed different intensity in mRNA expression at all time points in both clinical and environmental strains. The transcript of csp3 was induced at 25?C in both C7184Tr and VV3, but not in ATCC 27562. After a change from 15?C to 4?C, this gene was weakly induced in both ATCC 27562 and C7184Tr, while it was not expressed in VV3. The induction of csp4 mRNA was varied among strains. In ATCC 27562, this gene was induced at 25?C, after 36 h incubation at 15?C and the early incubation at 4?C. In C7184Tr, this gene was expressed at 25?C, the early period at 15?C, and during the beginning at 4?C, while this gene was discontinuously induced after the oysters were incubated at 15?C and 4?C in VV3. The transcription of csp5 was not continuously expressed at all time points in clinical strains, whereas this transcript was detected and showed discontinuous intensity in mRNA abundance after 1h change to an acclimation temperature (15?C) in environmental strain (VV3). The synthesis of csp1, oxyR, and katG were down-regulated throughout the study period in both clinical and environmental strains. 104 DISCUSSION Cold shock in microorganisms results in unstable metabolism and a longer lag phase. However, microbial growth after cold shock incubation for a lengthy period of time suggested cellular adaptive mechanisms with an altered or unaltered metabolic recovery (143). Therefore, the continued viability of V. vulnificus after incubation at 4?C demonstrated that V. vulnificus has the metabolic and genetic potential to survive at refrigerated temperatures. This research showed that upon cold shock some genes underwent a transient induction, some of which were found to decrease after cold shock. Jones et al. (79) hypothesized that the level of transcription is an important mechanism of regulation of cold shock protein synthesis. There was evidence that V. vulnificus survival and tolerance at cold temperature could be due to the expression of cold adaptive proteins (17). In the present study, the csp genes, encoding cold shock proteins were differentially expressed in response to cold shock in vivo. Some csp s genes were repressed whereas others were induced after temperature downshift meaning that these small cold shock family genes include non-cold and cold shock inducible genes. The repression of csp1, which has high sequence identity with cspA of V. cholerae occurred throughout the study period. This was in contrast to the findings obtained in V. cholerae when cold shocked in vitro in which cspA was highly induced during exposure to a low temperature and weakly induced at 37?C (36). In V. vulnificus the expression of csp3 and csp4 varied among strains from treatment to treatment. The synthesis of csp3 (encoded CspD), was repressed when oysters were transferred from 25?C to 35?C and abruptly shifted to 4?C. However, this gene was weakly induced at 25?C and at 4?C in C7184Tr but only induced 105 after a shift to 4?C in ATCC 27562. The expression of csp4, 86% identical to cold shock DNA-binding domain protein obtained from V. cholerae, occurred at both optimal temperature and at low temperature, while csp5, 84% identical to CspV encoded cspV of V. cholerae was induced during cold shock in vivo. This gene was detected at all time points but a discontinuous pattern was observed during cells adaptation to 15?C and to the cold. This result is in agreement with Datta and Bhadra (36). These authors reported that cspV was induced only after exposure to low temperature (15?C) in vitro and concluded that it was truly a product of cold shock inducible gene. V. vulnificus possess one catalase gene: katG which is induced by H 2 O 2 . KatG is regulated by alternative sigma factor, a product of the rpoS gene under control of oxyR (94). It appeares that rpoS mRNA was expressed by all three V. vulnificus strains upon cold shock and growth at 35 or 15?C. RpoS is involved in bacterial survival under stress conditions (70, 154, 192). These results were similar as those previously reported during in vitro and in situ study (160, 161) suggesting that rpoS transcript plays a role in cell survival during both cold and optimal growth conditions. KatG and oxyR genes, which are involved in oxidative stress, were not detected at all time points under any condition. Jenkins et al. and Altuvia et al. (2, 76) stated that the loss of oxyR induction was accompanied by decreased in katG expression and increased cellular sensitivity to H 2 O 2 , consistent with inadequate antioxidant defenses in these cells. Also, low temperature has been reported in inhibiting oxyR mediated catalase activity in V. vulnificus (94). Smith and Oliver (160, 161) investigated gene expression in vitro and in situ and reported that the expression of katG was decreased during winter months in the VBNC state of V. vulnificus. These authors showed that when V. vulnificus was restored to higher 106 temperatures (>20?C), it resulted in the induction of this gene. The cold shock found in V. vulnificus has been described as a ?peroxide sensitive?(14), however, the disappearance of katG and oxyR transcript of V. vulnificus may be due to cells experience the absence of oxygen during persistence in oyster tissue. Last but not least, the constitutively expression of 16SrRNA of V. vulnificus was observed in this study. The presence of this gene indicated viability of V. vulnificus cells. Oliver (131) indicated this is a key test to determine whether such cells were dead or alive. mRNA transcription has been reported to be a reliable marker of viable cells due to its short half life and its role in bacterial cell physiology (158). Several studies have demonstrated continued gene expression by cells in the VBNC state in vitro (34, 46, 102, 191). The current findings lead to the question if the decline in cell culturablility observed during cold shock was caused by dead cells or by cells entering into VBNC state. The expression of 16SrRNA provided evidence that during cold shock, V. vulnificus cells were alive and suggested that cells might enter into the VBNC state resulting in the decline in CFU on media. The difference in CFU between these two temperature profiles may be due to the cold adaptive mechanism that enhanced cell culturability. In addition, variations of V. vulnificus gene regulation may enhance bacterial survival during cold shock in vivo. Expression of 16SrRNA gene under cold shock should be quantified by using real-time RT-PCR or other quantification technique, in order to determine if a decrease in net gene expression occurred under cold temperatures. The current study is the first report on V. vulnificus gene expression during cold shock in vivo. Cold adaptive response in V. vulnificus was observed at the molecular 107 level when cells were exposed to intermediate temperatures. However, further research is need to confirm this cold adaptation is truly compensated their survivals. The loss of culturable cells due to cells entering the VBNC state using cooling strategy under oyster industry practices is another aspect of great concern for their virulence. Investigation of the role of cold shock genes and other stress-related genes and their regulation in V. vulnificus may explain the survival of this pathogen during cold shock. Understanding the impact of cold stress response on the physiology of this pathogenic microorganism in shellfish, particularly in regard to their survival will improve strategies control of V. vulnificus risk. 108 TABLE 4. Primers used for RT-PCR of cold shock genes and stress-related genes for the in vivo study Gene GenBank Accession Number Characteristics Size of amplified fragment (bp) Tm (?C) used in RT-PCR Primer sequence References csp1 VV12757 VV1504 Cold shock protein, 213 50 ATGTCTGGTAAAATGACTGG TTAGATAACTGTAACGTTAGAA This study csp3 VV12532 Cold shock-like protein (CspD) 231 50 ATGTATAGCATGGCTACAG TTACTTCGTTGCTTGTACTT This study csp4 VV12119 VV2324 Cold shock protein 210 50 ATGTCTAACACAGTAACCG TTAAAGCAGTGTGATCTCTT This study csp5 VV20519 VVA1067 Cold shock protein 213 50 ATGTCTAATAAAGTAACTGGTT TTATAGCGCAGTTACGTTTG This study oxyR AY102627 Control katG expression, mediated catalase activity. 496 57 CGAATACCTAGTGGCTTTGGC GCATGGCACAGCAACACTTA This study rpoS AY187681 Stress regulator 391 50 GTGCATTACGAGGCGATGAA ATTTTGCCCACATCGTCAAC This study katG VV12755 Catalase gene 504 57 GTTGGTACCTCAACTACCGC GGATGTCTAGATTGAGGGCC (142) 16SrRNA X74726 Housekeeping gene Vv Species-specific 273 55 TCTAGCGGAGACGCTGGA GCTCACTTTCGCAAGTTGGCC (87, 108) 16SrRNA X76333 X76334 Housekeeping gene 492 57 GCCTAACACATGCAAGTCGA CGTATTACCGCGGCTGCTGG (25) 108 109 16SrRNA M + - T 0 T 1 30m 60m 6h 12h 24h 48h 72h 120h 168h ATCC 27562 C7184Tr VV3 csp1 M + - T 0 T 1 30m 60m 6h 12h 24h 48h 72h 120h 168h ATCC 27562 C7184Tr VV3 csp3 M + - T 0 T 1 30m 60m 6h 12h 24h 48h 72h 120h 168h ATCC 27562 C7184Tr VV3 csp4 M + - T 0 T 1 30m 60m 6h 12h 24h 48h 72h 120h 168h ATCC 27562 C7184Tr VV3 csp5 M + - T 0 T 1 30m 60m 6h 12h 24h 48h 72h 120h 168h ATCC 27562 C7184Tr VV3 FIG. 25. Vibrio vulnificus gene expression profiles: internal control: 16SrRNA, and cold shock genes: csp1, csp3, csp4, and csp5 during oysters were cold shocked from 35?C 10 h to 4?C 168 h (T 0 : after overnight incubation at 25?C, T 1 : after 10 h incubation at 35?C, vertical arrow indicated cold shock at 4?C). 110 oxyR M + - T 0 T 1 30m 60m 6h 12h 24h 48h 72h 120h 168h ATCC 27562 C7184Tr VV3 katG M + - T 0 T 1 30m 60m 6h 12h 24h 48h 72h 120h 168h ATCC 27562 C7184Tr VV3 rpoS M + - T 0 T 1 30m 60m 6h 12h 24h 48h 72h 120h 168h ATCC 27562 C7184Tr VV3 FIG. 26. Vibrio vulnificus gene expression profiles: stress-related genes: oxyR, katG, and rpoS during oysters were cold shocked from 35?C 10 h to 4?C 168 h (T 0 : after overnight incubation at 25?C, T 1 : after 10 h incubation at 35?C, vertical arrow indicated cold shock at 4?C). 111 16SrRNA M + - T0 A 30m 60m 6h 12h 24h T0 B 30m 60m 6h 12h 24h 48h 72h 120h168h ATCC 27562 C7184Tr VV3 csp1 M + - T0 A 30m 60m 6h 12h 24h T0 B 30m 60m 6h 12h 24h 48h 72h 120h 168h ATCC 27562 C7184Tr VV3 csp3 M + - T0 A 30m 60m 6h 12h 24h T0 B 30m 60m 6h 12h 24h 48h 72h 120h 168h ATCC 27562 C7184Tr VV3 csp4 M + - T0 A 30m 60m 6h 12h 24h T0 B 30m 60m 6h 12h 24h 48h 72h 120h 168h ATCC 27562 C7184Tr VV3 csp5 M + - T0 A 30m 60m 6h 12h 24h T0 B 30m 60m 6h 12h 24h 48h 72h 120h 168h ATCC 27562 C7184Tr VV3 FIG. 27. Vibrio vulnificus gene expression profiles: internal control: 16SrRNA, and cold shock genes: csp1, csp3, csp4, and csp5 during oysters were cold shocked from 25?C overnight to 15?C 36 h , and from 15?C 36 h to 4?C 168 h (T0 A : after overnight incubation at 25?C, T0 B : after 36 h incubation at 15?C, the 1 st vertical arrow indicated cold shock at 15?C, the 2 nd vertical arrow indicated cold shock at 4?C). 112 oxyR M + - T0 A 30m 60m 6h 12h 24h T0 B 30m 60m 6h 12h 24h 48h 72h 120h 168h ATCC 27562 C7184Tr VV3 katG M + - T0 A 30m 60m 6h 12h 24h T0 B 30m 60m 6h 12h 24h 48h 72h 120h 168h ATCC 27562 C7184Tr VV3 rpoS M + - T0 A 30m 60m 6h 12h 24h T0 B 30m 60m 6h 12h 24h 48h 72h 120h 168h ATCC 27562 C7184Tr VV3 FIG. 28. Vibrio vulnificus gene expression profiles: stress-related genes: oxyR, katG, and rpoS during oysters were cold shocked from 25?C overnight to 15?C 36 h , and from 15?C 36 h to 4?C 168 h (T0 A : after overnight incubation at 25?C, T0 B : after 36 h incubation at 15?C, the 1 st vertical arrow indicated cold shock at 15?C, the 2 nd vertical arrow indicated cold shock at 4?C). 113 VII. CONCLUSIONS 114 Cold shock response may significantly increase survival in bacteria exposed to low temperatures including shellfish-borne pathogens such as V. vulnificus. In this dissertation, the primary objective addressed issues relating to cold shock adaptive mechanism in V. vulnificus both under laboratory conditions (in vitro) and in shellstock oysters (in vivo) based on the use of current shellstock temperature guidelines recommended by ISSC. The molecular mechanism underlying this cold shock phenomenon was also investigated at the transcriptional level. Overall results can be summarized as follows: 1) The negative effect on V. vulnificus cell culturabilty caused by 4?C is significantly greater than by 7.2?C. However, when V. vulnificus was exposed to optimal or suboptimal temperatures, neither 7.2?C nor 4?C reduced V. vulnificus cell culturability by less than factor of 100 during at a one-week period. 2) Cold shock adaptation mediated survival of V. vulnificus was found for both in vitro and in vivo experiments. - Cells adaptation to 15?C prior to cold shock to 7.2?C enhanced V. vulnificus survival. - Acclimation of infected oysters at 15?C resulted in V. vulnificus greater survival during refrigeration at 4?C. 115 3) The expression of cold shock genes (csp s ) of V. vulnificus varied among strains and differed between the in vitro and the in vivo conditions. Cold shock genes of V. vulnificus tested in this study (csp1, csp3, csp4, csp5) can not be considered cold shock inducible genes based on in vitro experiments. However, some weak inductions of csp3, csp4, and csp5 were observed under in vivo conditions. 4) The continuous synthesis of rpoS in both in vitro and in vivo studies highlights the importance of this gene for cells viability. 5) The induction of oxyR and katG suggested that an oxidative stress response occurred during the in vitro study but not during the in vivo study. 6) The constant expression of 16SrRNA in both in vitro and in vivo experiments indicated the viability of V. vulnificus cells through out the cold shock period. From these cumulative findings I suggest that: 1) The stronger negative effects caused by 4?C versus 7.2?C might warrant the use of stringent refrigeration conditions by the shellfish industry. 2) Vibrio vulnificus should be kept below the growing temperature range (<15?C) at all times to prevent its multiplication. 3) Conditions that favor adaptation to cold temperatures should be avoided in V. vulnificus control strategies. 4) Minimizing V. vulnificus growth by limiting the time from harvest to refrigeration should be strictly enforced. 5) Whether net reduction of V. vulnificus cold shocked cells is due to cells entering the VBNC state need to be further evaluated. Analysis of 16SrRNA gene 116 expression should be accurately quantified by real-time RT-PCR or other sensitive technique, in order to determine if a decrease in net gene expression occurred under cold temperatures. Additional studies on the VBNC state under ISSC guidelines conditions will probably unravel the risk of V. vulnificus infections due to VBNC cells. 6) Vibrio vulnificus response to cold shock is different from the elucidated in V. cholerae, since the homologous csp s investigated were not cold inducible under in vitro conditions. Vibrio vulnificus has to harbor additional cold response mechanism that should be further identified and investigated. Noteworthy, my dissertation is the first report examined the expression of cold shock genes (csp s ) in V. vulnificus under laboratory condition (in vitro), and while it was embedded in oysters (in vivo). The cumulative results obtained from this study highlight some very important aspects that need to be addressed from the perspective of seafood safety. My investigation ?In vitro and in vivo cold shock response in V. vulnificus? may provide some accurate scientific data for control analysis of V. vulnificus septicemia associated with the consumption of raw oysters. 117 CUMULATIVE BIBLIOGRAPHY 118 1. Altschul, S. F., W. Gish, W. Miller, E. W. Myers, and D. J. Lipman. 1990. Basic Local Alignment Search Tool. J. Mol. Biol. 215:403-410. 2. Altuvia, S., M. Almiron, G. Huisman, R. Kolter, and G. Storz. 1994. The Dps promoter is activated by oxyR during growth and by Ihf and a sigma(S) in stationary-phase. Mol. Microbiol. 13:265-272. 3. Amaro, C., and E. Biosca. 1996. Vibrio vulnificus biotype 2, pathogenic for eels, is also an opportunistic pathogen for humans. Appl. Environ. Microbiol. 62:1454- 1457. 4. Anderson, D. W., R. C. Lindrooth, B. C. Murray, and J. L. Teague. 1996. Cost of restrictions on Gulf oyster harvesting for control of Vibrio vulnificus- caused disease. Research Triangle Institute, NC. 5. Andrews, L. S., D. L. Park, and Y. P. Chen. 2000. Low temperature pasteurization to reduce the risk of vibrio infections from raw shell-stock oysters. Food Addit. Contam. 17:787-791. 6. Bayles, D. O., B. A. Annous, and B. J. Wilkinson. 1996. Cold stress proteins induced in Listeria monocytogenes in response to temperature downshock and growth at low temperatures. Appl. Environ. Microbiol. 62:1116-1119. 7. Berger, F., N. Morellet, F. Menu, and P. Potier. 1996. Cold shock and cold acclimation proteins in the psychrotrophic bacterium Arthrobacter globiformis SI55. J. Bacteriol. 178:2999-3007. 8. Berger, F., P. Normand, and P. Potier. 1997. capA, a cspA-like gene that encodes a cold acclimation protein in the psychrotrophic bacterium Arthrobacter globiformis SI55. J. Bacteriol. 179:5670-5676. 119 9. Berlin, D. L., D. S. Herson, D. T. Hicks, and D. G. Hoover. 1999. Response of pathogenic Vibrio species to high hydrostatic pressure. Appl. Environ. Microbiol. 65:2776-2780. 10. Biosca, E., J. Oliver, and C. Amaro. 1996. Phenotypic characterization of Vibrio vulnificus biotype 2, a lipopolysaccharide-based homogeneous O serogroup within Vibrio vulnificus. Appl. Environ. Microbiol. 62:918-927. 11. Birkenhauer, J. M., and J. D. Oliver. 2003. Use of diacetyl to reduce the load of Vibrio vulnificus in the eastern oyster, Crassostrea virginica. J. Food Prot. 66:38- 43. 12. Bisharat, N., V. Agmon, R. Finkelstein, R. Raz, G. Ben-Dror, L. Lerner, S. Soboh, R. Colodner, D. N. Cameron, D. L. Wykstra, D. L. Swerdlow, and J. J. Farmer. 1999. Clinical, epidemiological, and microbiological features of Vibrio vulnificus biogroup 3 causing outbreaks of wound infection and bacteraemia in Israel. Lancet 354:1421-1424. 13. Bleve, G., L. Rizzotti, F. Dellaglio, and S. Torriani. 2003. Development of reverse transcription (RT)-PCR and real-time RT-PCR assays for rapid detection and quantification of viable yeasts and molds contaminating yogurts and pasteurized food products. Appl. Environ. Microbiol. 69:4116-4122. 14. Bogosian, G., N. D. Aardema, E. V. Bourneuf, P. J. L. Morris, and J. P. O'Neil. 2000. Recovery of hydrogen peroxide-sensitive culturable cells of Vibrio vulnificus gives the appearance of resuscitation from a viable but nonculturable state. J. Bacteriol. 182:5070-5075. 120 15. Brayton, P. R., P. A. West, E. Russek, and R. R. Colwell. 1983. New selective plating medium for isolation of Vibrio vulnificus biogroup 1. J. Clin. Microbiol. 17:1039-1044. 16. Brisco, P., B. Anderson, and C. Smith. 1996. A rapid protocol for the isolation of total RNA suitable for RT-PCR, Promega Notes, vol. 55. 17. Bryan, P. J., R. J. Steffan, A. DePaola, J. W. Foster, and A. K. Bej. 1999. Adaptive response to cold temperatures in Vibrio vulnificus. Curr. Microbiol. 38:168-175. 18. Campbell, M. S., and A. C. Wright. 2003. Real-time PCR analysis of Vibrio vulnificus from oysters. Appl. Environ. Microbiol. 69:7137-7144. 19. CDC. 1996. Vibrio vulnificus associated with eating raw oyster - Los Angeles. MMWR CDC Surveill. Summ. 45:621-624. 20. Cerda-Cuellar, M., J. Jofre, and A. R. Blanch. 2000. A selective medium and a specific probe for detection of Vibrio vulnificus. Appl. Environ. Microbiol. 66:855-859. 21. Cerda-Cuellar, M., L. Permin, J. L. Larsen, and A. R. Blanch. 2001. Comparison of selective media for the detection of Vibrio vulnificus in environmental samples. J. Appl. Microbiol. 91:322-327. 22. Chakraborty, S., G. B. Nair, and S. Shinoda. 1997. Pathogenic vibrios in the natural aquatic environment.63-80, 1997 Apr-Jun. 121 23. Chen, C.-Y., K.-M. Wu, Y.-C. Chang, C.-H. Chang, H.-C. Tsai, T.-L. Liao, Y.-M. Liu, H.-J. Chen, A. B.-T. Shen, J.-C. Li, T.-L. Su, C.-P. Shao, C.-T. Lee, L.-I. Hor, and S.-F. Tsai. 2003. Comparative genome analysis of Vibrio vulnificus, a marine pathogen. Genome Res. 13:2577-2587. 24. Chen, C. Y., L. Eckmann, S. J. Libby, F. C. Fang, S. Okamoto, M. F. Kagnoff, J. Fierer, and D. G. Guiney. 1996. Expression of Salmonella typhimurium rpoS and rpoS-dependent genes in the intracellular environment of eukaryotic cells. Infect. Immun. 64:4739-4743. 25. Chen, Y. C., M. C. Chang, Y. C. Chuang, and C. L. Jeang. 2004. Characterization and virulence of hemolysin III from Vibrio vulnificus. Curr. Microbiol. 49:175-179. 26. Chen, Y. P. 1996. Reduction and management of Vibrio vulnificus in Gulf Coast oyster (Crassostrea virginica). Louisiana State University, Louisiana. 27. Christman, M. F., R. W. Morgan, F. S. Jacobson, and B. N. Ames. 1985. Positive control of a regulon for defenses against oxidative stress and some heat- shock proteins in Salmonella typhimurium. Cell 41:753-762. 28. Cloutier, J., D. Prevost, P. Nadeau, and H. Antoun. 1992. Heat and cold shock protein synthesis in arctic and temperate strains of Rhizobia. Appl. Environ. Microbiol. 58:2846-2853. 29. Connell, J. O. 2002. RT-PCR in biomedicine, vol. 193. Humana Press, New Jersy. 122 30. Cook, D. W. 1994. Effect of time and temperature on multiplication of Vibrio vulnificus in postharvest Gulf Coast shellstock oysters. Appl. Environ. Microbiol. 60:3483-3484. 31. Cook, D. W. 1997. Refrigeration of oyster shellstock: conditions which minimize the out growth of Vibrio vulnificus. J. Food Prot. 60:349-352. 32. Cook, D. W., P. O Leary, J. C. Hunsucker, E. M. Sloan, J. C. Bowers, R. J. Blodgett, and A. De Paola. 2002. Vibrio vulnificus and Vibrio parahaemolyticus in U.S. retail shell oysters: A national survey from June 1998 to July 1999 (vol 65, pg 79, 2002). J. Food Prot. 65:445-445. 33. Cook, D. W., and A. D. Ruple. 1992. Cold-storage and mild heat-treatment as processing aids to reduce the numbers of Vibrio vulnificus in raw oysters. J. Food Prot. 55:985-989. 34. Coutard, F., M. Pommepuy, S. Loaec, and D. Hervio-Heath. 2005. mRNA detection by reverse transcription-PCR for monitoring viability and potential virulence in a pathogenic strain of Vibrio parahaemolyticus in viable but nonculturable state. J. Appl. Microbiol. 98:951-961. 35. Craig, J. E., D. Boyle, K. P. Francis, and M. P. Gallagher. 1998. Expression of the cold-shock gene cspB in Salmonella typhimurium occurs below a threshold temperature. Microbiology-Uk 144:697-704. 36. Datta, P. P., and R. K. Bhadra. 2003. Cold shock response and major cold shock proteins of Vibrio cholerae. Appl. Environ. Microbiol. 69:6361-6369. 37. Dawe, L. L., and W. R. Penrose. 1978. Bactericidal property of seawater - death or debilitation. Appl. Environ. Microbiol. 35:829-833. 123 38. DePaola, A., G. M. Capers, and D. Alexander. 1994. Densities of Vibrio vulnificus in the intestines of fish from the U.S. Gulf Coast. Appl. Environ. Microbiol. 60:984-988. 39. DePaola, A., M. L. Motes, A. M. Chan, and C. A. Suttle. 1998. Phages infecting Vibrio vulnificus are abundant and diverse in oysters (Crassostrea virginica) collected from the Gulf of Mexico. Appl. Environ. Microbiol. 64:346- 351. 40. Derzelle, S., B. Hallet, K. P. Francis, T. Ferain, J. Delcour, and P. Hols. 2000. Changes in cspL, cspP, and cspC mRNA abundance as a function of cold shock and growth phase in Lactobacillus plantarum. J. Bacteriol. 182:5105-5113. 41. Etchegaray, J. P., and M. Inouye. 1999. CspA, CspB, and CspG, major cold shock proteins of Escherichia coli, are induced at low temperature under conditions that completely block protein synthesis. J. Bacteriol. 181:1827-1830. 42. Fang, F. C., S. J. Libby, N. A. Buchmeier, P. C. Loewen, J. Switala, J. Harwood, and D. G. Guiney. 1992. The alternative sigma-factor katf (RpoS) regulates Salmonella virulence. Proc. Natl. Acad. Sci. U. S. A. 89:11978-11982. 43. FDA. 2004. Vibrio. Bacteriological Analytical Manual online.Chapter 9:26 p. 44. FDACFSAN. 2003. Issue relating to a Vibrio vulnificus risk management plan for oysters. Interstate Shellfish Sanitation Conference, Columbia, SC., Columbia, SC. 45. FDACFSAN. 2005. National shellfish sanitation program: Guide for the control of molluscan shellfish 2005. Food and Drug Adminstration. 124 46. Fischer-Le Saux, M., D. Hervio-Heath, S. Loaec, R. R. Colwell, and M. Pommepuy. 2002. Detection of cytotoxin-hemolysin mRNA in nonculturable populations of environmental and clinical Vibrio vulnificus strains in artificial seawater. Appl. Environ. Microbiol. 68:5641-5646. 47. Foley, D. A., and T. C. Cheng. 1975. Quantitative study of phagocytosis by hemolymph cells of pelecypods Crassostrea virginica and Mercenaria mercenaria. J. Invertebr. Pathol. 25:189-197. 48. Gholami, P., S. Q. Lew, and K. C. Klontz. 1998. Raw shellfish consumption among renal disease patients - A risk factor for severe Vibrio vulnificus infection. Am. J. Prev. Med. 15:243-245. 49. Goatcher, L. J., S. E. Engler, D. C. Wagner, and D. C. Westhoff. 1974. Effect of storage at 5?C on survival of Vibrio parahaemolyticus in processed Maryland oysters (Crassostrea virginica). Journal of Milk and Food Technology. 37:74-77. 50. Goldstein, J., N. S. Pollitt, and M. Inouye. 1990. Major cold shock protein of Escherichia coli. Proc. Natl. Acad. Sci. U. S. A. 87:283-287. 51. Golovlev, E. L. 2003. Bacterial cold shock response at the level of DNA transcription, translation, and chromosome dynamics. Microbiology 72:1-7. 52. GonzalezFlecha, B., and B. Demple. 1997. Transcriptional regulation of the Escherichia coli oxyR gene as a function of cell growth. J. Bacteriol. 179:6181- 6186. 53. Gooch, J. A. 2000. Factors affecting quantitative and qualitative changes in Vibrio parahaemolyticus and Vibrio vulnificus population in oysters. Mississippi State University, Mississippi. 125 54. Graumann, P., and M. A. Marahiel. 1996. Some like it cold: Response of microorganisms to cold shock. Arch. Microbiol. 166:293-300. 55. Graumann, P., K. Schroder, R. Schmid, and M. A. Marahiel. 1996. Cold shock stress-induced proteins in Bacillus subtilis. J. Bacteriol. 178:4611-4619. 56. Graumann, P., T. M. Wendrich, M. H. W. Weber, K. Schroder, and M. A. Marahiel. 1997. A family of cold shock proteins in Bacillus subtilis is essential for cellular growth and for efficient protein synthesis at optimal and low temperatures. Mol. Microbiol. 25:741-756. 57. Graumann, P. L., and M. A. Marahiel. 1999. Cold shock proteins CspB and CspC are major stationary-phase-induced proteins in Bacillus subtilis. Arch. Microbiol. 171:135-138. 58. Grunstein, M., and D. S. Hogness. 1975. Colony hybridization - method for isolation of cloned DNAs that contain a specific gene. Proc. Natl. Acad. Sci. U. S. A. 72:3961-3965. 59. Gumley, A. W., and W. E. Inniss. 1996. Cold shock proteins and cold acclimation proteins in the psychrotrophic bacterium Pseudomonas putida Q5 and its transconjugant. Can. J. Microbiol. 42:798-803. 60. Harwood, V. J., J. P. Gandhi, and A. C. Wright. 2004. Methods for isolation and confirmation of Vibrio vulnificus from oysters and environmental sources: a review. J. Microbiol. Methods 59:301-316. 61. Hebraud, M., E. Dubois, P. Potier, and J. Labadie. 1994. Effect of growth temperatures on the protein-levels in a psychrotrophic bacterium, Pseudomonas fragi. J. Bacteriol. 176:4017-4024. 126 62. Herigstad, B., M. Hamilton, and J. Heersink. 2001. How to optimize the drop plate method for enumerating bacteria. J. Microbiol. Methods 44:121-129. 63. Hesselman, D. M., M. L. Motes, and J. P. Lewis. 1999. Effects of a commercial heat-shock process on Vibrio vulnificus in the American oyster Crassostrea virginica harvested from the Gulf Coast. J. Food Prot. 62:1266-1269. 64. Hlady, W. G. 1997. Vibrio infections associated with raw oyster consumption in Florida, 1981-1994. J. Food Prot. 60:353-357. 65. Hoi, L., I. Dalsgaard, and A. Dalsgaard. 1998. Improved isolation of Vibrio vulnificus from seawater and sediment with Cellobiose colistin agar. Appl. Environ. Microbiol. 64:1721-1724. 66. Hoi, L., J. L. Larsen, I. Dalsgaard, and A. Dalsgaard. 1998. Occurrence of Vibrio vulnificus biotypes in Danish marine environments. Appl. Environ. Microbiol. 64:7-13. 67. Houghtby, G., and J. Liston. 1965. Effect of heating on Staphylococcus aureus in frozen precooked seafoods. Food Technology. 19:874-&. 68. Hsu, W. Y., C. I. Wei, and M. L. Tamplin. 1998. Enhanced broth media for selective growth of Vibrio vulnificus. Appl. Environ. Microbiol. 64:2701-2704. 69. Hu, X. P., P. Mallikarjunan, J. Koo, L. S. Andrews, and M. L. Jahncke. 2005. Comparison of kinetic models to describe high pressure and gamma irradiation used to inactivate Vibrio vulnificus and Vibrio parahaemolyticus prepared in buffer solution and in whole oysters. J. Food Prot. 68:292-295. 127 70. Hulsmann, A., T. M. Rosche, I. S. Kong, H. M. Hassan, D. M. Beam, and J. D. Oliver. 2003. RpoS-dependent stress response and exoenzyme production in Vibrio vulnificus. Appl. Environ. Microbiol. 69:6114-6120. 71. Imbert, M., and F. Gancel. 2004. Effect of different temperature downshifts on protein synthesis by Aeromonas hydrophila. Curr. Microbiol. 49:79-83. 72. Inouye, M. a. K. Y. 2000. Cold shock response and adaptation. JMMB Symp Ser 2:147p. 73. Jakabi, M., D. S. Gelli, J. Torre, M. A. B. Rodas, B. Franco, M. T. Destro, and M. Landgraf. 2003. Inactivation by ionizing radiation of Salmonella enteritidis, Salmonella infantis, and Vibrio parahaemolyticus in oysters (Crassostrea brasiliana). J. Food Prot. 66:1025-1029. 74. Jamet, A., E. Kiss, J. Batut, A. Puppo, and D. Herouart. 2005. The katA catalase gene is regulated by OxyR in both free-living and symbiotic Sinorhizobium meliloti. J. Bacteriol. 187:376-381. 75. Jeffreys, A. G., K. M. Hak, R. J. Steffan, J. W. Foster, and A. K. Bej. 1998. Growth, survival and characterization of cspA in Salmonella enteritidis following cold shock. Curr. Microbiol. 36:29-35. 76. Jenkins, D. E., J. E. Schultz, and A. Matin. 1988. Starvation-induced cross protection against heat or H 2 O 2 challenge in Escherichia coli. J. Bacteriol. 170:3910-3914. 77. Jiang, W. N., Y. Hou, and M. Inouye. 1997. CspA, the major cold-shock protein of Escherichia coli, is an RNA chaperone. J. Biol. Chem. 272:196-202. 128 78. Jones, P. G., and M. Inouye. 1994. The Cold-Shock Response - a Hot Topic. Mol. Microbiol. 11:811-818. 79. Jones, P. G., R. Krah, S. R. Tafuri, and A. P. Wolffe. 1992. DNA gyrase, cs7.4, and the cold shock response in Escherichia coli. J. Bacteriol. 174:5798- 5802. 80. Jones, P. G., R. A. Vanbogelen, and F. C. Neidhardt. 1987. Induction of proteins in response to low-temperature in Escherichia coli. J. Bacteriol. 169:2092-2095. 81. Kaspar, C. W., and M. L. Tamplin. 1993. Effects of temperature and salinity on the survival of Vibrio vulnificus in seawater and shellfish. Appl. Environ. Microbiol. 59:2425-2429. 82. Kaysner, C. A., and A. De Paola. 2004. Vibrio cholerae, V. parahaemolyticus, V. vulnificus, and other Vibrio spp. Bacteriological Analytical Manual, 8th ed. 83. Kaysner, C. A., M. L. Tamplin, M. M. Wekell, R. F. Stott, and K. G. Colburn. 1989. Survival of Vibrio vulnificus in shellstock and shucked oysters (Crassostrea gigas and Crassostrea virginica) and effects of isolation medium on recovery. Appl. Environ. Microbiol. 55:3072-3079. 84. Keller, G. H., and M. M. Manak. 1989. DNA probes. Stockton Press, NewYork. 85. Kelly, M. T. 1982. Effect of temperature and salinity on Vibrio (Beneckea) vulnificus occurrence in a Gulf Coast environment. Appl. Environ. Microbiol. 44:820-824. 129 86. Kelly, M. T., and A. Dinuzzo. 1985. Uptake and clearance of Vibrio vulnificus from Gulf coast oysters (Crassostrea virginica). Appl. Environ. Microbiol. 50:1548-1549. 87. Kim, M. S., and H. D. Jeong. 2001. Development of 16S rRNA targeted PCR methods for the detection and differentiation of Vibrio vulnificus in marine environments. Aquaculture 193:199-211. 88. Kim, S. O., K. Merchant, R. Nudelman, W. F. Beyer, T. Keng, J. DeAngelo, A. Hausladen, and J. S. Stamler. 2002. OxyR: A molecular code for redox- related signaling. Cell 109:383-396. 89. Kim, W. S., and N. W. Dunn. 1997. Identification of a cold shock gene in lactic acid bacteria and the effect of cold shock on cryotolerance. Curr. Microbiol. 35:59-63. 90. Kim, W. S., N. Khunajakr, and N. W. Dunn. 1998. Effect of cold shock on protein synthesis and on cryotolerance of cells frozen for long periods in Lactococcus lactis. Cryobiology 37:86-91. 91. Kim, Y. R., S. E. Lee, C. M. Kim, S. Y. Kim, E. K. Shin, D. H. Shin, S. S. Chung, H. E. Choy, A. Progulske-Fox, J. D. Hillman, M. Handfield, and J. H. Rhee. 2003. Characterization and pathogenic significance of Vibrio vulnificus antigens preferentially expressed in septicemic patients. Infect. Immun. 71:5461- 5471. 130 92. Kitaura, T., S. Doke, I. Azuma, M. Imaida, K. Miyano, K. Harada, and E. Yabuuchi. 1983. Halo production by sulfatase activity of V. vulnificus and V. cholerae O1 on a new selective sodium dodecyl sulfate-containing agar medium - a screening marker in environmental surveillance. FEMS Microbiol. Lett. 17:205- 209. 93. Klein, P. G., and V. K. Juneja. 1997. Sensitive detection of viable Listeria monocytogenes by reverse transcription-PCR. Appl. Environ. Microbiol. 63:4441- 4448. 94. Kong, I. S., T. C. Bates, A. Hulsmann, H. Hassan, B. E. Smith, and J. D. Oliver. 2004. Role of catalase and oxyR in the viable but nonculturable state of Vibrio vulnificus. FEMS Microbiol Ecol 50:133-142. 95. Koo, J., A. DePaola, and D. L. Marshall. 2000. Effect of simulated gastric fluid and bile on survival of Vibrio vulnificus and Vibrio vulnificus phage. J. Food Prot. 63:1665-1669. 96. Lange, R., and R. Henggearonis. 1994. The cellular concentration of the sigma(S) subunit of RNA-polymerase in Escherichia coli Is controlled at the levels of transcription translation, and protein stability. Genes Dev 8:1600-1612. 97. Lee, S. J., A. G. Xie, W. N. Jiang, J. P. Etchegaray, P. G. Jones, and M. Inouye. 1994. Family of the major cold-shock protein, CspA (Cs7.4), of Escherichia coli, whose members show a high sequence similarity with the eukaryotic Y-Box binding-proteins. Mol. Microbiol. 11:833-839. 98. Leonard, D. L. 2003. Managing public health risk : Vibrio vulnificus. 4th International Conference on Molluscan Shellfish Safety, Spain:287-302. 131 99. Levin, R. E. 2005. Vibrio vulnificus, a notably lethal human pathogen derived from seafood: A review of its pathogenicity, subspecies characterization, and molecular methods of detection. Food Biotechnology 19:69-94. 100. Linkous, D. A., and J. D. Oliver. 1999. Pathogenesis of Vibrio vulnificus. FEMS Microbiol. Lett. 174:207-214. 101. Lipp, E. K., and J. B. Rose. 1997. The role of seafood in foodborne diseases in the United States of America. Revue Scientifique Et Technique De L Office International Des Epizooties 16:620-640. 102. Lleo, M. D., S. Pierobon, M. C. Tafi, C. Signoretto, and P. Canepari. 2000. mRNA detection by reverse transcription-PCR for monitoring viability over time in an Enterococcus faecalis viable but nonculturable population maintained in a laboratory microcosm. Appl. Environ. Microbiol. 66:4564-4567. 103. Loewen, P. C., B. Hu, J. Strutinsky, and R. Sparling. 1998. Regulation in the rpoS regulon of Escherichia coli. Can. J. Microbiol. 44:707-717. 104. Lorio, J. W., and C. Petrone. 1994. The cultivation of American oysters (Crassostrea virginica), Sothern Regional Aquaculture Center, SRAC Publication No. 432. 105. Lutz, C. G., P. Sambidi, and R. W. Harrison. 2005. Oyster profile. Agricultural Marketing Resource Center, Iowa State University. 106. Maas, R. 1983. An improved colony hybridization method with significantly increased sensitivity for detection of single genes. Plasmid 10:296-298. 107. Massad, G., and J. D. Oliver. 1987. New selective and differential medium for Vibrio cholerae and Vibrio vulnificus. Appl. Environ. Microbiol. 53:2262-2264. 132 108. Maugeri, T. L., M. Carbone, M. T. Fera, and C. Gugliandolo. 2006. Detection and differentiation of Vibrio vulnificus in seawater and plankton of a coastal zone of the Mediterranean Sea. Res. Microbiol. 157:194-200. 109. Mayr, B., T. Kaplan, S. Lechner, and S. Scherer. 1996. Identification and purification of a family of dimeric major cold shock protein homologs from the psychrotrophic Bacillus cereus WSBC 10201. J. Bacteriol. 178:2916-2925. 110. McGovern, V., and J. Oliver. 1995. Induction of cold-responsive proteins in Vibrio vulnificus. J. Bacteriol. 177:4131-4133. 111. Meynell, G. G. 1958. The effect of sudden chilling on Escherichia coli. J. Gen. Microbiol. 19:380-389. 112. Miceli, G. A., W. D. Watkins, and S. R. Rippey. 1993. Direct plating procedure for enumerating Vibrio vulnificus in oysters (Crassostrea virginica). Appl. Environ. Microbiol. 59:3519-3524. 113. Michel, V., J. Labadie, and M. Hebraud. 1996. Effect of different temperature upshifts on protein synthesis by the psychrotrophic bacterium Pseudomonas fragi. Curr. Microbiol. 33:16-25. 114. Michel, V., I. Lehoux, G. Depret, P. Anglade, J. Labadie, and M. Hebraud. 1997. The cold shock response of the psychrotrophic bacterium Pseudomonas fragi involves four low-molecular-mass nucleic acid binding proteins. J. Bacteriol. 179:7331-7342. 133 115. Morgan, R. W., M. F. Christman, F. S. Jacobson, G. Storz, and B. N. Ames. 1986. Hydrogen peroxide-inducible proteins in Salmonella typhimurium overlap with heat-shock and other stress proteins. Proc. Natl. Acad. Sci. U. S. A. 83:8059- 8063. 116. Morris, J. G., Jr, A. C. Wright, D. M. Roberts, P. K. Wood, L. M. Simpson, and J. D. Oliver. 1987. Identification of environmental Vibrio vulnificus isolates with a DNA probe for the cytotoxin-hemolysin gene. Appl. Environ. Microbiol. 53:193-195. 117. Motes, M. L., and A. DePaola. 1996. Offshore suspension relaying to reduce levels of Vibrio vulnificus in oysters (Crassostrea virginica). Appl. Environ. Microbiol. 62:3875-3877. 118. Murphy, S. K., and J. D. Oliver. 1992. Effects of temperature abuse on survival of Vibrio vulnificus in oysters. Appl. Environ. Microbiol. 58:2771-2775. 119. Muth, K. M., D. W. Anderson, S. A. Karns, B. C. Murray, and J. L. Domanico. 2000. Economic impacts of requiring post-harvest treatment of oysters. Research Triangle Institute, RTI Project Number 7466000. 120. Nakashima, K., K. Kanamaru, T. Mizuno, and K. Horikoshi. 1996. A novel member of the cspA family of genes that is induced by cold shock in Escherichia coli. J. Bacteriol. 178:2994-2997. 121. National Marine Fisheries Service, F. S. a. E. D., Commercial Fisheries May, 2007, posting date. http://www.st.nmfs.gov/st1/commercial/landings/annual_lanndings.html. [Online.] 134 122. National Marine Fisheries Service, F. S. a. E. D., Foreign Trade Information March, 6 2007, posting date. http://www.st.nmfs.gov/st1/trade/cumulative_data/TradeDataProduct.html. [Online.] 123. O'Connell, K. P., A. M. Gustafson, M. D. Lehmann, and M. F. Thomashow. 2000. Identification of cold shock gene loci in Sinorhizobium meliloti by using a luxAB reporter transposon. Appl. Environ. Microbiol. 66:401-405. 124. O'Connell, K. P., and M. F. Thomashow. 2000. Transcriptional organization and regulation of a polycistronic cold shock operon in Sinorhizobium meliloti RM1021 encoding homologs of the Escherichia coli major cold shock gene cspA and ribosomal protein gene rpsU. Appl. Environ. Microbiol. 66:392-400. 125. O'Neill, K. R., S. H. Jones, and D. J. Grimes. 1992. Seasonal incidence of Vibrio vulnificus in the Great Bay estuary of New Hampshire and Maine. Appl. Environ. Microbiol. 58:3257-3262. 126. Oliver, J., F. Hite, D. McDougald, N. Andon, and L. Simpson. 1995. Entry into, and resuscitation from, the viable but nonculturable state by Vibrio vulnificus in an estuarine environment. Appl. Environ. Microbiol. 61:2624-2630. 127. Oliver, J. D. 1993. Formation of viable but nonculturable cells p. 239-242. In S. Kjelleberg (ed.), Starvation in bacteria, Prenum Press, New York 128. Oliver, J. D. 1981. Lethal cold stress of Vibrio vulnificus in oysters. Applied and Environmental Microbiology 41:710-717. 129. Oliver, J. D. 1981. Lethal cold stress of Vibrio vulnificus in oysters. Appl. Environ. Microbiol. 41:710-717. 135 130. Oliver, J. D. 1995. The viable but non-culturable state in the human pathogen Vibrio vulnificus. FEMS Microbiol. Lett. 133:203-208. 131. Oliver, J. D. 2005. The viable but nonculturable state in bacteria. Journal of Microbiology 43:93-100. 132. Oliver, J. D., and J. B. Kaper. 2001. Vibrio species, 2nd edition ed. American Society for Microbiology, Washington, DC. 133. Oliver, J. D., R. A. Warner, and D. R. Cleland. 1982. Distribution and ecology of Vibrio vulnificus and other lactose-fermenting marine vibrios in coastal waters of the Southeastern United States. Appl. Environ. Microbiol. 44:1404-1414. 134. Oliver, J. D., R. A. Warner, and D. R. Cleland. 1983. Distribution of Vibrio vulnificus and other lactose-fermenting Vibrios in the marine environment. Appl. Environ. Microbiol. 45:985-998. 135. Olsen, A., A. Arnqvist, M. Hammar, S. Sukupolvi, and S. Normark. 1993. The RpoS sigma factor relieves H-NS-mediated transcriptional repression of csgA, the subunit gene of fibronectin-binding curli in Escherichia coli. Mol. Microbiol. 7:523-536. 136. Palmer, C. J. a. C. P.-K. 2004. Polymerase chain reaction (PCR), vol. 70. Elesevier Academic Press, UK. 137. Panek, H. R., and M. R. O'Brian. 2004. KatG is the primary detoxifier of hydrogen peroxide produced by aerobic metabolism in Bradyrhizobium japonicum. J. Bacteriol. 186:7874-7880. 136 138. Panicker, G., M. L. Myers, and A. K. Bej. 2004. Rapid detection of Vibrio vulnificus in shellfish and Gulf of Mexico water by real-time PCR. Appl. Environ. Microbiol. 70:498-507. 139. Panoff, J. M., D. Corroler, B. Thammavongs, and P. Boutibonnes. 1997. Differentiation between cold shock proteins and cold acclimation proteins in a mesophilic gram-positive bacterium, Enterococcus faecalis JH2-2. J. Bacteriol. 179:4451-4454. 140. Panoff, J. M., S. Legrand, B. Thammavongs, and P. Boutibonnes. 1994. The cold shock response in Lactococcus lactis subsp Lactis. Curr. Microbiol. 29:213- 216. 141. Panoff, J. M., B. Thammavongs, M. Gueguen, and P. Boutibonnes. 1998. Cold stress responses in mesophilic bacteria. Cryobiology 36:75-83. 142. Park, K.-J., M.-J. Kang, S. H. Kim, H.-J. Lee, J.-K. Lim, S. H. Choi, S.-J. Park, and K.-H. Lee. 2004. Isolation and characterization of rpoS from a pathogenic bacterium, Vibrio vulnificus: Role of {sigma}S in survival of exponential-phase cells under oxidative stress. J. Bacteriol. 186:3304-3312. 143. Patterson, T. E., and H. Jackson. 1979. Effect of storage at 1?C and 4?C on viability and Injury of Staphylococcus aureus, Escherichia coli and Streptococcus faecalis. J. Appl. Bacteriol. 46:161-167. 144. Perkins, E. B. 1995. Aquacultured oyster products: inspection, quality, handling, storage, safety, Sothern Regional Aquaculture Center, SRAC Publication No. 434 137 145. Pfeffer, C. S., M. F. Hite, and J. D. Oliver. 2003. Ecology of Vibrio vulnificus in estuarine waters of Eastern North Carolina. Appl. Environ. Microbiol. 69:3526- 3531. 146. Phadtare, S., J. Alsina, and M. Inouye. 1999. Cold shock response and cold- shock proteins. Curr. Opin. Microbiol. 2:175-180. 147. Phanthanh, L., and T. Gormon. 1995. Analysis of heat and cold shock proteins in Listeria by 2-dimensional electrophoresis. Electrophoresis 16:444-450. 148. Potasman, I., A. Paz, and M. Odeh. 2002. Infectious outbreaks associated with bivalve shellfish consumption: A worldwide perspective. Clin. Infect. Dis. 35:921-928. 149. Quevedo, A. C., J. G. Smith, G. E. Rodrick, and A. C. Wright. 2005. Ice immersion as a postharvest treatment of oysters for the reduction of Vibrio vulnificus. J. Food Prot. 68:1192-1197. 150. Ren, T. T., and Y. C. Su. 2006. Effects of electrolyzed oxidizing water treatment on reducing Vibrio parahaemolyticus and Vibrio vulnificus in raw oysters. J. Food Prot. 69:1829-1834. 151. Rippey, S. R. 1994. Infectious-diseases associated with molluscan shellfish consumption. Clin. Microbiol. Rev. 7:419-&. 152. Roberts, M. E., and W. E. Inniss. 1992. The synthesis of cold shock proteins and cold-acclimation proteins in the psychrophilic bacterium Aquaspirillum arcticum. Curr. Microbiol. 25:275-278. 153. Roberts, N. C., R. J. Siebeling, J. B. Kaper, and H. B. Bradford. 1982. Vibrios in the Louisiana Gulf-coast environment. Microb. Ecol. 8:299-312. 138 154. Rosche, T. M., D. J. Smith, E. E. Parker, and J. D. Oliver. 2005. RpoS involvement and requirement for exogenous nutrient for osmotically induced cross protection in Vibrio vulnificus. FEMS Microbiol Ecol 53:455-462. 155. Schellhorn, H. E., and V. L. Stones. 1992. Regulation of katF and katE in Escherichia coli K-12 by weak acids. J. Bacteriol. 174:4769-4776. 156. Science, N. A. o. 1991. Seafood safety reports. Microbiological and parasitic exposures and health effects.p 33-94. 157. Shapiro, R. L., S. Altekruse, L. Hutwagner, R. Bishop, R. Hammond, S. Wilson, B. Ray, S. Thompson, R. V. Tauxe, and P. M. Griffin. 1998. The role of Gulf Coast oysters harvested in warmer months in Vibrio vulnificus infections in the United States, 1988-1996. J. Infect. Dis. 178:752-759. 158. Sheridan, G. E. C., C. I. Masters, J. A. Shallcross, and B. M. Mackey. 1998. Detection of mRNA by reverse transcription PCR as an indicator of viability in Escherichia coli cells. Appl. Environ. Microbiol. 64:1313-1318. 159. 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. 176:1729-1737. 160. Smith, B., and J. D. Oliver. 2006. In situ and in vitro gene expression by Vibrio vulnificus during entry into, persistence within, and resuscitation from the viable but nonculturable state. Appl. Environ. Microbiol. 72:1445-1451. 161. Smith, B., and J. D. Oliver. 2006. In situ gene expression by Vibrio vulnificus. Appl. Environ. Microbiol. 72:2244-2246. 139 162. Stanford, S. E. 2000. Factors influence pseudodormancy and dormancy in Vibrio vulnificus. University of South Alabama, Alabama. 163. Strange, R. E., and F. A. Dark. 1962. Effect of chilling on Aerobacter aerogenes in aqueous suspension. J. Gen. Microbiol. 29:719-&. 164. Strange, R. E., and A. G. Ness. 1963. Effect of chilling on bacteria in aqueous suspension. Nature 197:819-&. 165. Strom, M. S., and R. N. Paranjpye. 2000. Epidemiology and pathogenesis of Vibrio vulnificus. Microbes Infect. 2:177-188. 166. Sun, Y., and J. D. Oliver. 1994. Antimicrobial action of some Gras compounds against Vibrio vulnificus. Food Addit. Contam. 11:549-558. 167. Sun, Y., and J. D. Oliver. 1994. Effects of Gras compounds on natural Vibrio vulnificus populations in oysters. J. Food Prot. 57:921-923. 168. Sun, Y., and J. D. Oliver. 1995. Hot sauce - no elimination of Vibrio vulnificus in oysters. J. Food Prot. 58:441-442. 169. Tamplin, M., G. E. Rodrick, N. J. Blake, and T. Cuba. 1982. Isolation and characterization of Vibrio vulnificus from two Florida estuaries. Appl. Environ. Microbiol. 44:1466-1470. 170. Tamplin, M. L., and G. M. Capers. 1992. Persistence of Vibrio vulnificus in tissues of Gulf Coast oysters, Crassostrea virginica, exposed to seawater disinfected with UV light. Appl. Environ. Microbiol. 58:1506-1510. 171. Tamplin, M. L., A. L. Martin, A. D. Ruple, D. W. Cook, and C. W. Kaspar. 1991. Enzyme immunoassay for identification of Vibrio vulnificus in seawater, sediment, and oysters. Appl. Environ. Microbiol. 57:1235-1240. 140 172. Tarnuzzer, R. W. 1998. Design and use of an exogenously added competitive RNA multiplex template for measurement of mRNA levels by RT-PCR. Eaton Publishing, US. 173. Thieringer, H. A., P. G. Jones, and M. Inouye. 1998. Cold shock and adaptation. Bioessays 20:49-57. 174. Triggsraine, B. L., and P. C. Loewen. 1987. Physical characterization of katG, encoding catalase Hpi of Escherichia coli. Gene 52:121-128. 175. Ulusarac, O., and E. Carter. 2004. Varied clinical presentations of Vibrio vulnificus infections: A report of four unusual cases and review of the literature. South. Med. J. 97:163-168. 176. USFDA 2004, posting date. Bacteriological Analytical Manual Online: Chapter 9 Vibrio. [Online.] 177. USFDA 2001, posting date. Bacteriological analytical manual online: Chapter 24 Identification of foodborne bacterial pathogens by gene probes. [Online.] 178. Vargas, M. D., S. Encarnacion, A. Davalos, A. Reyes-Perez, Y. Mora, A. Garcia-de los Santos, S. Brom, and J. Moral. 2003. Only one catalase, katG, is detectable in Rhizobium etli, and is encoded along with the regulator OxyR on a plasmid replicon. Microbiology-Sgm 149:1165-1176. 179. Veres, G., R. A. Gibbs, S. E. Scherer, and C. T. Caskey. 1987. The molecular- basis of the sparse fur mouse mutation. Science 237:415-417. 180. Wang, N., K. Yamanaka, and M. Inouye. 1999. CspI, the ninth member of the CspA family of Escherichia coli, is induced upon cold shock. J. Bacteriol. 181:1603-1609. 141 181. Weichart, D., J. D. Oliver, and S. Kjelleberg. 1992. Low-temperature induced nonculturability and killing of Vibrio vulnificus. FEMS Microbiol. Lett. 100:205- 210. 182. Whitesides, M., and J. Oliver. 1997. Resuscitation of Vibrio vulnificus from the viable but nonculturable state. Appl. Environ. Microbiol. 63:1002-1005. 183. WHO. 2005. Risk assessment of Vibrio vulnificus, MRA series 8, vol. Microbiological risk assessment series, no8. 184. Whyte, L. G., and W. E. Inniss. 1992. Cold shock proteins and cold-acclimation proteins in a psychrotrophic bacterium. Can. J. Microbiol. 38:1281-1285. 185. Wolf, P. W., and J. D. Oliver. 1992. Temperature effects on the viable but non- culturable state of Vibrio vulnificus. FEMS Microbiol Ecol 101:33-39. 186. Wouters, J. A., B. Jeynov, F. M. Rombouts, W. M. de Vos, O. P. Kuipers, and T. Abee. 1999. Analysis of the role of 7 kDa cold-shock proteins of Lactococcus lactis MG1363 in cryoprotection. Microbiology-Uk 145:3185-3194. 187. Wright, A., R. Hill, J. Johnson, M. Roghman, R. Colwell, and J. Morris, Jr. 1996. Distribution of Vibrio vulnificus in the Chesapeake Bay. Appl. Environ. Microbiol. 62:717-724. 188. Wright, A. C., G. A. Miceli, W. L. Landry, J. B. Christy, W. D. Watkins, and J. G. Morris, Jr. 1993. Rapid identification of Vibrio vulnificus on nonselective media with an alkaline phosphatase-labeled oligonucleotide probe. Appl. Environ. Microbiol. 59:541-546. 189. Wu, G., M.J. Welsh, P.B. Kaufman, and H.H. Zhang. 2004. Gene biotechnology(2nd ed.). CRC Press, Washington, D.C. 142 190. Yamanaka, K., L. Fang, and M. Inouye. 1998. The CspA family in Escherichia coli: multiple gene duplication for stress adaptation. Mol. Microbiol. 27:247-255. 191. Yaron, S., and K. R. Matthews. 2002. A reverse transcriptase-polymerase chain reaction assay for detection of viable Escherichia coli O157 : H7: investigation of specific target genes. J. Appl. Microbiol. 92:633-640. 192. Yildiz, F. H., and G. K. Schoolnik. 1998. Role of rpoS in stress survival and virulence of Vibrio cholerae. J. Bacteriol. 180:773-784. 193. Zheng, M., F. Aslund, and G. Storz. 1998. Activation of the OxyR transcription factor by reversible disulfide bond formation. Science 279:1718-1721. 143 APPENDICES 144 0 0.05 0.1 0.15 0.2 0.25 024681012 log C FU/ml Ab s o r b a n c e Appendix 1. The correlation between the numbers of V. vulnificus (log CFU/ml) and the absorbance of V. vulnificus at 450 nm. 145 1- M S G K M T G T V K W F N E T K G atg tct ggt aaa atg act ggt aca gta aaa tgg ttc aac gag act aaa ggt -1 18- F G F I S Q D N G G K D V F V H F ttt ggt ttt att tct caa gac aac ggc ggc aaa gat gta ttc gta cac ttc -52 35- R A I V S E G F K T L A E G Q K V cgt gct atc gtt tct gaa ggc ttc aaa act ctt gct gaa ggt cag aaa gta -103 52- S F N V E E G Q K G P Q A S N V T agc ttc aac gtt gaa gaa ggt caa aaa ggc cca caa gct tct aac gtt aca -154 69- V I * gtt atc taa -205 Appendix 2. Nucleotide and deduced amino acid sequence of cold shock protein V. vulnificus CMCP6 (GenBank accession number VV12757 and NP761574) and V. vulnificus YJ016 (GenBank accession number VV1504 and NP934297), designated csp1 in this study. The nucleotides are numbered on the right and the amino acids are numbered on the left. An asterisk above the last three nucleotides indicates a stop codon. 146 1- M K G K I L R W V D E R G F G F I atg aaa ggg aag att ctt cgc tgg gtt gat gaa cgt ggg ttt ggg ttc att -1 18- K S D E L D G D V F V H I S K F P aaa tct gat gag ctt gat ggc gac gta ttt gtt cat ata tcc aag ttt ccc -52 35- Q G Y R R P Q V G D H V E F H L A caa ggc tat cgt cgc cca caa gta ggt gat cat gta gag ttt cat ctg gca -103 52- N N Q P K L S A A S A R L I G V E aat aat cag cct aaa tta agt gca gca agt gct cgt tta att ggc gtg gaa -154 69- P Q K T N S L S V V I S A A V L G cct caa aaa aca aac tct ctt tca gtt gta ata tca gca gct gtc cta ggt -205 86- L L G A G L Y V F L I E P K L N P tta tta ggt gct gga tta tac gtg ttt ctt atc gaa cct aaa ctg aac cca -256 103- A Y E N M G F S C E G K T Y C S E gcg tat gaa aat atg ggt ttt agt tgt gaa ggt aaa acg tac tgt agc gaa -307 120- M V S C D E A K F Y L S N C P N V atg gtg tct tgt gac gaa gct aag ttt tat ttg tca aac tgc ccg aac gta -358 137- K I D G D R D G I P C E S Q L C G aaa ata gac ggc gac cgt gat ggc atc cct tgt gag agc cag tta tgt ggc -404 154- G W * ggc tgg tag -460 Appendix 3. Nucleotide and deduced amino acid sequence of cold shock protein V. vulnificus CMCP6 (GenBank accession number VV12119 and NP761361), designated csp2 in this study. The nucleotides are numbered on the right and the amino acids are numbered on the left. An asterisk above the last three nucleotides indicates a stop codon. 147 1- M Y S M A T G T V K W F N N A K G atg tat agc atg gct aca ggt aca gta aag tgg ttc aat aac gcc aaa gga -1 18- F G F I C S D Q E E G D I F A H Y ttt ggt ttt atc tgt tca gat caa gag gaa ggc gac atc ttc gcc cac tac -51 35- S T I Q M D G Y R T L K A G Q Q V tca acg att cag atg gac ggt tat cgt aca ctg aaa gcg ggc caa caa gtc -103 52- T Y E I E K G P K G C H A S S V V acg tat gag att gag aaa ggt ccg aaa ggg tgt cac gct agc agc gtc gtt -154 69- P L E V Q A T K * cca ctc gaa gta caa gca acg aag taa -205 Appendix 4. Nucleotide and deduced amino acid sequence of cold shock protein V. vulnificus CMCP6 (GenBank accession number VV12119 and NP760979) and V. vulnificus YJ016 (GenBank accession number VV2324 and NP935117), designated csp3 in this study. The nucleotides are numbered on the right and the amino acids are numbered on the left. An asterisk above the last three nucleotides indicates a stop codon. 148 1- M S N T V T G T V K W F N E T K G atg tct aac aca gta acc ggc acc gta aaa tgg ttt aac gaa act aaa ggc -1 18- F G F I K Q E N G P D V F A H F S ttt ggc ttc atc aag caa gaa aac ggc ccg gac gtt ttt gca cac ttc tct -51 35- A I K G D G F R T L A E G Q K V S gct atc aaa ggt gat ggc ttc cgc act ctt gct gaa ggt cag aaa gtt tct -103 52- F V I S Q G Q K G P Q A E E I T L ttc gtc atc tct caa ggc caa aaa ggt cct caa gca gaa gag atc aca ctg -154 69- L * ctt taa -205 Appendix 5. Nucleotide and deduced amino acid sequence of cold shock protein V. vulnificus CMCP6 (GenBank accession number VV20519 and NP762479) and V. vulnificus YJ016 (GenBank accession number VVA1067 and NP937123), designated csp4 in this study. The nucleotides are numbered on the right and the amino acids are numbered on the left. An asterisk above the last three nucleotides indicates a stop codon. 149 1- M S N K V T G S V K W F N E T K G atg tct aat aaa gta act ggt tca gta aaa tgg ttt aac gaa acg aaa ggt -1 18- F G F I S Q D N G G N D V F V H F ttc ggt ttt att tct caa gac aac ggt ggt aac gat gtg ttt gtt cac ttc -51 35- N S I V S T G F K T L A E G Q R V aac tct atc gtt tca aca ggt ttt aaa act ttg gca gaa ggc caa cgc gta -103 52- T F E V E Q G K K G P Q A A N V T aca ttt gaa gtt gag caa ggc aaa aaa ggc cca caa gct gca aac gta act -154 69- A L * gcg cta taa -205 Appendix 6. Nucleotide and deduced amino acid sequence of cold shock protein V. vulnificus CMCP6 (GenBank accession number VV20503 and NP762464) and V. vulnificus YJ016 (GenBank accession number VVA1052 and NP937108), designated csp5 in this study. The nucleotides are numbered on the right and the amino acids are numbered on the left. An asterisk above the last three nucleotides indicates a stop codon. 150 Appendix 7. Similarities between V. vulnificus cold shock protein obtained from V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4, designated csp1 in this study. Protein Species % identity Cold shock protein V. vulnificusCMCP6 V. vulnificusYJ016 100 Cold shock transcriptional regulator CspA V. cholerae 81 Cold shock transcriptional regulator CspA V. parahaemolyticus 80 Cold shock transcriptional regulator CspA V. alginolyticus 83 Putative cold shock-like protein Photobacterium profundum 80 Cold shock protein, DNA binding Shewanella denitrifican 80 151 Appendix 8. Similarities between V. vulnificus cold shock protein obtained from V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4, designated csp3 in this study. Protein Species % identity Cold shock-like protein CspD V. vulnificusCMCP6 V. vulnificusYJ016 98 Cold shock-like protein CspD V. parahaemolyticus 92 Cold shock-like protein CspD V. alginolyticus 94 Cold shock-like protein CspD V. cholerae 78 Cold shock-like protein CspD Vibrio sp. 80 Cold shock-like protein CspD V. splendidus 80 152 Appendix 9. Similarities between V. vulnificus cold shock protein obtained from V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4, designated csp4 in this study. Protein Species % identity Cold shock protein V. vulnificusCMCP6 V. vulnificusYJ016 98 Cold shock DNA-binding domain protein V. parahaemolyticus 86 Cold shock DNA-binding domain protein V. cholerae 86 Cold shock protein Pseudoalteromonas tunicata 81 Cold shock DNA-binding domain protein V. splendidus 81 Cold shock protein Hahella chejuensis 82 153 Appendix 10. Similarities between V. vulnificus cold shock protein obtained from V. vulnificus strain ATCC 27562, C7184Tr, VV3, and VV4, designated csp5 in this study. Protein Species % identity Cold shock protein V. vulnificusCMCP6 V. vulnificusYJ016 98 Cold shock transcriptional regulator CspD V. cholerae 87 Cold shock domain family protein V. cholerae 84 Cold shock protein CspV V. cholerae 84 Cold shock transcriptional regulator CspD V. parahaemolyticus 80 Cold shock transcriptional regulator CspD V. alginolyticus 82