ENVIRONMENTAL REGULATION OF CYCLIC DI-GMP TURNOVER IN VIBRIO CHOLERAE By Benjamin J. Koestler A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Microbiology and Molecular Genetics - Doctor of Philosophy 2014 ABSTRACT ENVIRONMENTAL REGULATION OF CYCLIC DI-GMP TURNOVER IN VIBRIO CHOLERAE By Benjamin J. Koestler 3’, 5’-cyclic diguanylic acid (c-di-GMP) is an intracellular bacterial second messenger that mediates the transition between a sessile, biofilm forming lifestyle to a motile, virulent lifestyle. Diguanylate cyclase (DGC) enzymes synthesize c-diGMP, whereas c-di-GMP specific phosphodiesterase (PDE) enzymes hydrolyze the second messenger. Although numerous bacterial behaviors are regulated by c-diGMP, the regulatory inputs of this system remain mostly undefined. Here, I examine how the marine bacterium and human pathogen Vibrio cholerae utilizes c-di-GMP signaling to interpret and respond to environmental cues. The central hypothesis that underpins my research is that V. cholerae senses environmental signals with DGCs and PDEs to modulate c-di-GMP concentrations in different environments. As c-di-GMP is a widely conserved second messenger utilized by many different bacteria, the mechanisms by which V. cholerae utilizes cdi-GMP in different environments can be applied to other bacterial systems to further comprehend how they behave in and adapt to various surroundings. To examine the influence of environmental factors on c-di-GMP synthesis and hydrolysis, I have developed a novel method named The Ex vivo Lysate c-di-GMP Assay (TELCA) that systematically measures total DGC and PDE cellular activity. I have shown that V. cholerae grown in different environments exhibits distinct intracellular levels of c-di-GMP, and using TELCA have determined that these differences correspond to changes in both c-di-GMP synthesis and hydrolysis. These findings highlight that modulation of both total DGC and PDE activity alters the intracellular concentration c-di-GMP. While searching for specific environmental cues that regulate c-di-GMP, I have found that bile acids, a prevalent constituent of the human proximal small intestine, increase intracellular c-di-GMP in V. cholerae. This bile-mediated increase in c-di-GMP is quenched by bicarbonate, the intestinal pH buffer secreted by intestinal epithelial cells. These findings lead me to propose that V. cholerae senses distinct microenvironments within the small intestine, using bile and bicarbonate as chemical cues, and responds by modulating intracellular c-di-GMP. In addition to its function as a bacterial second messenger, c-di-GMP has potent immunostimulatory properties in eukaryotes; these properties make c-diGMP a prime candidate for use as a vaccine adjuvant. Here, I present a novel method of delivering c-di-GMP into eukaryotic cells using adenovirus. I have demonstrated that c-di-GMP can be synthesized in vivo by delivering DGC DNA into the cell, and that this c-di-GMP increases the secretion of numerous cytokines and chemokines. This novel adenovirus c-di-GMP delivery system offers a more efficient and cost-effective method to administer c-di-GMP as an adjuvant to stimulate innate immunity. ACKNOWLEDGEMENTS I would like to acknowledge and thank C.M. Waters for providing mentorship and guidance. I would also like to acknowledge T.M. Schmidt, R.A. Britton, and S.H. Shiu for their guidance. I thank E.L. Reynolds and J.P. Massie for assistance in data collection (Chapter 4). I also acknowledge A. Amalfitano for guidance and S.S. Seregin, Y.A. Aldhamen, D.P.W. Rastall, S. Godbehere, J. Collins, and A.M.A. Hunt for assistance in collecting data (Chapter 5). I am grateful for assistance from L. Chen and A.D. Jones from the mass spectrometry facility. I would also like to thank K. Whitehead for providing the mixture for SHB, B.R. Pursley for constructing pBRP2, and C. Chan for constructing pANDA2. I thank S.M. Butler for help proofreading and formatting this dissertation. And finally, I would like to acknowledge the Rudolph Hugh Fellowship, the Russell B. DuVall Scholarship, and the Marvin Hensley Fellowship. iv TABLE OF CONTENTS LIST OF TABLES .................................................................................................... viii LIST OF FIGURES.....................................................................................................ix LIST OF ABBREVIATIONS .................................................................................... xii CHAPTER 1 – Background .........................................................................................1 The bacterial second messenger cyclic di-GMP ...............................................1 Vibrio cholerae: marine bacterium and human pathogen...............................3 Innate Immune response to c-di-GMP .............................................................5 Summary ...........................................................................................................8 CHAPTER 2 - Exploring Environmental Control of Cyclic di-GMP Signaling in Vibrio cholerae by Using the Ex Vivo Lysate Cyclic di-GMP Assay (TELCA)........10 Preface .............................................................................................................10 Bacterial strains and growth conditions ........................................................12 Synthesizing [13C]-c-di-GMP ..........................................................................13 Generation of lysates ......................................................................................14 Detection and quantification of c-di-GMP......................................................15 TELCA quantification of whole cell DGC activity .........................................17 TELCA quantification of whole cell PDE activity .........................................22 The local environment alters the intracellular concentration of c-di-GMP in V. cholerae .......................................................................................................23 Growth environments modulate both c-di-GMP synthesis and hydrolysis ........................................................................................................26 QS controls both total DGC and PDE activity in V. cholerae .......................29 CdgA contributes to the increased DGC activity at low cell density ............31 Discussion ........................................................................................................32 CHAPTER 3 – Bile Acids and Bicarbonate Inversely Regulate Intracellular Cyclic di-GMP in Vibrio cholerae ..............................................................................37 Preface .............................................................................................................37 Growth conditions and molecular methods ...................................................38 Detection and quantification of intracellular c-di-GMP ................................40 Systematic screen of DGC and PDE activity .................................................41 Systematic screen of DGC and PDE expression ............................................41 Biofilm quantification .....................................................................................42 Bile increases the intracellular concentration of c-di-GMP in V. cholerae .......................................................................................................43 v The bile mediated increase of intracellular c-di-GMP is growth-phase dependent ........................................................................................................44 Three DGCs are more active in the presence of bile .....................................45 The activity of PDEs in V. cholerae is not altered by bile .............................48 The three bile-inducible DGCs are inner membrane proteins ......................50 Three DGCs increase c-di-GMP synthesis in the presence of bile acids.......52 A toxR deletion mutant partially loses bile-mediated c-di-GMP increase ...53 Transcription of VC1295 is inhibited by bile acids ........................................55 Three DGCs and one HD-GYP account for bile-mediated c-di-GMP induction ..........................................................................................................59 Deletion of the bile-responsive DGCs and PDE reduces bile induction of V. cholerae biofilm formation ..............................................................................61 Bicarbonate decreases the bile-mediated increase of c-di-GMP in V. cholerae .......................................................................................................62 Discussion ........................................................................................................65 Future directions .............................................................................................71 CHAPTER 4 – Intracellular c-di-GMP Negatively Correlates with the Growth Rate and Ribosomal RNA of V. cholerae ...........................................................................76 Preface .............................................................................................................76 Measuring growth rate of V. cholerae ............................................................77 RNA extractions and quantification...............................................................78 Intracellular c-di-GMP and growth rate of V. cholerae are correlated .........78 Growth rate inhibition caused by ectopic DGC expression is dose-dependent ................................................................................................79 Intracellular c-di-GMP concentration required for growth inhibition varies for different DGCs ...........................................................................................83 Increased intracellular c-di-GMP reduces the expression of rRNA ..............86 Discussion ........................................................................................................88 Future directions .............................................................................................91 CHAPTER 5 - A novel Adenovirus Adjuvant Stimulates Innate Immunity by in vivo cyclic di-GMP Synthesis ............................................................................................95 Preface .............................................................................................................95 DNA manipulation and virus construction ....................................................96 HeLa cell transformation ................................................................................97 Animal protocols .............................................................................................98 Quantitative PCR............................................................................................99 Cytokine and chemokine quantification ......................................................100 ELISPOT analysis.........................................................................................101 C. difficile challenge ......................................................................................101 Generating an adenovirus harboring a V. cholerae DGC ...........................102 Synthesis of c-di-GMP in a murine system ..................................................105 vi C-di-GMP synthesized in vivo stimulates innate immunity in a mouse model .............................................................................................................106 C-di-GMP synthesized in vivo lowers the effective antigen dose of Ad5-TA...........................................................................................................110 Ad5-VCA0956 does not significantly enhance protection of Ad-TA against C. difficile.......................................................................................................114 Discussion ......................................................................................................119 Future directions ...........................................................................................122 CHAPTER 6 – Conclusions .....................................................................................124 APPENDICES..........................................................................................................127 REFERENCES ........................................................................................................138 vii LIST OF TABLES Table 1 Description of DGC expression strain intracellular c-di-GMP levels and growth inhibition ......................................................................... 82 Table 2 Cytokine and chemokine secretion of mice infected with Ad5 vectors after 6 hours ...................................................................................... 111 Table 3 Cytokine and chemokine secretion of mice infected with Ad5 vectors after 24h hours.................................................................................. 112 Table A1 Bacteria strain list ..................................................................................... 128 Table A2 Plasmid list ................................................................................................ 129 Table A3 Primer list .................................................................................................. 134 viii LIST OF FIGURES Figure 1 LC-MS/MS differentiation of [12C]-c-di-GMP and [13C]-c-di-GMP ....19 Figure 2 TELCA of [13C]-c-di-GMP synthesis of DGC and PDE ectopic expression strains ...............................................................................21 Figure 3 TELCA of [13C]-c-di-GMP hydrolysis of DGC and PDE ectopic expression strains ...............................................................................23 Figure 4 Growth curves of V. cholerae grown in different media....................25 Figure 5 Intracellular c-di-GMP concentrations of V. cholerae grown in different media ....................................................................................26 Figure 6 TELCA of V. cholerae grown in different media ...............................28 Figure 7 TELCA of V. cholerae QS mutants ....................................................30 Figure 8 Intracellular c-di-GMP concentrations of V. cholerae grown with different bile additives ........................................................................44 Figure 9 Intracellular c-di-GMP concentrations of V. cholerae during the course of growth with bile ..............................................................................46 Figure 10 High-throughput screen of the in vivo activity of each of the 40 V. cholerae DGCs with bile .....................................................................48 Figure 11 High-throughput screen of the in vivo activity of each of the 40 V. cholerae PDEs with bile .....................................................................50 Figure 12 The amino acid sequence hydrophobicity and predicted signaling domains of the bile-responsive DGCs ................................................51 Figure 13 Intracellular c-di-GMP concentrations of V. cholerae DGC expression strains in the presence and absence of SHB .....................................54 ix Figure 14 Intracellular c-di-GMP concentrations of V. cholerae ΔtoxR and ΔtoxT deletion mutants in the presence of bile ............................................55 Figure 15 Relative transcription of DGCs and PDEs in the presence and absence of bile ...................................................................................................57 Figure 16 Intracellular c-di-GMP concentrations of V. cholerae DGC and PDE mutant strains in the presence and absence of bile ..........................60 Figure 17 Biofilm formation of V. cholerae DGC and PDE mutants in the presence of bile ....................................................................................63 Figure 18 Intracellular c-di-GMP concentrations of V. cholerae grown in the presence of bile and bicarbonate ........................................................64 Figure 19 Proposed model of V. cholerae c-di-GMP regulation in the human small intestine ....................................................................................70 Figure 20 Linear regression of intracellular c-di-GMP concentrations on growth rates of V. cholerae grown in different media ...................................79 Figure 21 Growth rates of V. cholerae DGC expression strains........................81 Figure 22 Linear regression of intracellular c-di-GMP concentrations on growth rates of V. cholerae DGC expression strains .....................................84 Figure 23 Intracellular c-di-GMP concentrations of V. cholerae DGC expression strains .................................................................................................86 Figure 24 Growth rates of V. cholerae DGC expression strains........................87 Figure 25 Relative rRNA abundance of V. cholerae DGC expression strains ..89 Figure 26 C-di-GMP synthesized in vivo by VCA0956 in HeLa cells ..............104 Figure 27 Infection of Ad5-VCA0956 in a murine model system ....................107 Figure 28 qRT-PCR of mouse liver gene transcripts 24 hours after infection with Ad5 vectors ........................................................................................109 x Figure 29 IFNγ ELISPOT analysis of mice vaccinated with Ad5-TA and Ad5 vectors ...............................................................................................115 Figure 30 C. difficile challenge of BALB/c mice ...............................................116 Figure 31 C. difficile challenge of BALB/c mice vaccinated with Ad5-TA and Ad5-VCA0956....................................................................................118 xi LIST OF ABBREVIATIONS Ad5 ..................................................................................... Adenovirus type 5-derived AI ....................................................................................Quorum sensing autoinducer Bd ........................................................................................................ Below detection BiC ................................................................................................. Sodium bicarbonate BV ................................................................................................................ Bovine bile CDAD .......................................................................... C. difficile associated diarrhea c-di-AMP ............................................. Cyclic dimeric adenosine 3',5'-monophosphate c-di-GMP ............................................ Cyclic dimeric guanosine 3',5'-monophosphate cfu ................................................................................................. Colony forming unit CHAPS .................... 3-[(3-Cholamidopropyl)dimethylammonio]-1-propanesulfonate CV ............................................................................................................ Crystal violet CT .................................................................................................. Cholera enterotoxin DGC ................................................................................. Diguanylate cyclase enzyme GTP..................................................................................... Guanosine-5’ triphosphate h.p.i ............................................................................................... Hours post injection IM .......................................................................................................... Intramuscular IPTG ................................................................. Isopropyl β-D1-thiogalactopyranoside IV ............................................................................................................ Intravenously LB ................................................................................................ Luria-Bertani media LC-MS/MS ........... Liquid chromatography coupled with tandem mass spectrometry xii M.O.I........................................................................................ Multiplicity of infection m/z ................................................................................................ Mass to charge ratio OD......................................................................................................... Optical density ORF .............................................................................................. Open reading frame PBS ......................................................................................Phosphate buffered saline PDE ................................................................................... Phosphodiesterase enzyme ppGpp ..................................................................... Guanosine 3',5'-bispyrophosphate pppGpp ..................................................... Guanosine 3'-diphosphate, 5'-triphosphate QS ....................................................................................................... Quorum sensing R.L.U ............................................................................... Relative luminescence units rRNA ................................................................................................... Ribosomal RNA SDS .......................................................................................... Sodium dodecyl sulfate SFCs ................................................................................................. Spot forming cells SHB ............................................................................................ Synthetic human bile TELCA................................................................ The Ex vivo Lysate c-di-GMP Assay TCP ......................................................................................... Toxin co-regulated pilus TLR .................................................................................................... Toll-like receptor Tris ......................................................... 2-Amino-2-hydroxymethyl-propane-1,3-diol Vp ........................................................................................................... Viral particles WT .................................................................................................................. Wild type xiii CHAPTER 1 – Background The bacterial second messenger cyclic di-GMP Nucleotide small molecules are utilized by organisms from across the tree of life to relay signals from sensor proteins to target receptor proteins within the cell; known bacterial second messengers include cyclic adenosine 3',5'- monophosphate (cAMP), cyclic guanosine 3',5'-monophosphate (cGMP), guanosine 3',5'bispyrophosphate (ppGpp), guanosine 3'-diphosphate, 5'-triphosphate (pppGpp), cyclic dimeric adenosine 3',5'-monophosphate (c-di-AMP), and cyclic dimeric guanosine 3',5'-monophosphate (c-di-GMP). C-di-GMP was first discovered in 1985 by Benzimen et al. [1], and since has been predicted to be utilized in >75% of all bacteria in representatives from every major bacterial phyla [2]. Diguanylate cyclase enzymes (DGCs) synthesize c-di-GMP from two GTP molecules. These enzymes contain a conserved C-terminal GGDEF domain with a GG[D/E]EF active site motif and a highly variable N-terminus that often encodes conserved signal recognition domains such as GAF, PAS, or receiver domains [3-6]. Conversely, c-di-GMP specific phosphodiesterase enzymes (PDEs) hydrolyze c-diGMP and contain either a conserved C-terminal EAL or HD-GYP domain [7, 8]. Similar to DGCs, PDE domains are often linked to N-terminal signal recognition domains. C-di-GMP regulates numerous phenotypes in different bacteria (for a comprehensive review of c-di-GMP controlled phenotypes, please refer to Romling et 1 al. [9]). Motility, the ability of a bacterial cell to move, has been shown to be negatively regulated by c-di-GMP in many organisms [10-13]. Conversely, biofilm formation, the bacterial behavior where cells excrete exopolysaccharides to form a three-dimensional surface-associated community [14], is positively regulated by cdi-GMP. Biofilm-forming communities have been shown to play important roles in environmental persistence and virulence [15]. C-di-GMP has been shown to positively regulate biofilm formation in several pathogens including Pseudomonas aeruginosa [11], Salmonella enterica [11], and Vibrio cholerae [16]. The differential regulation of motility and biofilm formation by c-di-GMP indicates that c-di-GMP mediates the transition between sessile and motile, virulent lifestyles. A characteristic feature of c-di-GMP signaling networks is the relatively large number of DGCs and PDEs encoded in individual bacterial genomes that theoretically contribute to changes in the intracellular concentration of c-di-GMP. For example, the bacteria V. cholerae encodes 61 distinct proteins predicted to modulate c-di-GMP [17]. Because of this complexity, c-di-GMP signaling systems are typically redundant, and it is challenging to understand on a systems level how the concentration of c-di-GMP is controlled. Changes in c-di-GMP could be mediated by modulation of total DGC activity, PDE activity, or both. It is not well understood why some bacteria encode so many DGCs and PDEs, and the functions of the variable N-terminal domains for most of these enzymes have not been characterized. One hypothesis is that each DGC or PDE senses and responds to a 2 specific environmental signal by altering c-di-GMP synthesis or degradation activity in the C-terminal domain. A handful of environmental cues have been shown to influence intracellular c-di-GMP in various bacteria. For example, specific environmental signals that interact with DGCs and PDEs have been identified such as zinc in E. coli [18], amino acids in P. aeruginosa [19], nitric oxide in Legionella pneumophila [20], and norspermidine in V. cholerae [21]. Additionally, a DGC from Bordetella pertussis and a DGC/PDE complex from E. coli bind oxygen to alter c-di-GMP turnover [22, 23]. There is also evidence that phototrophic bacteria utilize c-di-GMP to modulate intracellular c-di-GMP in different light conditions [24, 25]. DGCs and PDEs have also been identified in Xanthomonas campestris and Vibrio parahaemolyticus that bind molecules associated with quorum sensing networks to modulate c-di-GMP turnover [8, 26, 27]. However, these enzymes with known ligands represent only a small fraction of the known DGCs and PDEs, and the environmental signals recognized by the vast majority of DGCs and PDEs remain unidentified. Vibrio cholerae: marine bacterium and human pathogen One bacterium in which c-di-GMP signaling has been studied is V. cholerae. V. cholerae, a gram-negative marine bacterium, is the causative agent of the human diarrheal disease cholera. In aquatic environments, V. cholerae favors a sessile lifestyle by preferentially forming biofilms on chitineous surfaces [28-30]. Upon ingestion into a human host, V. cholerae expresses virulence and colonization 3 factors that lead to severe diarrhea, subsequently reseeding the bacteria back into the environment. The pathogenesis of V. cholerae is dependent on two virulence factors, cholera enterotoxin (CT) and the toxin c-regulated pilus (TCP) [31]. The expression of both CT and TCP is controlled by the transcriptional regulators ToxR/S, which coordinate with the transcriptional regulator ToxT to optimize virulence expression in the human host [32, 33]. The transition of V. cholerae between aquatic environments and the human host is mediated by c-di-GMP, which regulates biofilm formation, motility, and virulence gene expression [16, 34-36]. It has been proposed that the intracellular cdi-GMP concentration in V. cholerae is relatively high in aquatic environments leading to a sessile biofilm forming lifestyle, whereas in the human host, reduced cdi-GMP concentrations stimulate virulence factor expression [34]. Due to the technical limitations of measuring intracellular c-di-GMP, this model has not been directly tested, and the specific environmental cues that regulate the transition between these two niches remain unknown. Upon entering the human host, it is imperative that V. cholerae recognizes environmental signals to mediate the transition from an aquatic bacterium to a human pathogen. One signal prevalent in the small intestine is bile. Bile is an antimicrobial substance secreted from the liver into the proximal small intestine that aids in digestion by emulsifying lipids. The composition of secreted bile is heterogeneous and includes inorganic salts, cholesterol, phospholipids, pigments, and bile acids [37]. Bile acids are derived from cholesterol in the liver and are 4 processed by hepatic cells to produce a mixture dominated by taurine and glycine conjugates [37, 38]. Bile acids have detergent properties that enable the interaction between bile and digestive lipids. For a complete review on the properties of bile, refer to Begley et al. [39]. The interplay between enteric bacteria and bile is extensive. In bacteria, there is evidence that bile causes oxidative stress, DNA damage, and perturbs the cell membrane [40-42]. To counteract this stress, V. cholerae upregulates the porin protein ompU in a toxR-dependent manner to increase bile resistance [43]. V. cholerae also employs six resistance-nodulation-division family efflux pumps that have been implicated in bile resistance [44-46]. Additionally, V. cholerae increases biofilm formation in a vpsR dependent manner in the presence of bile acids, presumably to increase resistance to the deleterious effects of bile [47]. It is clear that the sensing of bile by V. cholerae leads to distinct physiological changes, but a connection between bile and c-di-GMP has not been described in V. cholerae or any other bacterium. Innate Immune response to c-di-GMP In addition to its role regulatory role as a bacterial second messenger, c-diGMP is a potent stimulator of innate immunity in eukaryotic organisms. Presumably, as c-di-GMP is an intracellular second messenger unique to bacteria, it is recognized by eukaryotic organisms as a signal of bacterial invasion. This stimulation of innate immunity occurs at least in part through the eukaryotic 5 protein STING, which senses pathogen derived nucleic acids in the cytoplasm and subsequently activates a signaling cascade to stimulate a type-I interferon response [48, 49]. Studies show that the presence of c-di-GMP can trigger the production of IL-2, IL-4, IL-5, IL-6, IL-8, IL-12p40, IL-17, IP-10, TNF-α, KC, MIP-1α, MIP-1β, MIP-2, MCP-1, RANTES, IFN-β, IFN-γ, stimulate the NLRP3 inflammasome pathway, and promote the recruitment and activation of macrophages, NK-cells, αβ conventional T-cells, and enhance DC maturation [49-57]. Furthermore, in vivo studies have shown that co-administration of purified c-di-GMP with an antigen confers increased protection of animals in several different murine challenge models, including those utilizing Staphylococcus aureus, Klebsiella pneumoniae, and Streptococcus pneumoniae [53-55, 58]. Because c-di-GMP activates a robust immune response, there has been ongoing focus on using c-di-GMP as an adjuvant to improve vaccine efficacy [59]. Adjuvants are compounds administered alongside vaccine antigens for the purpose of enhancing the longevity, potency, or reducing the effective dose of the antigen without introducing toxic side effects. This is accomplished by stimulating the innate arm of the immune system, resulting in increased cytokine and chemokine production and upregulation of proinflammatory genes [60], which then enhances antigen recognition and response [61]. The development of novel adjuvants may be critical to the success of vaccines targeting diseases for which vaccinations have previously failed such as Clostridium difficile, human immunodeficiency virus, malaria, and cancer. Despite the demand, currently there are few adjuvants 6 approved for human use. The most commonly used adjuvants are aluminum-salt (Alum)-based; however, these adjuvants suffer drawbacks including local reactions to administration, inadequate T-cell responses, allergic IgE-type responses, and are ineffective with specific types of antigens [62]. Other less-commonly utilized adjuvants include oil and water emulsions, lipopolysacharide derivatives, selfassembling viral nanoparticles, and cholera toxin B subunit [62]. While each adjuvant offers different advantages and disadvantages, there is a large demand for novel adjuvants that can be paired with and improve vaccine antigens. Because cdi-GMP has effective immunostimulatory properties, it makes an ideal candidate for a novel adjuvant [59]. Despite the properties that make c-di-GMP a promising adjuvant, there are currently practical limitations to using c-di-GMP as a vaccine adjuvant. The studies examining the adjuvant properties of c-di-GMP utilized mucosal, intramuscular (IM), or systemic administration of chemically synthesized c-di-GMP [53-55, 58]. While this method has been shown to effectively activate innate immunity, it lacks efficiency and the mechanism by which c-di-GMP enters the cytoplasm of cells to stimulate STING is unknown. Furthermore, chemical synthesis of pharmaceutical grade c-di-GMP in large quantities is currently cost prohibitive. There is a need for an efficient and cost effective method of delivering c-di-GMP in vivo for use as an adjuvant. 7 Summary The broad goal of this study is to understand how the model bacterium V. cholerae uses c-di-GMP signaling to interpret and respond to changing environmental signals. The central hypothesis that underlies my research is that V. cholerae senses environmental signals with different DGC and PDE proteins to modulate c-di-GMP concentrations in different environments. The rationale for this study is that c-di-GMP is a conserved second messenger utilized by a broad spectrum of bacteria, and the mechanisms by which V. cholerae utilizes c-di-GMP signaling in different environments can be applied to other bacterial systems to further comprehend how they similarly respond and adapt to their environments. Furthermore, as c-di-GMP has potent immunostimulatory properties, knowledge of this signaling system can be appropriated to develop new applications for clinical settings. To show that a changing environment impacts c-di-GMP signaling globally, I have demonstrated that intracellular c-di-GMP concentrations of V. cholerae vary when V. cholerae is grown in different environments. I developed The Ex Vivo Lysate Cyclic di-GMP Assay (TELCA) to examine total cellular DGC and PDE activity, and used TELCA to show that these variations of c-di-GMP concentrations in different environments are due to changes in both c-di-GMP synthesis and hydrolysis. Using TELCA, I was also able to show that quorum sensing regulation of c-di-GMP levels in V. cholerae is due primarily to decreased DGC activity during transitions from low- to high-cell-density. 8 I have also begun to explore specific environmental cues that modulate c-diGMP in V. cholerae, and I have shown that bile acids increase intracellular c-diGMP in V. cholerae. I determined that this increase is due to multiple factors including altered c-di-GMP synthesis activity of three DGCs and reduced expression of one PDE. Additionally, this bile-mediated increase in intracellular c-di-GMP is quenched by bicarbonate, another common component of the human digestive tract. Finally, I have applied the knowledge I have gained regarding active DGCs of V. cholerae to develop a novel adenovirus adjuvant that delivers DGC DNA into eukaryotic cells, resulting in c-di-GMP synthesis in vivo and a robust innate immune response. Additionally, this adenovirus construct harboring a DGC gene is able to reduce the effective dose of a C. difficile vaccine antigen to stimulate IFNγ production of splenocytes. This novel method of in vivo c-di-GMP synthesis offers a more efficient and cost-effective alternative to using c-di-GMP as a vaccine adjuvant. 9 CHAPTER 2 - Exploring Environmental Control of Cyclic didi-GMP Signaling in Vibrio cholerae by Using the Ex Vivo Lysate Cyclic didi-GMP Assay (TELCA) Preface Quorum sensing (QS), the process of bacterial cell-cell communication mediated by AIs that are constitutively secreted and accumulate with cell density, is a fundamental chemical signaling system in bacteria. Sensing of AIs by bacteria provides information on the local population density and composition to appropriately regulate bacterial behaviors such as competence, biofilm formation, and virulence [63-68]. In V. cholerae, QS is mediated by two AI molecules (CAI-1 and AI-2) that are produced in tandem [67]. In the low-cell-density state, AI concentrations are low and their cognate receptors function as kinases ultimately leading to phosphorylation of the response regulator LuxO. Phosphorylated LuxO upregulates the transcription of four small regulatory RNAs, which subsequently suppress translation of the master QS regulator HapR [69, 70]. When the concentration of AIs increase, they bind the AI receptors switching them to phosphatases, leading to dephosphorylation of LuxO, deactivation of the sRNAs, and induction of HapR expression [71]. In V. cholerae, biofilm formation is repressed in the high-cell-density QS state via direct repression of extracellular polysaccharide synthesis genes by HapR [66, 72]. Both the QS and c-di-GMP signaling systems are inextricably linked in V. cholerae [73]. At low cell density, the intracellular c-di-GMP concentration of V. 10 cholerae is relatively high, promoting biofilm formation and inhibiting motility. Upon transition to the high-cell-density state the intracellular c-di-GMP concentration is suppressed, decreasing biofilm formation [66, 72]. In V. cholerae the transcription of 18 of the 61 predicted c-di-GMP turnover enzymes are differentially regulated at low and high cell density [66, 72, 74]. However, because of the complexity of this signaling system it remains unclear if the QS-mediated changes in c-di-GMP levels in V. cholerae are being driven by changes in c-di-GMP synthesis, hydrolysis or both. To further the exploration of the environmental modulation of c-di-GMP and the regulatory connections between QS, c-di-GMP, and biofilm formation in V. cholerae, I have developed a novel approach, hereby named The Ex vivo Lysate c-diGMP Assay (TELCA), which enables the quantification of c-di-GMP synthesis and hydrolysis of a whole-cell lysate. I have shown that growing V. cholerae in divergent environments can lead to a 20-fold difference in the intracellular concentration of cdi-GMP, and these distinct c-di-GMP concentrations can be attributed to both changes in c-di-GMP synthesis and hydrolysis. I also used TELCA to show that alteration of both c-di-GMP synthesis and hydrolysis between the low- and high-cell density QS states contributes to the QS-mediated change of intracellular c-di-GMP. I further show that deletion of the DGC CdgA, which is transcriptionally regulated by HapR and has been predicted to be a key DGC in the induction of biofilm formation [72, 74-76], significantly reduces the elevated DGC activity associated with low cell density, highlighting the sensitivity of TELCA. TELCA provides a new 11 method to study complex c-di-GMP signaling networks from a systems perspective. The findings of this chapter have been published in the Journal of Applied and Environmental Microbiology [77]. Bacterial strains and growth conditions The V. cholerae El Tor biotype strain C6706str2 was used for all experiments [78]. All strains, plasmids, and primers used are listed in Tables A1-3. The construction of the expression plasmids for VCA0956 and VC1086 are described elsewhere [72, 79]. Unless otherwise specified, cultures were grown in LuriaBertani media (LB, 1.0% tryptone, 0.5% yeast extract, 1.0% NaCl, Accumedia) at 35°C with shaking at 220 rpm. MOPS media was purchased from Teknova and prepared as per manufacturer’s instruction. The composition of AKI, AB, ASW, and M9 media are described elsewhere [80-83]. MOPS, AB, and M9 were supplemented with glucose as a carbon source (0.2% for MOPS and AB, 0.4% for M9). For experiments involving the expression of DGCs or PDEs from plasmids, media was supplemented with kanamycin (Sigma) at 100 µg/mL and the inducer isopropyl βD1-thiogalactopyranoside (IPTG) at 0.1 mM. The construction of the ∆hapR and ∆luxO strains has been described previously [67, 68]. To generate the ∆cdgA strain, natural transformation and homologous recombination were used. Briefly, Phusion DNA polymerase (NEB) was used to amplify approximately 500 bp upstream and downstream of cdgA (VCA0074) using the primers VCA0047KO-1 (upstream) and VCA0047KO-2 12 (downstream), and VCA0047KO-3 (upstream) and VCA0047KO-4 (downstream) (Table A3). Fusion PCR was then used to join these products upstream and downstream of a chloramphenicol resistance cassette (cat) bordered by FLP recombination target (FRT) sites from the plasmid pKD3 [84]. This DNA was introduced to V. cholerae by inducing natural transformation using a modified protocol from Marvig and Blokesch [85]. V. cholerae was grown on chitin flakes in modified M9 media [82]. This media (M9-T) contained the standard concentration of M9 salts supplemented with 5 mM CaCl2, 3.6 mM MgSO4, and 4.5 mM N-acetyl-Dglucosamine as a carbon source and was balanced to a pH of 6.0. Concentrated DNA was introduced to the culture, and homologous recombination events were selected by growing the culture on LB supplemented with 1 µg/mL chloramphenicol. The cat gene was excised by ectopically expressing the FLP recombinase from the vector pTL17 [86]. Synthesizing [13C]-c-di-GMP [13C]-c-di-GMP was synthesized from guanosine-[13C]-5’ triphosphate ([13C]GTP, Sigma) using the purified DGC WspR-R242A as previously described [87]. The reaction consisted of 56 µM WspR-R424A, 20 µM [13C]-GTP, 24 mM Tris-Cl, 5 mM MgCl2, 45 mM NaCl, and EnzCheck Phosphate Assay Kit reaction buffer mix (Invitrogen) at a total volume of 100 µL, hypothetically yielding 10 µM [13C]-c-diGMP (two GTP molecules per molecule c-di-GMP). The reaction was run overnight followed by treatment for 30 minutes at 37°C with Antarctic phosphatase (NEB) in its corresponding reaction buffer to remove any excess [13C]-GTP. The reaction was 13 heated at 100°C for 10 minutes to inactivate the Antarctic phosphatase and precipitate protein. After cooling to room temperature, the reaction was centrifuged at max speed for 10 minutes to remove the precipitated protein, and the supernatants were removed and stored at -20°C until use. [13C]-c-di-GMP synthesis was confirmed using liquid chromatography with mass spectrometry as described below. Generation of lysates Cell lysates for TELCA were generated by inoculating overnight cultures of V. cholerae 1:1000 into 30 mL of the specified media with the appropriate antibiotics or inducer. Cultures grown to the desired OD600 as indicated were centrifuged at 7800 rpm to pellet the cells. The pellets were then resuspended in 15 mL TELCA reaction buffer, which consisted of 25 mM Tris, 5 mM MgCl2, and 5 mM NaCl balanced to a pH of 7.5. The cultures were then processed three times using a M110-P processor (Microfluidics) at 20,000 psi producing a clear lysate. A protein assay (Bio-Rad) was performed according to the manufacturer’s instructions to quantify the total protein content, and each lysate was diluted to 0.1 mg/mL in TELCA reaction buffer. To measure c-di-GMP synthesis, 100 mM [13C]-GTP stock was added to a final concentration of 1 mM (10 µL [13C]-GTP stock in 990 µL lysate diluent for timecourse experiments, 1 µL [13C]-GTP stock in 99 µL lysate diluent for endpoint experiments). To measure c-di-GMP hydrolysis, 10 µM [13C]-c-di-GMP stock was added to a final concentration of 100 nM (10 µL [13C]-c-di-GMP stock in 990 µL lysate diluent for timecourse experiments, 1 µL [13C]-c-di-GMP stock in 99 14 µL lysate diluent for endpoint experiments). The concentrations are within the physiological ranges of the natural concentrations of GTP and c-di-GMP found in a bacterial cell [79, 88]. The reactions were then mixed and incubated at 37°C until the experimental endpoint, and then the reaction was stopped by mixing 100 µL reaction to 100 µL of phenol:chloroform:isoamyl alcohol (25:24:1). The mixture was centrifuged at max speed for 5 minutes, and then the aqueous phase was removed and stored at -80°C until analysis. Detection and quantification of c-di-GMP For the quantification of intracellular c-di-GMP, overnight cultures grown in the corresponding media were used to inoculate 2 mL cultures (1:1000). Cultures were grown to the appropriate OD600 (OD600, AKI: 0.5-1.0; LB: 0.5-0.8; ASW: 0.2-0.4; MOPS: 0.4-0.8; AB: 0.3-0.7; M9: 0.2-0.4), and a 1.5 mL aliquot of culture was centrifuged at maximum speed for 30 seconds. The supernatant was removed, and the pellet was resuspended in 100 µL of cold extraction buffer (40% acetonitrile/40% methanol/0.1 N formic acid). The slurry was incubated at -20° C for 20 minutes, and then the insoluble fraction was pelleted by centrifugation for 10 minutes at maximum speed. The supernatant was collected and stored at -80°C. Prior to analysis, the extraction buffer was evaporated using a vacuum manifold, and the pellet was resuspended in 100 µL water. [12C]-c-di-GMP and [13C]-c-di-GMP was quantified using an Acquity Ultra Performance liquid chromatography system (Waters) coupled with a Quattro Premier XE mass spectrometer (Waters) as previously described [79]. Chromatographic separation of 10 µL sample was 15 accomplished using reverse phase separation on a Waters BEH C18 2.1 x 50 mm column and a flow rate of 0.3 mL/minute, using the following gradient: solvent A (10 mM tributylamine and 15 mM acetic acid in 97:3 water:methanol) to solvent B (methanol), 0 minutes; A=99%:B=1%, 2.5 minutes; A=80%:B=20%, 7.0 minutes; A=35:B=65%, 7.5 minutes; A=5%:B=95%, 9.01 minutes; 99%:B=1%, 10 minutes; gradient ends. After separation, the detection of c-di-GMP was carried out using electrospray ionization with multiple reaction monitoring in negative ion mode where [13C]-c-di-GMP was detected with a m/z of 709.16 -> 354.31 and [12C]-c-diGMP with a m/z of 689.16 -> 344.31. The MS detection was optimized using the following parameters: capillary voltage, 3.5 KV; cone voltage, 50 V; collision energy, 34 V; source temperature, 120°C; desolvation temperature, 350°C; cone gas flow, 50 L/h (nitrogen), 800 L/h; collision gas flow (nitrogen), 0.15 mL/min; and multiplier voltage, 650 V. The concentration of c-di-GMP was determined by generating an 8point standard curve (1:2 dilutions) of chemically synthesized c-di-GMP (Biolog) ranging from 1.9 nM to 250 nM. The intracellular c-di-GMP concentration was calculated by dividing the c-di-GMP concentration of the sample by the total intracellular volume of bacteria extracted. This was estimated by multiplying the average cell volume of the bacterial cell, determined by measuring individual cell dimensions using differential image contrast microscopy, and the number of bacterial cells in the extract. 16 TELCA quantification of whole cell DGC activity I sought to develop a systems approach to examine if changes in the levels of c-di-GMP in V. cholerae could be attributed to the modulation of net DGC and/or PDE activity. The rationale for developing this method is that current approaches to directly measure c-di-GMP primarily utilize liquid chromatography coupled with tandem mass spectrometry (LC-MS/MS) of cell extracts [89-91] or an in vivo FRET biosensor to quantify c-di-GMP at the single cell level [92]. While these methods allow accurate quantification of the average intracellular c-di-GMP in the cell, they cannot distinguish if differences in the intracellular c-di-GMP are being driven by changes in c-di-GMP synthesis or hydrolysis. Consider a case where the levels of cdi-GMP are elevated by growth in a different environment. This increase in c-diGMP can be due to increased DGC activity, decreased PDE activity, or modulation of both. One cannot predict a priori solely based on the concentration of c-di-GMP how this change is mediated. My method, named TELCA, quantifies total DGC or PDE activity of cell lysates using [13C]-containing substrates to allow the differentiation of newly synthesized or degraded c-di-GMP from natural occurring cdi-GMP by mass spectrometry. The substrate for the DGC assay measures synthesis of [13C]-c-di-GMP from exogenous [13C]-GTP whereas the PDE assay monitors the degradation of exogenous [13C]-c-di-GMP. To assess how effectively TELCA measures cellular DGC activity, extracts were generated from cultures of wild type (WT) V. cholerae ectopically expressing an active DGC (VCA0956), an active PDE (VC1086), or the vector control. Note that 17 WT V. cholerae has a relatively low intracellular concentration of c-di-GMP at the growth state used for these experiments (mid-exponential growth, OD600 of 0.5-0.8) due to QS mediated reduction in c-di-GMP levels at high cell density. I hypothesized that lysates produced from cells expressing VCA0956 would more rapidly synthesize [13C]-c-di-GMP from [13C]-GTP than lysates with the empty vector control or overproducing the PDE VC1086. The cells from all three cultures were harvested and lysed using a Microfluidics M110-P Processor, which exposes the cells to high shear and impact force resulting in a uniform lysate. This process causes minimum disruption of protein conformation and thus is well suited to generate lysates for this assay. [13C]-GTP was added to 0.1 mg/mL total protein from each extract to a final concentration of 1.0 mM, and aliquots of the reaction were removed at specified timepoints followed by quenching with a phenolchloroform extraction. The concentration of newly synthesized [13C]-c-di-GMP was quantified using LC-MS/MS. To differentiate [13C]-c-di-GMP from naturally occurring [12C]-c-di-GMP, I measured a mass to charge ratio (m/z) of 709.16 > 354.31 to detect [13C]-c-di-GMP and 689.16 > 344.31 for [12C]-c-di-GMP as previously described (Fig. 1A) [79]. The TELCA reaction generated a signal at the same retention time of chemically synthesized [12C]-c-di-GMP standards at a m/z of 709.16 > 354.31 with no signal at 689.16 > 344.31, indicating that this detection method is sufficient to discriminate between these two carbon isotope forms of c-diGMP (Fig. 1B). 18 Figure 1. LCLC-MS/MS differentiation differentiation of [12C]C]-c-didi-GMP and [13C]C]-c-didi-GMP. (A) Molecular structure of c-di-GMP with the m/z used for detection and quantification. The top m/z describes the half of the molecule observed after fragmentation (as indicated by the line) while the bottom m/z represent the entire molecule. The values for both [12C]- and [13C]-c-di-GMP are indicated. (B) The spectrum of [13C]-cdi-GMP. The top contains a peak at 5.4 minutes with a m/z of 709.16 > 354.31, indicating the presence of [13C]-c-di-GMP. The bottom contains no peak at 5.4 minutes with a m/z of 689.16 > 354.31, indicating that no [12C]-c-di-GMP is present. 19 A non-linear regression with a one-phase decay model (Graphpad Prism) was used to analyze the increase in [13C]-c-di-GMP concentrations over time. I constrained the maximum yield of the model to 0.5 mM [13C]-c-di-GMP as this is the maximum theoretical yield of the reaction. The lysate from V. cholerae expressing the DGC produced significantly more (two tailed Student’s T-Test, p < 0.05) [13C]-cdi-GMP over time, yielding 10.2 fold and 17.1 fold more c-di-GMP than the vector control and PDE lysates, respectively, after 120 minutes (Fig. 2). Furthermore, the rate constant for the DGC lysate, which represents the increase of c-di-GMP synthesis per unit time expressed as inverse minutes (6.9±1.1x10-6 min-1, R2=0.76) was significantly different (sum-of-squares F-test, p < 0.05) than those of the vector control lysate (5.9±0.5x10-7 min-1, R2=0.90) and the PDE lysate (4.1±0.4x10-7 min-1, R2=0.87). This result indicates that TELCA is able to differentiate the cellular DGC activity of lysates containing a higher concentration of active DGCs. Furthermore, I found that lysates generated from V. cholerae expressing the PDE produced 1.7 fold less [13C]-c-di-GMP than the vector control lysate after 120 minutes and exhibited a lower rate constant, indicating that additional PDEs in the lysate can reduce the net c-di-GMP synthesis by hydrolyzing [13C]-c-di-GMP post synthesis. Of note, [13C]-c-di-GMP accumulated in all three lysates including the PDE overexpression strain suggesting that given excess GTP, DGC activity supersedes PDE activity in these conditions over the time course examined. A subset of DGCs exhibit feedback inhibition by an allosteric RXXD inhibition site in close proximity to the DGC active site [10]. VCA0956 does not have a predicted inhibition site; 20 600 [ 13 C]-c-di- GMP (nM) Vector PDE 400 DGC 200 0 1 10 100 Time (minutes) C]--c-di di--GMP synthesis of DGC and PDE ectopic expression Figure 2. TELCA of [13C] strains. TELCA was applied to lysates of V. cholerae strains ectopically expressing an active DGC (VCA956, solid squares), an active PDE (VC1086, empty squares), and an empty vector control (empty circles). [13C]-GTP was added to each lysate and [13C]-c-di-GMP synthesis was tracked over time using LC-MS/MS. A non-linear regression using a two-phase exponential decay model was applied to each lysate. however, it is possible that the c-di-GMP synthesis of another DGC subject to feedback inhibition would have an altered synthesis profile. Additionally, it should be noted that I was not able to detect any hybrid c-di-GMP composed of one [12C]GTP and one [13C]-c-di-GMP (detection m/z of 709.16 > 344.31), indicating that the native [12C]-GTP is depleted upon addition of [13C]-GTP. Furthermore, there was no [12C]-c-di-GMP detected in the vector control lysate or the PDE lysate at any timepoint. However, in the DGC lysate, I did detect [12C]-c-di-GMP in low abundance alongside synthesized [13C]-c-di-GMP suggesting that if [12C]-GTP is to 21 be used in lieu of [13C]-GTP, additional measures would be required to compensate for the existence of native c-di-GMP in certain strains. TELCA quantification of whole cell PDE activity I performed a similar experiment to determine if TELCA is effective at differentiating the PDE activity of different bacterial strains; however, instead of using [13C]-GTP as the initial substrate, [13C]-c-di-GMP was added. To generate [13C]-c-di-GMP, a c-di-GMP synthesis reaction was performed using [13C]-GTP and the DGC WspR that contained a mutation in the RXXD allosteric inhibition site (R242A, [87]). The remaining labeled GTP was removed by treating the reaction with Antarctic phosphatase, although quantification of enzymatically synthesized [13C]-c-di-GMP using LC-MS/MS indicated that the reaction went to completion, generating the expected hypothetical yield. I observed that the PDE lysate more rapidly degraded [13C]-c-di-GMP than both the vector control lysate and the DGC lysate (Fig. 3), as the PDE lysate retained an average of 20.5% [13C]-c-di-GMP after 120 minutes whereas the vector control lysate and the DGC lysate retained averages of 69.6% and 73.9% of the [13C]-c-di-GMP, respectively. The difference between the PDE lysate and both the vector control and DGC lysates were significantly different (two tailed Student’s T-Test, p < 0.05). Likewise, a non-linear regression with a one-phase decay model with a plateau constraint of zero was applied to the data, and I found that the rate constant of the PDE lysate (1.9±0.6x10-2 min-1, R2=0.66) was significantly higher (sum-of-squares F-test, p < 0.05) than both the vector control lysate (2.5±1.0x10-3 min-1, R2=0.35) and the DGC 22 lysate (2.0±0.9x10-3 min-1, R2=0.27). As expected, there were no notable differences between the DGC-expressing lysate and the vector control lysate suggesting that a difference in DGC activity does not impact the PDE activity. From these data, I conclude that TELCA is sufficient to differentiate both the c-di-GMP synthesis and hydrolysis potential of whole-cell lysates. [ 13 C]-c-di-GMP (nM) 100 Vector PDE DGC 50 0 1 10 100 Time (minutes) Figure 3. TELCA of [13C]C]-c-didi-GMP hydrolysis of DGC and PDE ectopic expression strains. TELCA was applied to lysates of V. cholerae strains ectopically expressing an active DGC (VCA956, solid squares), an active PDE (VC1086, empty squares), and an empty vector control (empty circles). [13C]-c-di-GMP was added to each lysate and the hydrolysis was tracked over time using LC-MS/MS. A non-linear regression using a two-phase exponential decay model was applied to each lysate. The local environment alters the intracellular concentration of c-di-GMP in V. cholerae How the extracellular environment impacts the synthesis and degradation of intracellular c-di-GMP is poorly understood. It has been proposed that extracellular 23 stimuli, such as oxygen, light, amino acids, and AIs, regulate intracellular c-di-GMP by altering the activity of DGCs and PDEs [8, 19-24, 26, 27]. Also, many DGCs and PDEs have domains predicted to be involved in sensing extracellular ligands [93]. I hypothesized that V. cholerae growing in different environments would exhibit distinct concentrations of c-di-GMP due to alterations in the environmental inputs controlling DGC and PDE activity. To test this hypothesis, I quantified the intracellular c-di-GMP concentration of V. cholerae in six different growth media using LC-MS/MS. For the purposes of this study, I consider AKI [80], LB, and ASW complex, carbon rich media, while MOPS, AB [81] and M9 [82] are considered defined, carbon poor media. I sampled from each media during the latter half of steady-state growth, which I define as 1-2 doublings from stationary phase. This time point was determined by generating growth curves for V. cholerae in each different media (Fig. 4). I found that V. cholerae demonstrated differences of up to 20-fold in the intracellular concentration of c-di-GMP when grown in these different media (Fig. 5). Consistent with prior studies, I measured the concentration of c-di-GMP in LB to be 101.2±52.7 nM [79]. The complex, carbon rich media AKI also produced relatively low intracellular c-di-GMP that was statistically indistinguishable from LB, measuring as 113.9±43.3 nM. In contrast, the carbon poor media AB and M9 exhibited significantly higher levels of intracellular c-di-GMP at 529.4±307.2 nM and 2379.9±745.8 nM, respectively, and M9 media had significantly higher intracellular c-di-GMP than AB media. Interestingly, the intracellular c-di-GMP of 24 10 LB AKI OD 600 1 ASW 0.1 MOPS AB 0.01 M9 0.001 0 200 400 Time (minute s) 600 800 Figure 4. Growth curves of V. cholerae grown in different media. Cultures of V. cholerae were inoculated 1/1000 into 50 mL of the designated media, and OD600 readings were taken at regular time intervals. Each point represents the mean of three replicates, and the error bars indicate standard error of the mean. V. cholerae grown in MOPS more resembled LB and AKI, measuring at 126.2±41.5 nM; in contrast, intracellular c-di-GMP of V. cholerae grown in ASW was modestly higher, measuring at 251.6±75.5 nM. Thus, these results support the idea that c-diGMP signaling systems primarily function as environmental sensors and modulate c-di-GMP accordingly. Moreover, the high concentration of c-di-GMP observed in M9 media suggests the range of physiologically relevant concentrations of c-di-GMP experienced by V. cholerae is greater than currently appreciated. 25 4000 Intrace llular c-di-GMP (nM) 3000 2000 1000 800 600 400 200 0 AKI α LB β ASW χ MOPS δ AB ε M9 φ χεφ χεφ αβδφ χεφ αβδφ αβχδε Figu Figure 5. Intracellular cc-didi-GMP concentrations concentrations of V. cholerae grown in different media. media. V. cholerae was grown in 6 different media, and cells were harvested during late-exponential growth. The intracellular concentration was quantified using LCMS/MS and calculated using cell counts, determined by OD600 and plating, and cell size, determined by microscopy. Error bars represent standard deviation and corresponding symbols indicate statistical significance, which was determined using a two-tailed Student’s T-Test (p < 0.05). Growth environments modulate both c-di-GMP synthesis and hydrolysis While it is clear that the different growth environments alter intracellular cdi-GMP concentrations of V. cholerae, it is unclear if these changes are being driven by modulation of DGC and/or PDE activity. I addressed this question using TELCA on cultures that were grown in four of the six different growth media examined in Fig. 5. I hypothesized that [13C]-c-di-GMP synthesis would be greater when cells 26 were grown in both AB and M9 media, as the intracellular c-di-GMP of cells grown in these media is elevated. [13C]-GTP was added to lysates generated from V. cholerae grown in each environmental condition, and [13C]-c-di-GMP synthesis was quantified after 60 minutes using LC-MS/MS (Fig. 6A). Analogous to the intracellular c-di-GMP measurements, V. cholerae grown in AKI and LB exhibited similar c-di-GMP synthesis. Consistent with my hypothesis, growth of V. cholerae in AB media exhibited significantly elevated c-di-GMP synthesis, corresponding to its elevated intracellular c-di-GMP concentration. Interestingly, however, cells grown in M9 media showed a lower amount of c-di-GMP synthesis than those grown in AB, even though M9 produced the highest intracellular concentration of c-di-GMP that I observed. Based on my analysis of DGC activity I hypothesized that cells grown in M9 must exhibit lower PDE activity relative to the other media to account for the high concentration of c-di-GMP. To test this hypothesis, TELCA was used to quantify [13C]-c-di-GMP hydrolysis of cell lysates (Fig. 6B). Consistent with the intracellular c-di-GMP measurements, cells grown in AKI and LB had similar PDE activity; cells grown in AB media also had similar c-di-GMP hydrolysis to AKI and AB. However, in support of my hypothesis, cells grown in M9 media had significantly lower PDE activity compared to cells grown in the other media conditions. TELCA analysis of V. cholerae grown in these four different media suggests that suppressed PDE activity increases intracellular c-di-GMP in M9 media, while elevated DGC activity increases intracellular c-di-GMP in AB media. 27 A 13 [ C]-c-di-GMP S ynthesized (nM ) 100 80 60 40 20 0 AKI LB AB M9 AKI LB AB M9 B [ 13 C]-c-di-GMP Percent Los s 100 80 60 40 20 0 Figure 6. TELCA of V. cholerae grown in different media. (A) TELCA was applied to lysates of WT V. cholerae grown in different media. [13C]-GTP was added to each lysate and the amount of [13C]-c-di-GMP synthesis was quantified using LC-MS/MS. (B) TELCA was applied to lysates of WT and QS mutants of V. cholerae. [13C]-c-diGMP was added to each lysate, and the amount of [13C]-c-di-GMP hydrolysis was quantified using LC-MS/MS. The percent loss was calculated by comparing each lysate to a no-protein control reaction. All cultures were grown in triplicate, and the error bars indicate the standard deviation from the mean. Brackets indicate statistical significance using a one-tailed Student’s T-Test (p<0.05). 28 QS controls both total DGC and PDE activity in V. cholerae As mentioned above, QS modulates the intracellular concentration of c-diGMP in V. cholerae to regulate biofilm formation, but the specific DGCs and PDEs involved in this regulation are not fully understood. This question is challenging to address as QS controls the transcription of 18 proteins that potentially alter c-diGMP levels [66, 72]. Moreover, DGCs and PDEs could have differential enzymatic activities at low versus high cell density. To address this question, I used TELCA to determine if QS modulates the levels of c-di-GMP in V. cholerae by altering DGC and/or PDE activity. To analyze DGC and PDE activity of V. cholerae in different QS states, the WT, ∆hapR, and ∆luxO mutant strains were analyzed. The ∆hapR strain is a QS mutant that is locked in low cell density as HapR is the master high cell density transcriptional regulator, whereas the ∆luxO strain is a QS mutant locked in high cell density as the Qrr sRNAs are never expressed. Importantly, cultures were grown and harvested at an OD600 of 0.2-0.3, corresponding to the lowcell-density state. Therefore, I expected the WT strain to be in low cell density and mimic the ∆hapR mutant. Indeed, TELCA analysis indicated that the WT and the ∆hapR lysates exhibited significantly elevated c-di-GMP synthesis compared to the ∆luxO mutant (Fig. 7A), consistent with these strains having a higher concentration of c-di-GMP [66, 72]. Analysis of the DGC activity of WT cells grown to high cell density (compare Fig. 6A, LB to Fig. 7A, WT) indicates that the DGC activity decreases as cells reach a quorum. In fact, the DGC activity of the ∆luxO mutant at 29 A 13 [ C]-c-di-GM P S ynthesized (nM) 100 80 60 40 20 0 WT ∆ hapR ∆ luxO ∆ cdgA WT ∆ hapR ∆ luxO ∆ cdgA B [ 13 C]-c-di-GMP Percent Los s 100 80 60 40 20 0 Figure 7. TELCA of V. cholerae QS mutants. (A) [13C]-GTP was added to each lysate and the amount of [13C]-c-di-GMP synthesis was quantified using LC-MS/MS. (B) [13C]-c-di-GMP was added to each lysate, and the amount of [13C]-c-di-GMP hydrolysis was quantified using LC-MS/MS. The percent loss was calculated by comparing each lysate to a no-protein control reaction. All cultures were grown in triplicate, and the error bars indicate the standard deviation from the mean. Brackets indicate statistical significance using a one-tailed Student’s T-Test (p<0.05). 30 low cell density (Fig. 7A) is quite similar to that of WT V. cholerae grown to high cell density (Fig. 6A, LB). I performed an analogous experiment to quantify the impact of QS on net PDE activity. The relative PDE activity was determined for each of the WT, ∆hapR, and ∆luxO lysates by measuring the percent loss of [13C]-c-di-GMP. I found that the percent loss of [13C]-c-di-GMP was significantly higher in the ∆luxO strain compared to the ∆hapR mutant and WT strain (Fig. 7B). The PDE activity of WT cells harvested at low cell density was indistinguishable from the ∆hapR mutant. This result would suggest that PDE activity is increased as cells transition from low to high cell density. However, a comparison the WT strain grown to high cell density (40.6 ± 8.3 nM; Fig. 6B, LB) and the WT strain grown at low cell density (46.6 ± 8.7 nM; Fig. 7B) reveals a similar PDE activity. Indeed, the PDE activity of WT grown to high cell density (40.6 ± 8.3 nM; Fig. 6B, LB) was lower than the ∆luxO mutant (72.2 ± 8.0 nM; Fig. 7B). This difference in PDE activity between the WT strain and ∆luxO mutant was evident even if the WT strain was grown to stationary phase (data not shown). Thus, in regards to PDE activity the ∆luxO strain has a more dramatic phenotype than WT V. cholerae growing in LB. CdgA contributes to the increased DGC activity at low cell density The above experiments suggest that increased DGC activity at low cell density contributes to the higher concentration of c-di-GMP. CdgA (VCA0074) is a well-studied DGC that is repressed in the high-cell-density state and is important 31 for biofilm formation [72, 75, 94-97]. Based on a large body of evidence that indicates CdgA plays an important role in bridging c-di-GMP signaling and QS (see discussion), I hypothesized that CdgA significantly contributed to the increased DGC activity I observed at low cell density. To test this hypothesis and determine if TELCA can detect the contribution of a single DGC to total cellular activity, lysates from the ∆cdgA mutant grown to low cell density were analyzed. I observed a 36.4% decrease in DGC activity in the ∆cdgA strain compared to the WT strain, although the differences did not quite reach statistical significance (p=0.06) (Fig. 7A). As expected, there was no significant difference in the net PDE activity of the ∆cdgA strain and the WT strain (Fig. 7B). These results confirm previous genetic studies that implicate CdgA as a major contributor to the increased DGC activity observed in the low-cell-density state in V. cholerae, and further illustrate the applicability of TELCA to analyze complex c-di-GMP signaling pathways. Discussion The bacterial second messenger c-di-GMP is a central regulator of biofilm formation and motility in bacteria. Indeed, bioinformatic studies indicate that DGCs and PDEs are encoded by more than 80 percent of all sequenced bacteria [2]. C-diGMP controls many important phenotypes in bacteria including biofilm formation, motility, production of virulence factors, and lifecycle progression, and this list continues to grow [11, 13, 16, 98-100]. 32 Net changes in DGC or PDE activity leading to different intracellular c-diGMP concentrations can be difficult to parse as bacteria often encode multiple DGCs and PDEs. Furthermore, it is often challenging to resolve differences in c-diGMP signaling pathways in single DGC or PDE mutants due to this redundancy and the inherently low concentration of c-di-GMP in many bacterial species (for example, see [101]). To circumvent these challenges, I developed TELCA to quantify total cellular c-di-GMP synthesis and hydrolysis activities. The [13C]-labeled substrates TELCA utilizes are stable, safe to use, and can be distinguished from naturally occurring [12C]-c-di-GMP using mass spectrometry. To my knowledge, this is the first description of an in vitro assay to measure cellular DGC or PDE activity. To test TELCA, I examined strains of V. cholerae ectopically expressing either a GGDEF or an EAL enzyme and observed enhanced DGC and PDE activity, respectively. These experiments validated the approach and demonstrated that I can differentiate lysates with increased synthesis and hydrolysis activities. As an in vitro assay, disruption of the cell might alter the activity of localized or membrane bound proteins in the lysate. However, it should be noted that CdgA is a predicted integral inner membrane protein yet I was able to quantify a loss in cellular DGC activity of a ∆cdgA mutant using TELCA, showing that CdgA remained active during the assay. Regulatory inputs of the c-di-GMP signaling network remain undercharacterized. The consensus of the field is that c-di-GMP turnover enzymes bind and respond to environmental ligands, host derived metabolites, or are directly 33 modified by post-translational regulatory mechanisms to regulate c-di-GMP synthesis or hydrolysis, thus enabling the bacterium to fine tune intracellular c-diGMP to its environment [6]. Indeed, I showed that when V. cholerae is cultured in different media, the intracellular c-di-GMP varies significantly (Fig. 5). Surprisingly, the changes in the c-di-GMP levels were dramatic with cells grown in M9 exhibiting 20 times more c-di-GMP than cells grown in LB. This result suggests that V. cholerae, and potentially other bacteria, experience a much wider range of cdi-GMP concentrations than previously appreciated. Moreover, cells grown in defined, carbon poor media demonstrated slower growth rates and higher levels of cdi-GMP compared with complex, carbon rich media. The increased c-di-GMP observed in AB and M9 could be generated by three distinct processes: 1) increased DGC activity, 2) decreased PDE activity, or 3) modulation of both DGC and PDE activity. TELCA analysis revealed that c-di-GMP in AB media was driven by increased DGC activity while the c-di-GMP increase in M9 was driven by reduced PDE activity. This result highlights that bacteria can and do control intracellular c-di-GMP by regulating both c-di-GMP synthesis and hydrolysis. Moreover, one cannot distinguish the relative DGC and PDE inputs based solely on measuring the intracellular concentration of c-di-GMP. I applied TELCA to further understand how QS controls intracellular c-diGMP in V. cholerae. By analyzing mutants that lock the bacteria in the low- or high-cell-density state, I determined that DGC activity is increased at low cell density and decreases as the cells enter a quorum. Alternatively, I observed 34 increased PDE activity of the ∆luxO high-cell-density mutant, compared to a locked low-cell-density mutant and the WT strain. However, I did not observe changes in PDE activity of the WT strain at different cell densities; thus, the ∆luxO mutant exhibits a more pronounced modulation of PDE activity compared to the WT strain. A discrepancy in phenotypic expression of the WT and ∆luxO mutant is not unprecedented. For example, the ∆luxO mutant is deficient for colonization of the infant mouse whereas the WT is proficient [68]. Therefore, these strains should not be considered equivalent. The analysis of the ∆luxO mutant does indicate that QS in V. cholerae has the potential to modulate cellular PDE activity. I speculate that in other conditions not examined here, in which the Qrr sRNAs are more fully repressed at high cell density, the WT strain would exhibit density dependent PDE activity. My results clearly show that in LB V. cholerae QS control of c-di-GMP occurs through modulation of c-di-GMP synthesis but not c-di-GMP degradation. These results experimentally validate previous transcriptional studies that have shown that the expression of both DGCs and PDEs are differentially regulated by QS [66, 72, 74]. One protein that has been implicated in QS-mediated control of c-di-GMP is the DGC CdgA. It has been previously shown that CdgA is capable of synthesizing c-di-GMP in V. cholerae [79]. The regulation of cdgA is controlled by QS, as cdgA is repressed at high cell density via direct binding of HapR to the promoter region [72, 74, 75]. Furthermore, CdgA actively regulates biofilm formation as a ∆cdgA mutant strain demonstrates decreased colony rugosity [76] and ectopic expression of cdgA 35 leads to increased biofilm formation [79]. Also of note, a ∆hapR∆cdgA double mutant strain has reduced biofilm formation compared to the ∆hapR mutant, indicating that CdgA is involved in the HapR mediated increase in biofilm formation [102]. Taken together, this evidence suggested that repression of cdgA by HapR is one mechanism by which QS controls the intracellular concentration of c-di-GMP and biofilm formation. To confirm these previous genetic studies and quantify the contribution of CdgA to total DGC activity, I created a deletion mutant of cdgA and measured the relative c-di-GMP synthesis and hydrolysis potential using TELCA. My results are consistent with previous findings as the ∆cdgA strain had reduced cdi-GMP synthesis compared to the ∆hapR and WT strains while having similar c-diGMP synthesis as the ∆luxO mutant (Fig. 7). Importantly, this finding demonstrates that TELCA is able to distinguish the input of one DGC to total cellular DGC activity. TELCA is a straightforward technique that can be utilized to quantify total DGC or PDE activity for any bacterium. This method allows a systems level analysis of total c-di-GMP synthesis or degradation activity and can serve as a complementary approach to the current methods of c-di-GMP detection to further unravel complex c-di-GMP signaling networks. 36 CHAPTER 3 – Bile Acids and Bicarbonate Inversely Regulate Intracellular Cyclic didi-GMP in Vibrio cholerae Preface I hypothesized that bile is an environmental signal sensed by V. cholerae to recognize and adapt to growth in the human environment by modulating c-di-GMP signaling pathways. Based on current models postulating that V. cholerae reduces intracellular c-di-GMP in the human host, I predicted that bile acids would reduce global c-di-GMP concentrations. Surprisingly, I discovered that bile acids increase intracellular c-di-GMP. A screen of the activity of all 61 V. cholerae c-di-GMP turnover enzymes in the presence and absence of bile acids identified three DGCs that showed increased c-di-GMP synthesis in the presence of bile acids. Furthermore, a screen on the expression of all 61 V. cholerae c-di-GMP turnover enzymes revealed that bile acids inhibited the expression of one PDE. Deletion of these four enzymes abolished the induction of c-di-GMP by bile and negated the ability of V. cholerae to form biofilms in the presence of bile. Bicarbonate, a biological pH buffer secreted by intestinal epithelial cells in the small intestine, suppressed the bile-induced increases in intracellular c-di-GMP. I propose that bile and bicarbonate inversely control c-di-GMP levels in V. cholerae, allowing this bacterium to sense and adapt to local environmental niches within the small intestine. 37 Growth conditions and molecular methods All strains, plasmids, and primers used are listed in Tables A1-3. The V. cholerae El Tor biotype strain C6706str2 was used for all experiments [78] and Escherichia coli strains DH10B (Invitrogen) and S17-λpir [103] were used to harbor and conjugate plasmid DNA into V. cholerae. The construction of the ∆vpsL, ∆luxO, and ∆hapR strains has been described elsewhere [67, 68, 72]. For all experiments, unless otherwise specified, cultures were grown in LB media at 35°C with shaking at 220 rpm. When necessary, media was supplemented with kanamycin (Sigma) at 100 µg/mL or chloramphenicol (Sigma) at 10 µg/mL. The inducer isopropyl β-D1thiogalactopyranoside (IPTG) was added at 0.1 mM when required. Synthetic human bile (SHB) is a mixture of six purified conjugated bile acids added to LB media at physiologically relevant concentrations to mimic the human small intestine [37, 38, 104, 105]. All bile acids were purchased from Sigma. The conjugated bile acids added were taurocholate (0.46 mM), glycocholate (0.93 mM), taurochenodeoxycholate (0.46 mM), glycochenodeoxycholate (0.93 mM), taurodeoxycholate (0.32 mM), and glycodeoxycholate (0.64 mM). Bovine bile (BV, Sigma) was supplemented at 0.4% w/v. 3-[(3-Cholamidopropyl)dimethylammonio]-1propanesulfonate (CHAPS) was supplemented at 0.4% w/v, and sodium dodecyl sulfate (SDS) was supplemented at 0.01 % w/v due to its potent bactericidal activity. Taurine and glycine were each supplemented at 4 mM. Sodium bicarbonate (BiC) was supplemented when indicated at 0.3% w/v (49.2 mM), consistent with CT 38 inducing conditions [80], while 2-Amino-2-hydroxymethyl-propane-1,3-diol (Tris) was supplemented when indicated at 0.6% w/v (49.5 mM). All of the protocols used in this study for DNA manipulation and plasmid construction were performed as previously described [106]. The DNA polymerase Phusion (Thermo Scientific) was used for all PCR reactions. The expression plasmids for the DGCs and PDEs were constructed as described elsewhere [79]. Briefly, these plasmids allow controlled expression of each DGC encoded by V. cholerae via induction of the Ptac promoter with IPTG. To construct the HD-GYP expression plasmids, each HD-GYP gene was amplified from the V. cholerae chromosome, then inserted into the pEVS143 vector using the EcoRI and BamHI insertion sites, as previously described [72]. The identification of the c-di-GMP reporter plasmid 6:C9-lux is described elsewhere [107]. The DGC mutant allele plasmids were generated using the Lightning Site-Directed Mutagenesis kit (Agilent) with the DGC expression plasmid as template using the primers listed in Table A3. The VC2497-lux reporter was constructed by amplifying the promoter of VC2497 from the V. cholerae chromosome by PCR, then inserting it into the pBBRlux vector using the SacI and BamHI insertion sites. To generate the DGC and toxR deletion strains, natural transformation and homologous recombination were used. A PCR product was generated that contained a chloramphenicol resistance cassette (cat) bordered by FLP recombination target (FRT) sites from the plasmid pKD3 [84], flanked by 500 bp upstream and downstream of the targeted gene using the primers KO1/KO2 and KO3/KO4. The 39 PCR products generated with these promoters were fused to the cat gene using zipper PCR. Natural competence was induced by ectopically expressing tfoX (VC1153) [63] from the Ptac promoter using the plasmid pANDA2. pANDA2 was constructed by amplification of VC1153 with the primers CMW464 and CMW465 and insertion of this product into the EcoR1/BamH1 sites of the plasmid pEVS143. Homologous recombination events were selected by growing the culture on LB supplemented with chloramphenicol at 1 µg/mL. The cat gene was then removed by ectopically expressing a FLP recombinase on the vector pTL17 [86]. The toxT deletion mutant was constructed using allelic exchange as previously described via the plasmid pBK51, a derivative of the plasmid pKAS32 [108]. Detection and quantification of intracellular c-di-GMP All c-di-GMP quantifications were analyzed using liquid chromatography coupled with tandem mass spectrometry (LC-MS/MS) as described in Chapter 2. Unless otherwise specified, cultures were grown to an OD600 between 0.6 and 1, and a 1.5 mL aliquot of this culture was removed and centrifuged for 30 seconds at maximum speed. The supernatant was immediately removed and the pellet was resuspended in 100 µL of cold extraction buffer (40% acetonitrile/40% methanol/0.1 N formic acid) and incubated for 20 minutes at -20°C. The insoluble fraction was pelleted in a bench top centrifuge at 4°C for 5 minutes and the supernatant was collected and stored at -80°C. Prior to mass spectrometry, the extraction buffer was evaporated using a vacuum manifold. The pellet was then resuspended in 100 µL water. 40 Systematic screen of DGC and PDE activity To determine the in vivo activity of each V. cholerae DGC, each of the 40 DGC expression vectors was conjugated into a ∆vpsL strain of V. cholerae harboring a separate reporter vector encoding a luciferase-transcriptional fusion of a c-di-GMP inducible promoter located within the open reading frame (ORF) of VC1673 (6:C9- lux, [107]). These strains were then grown in solid white 96-well clear-bottom plates (Costar 3903) in 150 µL of LB or LB with SHB in the presence of IPTG inoculated as a 1/100 from an overnight culture. Luminescence and OD600 values were recorded in a Spectramax M5 Platereader (Molecular Devices) after 8 hours of growth, and were reported as relative luminescence units (R.L.U.). The screen for PDE activity was performed in the same manner as the DGC assay with the addition of a third vector, pBRP02, which carry the allele of the DGC qrgB under pTac control. pBRP02 has the Vibrio harveyi qrgB allele cloned into the vector pMMB67eh [109], allowing it to coexist with the PDE expression vectors and the 6:C9-lux reporter plasmid. Systematic screen of DGC and PDE expression DGC-gfp (GGDEF domain proteins) and PDE-gfp (GGDEF + EAL and EAL domain proteins) transcriptional fusion plasmids were constructed as previously described [72]. Strains containing the plasmids were grown in triplicate overnight and then inoculated 1/100 in 150 µL LB or LB with SHB in a Costar black, clearbottom 96 well plate (cat. # 3904). Cultures were grown for 8 hours, and then 41 fluorescence was quantified (excitation 475, emission 510) using a M5 Spectramax plate reader (Molecular Devices). Relative fluorescence was quantified by dividing the fluorescence by the OD600 reading of the culture. PDE-lux transcriptional fusion plasmids were constructed as previously described [66]. Strains containing the plasmids were grown in triplicate overnight and then inoculated 1/100 in 150 µL LB or LB with SHB in a Costar white, clear-bottom 96 well plate (cat. # 3903). Cultures were grown for 6 hours, and then luminescence was quantified using an Envision plate reader (PerkinElmer). Relative luminescence was quantified by dividing the luminescence by the OD600 reading of the culture. Biofilm quantification Static biofilm formation was determined using a modified protocol from [47]. Overnight planktonic cultures of V. cholerae and V. cholerae DGC mutants were grown in LB media, and cultures were diluted 1:1000 into 1 mL LB or LB with BV as specified in 17 x 100 mm polystyrene test tubes (BD Falcon). These tubes were incubated at 35°C for 24 hours without shaking. The supernatant was then removed, an absorbance reading of the supernatant was taken (OD600), and the biofilm was gently washed with approximately 2 mL of phosphate buffer saline (PBS). The biofilm was then stained with 0.41% crystal violet (CV) in 12% ethanol for 3 minutes followed by three washes with PBS, elution of the CV in 10 mL 95% ethanol, and OD570 measurement for each biofilm. Each OD570 biofilm measurement was normalized to the OD600 measurement of the planktonic culture to account for differences in growth. 42 Bile increases the intracellular concentration of c-di-GMP in V. cholerae It has been proposed that V. cholerae has a high intracellular c-di-GMP concentration in environmental reservoirs where it predominantly exists as surface associated biofilm communities and reduces c-di-GMP upon entrance into the human host, leading to the activation of virulence factor expression [35]. As bile is a prevalent constituent of the human small intestine, I hypothesized that V. cholerae would reduce c-di-GMP levels in response to bile. To test this hypothesis, I grew wild type V. cholerae in the presence of a crude bovine bile extract (BV) or synthetic human bile (SHB) and quantified the intracellular c-di-GMP concentration using LC-MS/MS. BV was supplemented to the media at 0.4% w/v, a concentration which is physiologically relevant to that found in the small intestine [39, 110]. SHB is a mixture of six purified bile acids that replicate the physiological concentrations of bile acids found in the human small intestine [37, 38, 104, 105]. To my surprise, growth in both BV and SHB-containing media significantly increased the intracellular concentration of c-di-GMP 4.4 and 3.6-fold, respectively (Fig. 8). To determine if the increase in intracellular c-di-GMP was due to the detergent activity of the bile acids, I grew V. cholerae in the presence of either the anionic detergent SDS (0.01% w/v) or the zwitterionic detergent CHAPS (0.4% w/v). I observed no significant difference in intracellular c-di-GMP upon addition of either detergent, suggesting that this increase in intracellular c-di-GMP is specific to bile acids. As the SHB mixture of bile acids contains bile acids conjugated to either taurine or a glycine, it was possible that these moieties were causing the increase in 43 intracellular c-di-GMP. However, there was no significant difference in the intracellular c-di-GMP concentration of V. cholerae grown with the addition of either taurine or glycine at equivalent concentrations (Fig. 8). Therefore, I conclude 1000 * * 800 600 400 200 + C H A LB PS + Ta ur in LB e + G ly ci ne SD S + LB LB + LB LB + B V SH B 0 LB Intracellular c-di-GMP (nM) that bile acids increase the intracellular concentration of c-di-GMP in V. cholerae. Figure 8. Intracellular cc-didi-GMP concentrations of V. cholerae grown with different bile additives. V. cholerae was grown in LB media and LB media with bovine bile (BV), synthetic human bile (SHB), the detergents SDS and CHAPS, and the amino acids taurine and glycine. Intracellular c-di-GMP was measured using LC-MS/MS. The reported values indicate the mean. Error bars indicate standard deviation, and * indicates statistical significance compared to LB as determined by a Student’s two-tailed T-Test (n=3, p < 0.05). The bile mediated increase of intracellular c-di-GMP is growth-phase dependent The experiments described thus far analyzed cultures grown to late exponential growth. I therefore examined the temporal dynamics of c-di-GMP induction by bile at various points during bacterial growth. When grown in 44 standard culture conditions with shaking, there is no significant difference in the doubling time of V. cholerae in the presence or absence of SHB (LB: 22.3 ± 0.3 minutes, LB + SHB: 21.6 ± 0.6 minutes). I measured intracellular c-di-GMP in three growth phases in the presence and absence of SHB: early exponential growth (OD600 0.2-0.3), late exponential growth (OD600 0.6-0.8), and stationary phase (OD600 1.01.2). When grown in LB alone, I observed elevated c-di-GMP in early exponential growth, and the concentrations of c-di-GMP decreased as the cell density increased. These observations are consistent with previous studies showing that c-di-GMP is elevated in the low-cell-density QS state [66, 72]. I observed significant increases of the intracellular c-di-GMP concentration in the presence of bile in early exponential and late exponential growth phases (Fig. 9). Alternatively, no significant induction of c-di-GMP was observed in stationary phase. These results indicate that bile induction of c-di-GMP is growth phase dependent. I speculated that QS might be responsible for density-dependent bile induction of c-di-GMP. However, bile induction of c-di-GMP was unchanged in V. cholerae mutants locked in either the low- or high-cell density QS state showing this process is independent of QS (data not shown). Three DGCs are more active in the presence of bile As SHB increases intracellular c-di-GMP, I hypothesized that specific DGCs in V. cholerae would exhibit increased activity in the presence of SHB. To identify these DGCs, I developed a high-throughput screen where the in vivo activity of each of the 40 V. cholerae DGCs could be determined in the presence and absence of 45 Intracellular c-di-GMP (nM) 4000 3000 2000 1000 0 Early Exponential Growth Late Exponential Growth Stationary Phase Figure 9. Intracellular cc-didi-GMP concentrations concentrations of V. cholerae during the course of growth with bile. bile. All intracellular c-di-GMP concentrations were determined using LC-MS/MS. Black bars indicate strains grown in LB and grey bars indicate strains grown in LB + SHB. The reported values indicate the mean and the error bars indicate the standard deviation from the mean. Brackets indicate statistical significance as determined by a one-tailed Student’s T-Test (n=3, p > 0.05). SHB. This screen utilized a series of expression vectors carrying every V. cholerae DGC under the control of the IPTG-inducible Ptac promoter and a common ribosome binding site [79]. Therefore, I infer in this system that differences in activity are not due to changes in gene expression. In addition to the DGC expression vector, each strain contained a separate reporter plasmid that carried a transcriptional fusion of a c-di-GMP inducible promoter located within the ORF of VC1673 with the luciferase operon (named 6:C9-lux, [107]). These vectors were introduced into a ∆vpsL strain of V. cholerae which abrogates biofilm formation and eliminates interference of aggregate formation on the analysis of reporter gene expression. In the absence of additional c-di-GMP generated by exogenous 46 expression of DGCs, the reporter plasmid does not produce significant luminescence in the presence or absence of SHB (Fig. 10, vector control). C-di-GMP production by active DGCs leads to increased luciferase production, as can be seen by induction of luminescence by expression of the constitutively active DGC qrgB from V. harveyi (Fig. 10, qrgB). Induction of this transcriptional reporter is dependent on c-di-GMP synthesis as expression of qrgB*, an allele of qrgB harboring a mutation in the active site of the protein, does not induce luminescence (Fig. 10, qrgB*). 6:C9-lux expression was determined upon induction of each DGC in the presence of LB and LB with SHB. Seven DGCs showed reduced luminescence expression in the presence of SHB, although the extent of this reduction was not large (< 3.5-fold). It is not currently known if this reduction of expression is related to bile inhibition of the DGC activity of these proteins, unrelated transcriptional regulation of 6:C9-lux, or a non-specific effect of SHB on luminescence. Due to the large number of DGCs that are negatively affected (7, p <0.01), the relatively low reduction, and that bile increases the total intracellular c-di-GMP concentration, I favor the latter two possibilities. Nevertheless, three DGCs demonstrated more than a five-fold fold increase luminescence in the presence of SHB (p<0.01, Fig. 10); these DGCs are VC1067 (7.0 fold increase), VC1372 (47.6 fold increase), and VC1376 (6.7 fold increase). 47 1.5 1.5× ×10 7 1.0 1.0× ×10 7 5.0 5.0× ×10 6 * * QrgB QrgB* Vector Control VC0900 VC1029 VC1067 VC1104 VC1185 VC1216 VC1353 VC1367 VC1370 VC1372 VC1376 VC1593 VC1599 VC2224 VC2285 VC2370 VC2454 VC2697 VCA0049 VCA0074 VCA0165 VCA0217 VCA0557 VCA0560 VCA0697 VCA0848 VCA0939 VCA0956 VCA0960 VCA0965 VCA1082 VC0072 VCA0785 VC0398 VC0653 VC0658 VC0703 VC1934 VC2750 VCA0080 0 Relative Lumine scence Units * GGDEF GGDEF + EAL Figure 10. HighHigh-throughput screen of the in vivo activity of each of the 40 V. cholerae DGCs with bile. bile. V. cholerae strains containing a DGC expression plasmid and the 6:C9-lux reporter plasmid were measured for luminescence in LB (black) and LB with SHB (grey). Expression plasmids encoding qrgB and its mutant allele counterpart qrgB* were included as positive and negative controls, respectively. The vector control indicates expression of 6:C9-lux in the absence of protein induction. The error bars indicate standard deviation. Statistical significance (*) was determined for cultures exhibiting a positive fold change (LB + SHB / LB) greater than two as determined by a Student’s one tailed T-Test (n=3, p < 0.01). The activity of PDEs in V. cholerae is not altered by bile The hydrolysis of c-di-GMP is driven by c-di-GMP-specific PDE enzymes, which contain either a C-terminal EAL or HD-GYP domain. Like DGC enzymes, these proteins are also modular and the variable N-terminal domain is thought to 48 respond to environmental stimuli. I hypothesized that an increase in intracellular cdi-GMP in response to bile acids could be caused by a repression of PDE activity. To examine this possibility, I developed a second high-throughput screen to analyze the enzymatic activity of the 29 predicted PDEs of V. cholerae. The sequence of one EAL protein, VC0515, was highly divergent from the sequenced N16961 strain upon amplification from the genome, and was thus excluded from analysis; this is consistent with other studies reporting sequence variance [72]. Similar to the screen above, each V. cholerae PDE is expressed from an inducible expression vector that has been introduced into a ∆vpsL strain of V. cholerae containing the 6:C9-lux reporter vector. A third vector (pBRP02) was introduced containing the constitutively active DGC qrgB, effectively increasing the baseline c-di-GMP concentration and subsequently the baseline luminescence values. Robust reporter gene expression can be seen in a strain expressing qrgB with a vector control that has no exogenous PDE expression (Fig. 11, vector control). Strains of V. cholerae expressing each PDE alongside the DGC qrgB were grown in LB and LB with SHB, and the expression of 6:C9-lux was quantified. 22 of the 29 PDEs showed appreciable decreases (> 2-fold, p <0.01) in luminescence when grown in LB compared to the vector control (Fig. 11), indicating that these PDEs are actively hydrolyzing c-di-GMP. However, none of the PDEs demonstrated any appreciable difference when grown in LB with SHB compared to LB alone (Fig. 11). Thus, I conclude that SHB does not affect the c-di-GMP hydrolysis of PDEs in the conditions I examined here. 49 6.0 ×10 7 4.0 ×10 7 2.0 ×10 7 Vector Control VC0072 VCA0785 VC0398 VC0653 VC0658 VC0703 VC1934 VC2750 VCA0080 VC1087 VC1295 VC1348 VC2340 VC2497 VCA0210 VCA0681 VCA0895 VCA0931 VC0137 VC1086 VC1211 VC1592 VC1641 VC1652 VC1710 VC1851 VCA0101 VCA0536 VCA1083 0 Re lative Lumine sce nc e Units GGDEF + EAL HD-GYP EAL Figure 11. HighHigh-throughput screen of the in vivo activity of each of the 40 V. cholerae PDEs with bile. bile. V. cholerae strains containing a PDE expression plasmid, the 6:C9-lux reporter plasmid, and the qrgB expression plasmid were measured for luminescence in LB (black bars) and LB with SHB (grey bars). The vector control indicates expression of 6:C9-lux with qrgB induction and without PDE expression. The error bars indicate standard deviation (n=3). The three bile-inducible DGCs are inner membrane proteins I analyzed the predicted domain structure and subcellular localization of the three bile inducible DGCs to search for commonalities. An analysis of the relative hydrophobicity of the amino acid sequence of these three DGCs using the program toppred predicts that all contain transmembrane spanning domains (Fig. 12, [111113]). The DGC VC1067 (GI: 15641080), which has also been referred to as cdgH 50 B 2 0 -2 -4 0 4 2 0 -2 -4 0 200 400 600 Amino Ac id C Hydrophobicity Value 4 Hydrophobicity Value Hydro pho bic ity Value A 100 200 300 Amino Acid 4 2 0 -2 -4 0 200 400 Amino Acid Figure 12. The amino acid sequence hydrophobicity and predicted signaling domains of the bilebile-responsive DGCs. The amino acid sequence of (A) VC1067, (B) VC1372, and (C) VC1376 was used to predict transmembrane domains using the toppred program. The dotted line indicates the cutoff value for a potential transmembrane domain. A depiction of the potential N-terminal signaling domains and transmembrane domains (grey rectangles) and the C-terminal GGDEF domain of each DGC are shown below each respective hydrophobicity plot. The images were adapted from the SMART database [113]. 51 [94], is predicted to contain two sequential periplasmic substrate-binding domains (PBPb) in the N-terminus while VC1376 (GI: 15641388), which has also been referred to as cdgM [97], is predicted to contain a conserved CHASE domain. VC1372 (GI: 15641384) does not contain any conserved protein domains in the Nterminus; however, analysis of the relative hydrophobicity of the amino acid sequence using toppred predicts 6 sequential intra-membrane spanning domains. Three DGCs increase c-di-GMP synthesis in the presence of bile acids To directly examine if SHB increases the DGC activity of these three enzymes, I quantified the intracellular c-di-GMP concentration of strains ectopically expressing VC1067, VC1372, or VC1376 grown in the presence and absence of SHB. I similarly analyzed expression of active site mutant alleles of these genes (GG[D/E]EF -> AA[D/E]EF) that render them incapable of c-di-GMP synthesis. As addition of SHB increases the intracellular c-di-GMP concentration of V. cholerae (Fig. 8), I hypothesized that the strains expressing each AA[D/E]EF mutant would show increased intracellular c-di-GMP in the presence of bile similar to that of the WT strain. Moreover, induction of the bile-activated DGCs would lead to more c-diGMP synthesis in the presence of bile versus LB alone. When grown in LB media, there was no notable difference between strains expressing any of the AA[D/E]EF alleles and the vector control in LB media. As expected growth of the vector control or DGC mutants in bile showed similar increases in c-di-GMP. Induction of the DGCs VC1067 and VC1376 produced 52 significantly increased amounts of c-di-GMP in LB alone, leading to intracellular cdi-GMP concentrations of 48.4±7.0 µM and 19.3±10.0 µM, respectively (Fig. 13). Alternatively, the intracellular c-di-GMP of the VC1372 expression strain in the absence of bile (116.9±17.4 nM) was not significantly altered compared to the vector control (263.7±119.7 nM), indicating that this DGC does not produce c-di-GMP in LB media alone. Importantly, the intracellular concentration c-di-GMP was increased significantly upon ectopic expression of VC1067, VC1372, and VC1376 in the presence of LB with SHB versus LB alone resulting in intracellular concentrations of 89.2±31.5 µM, 4.3±0.4 µM, and 67.4±20.1 µM c-di-GMP, respectively (Fig. 13). These data together indicate that VC1067, VC1372, and VC1376 synthesize more c-di-GMP in the presence of bile acids. A toxR deletion mutant partially loses bile-mediated c-di-GMP increase One system that has been implicated in the bile response of V. cholerae is the ToxR virulence regulatory network. The transcriptional regulator ToxR controls the expression of the virulence regulator ToxT, as well as the expression of CT and TCP[114, 115]. Furthermore, ToxR mediates bile resistance in V. cholerae by altering the expression of ompU and ompT [43, 116]. It has also been shown that bile acids alter the protein activity of ToxR [117]. I hypothesized that bile acids act through ToxR or ToxT to alter the transcription of c-di-GMP turnover enzymes resulting in increased global c-di-GMP concentrations. To determine if bile mediated changes in ToxR or ToxT transcription regulation effect intracellular c-diGMP, I generated ∆toxR and ∆toxT deletion strains and measured intracellular c-di53 Intr acellular c -di-GMP (nM) 150000 100000 50000 6000 4000 2000 V ec to r C on tr ol V V C1 C1 06 06 7 7 (A AD E F) V V C C 13 13 72 72 (A A E E F) V V C C1 13 37 76 6 (A AD E F) 0 Figure 13. 13. Intracellular c-didi-GMP concentrations of V. cholerae DGC expression strains in the presence and absence of SHB. The c-di-GMP concentrations of the empty vector strain or the VC1067, VC1372, or VC1376 expression strains were quantified with LC-MS/MS. The intracellular c-di-GMP of strains expressing alleles containing mutations in the active site motif of each DGC were also quantified. The black bars indicate strains grown in LB, while the grey bars indicate strains grown in LB with SHB. Error bars indicate standard deviation. Brackets indicate statistical significance, which was determined using a Student’s one tailed T-Test (n=3, p < 0.05). GMP in the presence and absence of SHB. I found that while the ∆toxR strain showed no significant difference to the WT strain in LB alone, intracellular c-diGMP was 44.9% lower in the ∆toxR strain compared to WT in LB with SHB (Fig. 14). Similar to the ∆toxR strain, the intracellular c-di-GMP of the ∆toxT strain is 54 not discernibly different in LB. In contrast, the intracellular c-di-GMP of the ∆toxT strain is modestly reduced in LB with SHB compared to WT, however this difference is not statistically relevant. These results indicate that the transcriptional regulator ToxR plays a role in the bile-mediated increase in Intrace llular c-di-GMP (nM) intracellular c-di-GMP. 600 400 200 0 WT ∆ toxR ∆ toxT Figure 14. Intracellular cc-didi-GMP concentrations of V. cholerae ∆toxR and ∆toxT deletion mutants in the presence of bile. bile. LC-MS/MS was used to quantify intracellular c-di-GMP. Black bars indicate strains grown in LB and grey bars indicate strains grown in LB with SHB. The reported values indicate the mean and the error bars indicate the standard deviation from the mean. Brackets indicate statistical significance as determined by a one-tailed Student’s T-Test (n=6, WT and ∆toxR, n=3, ∆toxT, p > 0.05). Transcription of VC1295 is inhibited by bile acids I hypothesized that the presence of bile could also regulate the transcription of DGCs or PDEs to control intracellular c-di-GMP concentration. Specifically, bile could increase the expression of DGCs or decrease the expression of PDEs to 55 increase intracellular c-di-GMP. To examine this possibility, I measured the relative transcription of 51 GGDEF, GGDEF+EAL, and EAL proteins in LB and LB with SHB by growing strains of ∆vpsL V. cholerae containing transcriptional fusions of approximately 500 bp DNA upstream of each gene to gfp. After 8 hours of growth, I found that there was less than a 2-fold difference in fluorescence between all GGDEF, GGDEF + EAL, and EAL strains grown in the presence and absence of bile (Fig. 15 A and B) suggesting that in the conditions examined here, regulation of GGDEF and EAL gene transcription by bile does not play a significant role in the bile mediated increase of intracellular c-di-GMP. I also examined the relative transcription of the 8 HD-GYP proteins in LB and LB with SHB using transcriptional fusions of each promoter, defined as approximately 500 bp upstream of each HD-GYP gene, to the lux operon. The HDGYP gene VC1087 is encoded in a putative operon with the EAL gene VC1086, and thus the expression is presumed to be synonymous with VC1086 (Fig. 15 B). The luminescence of each reporter strain was determined after 6 hours of growth in either LB or LB with SHB. The expression of six of the genes encoding HD-GYP proteins was not significantly changed in LB with SHB compared to LB (> 2-fold), whereas the expression of VC2497 was modestly increased in the presence of SHB (2.2-fold, p < 0.05). Importantly, the expression of VC1295 (GI: 15641308) was decreased 2.8 fold when grown in the presence of SHB (Fig. 15C, p < 0.05). VC1295 appears to be an active PDE when ectopically expressed in V. cholerae (Fig. 11). 56 3000 2000 1000 VC0900 VC1029 VC1067 VC1104 VC1185 VC1216 VC1353 VC1367 VC1370 VC1372 VC1376 VC1593 VC1599 VC2224 VC2285 VC2454 VC2697 VCA0049 VCA0074 VCA0165 VCA0217 VCA0557 VCA0560 VCA0697 VCA0848 VCA0939 VCA0956 VCA0960 VCA0965 VC2370 0 Relative Fluorescence Units A GGDEF Figure 15. Relative transcription of DGCs and PDEs in the presence and absence of bile. (A) V. cholerae strains containing a transcriptional fusion of each GGDEF protein promoter to gfp was grown in LB (black) or LB with SHB (grey). The fluorescence in each environmental condition was quantified after 8 hours of growth (n=3). (B) V. cholerae strains containing a transcriptional fusion of each GGDEF + EAL or EAL protein promoter to gfp was grown in LB (black) or LB with SHB (grey). The fluorescence in each environmental condition was quantified after 8 hours of growth (n=3). (C) V. cholerae strains containing a transcriptional fusion of each promoter driving expression of an HD-GYP to luciferase were grown in LB (black) or LB with SHB (grey). Luminescence was quantified after 6 hours of growth (n=4) in each environmental condition. All cultures were normalized to an OD600 reading. Error bars indicate standard deviation. * indicates statistical significance from the LB condition, determined by a one-tailed Students T-Test (p < 0.05). 57 C 58 HD-GYP VCA0931 7.5 ×10 1.5 ×10 10 7 GGDEF + EAL VCA0895 VCA0681 VCA0210 VC2497 VC2340 VC1348 VC1295 0 10 6 Relative Luminescence Units B EAL * VCA0536 VCA0101 VC1851 VC1710 VC1652 VC1641 VC1592 VC1211 VC1086 VC0515 VC0137 VCA0080 VC2750 VC1934 VC0703 VC0658 VC0653 VC0398 VCA0785 VC0072 VCA1082 0 12500 25000 Relative Fluorescence Units Figure 15 (cont’d) This result indicates that bile acids decrease the expression of VC1295, possibly resulting in decreased c-di-GMP hydrolysis contributing to increased intracellular cdi-GMP. Three DGCs and one HD-GYP account for bile-mediated c-di-GMP induction To determine if the DGCs VC1067, VC1372, and VC1376 contribute to the bile-mediated increase of intracellular c-di-GMP in V. cholerae, I constructed unmarked V. cholerae DGC deletion mutants and examined SHB mediated induction of c-di-GMP (Fig. 16). Similar to my previous findings, the intracellular cdi-GMP concentration of V. cholerae increased 4.1 fold in the presence of SHB in the WT strain. Both the ∆VC1067 and ∆VC1376 single mutants showed a modest but significant reduction of c-di-GMP when grown in bile compared to the WT strain, having 29.6% and 24.9% less c-di-GMP, respectively. The ∆VC1372 single mutant and ∆VC1372∆1376 double mutant were not statistically different from the WT strain. Importantly, the ∆VC1067∆VC1372∆VC1376 triple mutant strain exhibited the greatest reduction of intracellular c-di-GMP in the presence of bile, losing 46.8% of the intracellular c-di-GMP compared to the WT strain in LB with SHB (p< 0.05). These results suggest that VC1067, VC1372, and VC1376 function redundantly in the bile-mediated c-di-GMP induction. As the expression of the HD-GYP VC1295 was inhibited by bile acids, I constructed an unmarked V. cholerae VC1295 deletion mutant and quantified intracellular c-di-GMP in the presence and absence of SHB. In LB alone, the 59 Intrac ellular c-di-GMP (nM) 1500 1000 500 ∆ VC1295 ∆ VC1067 ∆ VC1372 ∆ VC1376 ∆ VC1295 ∆ VC1067 ∆ VC1372 ∆ VC1376 ∆ VC1372 ∆ VC1376 ∆ VC1376 ∆ VC1372 ∆ VC1067 WT 0 Figure 16. Intracellular cc-didi-GMP concentrations of V. cholerae DGC and PDE mutant strains in the presence and absence of bile. LC-MS/MS was used to quantify c-di-GMP after growth in LB (black) or LB with SHB (grey). The reported values indicate the mean and the error bars indicate standard deviation. Brackets indicate statistical significance as determined by a one-tailed Students T-Test (n=9-10, p < 0.05). intracellular c-di-GMP was modestly increased 1.8 fold in the ∆VC1295 mutant compared to WT (Fig. 16, p < 0.05). This result is expected as deletion of an active PDE will increase intracellular c-di-GMP. However, the intracellular c-di-GMP concentrations of the ∆VC1295 strain grown in the presence of SHB were indistinguishable from WT. I hypothesized that both activation of DGC activity and transcriptional regulation of VC1295 contribute to bile induction of c-di-GMP. To test this, I created a quadruple ∆VC1295∆VC1067∆VC1372∆VC1376 mutant and 60 measured intracellular c-di-GMP in the presence and absence of SHB. Similar to the ∆VC1295 single mutant, when grown in LB alone the quadruple mutant had elevated intracellular c-di-GMP compared to the WT strain (2.4 fold, p < 0.05). Importantly, this strain showed no change in intracellular c-di-GMP in the presence of SHB. This indicates that these four proteins are responsible for the bile-mediated changes in intracellular c-di-GMP. Deletion of the bile-responsive DGCs and PDE reduces bile induction of V. cholerae biofilm formation It has been previously reported that BV (i.e., bovine bile) increases biofilm formation of V. cholerae, and this induction is dependent on the transcriptional regulator vpsR [47]. VpsR binds c-di-GMP to regulate the transcription of biofilm genes [107]. I wondered if the levels of c-di-GMP measured in the various DGC and PDE mutants strains with and without bile would correlate with biofilm formation. To test this, I performed a static biofilm assay where the wild type, ∆vpsL mutant, and the DGC and PDE V. cholerae mutant strains were grown in polystyrene test tubes containing LB or LB with BV without shaking followed by crystal violet (CV) staining of the resulting attached biofilm. The ∆vpsL mutant cannot produce exopolysaccharide and thus does not form biofilms. BV was used to induce biofilm formation to remain consistent with prior studies [47, 117] and because it induced more robust biofilm formation in this assay than SHB (data not shown). 61 I observed that all cultures of V. cholerae grew to a significantly lower optical density, as measured by OD600, after static growth in the presence of BV. To account for these growth differences, the biofilm formation of each culture was normalized to the OD600 of the planktonic culture. Consistent with previous reports [47], BV increased biofilm formation in the WT strain 2.7-fold, and this response was eliminated in the ∆vpsL strain (Fig. 17). While the ∆VC1067, ∆VC1376, and ∆VC1372∆1376 mutants all showed modest to no loss of bile induced c-di-GMP (Fig. 16), these mutants no longer exhibited significant bile-induced biofilm formation. Only the ∆VC1372 mutant induced biofilm formation in response to bile, indicating that this DGC contributes less to biofilm formation. As expected, the triple DGC mutant showed the lowest level of biofilm formation and was not responsive to bile addition. Although the ∆VC1295 strain had modestly elevated c-di-GMP in LB, deletion of VC1295 had no noticeable impact on biofilm formation. Like the triple mutant, the DGC/PDE quadruple mutant exhibited low biofilm formation that was unresponsive to bile addition. These results demonstrate that the bile responsive DGCs are required for c-di-GMP dependent biofilm formation in response to bile acids. However, the HD-GYP VC1295 did not contribute to biofilm formation in this assay. Bicarbonate decreases the bile-mediated increase of c-di-GMP in V. cholerae As c-di-GMP has been shown to repress virulence gene expression in V. cholerae, it remained puzzling as to why bile would increase c-di-GMP [35]. I hypothesized that additional signals in the small intestine could override this 62 0.8 0.6 0.4 ∆ VC1295 ∆ VC1067 ∆ VC1372 ∆ VC1376 ∆ VC1295 ∆ VC1067 ∆ VC1372 ∆ VC1376 ∆ VC1372 ∆ VC1376 ∆ VC1376 ∆ vpsL WT 0.0 ∆ VC1372 0.2 ∆ VC1067 Biofilm Fo rmatio n (OD 570 CV / OD 600 p lanktonic) 1.0 Figure 17. 17. Biofilm formation of V. cholerae DGC and PDE mutants in the presence of BV. BV. Biofilm formation was quantified in test tubes using CV. A strain containing a mutation in the vpsL gene was included as a negative control. The CV value was normalized by the OD600 value of the planktonic culture to account for differences in growth. The reported values indicate the mean, and error bars indicate standard deviation from the mean. Brackets indicate statistical significance as determined by a Student’s one tailed T-Test (n=3, p <0.05). induction. Another major component of the human small intestine is bicarbonate, which is secreted by the pancreas as well as the small intestinal epithelial cells. Bicarbonate acts as a pH buffer to neutralize acids secreted by the stomach [118, 119]. Bicarbonate has also been implicated in virulence gene regulation in V. cholerae [120]. As bicarbonate is abundant in the small intestine, I hypothesized that it may contribute to the regulation of intracellular c-di-GMP in V. cholerae. To examine if bicarbonate impacted bile induction of c-di-GMP, I measured 63 intracellular c-di-GMP of V. cholerae when grown in the presence and absence of SHB and bicarbonate. I observed no significant difference in intracellular c-di-GMP of V. cholerae in the presence of bicarbonate alone (Fig. 18). However, growth of V. cholerae in the presence of both SHB and bicarbonate completely abolished bile 500 * 400 300 200 100 Tr is B iC SH B + LB LB + SH B + + + LB LB + SH B B iC 0 LB Intracellular c-di-GMP ( µM) induction of c-di-GMP (Fig. 18). Figure 18. Intracellular Intracellular cc-didi-GMP concentrations of V. cholerae grown in the presence of bile and bicarbonate. LC-MS/MS was used to quantify c-di-GMP ofV. cholerae grown in LB, LB with bicarbonate (BiC), or LB with SHB and different supplements. The reported values indicate the mean and the error bars indicate the standard deviation. * indicates statistical significance compared to LB as determined by a one-tailed Student’s T-Test (n=3, p > 0.05). As bicarbonate acts as a buffer in the small intestine, I hypothesized that bicarbonate suppression of the bile induction of c-di-GMP was due to pH changes. I tested this idea by growing V. cholerae in the presence of SHB with the pH buffer 64 Tris. The pH of LB media and LB with SHB media was approximately the same and did not change significantly over the course of growth (pH LB, pre growth: 7.3, post growth: 7.0; pH LB with SHB, pre growth: 7.3, post growth: 7.1). Upon addition of bicarbonate, the pH increased substantially (pH LB with SHB and BiC, pre growth: 8.1, post growth: 8.7). The addition of Tris caused a similar increase in pH (pH LB with SHB and Tris, pre growth: 8.7, post growth: 8.6). Analogous to bicarbonate, the addition of Tris to LB with SHB media inhibited the normal induction of c-di-GMP by bile (Fig. 18), indicating that this inhibition of bile induced c-di-GMP is pH dependent. Discussion Bile is an abundant component of the human small intestine, and thus a probable physiological cue for V. cholerae to recognize upon entry into this environment. A number of lines of evidence suggest that bile is an important signal in the transition of V. cholerae between environmental and infectious lifestyles. Bile acids increase the expression of ompU in a toxR-dependent manner to increase bile resistance, indicating that the classical biotype of V. cholerae is capable of sensing the presence of bile [116]. There are conflicting reports regarding bile control of virulence factor expression. Bile acids were reported to negatively regulate the expression of the TCP and CT in a toxT-dependent manner in a classical biotype, demonstrating that there is a link between bile and virulence [121]. Contrary to this finding, bile acids were reported to induce CT and TCP in a toxR-dependent manner in a classical V. cholerae biotype and a tcpP-dependent manner in the El Tor 65 biotype used in this study [117, 122]. From these studies, it is clear that V. cholerae responds to bile to induce a number of physiological changes. In this work, I explore the connections between bile and c-di-GMP. Based on the prevailing V. cholerae disease model hypothesizing that c-diGMP levels are decreased upon infection, I predicted that bile acids would decrease intracellular c-di-GMP concentrations. Contrary to this prediction, I found that bile acids increase the intracellular c-di-GMP concentration of V. cholerae (Fig. 8). This finding suggests that the dynamics of the c-di-GMP signaling system in the human host are more complex than previously appreciated. Furthermore, I have found the difference in the bile mediated change of c-di-GMP is largest during exponential growth and that at stationary phase this difference was negated. Consistent with these findings, it has been shown that intracellular c-di-GMP is depleted at high cell density, and that these changes are due in part to quorum sensing [66, 72, 77]. My results indicate that this regulation is dominant over the increase in intracellular c-di-GMP caused by bile. To begin to understand how V. cholerae modulates its intracellular c-di-GMP in response to bile, I determined if the activity of any DGCs or PDEs were affected by bile using a novel high-throughput in vivo assay. This assay is easily adaptable to examine the response of DGCs and PDEs to any environmental cue, and it can be modified to explore other bacteria if a suitable in vivo reporter of c-di-GMP levels is available. Three DGCs (VC1067, VC1372, and VC1376) exhibited increased c-diGMP synthesis activity in the presence of SHB (Fig. 10, 13). The DGC activity of 66 VC1372 appeared to be absolutely dependent on bile whereas bile modulated the basal activities of VC1067 and VC1376. All of these DGCs are important for the bile-induced increase of intracellular c-di-GMP and biofilm formation of V. cholerae (Fig. 16, 17). The mechanisms by which these three DGCs respond to bile acids are currently unknown. All three DGCs are predicted to be associated with the inner membrane. VC1067 has been implicated in biofilm formation as it induces rugosity associated phenotypes in V. cholerae [94]. Furthermore, it has been shown that both VC1067 and VC1376 actively produce c-di-GMP and stimulate biofilm formation [79, 94], and a VC1376 mutant strain also demonstrates lower vpsL expression [97]. VC1376 has also been implicated in the indole-induced increase in biofilm formation [96]. Furthermore, all three DGCs also have been shown to repress motility when ectopically expressed in V. cholerae; this is particularly interesting as the levels of motility repression of the VC1372 and VC1376 expression strains do not seem to correlate with the intracellular c-di-GMP levels reported here (Fig. 13) [123]. Analysis by BLAST revealed that homologous sequences to VC1372 are only found in a few Vibrio species dominated by strains of V. cholerae. A previous study has also noted that VC1372 is unique to V. cholerae amongst the genus Vibrio [124]. The phylogenetic link of VC1372 with an enteric human pathogen, the domain structure of VC1372, and my observation that the activity of VC1372 is absolutely dependent on the presence of bile suggest the physiological cue which controls VC1372 is bile. Bile acids are known to interact with the cell membrane 67 due to their detergent activity [41]. Moreover, I observed bile-mediated activation of VC1372 in E. coli (data not shown), an orthologous system with no clear homolog to VC1372. Alternatively, as both VC1067 and VC1376 maintain robust activity upon exogenous expression even in the absence of bile, I postulate these DGCs might be controlled indirectly by bile through sensing perturbations in the membrane. Additionally, the expression of the PDE VC1295 was inhibited by bile acids (Fig. 15C). This HD-GYP actively hydrolyzes c-di-GMP in both LB and LB with SHB, but bile acids do not affect this activity (Fig. 11). VC1295 is predicted to be composed of 492 amino acids and contains 6 predicted N-terminal transmembrane domains preceding a HAMP domain linked to the HD-GYP domain [111-113]. The mechanism governing this transcriptional regulation of VC1295 by bile remains unknown. Analysis of the VC1295 promoter region reveals motifs resembling the ToxR binding site (-127 bp, TCAAA-N11-TTAAA [114]). While this gene is not listed amongst the known genes regulated by ToxR [33] there is evidence that the activity of ToxR is altered by bile [117], suggesting that the ToxR regulon could be altered when bile acids are present. As a ∆toxR strain of V. cholerae partially loses the bileinduced increase in c-di-GMP (Fig. 14), I hypothesize that in the presence of bile, ToxR inhibits the expression of VC1295. Although VC1295 contributed to bileinduced c-di-GMP, I did not observe any effect of VC1295 on biofilm formation. I speculate that this result is due to the distinct experimental differences in which cdi-GMP and biofilm formation were measured. 68 Another important host derived cue is bicarbonate, a biological pH buffer that is highly abundant in the human small intestine [119]. Bicarbonate has been shown to be important for virulence, as bicarbonate is critical for in vitro production of CT [80]. Bicarbonate is capable of activating V. cholerae virulence gene expression via the transcriptional regulator toxT [120]. I have shown that bicarbonate is able to suppress the bile-mediated induction of c-di-GMP in V. cholerae. Furthermore, this regulation is driven by changes in pH, as the bile-mediated induction is similarly repressed by Tris (Fig. 18). It is possible that the change in pH alters the structure of bile so that it no longer triggers the c-di-GMP synthesis activity of the DGCs. Alternatively, the bicarbonate could directly interact with DGCs or PDEs to competitively alter intracellular c-di-GMP. As bile has strong antimicrobial properties [125], it may be physiologically advantageous for V. cholerae to increase c-di-GMP to grant elevated tolerance to bile acids and other stresses associated with the intestinal environment. Other studies have indicated that biofilm formation is important for increased acid shock tolerance and protection from bile acids [47, 75], and biofilms increase infectivity and intestinal colonization in a mouse infection model [126]. Indeed, I confirmed that BV induces biofilm formation in V. cholera and showed that all three bileresponsive DGCs were required for bile-induction of biofilm formation. I propose that bile stimulates a high intracellular c-di-GMP concentration in the intestinal lumen (Fig. 19). Upon penetration of the mucosal layer where the bicarbonate concentration and thus the pH is elevated [127], the response to bile is 69 abrogated leading to a corresponding decrease in intracellular c-di-GMP. The physiological consequences of spatial alteration of c-di-GMP within the small intestine remain to be determined although I speculate c-di-GMP could be modulating biofilm formation, motility, and virulence gene expression. My model predicting high c-di-GMP concentrations in the lumen and decreased c-di-GMP proximal to the intestinal epithelial cells is consistent with previously described virulence gene regulatory models for V. cholerae and S. enterica [120, 128, 129]. Intestinal Lumen Bile ↑ Bicarbonate (pH) ↓ c-di-GMP ↑ Mucus Lining Bile ↓ Bicarbonate (pH) ↑ c-di-GMP ↓ Intestinal Epithelial Cells Intestinal Epithelial Cells Figure 19 19. Proposed model of V. cholerae c-didi-GMP regulation in the human small intestine. C-di-GMP is elevated in the lumen, where the concentration of bile is elevated and the concentration of bicarbonate is low. Upon entry into the mucosal layer where bile is low and the bicarbonate concentration is elevated, c-di-GMP is repressed.. 70 My findings indicate that both bile and bicarbonate are environmental cues that modulate c-di-GMP signaling in V. cholerae and facilitate the transition from aquatic environments to the human host. They suggest that modulation of c-diGMP levels by V. cholerae upon entry into the human host is more complex than previously appreciated, and that both bile and bicarbonate act together to inversely regulate the intracellular concentration of c-di-GMP to presumably enable the bacteria to adapt and thrive in the diverse intestinal environment. Future directions In the search to identify specific environmental signals that affect c-di-GMP levels, I found that bile acids increase intracellular c-di-GMP in V. cholerae. Furthermore, I showed that bile acids increase the c-di-GMP synthesis of three DGCs. However, the mechanisms that govern the behavior of these three enzymes are unknown. One common attribute that these proteins share is that they all have predicted trans-membrane domains. It is possible that changes in the cell membrane caused by bile acids could serve as a signal to alter DGC activity. In support of this, it has been shown that temperature mediated changes in membrane properties can alter protein behavior and initiate phenotypic changes [130, 131]. I hypothesize that bile acids are acting through DGC transmembrane domains to influence c-di-GMP synthesis by altering the physical properties of the lipid membrane. To test this hypothesis, I propose to examine the effect of temperature on the bile-induced increase c-di-GMP. V. cholerae cultures can be grown at either 35 ºC or 20 ºC in the presence and absence of bile acids. Membrane fluidity can be 71 assessed using fluorescence anisotropy as previously described [132] and intracellular c-di-GMP can be measured using LC-MS/MS. In addition, the VC1067, VC1372, and VC1376 ectopic expression strains and their mutant allele counterparts also should be examined at the two temperatures in the presence and absence of bile. If membrane fluidity is affecting the c-di-GMP synthesis of these proteins, I expect reduced c-di-GMP synthesis in the presence of bile at the lower temperatures. Alternatively, it is possible that bile induces composition changes of the lipid membrane of V. cholerae, which then alters the c-di-GMP synthesis of the three bile-responsive DGCs. This phenomenon has been observed in Lactobacillus reuteri, where bile acids alter the lipid membrane profile [133]. To test the effect of membrane composition on c-di-GMP synthesis, I propose ectopically express fatty acid biosynthesis enzymes to artificially modify the fatty acid composition of the cell membrane. Increased expression of the fatty acid biosynthesis gene fabF in B. subtilis promotes the proportion of long-chain fatty acids leading to decreased membrane fluidity, while expression of the gene fabH promotes branched-chain fatty acids and increases membrane fluidity [132, 134]. Either the fabF gene homolog (VC2019) or the fabH gene homologs (VC2023, VCA0751) can be ectopically expressed alongside each of the three bile-responsive DGCs, and intracellular c-diGMP can be measured using LC-MS/MS. Fluorescence anisotropy can confirm changes in membrane fluidity. I hypothesize that the DGCs expressed alongside fabF (reduced membrane fluidity) will have reduced c-di-GMP synthesis, whereas 72 DGCs expressed alongside fabH (increased membrane fluidity) will have increased c-di-GMP synthesis. In addition to the three DGCs that increased c-di-GMP synthesis in the presence of bile, I demonstrated that bile acids decrease the expression of one PDE, VC1295. However, the regulators driving bile-mediated expression changes of VC1295 are undefined. As I have identified putative ToxR binding sites in the promoter region of VC1295, and I have demonstrated that a ∆toxR mutant reduces induction of c-di-GMP in the presence of bile acids, I hypothesize that ToxR binds to the promoter region in the presence of bile acids to inhibit VC1295 expression. In support if this, it has been shown that the transcriptional regulation activity of ToxR is altered in the presence of bile acids, and that this regulation is dependent on the transmembrane domain of ToxR [117]. To test the hypothesis that bile and ToxR drive VC1295 expression, I propose introducing the VC1295-lux vector in the ∆toxR mutant strain and examining the expression of VC1295 in the presence and absence of bile. I expect that the luminescence will be unchanged in the presence of bile acids in the ∆toxR mutant strain. Electromobility gel-shift assays (EMSA) can be performed as previously described [107] using purified ToxR protein and DNA from the promoter region of VC1295 in the presence of increasing amounts of purified bile acids [135]. If ToxR is interacting with the VC1295 promoter in the presence of bile, I expect to see increased binding of DNA in the presence of bile acids. Finally, I propose creating mutations in the predicted ToxR binding sites on the VC1295-lux promoter vector, and examining luminescence in the WT strain in 73 the presence and absence of bile acids. These experiments will elucidate the role ToxR plays in altering the expression of VC1295 and increasing intracellular c-diGMP concentrations in the presence of bile acids. While bile acts to increase intracellular c-di-GMP in V. cholerae, I have shown that bicarbonate can quench this response; however, it remains unclear how this occurs. To explore this question, I propose to comprehensively examine the expression of genes in V. cholerae cultures grown in LB, LB with bicarbonate, LB with SHB, and LB with SHB and bicarbonate. To do this, RNA can be extracted from each culture and whole transcriptome shotgun sequencing (RNAseq) can be used to quantify gene expression. I expect that differences in gene expression will be observed in cultures grown in the presence of SHB versus SHB with bicarbonate. After genes have been identified that are differentially regulated in the presence of bicarbonate, genomic knockout mutant strains can be made of these target genes and the intracellular c-di-GMP of these strains can be determined when grown in the presence of SHB and SHB with bicarbonate. I demonstrated that the deletion of the enzymes responsible for the bilemediated c-di-GMP induction reduces bile-induced biofilm formation. However, the full physiological implications of increased c-di-GMP in the presence of bile and subsequent consequences on the pathogenesis of V. cholerae have yet to be determined. To ascertain if increased c-di-GMP is beneficial for bile tolerance, I propose competing the WT V. cholerae strain and the V. cholerae quadruple DGC/PDE mutant strain in the presence and absence of bile and measuring relative 74 fitness. Neutral markers can be inserted into the genomes of each strain, such as antibiotic resistance genes or fluorescent proteins, to differentiate the strains. I hypothesize that the quadruple mutant will have lower competitive fitness in the presence of bile acids compared to the WT strain. As controls, the ∆vpsL strain should be competed against the WT strain to determine the role of biofilm formation plays in competitive fitness. Likewise, as positive controls, ∆toxR and ∆ompU mutants should be competed against WT in the presence and absence of bile, as these genes are important for bile resistance and fitness [43, 116, 136]. To determine the role bile-induced c-di-GMP plays in vivo, I propose determining the competitive index of the quadruple DGC/PDE mutant to the WT strain of V. cholerae in an infant mouse model as previously described [35, 116]. While the bile composition of a mouse is different from that of a human [137], the infant mouse model has been used to study the effects of proteins related to bile resistance in pathogenesis [45, 46, 116]; thus I expect that this model will be sufficient to examine the role of c-di-GMP in bile-related fitness. I expect that the quadruple mutant has significantly reduced competitive index in a mouse model. These experiments will clarify the role c-di-GMP plays in vivo during V. cholerae pathogenesis. 75 CHAPTER 4 – Intracellular cc-didi-GMP Negatively Correlates with the Growth Rate and Ribosomal RNA of V. cholerae Preface It has been observed that ectopic DGC expression can cause a growth deficiency in both V. cholerae and E. coli [4, 79]. It remains unknown as to why ectopic DGC expression would affect growth rate. Many factors influence bacterial growth rate; the growth rate of bacteria is positively correlated with the ribosomal RNA (rRNA) gene copy number, and this phenotype is associated with how bacteria respond to resource availability [138]. Furthermore, ribosome abundance in the cell correlates with growth rate, although it remains unclear if ribosome depletion inhibits growth rate or if growth rate inhibition causes ribosome depletion [139146]. In support of the former, it has been shown that (p)ppGpp accumulates in the cell upon nutrient deprivation, altering GTP homeostasis and inhibiting RNA synthesis in the cell resulting in reduced growth rate [143, 147-151]. There are numerous phenotypes controlled by c-di-GMP including motility, biofilm formation, and virulence [11, 13, 16, 98]. Here, I examine the hypothesis that c-di-GMP negatively regulates the growth rate of V. cholerae. I have shown that there is a correlation between intracellular c-di-GMP and growth rate of V. cholerae grown in different media. Furthermore, I have shown that ectopic DGC expression represses the growth rate in V. cholerae in a dose-dependent manner, and that the efficiency of this repression varies between different DGCs. Finally, I 76 have shown that ectopic expression of the DGC VC1067 reduces rRNA in the cell. Together, I propose that c-di-GMP is negatively regulating growth rate in V. cholerae by reducing intracellular ribosomes via an unidentified molecular mechanism. Measuring growth rate of V. cholerae Growth curves were determined by taking OD600 readings of liquid cultures at regular intervals. For cultures grown in 150 µL volumes in clear 96-well plates (Greiner), OD600 readings were taken with a SpectraMax M5 plate reader (Molecular devices) every 10 minutes for a period of 15 hours. For cultures grown in 2 mL volumes in 14 mL round-bottom polystyrene tubes (Falcon), OD600 readings were taken with a DU 700 series spectrophotometer (Beckman Coulter) approximately every 30 minutes for a period of 6 hours. All cultures were grown at 35 ºC, with shaking at 220 rpm, and DGC expression strains were supplemented with kanamycin at 100 µg/mL. Relative growth rates were determined by plotting each curve in Graphpad Prism and calculating area under the curve. The area under the curve values for each strain at each inducer concentration were normalized to the area under the curve of a no-inducer control. For calculating specific growth rates, growth curves were analyzed with the grofit package in R [152, 153]. C-di-GMP was quantified using LC-MS/MS as described in Chapter 2. 77 RNA extractions and quantification For RNA quantification, 5mL of LB media supplemented with 100 µg/mL kanamycin and 0.1 mM IPTG in 25 mL Erlenmeyer flasks were inoculated with 5 µL overnight V. cholerae cultures and grown to an OD600 of approximately 0.6-0.8 (late exponential growth). Cultures were then diluted to an OD600 of 0.5, and RNA extractions were performed using Trizol reagent (Ambion, Life Technologies) as per the manufacturer’s instructions. Total RNA was quantified using a NanoDrop 1000 (Thermo Scientific). rRNA was then quantified in each sample using a 2100 Bioanalyzer (Agilent). Intracellular c-di-GMP and growth rate of V. cholerae are correlated In Chapter 2, I showed that intracellular c-di-GMP concentrations of V. cholerae vary when the bacteria are grown in different growth media (Fig. 5) and I observed that these concentrations tended to be higher in minimal media, where the growth rate was slower (Fig. 4). I hypothesized that intracellular c-di-GMP may be negatively correlated with bacterial growth rate. To examine this hypothesis that cdi-GMP and growth rate are connected, I determined the doubling times (in minutes) during exponential growth of V. cholerae grown in different media (Fig. 4) by selecting two timepoints during exponential growth in each environment and logtransforming the fold-change of OD600-2 and then multiplying this value by the time elapsed. The doubling time of V. cholerae grown in LB determined here is consistent with previous studies [154]. Growth rate (divisions per hour) was determined by 78 dividing 1 hour by the doubling time; these growth rates were then regressed on intracellular c-di-GMP concentrations (Fig. 20). I found that there was a negative trend of intracellular c-di-GMP and growth rate, although the slope bordered outside of statistical significance from zero (p=0.056). Furthermore, the goodness of fit (R2=0.163) indicates that only a modest portion of this relationship can be attributed to intracellular c-di-GMP. This analysis provides preliminary evidence Intr acellular c -di-GMP (nM) that there is a relationship between intracellular c-di-GMP and growth rate. 3000 2000 1000 0 1 2 3 4 Growth Rate (Doublings / Hour) Figure 20 20. Linear regression of intracellular cc-didi-GMP concentrations on growth rates of V. cholerae grown in different media. Dotted lines indicate 95% confidence intervals. Goodness of fit (R2) was 0.163, and non-zero significance (p) was 0.056. Growth rate inhibition caused by ectopic DGC expression is dose-dependent Previous studies have indicated that ectopic expression of a DGC can inhibit bacterial growth [4, 79], corroborating my hypothesis that c-di-GMP negatively affects growth rate. I wanted to further explore the dynamics of the relationship 79 between growth rate repression and ectopic DGC expression; specifically, I sought to confirm these previous observations and determine if this inhibition of growth is DGC dose-dependent. I utilized two DGC inducible expression vectors encoding VC1067 and VC2285 to alter the intracellular concentration of c-di-GMP. These two DGCs have been shown to synthesize high levels of intracellular c-di-GMP upon IPTG induction (Chapter 3, [79]). I measured the growth of V. cholerae strains harboring these DGC expression vectors at IPTG concentrations ranging from 0.002 mM to 0.1 mM by growing the strains in a 96-well plate and taking OD600 readings every 10 minutes for 15 hours. The growth rate of each strain at each IPTG concentration was determined by fitting each growth curve to an exponential growth model using the program grofit [152, 153], and determining the maximum slope of each curve (Fig. 21). In an empty vector negative control, I observed no change in growth rate across the inducer gradient. In contrast, both the VC1067 and VC2285 expression strains showed a dose-dependent reduction in growth rate. The VC2285 strain showed a greater reduction in growth at all timepoints compared to the VC1067 strain. This is consistent with observations that the VC2285 strain produces more c-di-GMP that the VC1067 strain [79]. I conclude from these data that ectopic DGC expression inhibits growth rate in a dosedependent manner. I next sought to determine what concentrations of c-di-GMP lead to growth inhibition in V. cholerae. I selected nine DGC expression strains and grew them in 8 different IPTG concentrations ranging from 0.0005 mM to 1.0 mM. The relative 80 Max Gro wth Rate (mu) 0.003 Vector Control VC1067 VC2285 0.002 0.001 0.000 0.001 0.01 0.1 IPTG Conce ntratio n (mM) Figure 21 21. Growth rates of V. cholerae DGC expression strains. strains. V. cholerae DGC expression cultures were grown in a 96-well plate and growth rate was determined at a range of IPTG concentrations. The max growth rate (mu) indicates the maximum slope during log-phase growth of OD600 curves fit to exponential growth models. Each point represents the mean of three replicates, and error bars indicate standard deviation. growth rate of each strain was determined by measuring OD600 at regular intervals, then determining the area under each curve. Intracellular c-di-GMP of each strain was determined at each inducer condition during late-stationary growth using LCMS/MS as previously described [79]. Then, the growth rate of each strain at each inducer concentration was regressed on the intracellular c-di-GMP concentrations. I found that 4 of the 9 regressed strains had slopes significantly different from zero, indicating that these DGCs are capable of inhibiting growth of V. cholerae upon ectopic expression (Table 1). 81 Expression strain Max intracellular c-di-GMP (nM) Min Area under the curve Slope Nonzero significance (p < 0.01) r2 -1.0e-5 * 0.88 ± 8.1e-7 -8.5e-5 VC1104 3476.9 0.64 ± 0.05 * 0.65 ± 1.3e5 -2.8e-5 VC1216 5840.2 0.86 ± 0.08 0.12 ± 2.0e-5 -7.6e-5 VC1353 1596.2 0.94 ± 0.07 0.14 ± 4.1e-5 1.5e-4 VC1599 888.7 0.91 ± 0.04 0.07 ± 1.1e-4 -2.4e-5 VC2224 22249.3 0.5 ± 0.05 * 0.83 ± 2.3e-6 -1.3e-4 VC2454 809.5 0.82 ± 0.03 0.03 ± 1.5e-4 -1.1e-4 VCA0074 3515.7 0.51 ± 0.09 * 0.59 ± 2.0e-5 -2.6e-4 VCA0165 541.0 0.88 ± 0.07 0.20 ± 1.1e-4 Table 1. Description of DGC expression strain intracellular cc-didi-GMP levels and growth inhibition. Max intracellular c-di-GMP indicates the highest level of intracellular c-di-GMP over the IPTG gradient. Min area under the curve indicates the lowest area under the curve value of each strain grown in the corresponding IPTG concentration. Linear regression analysis of c-di-GMP and growth rate was performed, and the resulting slopes are displayed here. The nonzero statistical significance and the r2 values are also shown. VC1067 98111.4 0.11 ± 0.03 To compare the concentrations of c-di-GMP required to inhibit growth rate across different DGCs, the maximum c-di-GMP produced by each DGC expression strain capable of inhibiting growth was regressed on the minimum growth rate observed for each strain. I separately analyzed the DGC expression strains that were unable to significantly alter growth rate (Fig. 22A). I found that strains expressing DGCs capable of inhibiting growth rate did so in a c-di-GMP dose82 dependent manner, where the slope demonstrated a significant negative deviation (p < 0.05). Interestingly, two DGCs (VC1104 and VCA0074) inhibited growth rate at c-di-GMP concentrations less than that of the DGC VC1216, which did not show any significant growth rate inhibition. I went on to analyze the dynamics of growth inhibition of the four DGCs that inhibited growth rate of V. cholerae. Two of the DGC expression strains, VC1104 and VCA0074, produced similar levels of c-di-GMP upon expression resulting in similar levels of growth rate inhibition (Fig. 22B). In contrast, the VC1067 and VC2224 strains required significantly more intracellular c-di-GMP to achieve the same level of growth rate inhibition as either the VC1104 or VCA0074 strains. From this, I conclude that certain DGCs are able to inhibit growth rate at lower concentrations of c-di-GMP than other DGCs, suggesting there is signaling specificity within this pathway. Intracellular c-di-GMP concentration required for growth inhibition varies for different DGCs The growth rate inhibition of ectopic DGC expression has been attributed to c-di-GMP toxicity [4]. However, my data indicate that DGCs inhibit growth rate at different concentrations of c-di-GMP, suggesting that this inhibition is occurring independent of generalized c-di-GMP toxicity. I hypothesize that c-di-GMP is regulating growth rate in V. cholerae through a specific, unknown mechanism. I selected two DGCs that inhibit growth rate at different c-di-GMP concentrations; 83 Re lative Gro wth Rate (Are a under the curve ) A 1.2 text 0.8 0.4 0.0 10 3 10 4 10 5 B Relative Growth Rat e (Area under the curve) Intrace llular c-di-GMP (nM) VC1067 VC1104 VC2224 VCA0074 1.0 0.5 0.0 10 2 10 3 10 4 Intrace llular c-di-GMP (nM) 10 5 Figure 22 22. Linear regression of intracellular cc-didi-GMP concentrations on growth rates of V. cholerae DGC expression strains. (A) Regression analysis of maximum intracellular c-di-GMP and maximum growth rate of different DGCs of V. cholerae. Dotted lines indicate linear regression of DGCs that significantly inhibit growth (black circles) or DGCs that do not significantly inhibit growth (grey squares). (B) The intracellular c-di-GMP concentrations of individual DGCs was regressed on relative growth rates over a range of IPTG concentrations. Each point represents the mean of three replicates. Error bars indicate standard deviation. 84 these DGCs are VC1067 and VCA0074. Intracellular c-di-GMP was measured using LC-MS/MS at an inducer concentration that was predicted to produce equal levels of growth rate inhibition at different c-di-GMP concentrations, which was 0.1 mM IPTG (Fig. 22B). I found that VC1067 produced 4.0-fold more intracellular c-diGMP at this inducer concentration than VCA0074 (p < 0.05, Fig. 23). In contrast, intracellular c-di-GMP concentrations of strains containing mutant alleles (GG[D/E]EF -> AA[D/E]EF) of either VC1067 or VCA0074 were indistinguishable from those of the vector control. I then measured growth of the VC1067 and VCA0074 strains grown at an IPTG concentration of 0.1 mM to confirm that ectopc expression of each of these DGCs inhibits growth rate at equivalent levels (Fig. 24A). These curves were used to calculate the doubling time of each strain by fitting each curve to an exponential growth equation using grofit [152, 153] and deriving the maximum slope. The doubling time was determined from the slope by dividing log10(2) by each slope value. I found that the doubling time of both the VC1067 and VCA0074 expression strains was significantly increased over the vector control and their respective active site mutant alleles (Fig. 24B). To account for growth rate deficiency caused by the burden of ectopically expressing each DGC, the doubling time of the VC1067 and VCA0074 expression strains was divided by the doubling time of their respective mutant allele counterpart expression strains. The VC1067 strain demonstrated a 1.47-fold increase in doubling time over its mutant allele, whereas the VCA0074 strain demonstrated a 1.45–fold increase in doubling time. I thus 85 50000 30000 * VC A 00 74 00 74 VC A VC 10 67 VC 10 C on tr ct or Ve 67 * 10000 ol Intrac ellular c-di-GMP (nM) 70000 DGC Expression Strain Figure 23 23. Intracellular cc-didi-GMP concentrations of V. cholerae DGC expression strains. strains. Cultures were grown in the presence of 0.1 mM IPTG. Bars indicate the mean of three replicates, and error bars indicate standard deviation. Brackets indicate statistical significance, which was determined using a two-tailed Students T-test (p < 0.05). conclude that while the VC1067 strain synthesizes significantly more c-di-GMP at this inducer concentration compared to the VCA0074 strain, the growth rate is virtually indistinguishable. Increased intracellular c-di-GMP reduces the expression of rRNA As my evidence indicates that intracellular c-di-GMP is negatively impacting growth rate, I sought the mechanism governing this phenomenon. Numerous studies have demonstrated that growth rate and ribosome abundance are inextricably linked [139-151]. I hypothesized that c-di-GMP negatively affects the transcription of rRNA in V. cholerae. This would lead to decreased ribosomes in the 86 A OD 600 1 Vector Control VC1067 VC1067* VCA0074 VCA0074* 0.1 0.01 0 100 100 Doubling Time (min) B 200 Time (min) 1.47-fold 80 300 1.45-fold * * 60 40 VC 10 67 V C 10 67 * V C A 00 74 V C A 00 74 * V ec to r C on tr ol 20 DGC Expre ssio n Strain Figure 24. 24. Growth rates of V. cholerae DGC expression strains. (A) Growth curves of DGC expression strains grown in the presence of 0.1 mM IPTG. OD600 readings were taken at regular intervals (B) Doubling times were determined from growth curves. Each bar represents the mean of three replicates, and error bars indicate standard deviation. * indicates statistical significance from their respective mutant alleles, as determined by a Students two-tailed T-test (p < 0.05). 87 cell, resulting in slower protein synthesis and subsequently slower cell division. To examine this hypothesis, RNA extractions were performed on the VC1067 expression strain and its mutant allele on cultures normalized to equal cell density. 16S and 23S rRNA was quantified using a 2100 Bioanalyzer (Agilent). Relative RNA abundance was determined by normalizing the respective RNA abundance by the initial RNA concentration of each sample. I found that strains ectopically expressing VC1067 strain had 39.6% and 32.2% less 16S and 23S rRNA, respectively, compared to the VC1067* mutant allele (p < 0.05, Fig. 25). These data support the hypothesis that c-di-GMP regulates growth rate by altering ribosome abundance in the cell. Discussion The regulation of bacterial growth rate has been the subject of study for over 50 years [146]. Indeed, the rate at which bacteria are able to reproduce has large implications for pathogenesis and epidemiology of disease, and involves core biological functions that are widely conserved amongst bacteria. The concept of a second messenger controlling bacterial growth rate is not novel. It has been shown that the small nucleotide signal (p)ppGpp regulates growth in response to metabolic stresses [143, 147-151]. Here, I explore the connection between growth rate regulation and c-di-GMP. I have shown that there is a correlation between intracellular c-di-GMP and growth rate in V. cholerae grown in different media, as well as V. cholerae ectopically expressing different DGCs (Fig. 20, 21, 22). It has 88 80000 60000 40000 16S VC1067 VC1067* 0 VC1067* 20000 VC1067 Re lative RNA Abundance 100000 23S Figure 25. 25. Relative rRNA abundance of V. cholerae DGC expression strains. The rRNA of RNA extractions from V. cholerae strains ectopically expressing VC1067 (black) or the active-site mutant allele VC1067* (grey) were quantified using a 2100 Bioanalyzer. Each bar represents the mean of five replicates, and error bars indicate standard deviation. Brackets indicate statistical significance, as determined using a Students two-tailed T-test (p < 0.05). been noted in previous studies that ectopic DGC expression leads to growth rate inhibition in E. coli and V. cholerae [4, 79]. I have shown that growth rate inhibition is dependent on the c-di-GMP synthesis of DGCs, as expression of DGC mutant alleles had no significant impact on growth rate (Fig. 24). This inhibition has been previously attributed to a c-diGMP toxicity associated with unnaturally high intracellular c-di-GMP concentrations in the cell. However, my data indicate that this is not necessarily the case, as growth rate inhibition was observed upon the expression of VC1104, while 89 expression of VC1216 and VC1353 produced equal or greater concentrations of c-diGMP while failing to inhibit growth (Fig. 22). Furthermore, I also observed that the DGCs VC1067 and VCA0074 produce different levels of intracellular c-di-GMP yet produce the same growth rate inhibition. These findings suggest that growth rate inhibition is subject to DGC signaling specificity, a phenomenon where c-di-GMP produced by certain DGCs effects downstream phenotypes at concentrations different from other DGCs. Other studies have validated the concept of high specificity c-di-GMP signaling for phenotypes such as biofilm formation [79, 91, 98, 155, 156]. Although the mechanisms by which c-di-GMP inhibits growth have not yet been fully characterized, I have shown that ectopic expression of a DGC can lead to decreased 16S and 23S rRNA in the cell (Fig. 25). This change in intracellular ribosomes could account for the decreased growth rate observed in cells containing elevated c-di-GMP (Fig. 24). Growth rate has been correlated with rRNA gene copy number in bacteria, and it is well established that ribosome abundance in the cell is correlated with growth rate [139-151]. Furthermore, it has been shown that disrupting the rRNA copy number in E. coli has negative implications for fitness [157]. Transcriptional regulation of rRNA has been associated with nutrient abundance; specifically, during periods of starvation, the stringent response is activated, resulting in increased (p)ppGpp and simultaneous repression of rRNA transcription [143, 147-151]. Thus it makes sense that c-di-GMP, a second 90 messenger affected by environmental factors, would regulate growth rate by altering ribosome abundance. From these studies, I postulate two hypotheses to explain the phenomenon of c-di-GMP-dependent growth rate inhibition I observe. The first hypothesis is that cdi-GMP is inhibiting rRNA transcription through some yet unidentified effector protein. I propose that a subset of DGCs, such as VC1104 and VCA0074, specifically regulate growth rate through this unknown factor. It is possible c-di-GMP acts through a similar mechanism as (p)ppGpp, which binds RNA polymerase to regulate transcription [158]. The second hypothesis is that DGC expression is leading to a depletion of GTP, which subsequently alters growth rate. Previous studies have shown that GTP depletion can lead to growth rate inhibition in B. subtilis [159]. Furthermore, the growth rate regulator (p)ppGpp has also been shown to be a critical regulator of GTP homeostasis [151]. I propose that DGCs that produce very high amounts of c-di-GMP, such as VC1067, inhibit growth by disrupting GTP homeostasis. These hypotheses have not yet been tested. Future directions C-di-GMP has been shown to regulate numerous bacterial phenotypes, and more have yet to be identified [73]. I have shown that growth rate and c-di-GMP are correlated and that ectopic expression of select DGCs results in c-di-GMP dependent growth rate inhibition. This is significant, as the growth rate has important implications in bacterial ecology and pathogenesis [138, 143, 157, 160]. The 91 mechanism governing the c-di-GMP mediated inhibition of growth rate is unknown. I have shown that the DGCs VC1067 and VCA0074 produce significantly different levels of c-di-GMP in the cell yet result in the same level of growth rate inhibition. In order to explore this relationship and determine how this growth rate inhibition is occurring, I propose extracting RNA from both the VC1067 and VCA0074 expression strains. RNAseq can then be performed on extracts from both cultures to identify differentially regulated transcripts. As controls, the vector control and the active site mutant allele strains should be included. Since these two DGCs inhibit growth at different concentrations of c-di-GMP, they may be acting through two distinct pathways. I propose focusing on differentially regulated genes targeting known growth rate factors, metabolic enzymes, and genes of unknown function. V. cholerae deletion mutants of promising targets will be made, and subsequently analyzed for growth in the presence and absence of elevated intracellular c-di-GMP. An alternative method for determining the mechanism of c-di-GMP mediated growth rate inhibition is to utilize experimental evolution to generate V. cholerae strains that do not exhibit growth rate inhibition at high c-di-GMP concentrations. To do this, I propose creating a c-di-GMP strain with a neutral-site chromosomal insertion of the VC2285 gene under the control of the pTac promoter. VC2285 will be used due to its ability to synthesize very high levels of c-di-GMP and its robust inhibition of growth rate (Fig. 21, [79]). As I have shown that expression of VC2285 inhibits growth, I postulate that if VC2285 is grown and subsequently passaged repeatedly, spontaneous suppressor mutants will arise in the population that will 92 be able to grow faster in the presence of high levels of c-di-GMP. These mutants will be able to out-compete their ancestors and overtake the population over the course of the experiment. This strain can be passaged in the presence of the inducer IPTG once a day; when cultures arise that exhibit faster growth rates than their ancestors, members of the culture population will be isolated and whole-genome sequencing will be used to identify where on the chromosomes the mutations occurred. Alternatively, it is possible that while c-di-GMP and growth rate are correlated, c-di-GMP is not directly regulating growth rate. An alternative hypothesis is that growth rate is regulating intracellular c-di-GMP. This could be occurring through protein dilution [161], where protein synthesis cannot keep up with cell division during rapid growth, resulting in different DGC and PDE abundance in the cell and subsequently altering intracellular c-di-GMP concentrations. To examine this hypothesis, I propose artificially altering the growth rate of V. cholerae using a continuous culture chemostat, and measuring intracellular c-di-GMP at different growth rates. I hypothesize that c-di-GMP would be lower at faster dilution rates where the culture is growing faster, and higher c-diGMP in the slower dilution rates where growth rate is lower. Another mechanism by which DGC expression could be altering growth rate is by interacting with the (p)ppGpp signaling network. It has been shown that (p)ppGpp negatively regulates growth rate in response to nutrient deprivation [143, 147-151] . I postulate that c-di-GMP regulates (p)ppGpp synthetases and hydrolases 93 to alter intracellular (p)ppGpp concentrations and subsequently regulates growth rate. To examine this hypothesis, I propose to use LC-MS/MS to quantify (p)ppGpp of the VC1067 DGC expression strain and its mutant allele counterpart over a range if inducer concentrations. I expect that the VC1067 strain would have elevated (p)ppGpp concentrations compared to the VC1067* mutant strain, and that this relationship would be dose dependent. One important regulatory target of (p)ppGpp is GTP homeostasis [151]; this is particular relevance, as c-di-GMP is synthesized from GTP. To determine if there is any link between intracellular c-diGMP, (p)ppGpp, and GTP, I propose to also measure the intracellular concentration of GTP and GTP precursors of cells expressing either VC1067 or the VC1067* mutant allele. Concentrations of GTP are approximately 5 mM in E. coli [88]; in contrast, intracellular c-di-GMP of the VC1067 expression strain in the presence of 0.1 mM IPTG (where growth rate is inhibited) is approximately 50 µM. Thus, as one c-di-GMP molecule is synthesized from two GTP molecules and assuming V. cholerae has similar intracellular GTP homeostasis, I would expect that ectopic expression of VC1067 could detract approximately 2% of the normal intracellular GTP pool. Whether the extent of these changes on intracellular GTP or fluxes of GTP precursors would be enough to effect growth rate is yet to be determined. 94 CHAPTER 5 - A novel Adenovirus Adjuvant Stimulates Innate Immunity by in vivo cyclic didi-GMP Synthesis Preface Thus far, I have described the interaction between environmental factors and intracellular c-di-GMP, and its impact on the physiology of V. cholerae. As a byproduct of this research, I have determined which of the V. cholerae DGCs actively produce c-di-GMP (Fig. 2, 10, 13, 22, 23). I sought to use this information to develop a clinical application. Other studies have demonstrated that c-di-GMP has potent immunostimulatory properties [49-57]. I postulated that a DGC from V. cholerae could be introduced into a eukaryotic cytoplasm to synthesize c-di-GMP in vivo and subsequently activate the innate arm of the immune system. One vehicle that has been used for antigen and adjuvant delivery is replication deficient adenovirus based vectors [162]. Adenovirus vectors transduce large fragments of DNA into a wide range of cells in order to synthesize proteins in vivo, and gene expression can be modulated and even localized to specific cell types. Unlike other types of viral delivery systems, DNA delivered by adenovirus vectors does not integrate into the genome and thus circumvents the danger of insertional mutagenesis [163]. Additionally, adenovirus vectors can be produced cost-efficiently in high abundance. Importantly, adenovirus vectors are currently being used in human clinical trials world-wide [164]. 95 Here I have investigated whether an adenovirus vector could be used to deliver DGC DNA into eukaryotic cells to synthesize c-di-GMP in vivo, manipulating a host-pathogen interaction to activate the innate arm of the immune system and improve the immune response against co-administered antigen. I have constructed an adenovirus containing an allele encoding the V. cholerae DGC VCA0956. I have shown that this DGC is capable of synthesizing c-di-GMP when virally transduced into cell culture lines. Furthermore, I have demonstrated that use of the VCA0956 adenovirus construct in a murine system results in c-di-GMP synthesis and subsequent upregulation of genes associated with innate immunity as well as increased secretion of numerous cytokines and chemokines. Finally, I have shown that c-di-GMP produced by VCA0956 in vivo functions as an adjuvant by reducing the effective dose of a C. difficile toxin A antigen necessary to generate a potent immune response. I propose that this novel adenovirus c-di-GMP delivery system offers a more efficient and cost-effective method for use of c-di-GMP as an adjuvant to stimulate innate immunity. DNA manipulation and virus construction All strains, plasmids, and primers used are listed in Tables A1-3. All of the DNA manipulation and plasmid construction was performed as previously described [82]. The DGC alleles were amplified from V. cholerae El tor strain C6706 or the plasmids pCMW75 and pCMW98 using the DNA polymerase Phusion (New England Biolabs) and the oligonucleotides listed in Table A3 (IDT). This product was then inserted into the plasmid pShuttle-CMV [165] by digesting with Kpn1 and 96 XhoI (Fermentas), and then ligated with a T4 DNA ligase (Invitrogen). E. coli strain DH10B (Invitrogen) was used for harboring plasmid DNA, and sequence fidelity was confirmed by sequencing (Genewiz). The active site mutant alleles were generated using the QuickChange Lightning site-directed mutagenesis kit (Agilent). A first-generation, human Adenovirus type 5-derived (Ad5) replication deficient vector (deleted for the E1 and E3 genes) was used in this study [166]. Recombination, viral propagation of the Ad5 vectors, and subsequent virus characterization was performed as previously described [166, 167]. Viral particle number and infectious units titer (TCID/ml) was determined by optical density measurement at 260 nm and validated as previously described [168].Construction of the Ad5-Null and Ad5-TA is described elsewhere [169, 170]. All virus constructs were confirmed to be RCA negative using RCA PCR and direct sequencing methods [166] and the bacterial endotoxin content was found to be < 0.15 EU per mL [166]. All procedures with recombinant adenovirus constructs were performed under BSL2 conditions. HeLa cell transformation All transfections of plasmid DNA into HeLa cells was performed with the TransIT-HeLaMONSTER transfection kit (Mirus) in 6-well plates with 2.5 µg plasmid DNA. For HeLa cell infections with adenovirus vectors, cells were infected with 2.0*109 viral particles (multiplicity of infection, M.O.I. of 500). Cell cultures were all checked for confluence and morphology before and after transfection and 97 infection using microscopy. After 24 hours of growth at 37 °C in 5% CO2, the cells were dissociated using 300 µL 0.25% trypsin, and then cells were resuspended in 4 mL PBS and then pelleted by centrifugation at 1600 RPM at 4 °C. Afterwards the cells were resuspended in 100 µL extraction buffer (40% acetonitrile, 40% methanol, and 0.1 N formic acid). The cell lysate was incubated at -20 °C for 30 minutes, and then centrifuged at max speed for 10 minutes. The extraction buffer was removed from the pelleted debris and stored at -80 °C until analysis. Immediately prior to analysis, the extraction buffer was evaporated using a vacuum manifold, and the samples were rehydrated in 100 µL water. C-di-GMP was quantified as described in Chapter 2. Animal protocols Adult BALB/c WT male mice (6-8 weeks old) were used for all animal experiments (Jackson Laboratory). For c-di-GMP quantification and innate studies, mice were anesthetized using isofluorane, and 2x1011 adenovirus viral particles (vp) per mouse (200 µL total volume, suspended in PBS) were administered intravenously (IV) via retro-orbital injection. After administration, mice were monitored every 6 hours by lab personnel for mortality and other health parameters in accordance with Michigan State EHS and IACUC. After 24 hours the mice were sacrificed, and the spleen and liver were isolated from each animal. Each tissue was placed in 500 µL PBS, and then the tissue suspension was homogenized using an Omni Tissue Homogenizer (Omni International). 300 µL of homogenate was added to an equal volume of equilibrated Phenol Solution (Sigma). The homogenate-phenol 98 solution was then vortexed and then centrifuged at 15,000 rpm for 10 minutes. The aqueous phase was removed and added to 500 µL chloroform. The mixture was vortexed and then centrifuged at 15,000 rpm for 10 minutes. The aqueous phase was then removed and stored at -80 °C until analysis. Quantitative PCR Quantitative PCR was used to determine adenovirus abundance from DNA extracted from liver tissue as previously described [171]. Ad5 genome copy numbers were quantified using an ABI 7900HT Fast Real-Time PCR system and the SYBR Green PCR Mastermix (Applied Biosystems) in a 15 µL reaction using a primer set for the Ad5 Hexon gene that has been previously described [172]. All PCRs were subjected to the following procedure: 95.0 °C for 10 minutes, followed by 40 cycles of 95.0 °C for 15 seconds and 60.0 °C for 1 minute. Standard curves to determine the number of viral genomes per liver cell were run in duplicate and consisted of 6 halflog dilutions using DNA extracted from purified Ad5 virus [166]. As an internal control, liver DNA was quantified using primers spanning the GAPDH gene [171] and standard curves were generated from total genomic DNA. Melting curve analysis was performed to confirm the quality and specificity of the PCR (data not shown). To determine relative abundance of specific liver-derived RNA transcript, reverse transcription was performed on RNA derived from the liver tissue using SuperScript III (Invitrogen) and random hexamers (Applied Biosystems) as per the 99 manufacturer’s instruction. RT reactions were diluted to a total volume of 60 µL, and 2 µL from each sample was used as template for subsequent PCR. Quantitative PCR was subsequently performed as described above using an ABI 7900HT Fast Real-Time PCR system and SYBR Green PCR Mastermix (Applied Biosystems) using primer sets that have been previously described [166]. The comparative Ct method was used to determine relative gene expression using GAPDH to standardize expression levels across all samples. Relative expression changes were calculated by comparing experimental levels of liver transcript to levels of liver transcript derived from mock-treated animals. Cytokine and chemokine quantification Cytokine and chemokine concentrations were quantified from plasma samples using a Bio-Plex multiplex bead array system (Bio-Rad). At 6 and 24 hours, blood samples were taken from mice using heparinized capillary tubes and EDTAcoated microvettes (Sarstedt). The samples were centrifuged at 3,400 rpm for 10 minutes to isolate plasma. Samples were assayed for 23 independent cytokines and chemokines (IL-1α, IL-1β, IL-2, IL-3, IL-4, IL-5, IL-6, IL-9, IL-10, IL12-p40, IL-13, IL-17, IFN-γ, G-CSF, GM-CSF, Eotaxin, KC, MCP-1, MIP-1α, MIP-1β, RANTES, and TNF-α) as per the manufacturer’s instructions (Bio-Rad) via Luminex 100 technology (Luminex). 100 ELISPOT analysis For adaptive immunity studies, mice were administered adenovirus ranging from 1x106 to 5x109 vp per mouse suspended in 25 µL PBS via IM injection into the tibialis anterior of the right hindlimb. To measure antigen specific recall responses, mice were sacrificed and the spleen was harvested after 14 days. Splenocytes were isolated and ex vivo stimulated with immunogenic peptides from C. difficile TA library as previously described [170]. ELISpot analysis for was performed as previously described [170] using 96-well multiscreen high-protein binding Immobilon-P membrane plates (Millipore) and the Ready-Set Go IFN-γ mouse ELISpot kit (eBioscience). Spots were photographed and counted using an automated ELISpot reader system (Cellular Technology). C. difficile challenge C. difficile mouse challenges were performed as previously described [173, 174]. Mice were administered Ad5 vectors; after 10 weeks, mice were administered cefoperazone in their drinking water at 0.5 mg / mL for 10 days, followed by 2 days with untreated drinking water. Immediately prior to the challenge, weights of each mouse were recorded. Mice were then administered 100 µL C. difficile CD2015 spores by oral gavage. Mice were monitored twice a day, and the endpoint was determined when mice lost ≥20% of their initial body weight. All animal procedures were reviewed and approved by the Michigan State University EHS and IACUC. Care for the mice was provided in accordance with 101 PHS and AAALAC standards. Plasma and tissue samples were collected and handled in accordance with the Michigan State University Institutional Animal Care and Use Committee. Generating an adenovirus harboring a V. cholerae DGC There is an abundance of literature demonstrating that c-di-GMP is able to act as a vaccine adjuvant by stimulating the innate immune system [59]. However, these studies utilized chemically synthesized c-di-GMP delivered intramuscularly or intravenously, which is cost prohibitive and inefficient in delivery. Because c-diGMP is synthesized from GTP and GTP is abundant in the cytoplasm of eukaryotic organisms, I postulated that a DGC expressed under the control of a strong eukaryotic promoter/enhancer element would lead to c-di-GMP synthesis within the eukaryotic cell and subsequent enhancement of downstream innate immune responses. To identify a DGC that would produce c-di-GMP in the cytoplasm of a eukaryotic cell, I examined the two DGCs VCA0848 and VCA0956 from V. cholerae, as V. cholerae is a well-studied model system for c-di-GMP signaling and many V. cholerae DGCs have been shown to synthesize c-di-GMP in high concentrations [79]. Additionally I examined the DGC qrgB from V. harveyi, which has been shown to synthesize c-di-GMP in V. cholerae [72]. I selected these three DGCs since they had no predicted N-terminal regulatory or trans-membrane domains and canonical GGDEF domains and active site motifs. Ectopic expression of VCA0956 has been shown to increase biofilm formation in both V. cholerae and Vibrio vulnificus [79, 102 175], repress motility in V. cholerae [123], and increase intracellular c-d-GMP in V. cholerae and Shewanella oneidensis [35, 77, 176]. To determine if any of these DGCs are able to synthesize c-di-GMP in a eukaryotic cytoplasm, I constructed plasmids containing each DGC under the control of the constitutive CMV promoter/enhancer in the plasmid pShuttleCMV. I also constructed corresponding vectors containing the same DGC alleles with a mutation in the active site of the GGDEF domain (GG[D/E]EF -> AA[D/E]EF). These plasmids were transfected into HeLa cells, and c-di-GMP levels were measured in cell lysates after 24 hours using LC-MS/MS. I found that eukaryotic cells transfected with both the VCA0848 and VCA0956 alleles produced detectable levels of c-di-GMP; in contrast, no detectable c-di-GMP was observed in cells transfected with the qrgB allele, or the active site mutant alleles and mock treatment controls (Fig. 26A). The VCA0848 vector produced 31.6-fold more c-diGMP than the VCA0956 vector. Cell cultures were checked by microscopy and I observed no discernible morphological differences between expression of VCA0848, VCA0956, and the controls. Furthermore, trypan blue staining indicated that treatment with VCA0848 or VCA0956 did not appear to impact overall cell viability. Additionally, HeLa cells transfected with the VCA0956 plasmid and measured 48 hours after had less intracellular c-di-GMP, suggesting that c-di-GMP synthesized intracellularly is transient (data not shown). These results indicate that VCA0848 and VCA0956 are capable of synthesizing c-di-GMP in the cytoplasm of a eukaryotic cell. 103 A 900 c-di-GMP (nM) 600 300 30 20 10 bd bd bd bd bd g qr B q * V CA 08 V 48 C A 08 48 V CA * 09 V 56 C A 09 56 * M oc k 0 B rg B c-di-GMP (nM) 30 20 10 bd bd bd Ad 5VC A 09 56 A d5 -V C A 09 56 * A d5 -N ul l M oc k 0 Figure 26. CC-didi-GMP synthesized in vivo by VCA0956 in HeLa cells. LC-MS/MS was used to quantify c-di-GMP in HeLa cells. (A) HeLa cells were transfected with plasmid vectors containing the VCA0956 allele or the active site mutant allele, VCA0956*. (B) HeLa cells were infected with Ad5 vectors. Bars represent the mean of 3 replicates; error bars indicate standard deviation. bd indicates below detection. 104 I selected VCA0956 for further analysis, as this DGC has been the most extensively characterized. The pShuttleCMV-VCA0956 plasmid and its mutant allele counterpart were used to construct and purify to high concentration the respective recombinant Ad5-based vectors. To confirm that the VCA0956 Ad5 construct, herein referred to as Ad5-VCA0956, was able to produce c-di-GMP in a eukaryotic cytoplasm, I infected HeLa cells (500 M.O.I.) with the Ad5-VCA0956 and Ad5-VCA0956 mutant allele (Ad5-VCA0956*) adenovirus vectors and measured cdi-GMP using LC-MS/MS after 24 hours. The Ad5-Null vector, an adenovirus construct carrying no transgene, was also included as a negative control. I found that cells infected with the Ad5-VCA0956 produced high concentrations of c-diGMP, whereas cells infected with the Ad5-VCA0956* or the Ad5-Null produced no detectable c-di-GMP (Fig. 26B). Importantly, similar to VCA0956 plasmid transfections, infection with Ad5-VCA0956 had no noticeable impact on cell morphology or viability. These results demonstrate that an adenovirus vector can be used to deliver a V. cholerae DGC into a eukaryotic cell to synthesize c-di-GMP. Synthesis of c-di-GMP in a murine system As the Ad5-VCA0956 vector is capable of producing c-di-GMP in HeLa cells in vitro, I next determined if my vector produces c-di-GMP in vivo in a murine model system. Therefore, BALB/c mice (n=3) were IV injected with the Ad5-Null, Ad5VCA0956, or the Ad5-VCA0956* vectors and quantitative PCR was utilized to measure adenovirus genomes in the spleen and liver of injected mice at 24 hours post injection (h.p.i.). I observed comparable Ad5 genome counts for each treatment 105 in both the liver and spleen (Fig. 27A). Consistent with previous reports that the predominant tropism of adenovirus is in the liver [172, 177, 178], there were significantly more Ad5 genomes in the liver cells than in the spleen cells. I then measured c-di-GMP in both the liver and spleen using LC-MS/MS, and found that the Ad5-VCA0956 vector produced detectable c-di-GMP in both tissues, whereas the Ad5-Null and Ad5-VCA0956* vectors produced no detectable c-di-GMP (Fig. 27B). The concentration of c-di-GMP was consistent with the abundance of Ad5-VCA0956 genomes per cell, as the amount of c-di-GMP was significantly higher in the liver tissue than the spleen. These data indicate that the Ad5-VCA0956 vector is capable of initiating c-di-GMP synthesis in a murine model system. C-di-GMP synthesized in vivo stimulates innate immunity in a mouse model It has been previously shown that adenovirus vectors stimulate several proinflammatory innate immune response genes [162, 166, 171]. To examine if the Ad5-VCA0956 alters the profile of innate immune gene expression compared to the Ad5 vector alone, Balb/c mice (n=3) were IV injected with Ad5-Null, Ad5-VCA0956, and Ad5-VCA0956* and qRT-PCR was utilized to quantify the expression levels of several liver gene transcripts at 24 h.p.i. Infection with Ad5-VCA0956 had no observable effect on the health of the mice. I found that the Ad5-Null treatment was able to stimulate 6 of the 12 markers examined (> 2-fold; ADAR, MCP-1, TLR2, IP10, Oas1a, RIG1) (Fig. 28). These results are consistent with previous studies demonstrating that the adenovirus vector alone is capable of altering gene expression in the liver [165, 166]. The expression of four genes was significantly 106 A Ad5 genomes / cell 200 Liver Spleen 150 100 50 A d5 -V C A 09 56 Ad 5V C A 09 56 * A d5 -N ul l N ai ve 0 B c-di-GMP (nM) 2000 Liver Spleen 1500 1000 500 bd bd bd A d5 -V C A 09 56 A d5 -V C A 09 56 * A d5 -N ul l N ai ve 0 Figure 27 27. Infection of Ad5Ad5-VCA0956 in a murine murine model system. Mice were infected with Ad5 vectors intraveneously. (A) After 24 hours qPCR was used to quantify Ad5 genomes in liver cells (black) or spleen cells (hashed). Data were normalized to internal GADPH control. (B) LC-MS/MS was used to quantify c-di-GMP extracted from the liver (black) or spleen (checkered). Bars represent the mean of 3 replicates; error bars indicate standard deviation. bd indicates below detection. 107 (p < 0.05) higher in the Ad5-VCA0956 treatment compared to the Ad5-VCA0956* treatment (Fig. 28A); these include the IFN-responsive gene ADAR, the monocyte and basophil chemotractant MCP-1, the toll-like receptor (TLR) signaling pathway gene MyD88, and the pattern recognition receptor TLR2. It is worth noting that cdi-GMP sensing in the cytoplasm is thought to be independent of TLRs [53]. Additionally, the expression of three genes was significantly (p < 0.05) repressed in the Ad5-VCA0956 treatment compared to the Ad5-VCA0956* treatment (Fig. 28B): the pro-inflammatory interleukin genes IL18 and IL1β, and the interferon transcription factor IRF3. Interestingly, IRF3 has been shown to interact with STING to initiate a c-di-GMP-mediated host type I interferon response [48, 179, 180]. In the cytoplasm, c-di-GMP interacts with STING to activate IRF3, NF-κβ, and the p38/JNK/ERK MAP kinase signaling pathways, resulting in increased production of numerous cytokines and chemokines characteristic of an innate immune response [48]. To determine if the Ad5-VCA0956 vector is capable of increasing cytokine and chemokine secretion, I directly quantified the abundance of cytokines and chemokines in the plasma of mice treated with Ad5-VCA0956 using a multiplexed assay system at 6 (Table 2) and 24 (Table 3) h.p.i. Consistent with prior studies showing that the adenovirus vector stimulates the secretion of proinflammatory cytokines and chemokines [166, 167], I observed 9 cytokines and 108 Fold c hange (Ad5 treated / Moc k) 10 5 TL R 2 M yD 88 P1 M C A D A R 0 IR F3 IL 18 IL 1β 2.5 2.0 1.5 1.0 0.5 0.0 TR IF TB K 1 O as 1a R IG 1 80 60 40 20 0 IP 10 C Fold c hange (Ad5 tre ated / Moc k) B 20 15 Fold change (Ad5 treated / Mock) A Figure 28 28. qRTqRT-PCR of mouse liver gene transcripts 24 hours after infection with Ad5 vectors. Ad5-Null (striped), Ad5-VCA0956 (black), or Ad5-VCA0956* (grey) were normalized to internal GADPH control. Fold change indicates each value normalized to values measured from mock treated mice. Results are separated into liver gene expression increased by Ad5-VCA0956 (A), decreased by Ad5-VCA0956 (B), or unaffected by Ad5-VCA0956 (C). Bars represent the mean of 3 replicates; error bars indicate standard deviation. Brackets indicate statistical significance, 109 which was determined using a two-tailed Student’s t-test (P < 0.05). chemokines that were modestly induced in the Ad5-Null treated mice compared to the naïve mice (IFN-γ, MCP-1, G-CSF, MIP-1α, IL-6, MIP-1β, IL-12p40, KC, RANTES; > 3-fold), and these differences were greatest at the 6-hour time point. I found that 15 cytokines and chemokines were significantly increased in the plasma of the Ad5-VCA0956 treated mice compared to the control Ad5-VCA0956* treated mice at one or both of the two timepoints. Furthermore, for the majority of cytokines and chemokines examined, the largest differences observed were at the 24 hour time point, indicating that the effect of Ad5-VCA0956 is both more potent and longer lasting than that of the adenovirus vector alone. The induction of most of these cytokines and chemokines are consistent with other studies examining the immunostimulatory effects of c-di-GMP [50, 52-56]. Interestingly, I observed increases in IL-1α, G-CSF, and Eotaxin levels in the Ad5-VCA0956 injected mice, which have not been previously reported to be induced by c-di-GMP. These data together indicate that the Ad5-VCA0956 vector is capable of inducing a robust innate response beyond that of the adenovirus vector alone in a murine model system. C-di-GMP synthesized in vivo lowers the effective antigen dose of Ad5-TA The purpose of an adjuvant is to enhance the efficacy of a paired antigen by increasing the longevity, potency, or reducing the effective dose. To determine how well the Ad5-VCA0956 construct functions as a vaccine adjuvant, I tested if Ad5VCA0956 could enhance the adaptive response to a C. difficile antigen. C. difficile, a 110 6 h.p.i Mock Ad5-Null Ad5-VCA0956 Ad5-VCA0956* 26.93 33.96 39.19 35.67 IL-1α ± 5.33 ± 3.90 ± 11.51 ± 18.20 1266.47 597.54 601.11 792.85 IL-1β ± 976.22 ± 54.53 ± 135.01 ± 177.15 2.70 4.26 4.11 3.51 IL-4 ± 0.39 ± 0.47 ± 0.33 ± 0.33 27.89 137.62 2836.67 135.95 ILIL-6 ± 31.57 ± 25.88 ± 313.53** ± 72.95 26.65 45.46 32.71 IL-10 bd ± 11.62 ± 3.5 ± 9.65 511.64 3351.06 3847.84 4165.28 IL-12(p40) ± 210.70 ± 368.94 ± 733.63 ± 601.5 164.30 171.12 231.81 156.88 IL-13 ± 205.70 ± 85.67 ± 44.87 ± 13.23 1033.46 1578.95 2582.84 2181.27 Eotaxin ± 286.53 ± 204.69 ± 189.22 ± 413.21 93.43 1265.03 3722.93 1173.55 G-CSF ± 100.88 ± 771.43 ± 577.59** ± 235.67 84.27 142.15 85.35 GM-CSF bd ± 6.55 ± 14.05 ± 41.35 3.66 28.14 452.73 37.45 IFNIFN-γ ± 4.81 ± 4.64 ± 90.62* ± 4.85 5.16 257.61 346.74 174.94 KC ± 5.48 ± 96.01 ± 39.01 ± 136.14 4309.67 7146.20 4816.93 MCPbd MCP-1 ± 213.69 ± 1166.34** ± 2357.41 7.98 47.50 61.10 48.89 MIP-1α ± 5.56 ± 10.61 ± 5.88 ± 20.07 73.61 582.11 1022.01 754.20 MIP-1β ± 92.93 ± 108.91 ± 105.93 ± 354.18 22.10 71.37 109.03 67.26 RANTES ± 11.30 ± 4.45 ± 10.51 ± 16.77 15.17 345.81 1091.77 463.09 ILIL-12(p70) ± 2.33 ± 81.94 ± 331.08** ± 106.69 Table 2. Cytokine and chemokine secretion of mice infected with Ad5 vectors after 6 hours. hours. Mice were infected IV with either Ad5-Null (n=2), Ad5-VCA0956 (n=3), or Ad5-VCA0956 (n=3). Values indicate the mean pg / mL ± standard deviation. Asterisks (*) indicate a statistically significant difference between Ad5-VCA0956 and Ad5-VCA0956*, which was determined using a two-way ANOVA test combined with a Bonferroni posttest (* p <0.05; ** p < 0.01). 111 24 h.p.i. Mock Ad5-Null Ad5-VCA0956 Ad5-VCA0956* 9.99 15.96 115.58 18.78 ILIL-1α ± 2.16 ± 4.12 ± 30.86** ± 5.57 1056.48 441.68 723.60 642.02 IL-1β ± 950.52 ± 6.21 ± 63.81 ± 259.42 2.88 2.79 6.08 3.25 ILIL-4 ± 0.13 ± 0.25 ± 0.87** ± 0.47 640.05 IL-6 4.18 bd 27.03 ± 291.76 6.22 212.84 11.91 ILIL-10 bd ± 4.09 ± 145.12* ± 5.82 309.78 989.84 11602.17 1069.21 ILIL-12(p40) ± 63.47 ± 20.73 ± 3585.45** ± 335.21 55.80 19.00 287.97 13.08 ILIL-13 ± 68.56 ± 16.51 ± 7.54* ± 8.15 836.39 3974.88 743.67 Eotaxin 539.42 ± 229.64 ± 894.88** ± 187.39 94.72 302.85 156212.26 300.03 G-CSF ± 46.39 ± 56.98 ± 59031.70** ± 38.42 216.35 GM-CSF 15.06 bd bd ± 16.78 17.10 IFN-γ bd bd bd ± 4.60 5.61 43.02 1159.46 61.28 KC ± 6.84 ± 11.94 ± 523.33* ± 57.88 169.58 55001.74 327.11 MCPbd MCP-1 ± 12.19 ± 19457.26** ± 38.26 12.85 303.94 14.24 MIPMIP-1α 2.45 ± 8.24 ± 27.35** ± 9.31 39.55 40.73 1291.42 92.01 MIPMIP-1β ± 50.19 ± 0.69 ± 360.00** ± 73.37 11.5 53.69 3657.80 68.55 RANTES ± 4.33 ± 3.49 ± 1455.71** ± 14.17 7.19 6.45 90.28 8.95 IL-12(p70) ± 4.39 ± 5.43 ± 6.22 ± 9.74 Table 3. Cytokine and chemokine secretion of mice infected with Ad5 vectors after 24 hours. Mice were infected IV with either Ad5-Null (n=2), Ad5-VCA0956 (n=3), or Ad5-VCA0956 (n=3). Values indicate the mean pg / mL ± standard deviation. Asterisks (*) indicate a statistically significant difference between Ad5-VCA0956 and Ad5-VCA0956*, which was determined using a two-way ANOVA test combined with a Bonferroni posttest (* p <0.05; ** p < 0.01). 112 Gram-positive spore-forming anaerobic bacteria, is the leading causative agent of noscomial infections leading to diarrheal disease in the developed world. C. difficile associated diarrhea (CDAD) represents nearly 1% of all hospital stays in the United States and can lead to septicemia, renal failure, and toxic megacolon [181]. Incidents and mortality of C. difficile infections are rising in the U.S., and the economic burden on the health care system is reported to be in the billions of dollars [181-185]. Furthermore, to date there are no approved effective vaccine treatments available for CDAD treatment or prevention [186]. A previously developed adenovirus vector that expresses the immunogenic portion of the C. difficile toxin A (Ad5-TA) protects mice from a toxin challenge by generating a humoral and T-cell response specific to toxin A in a murine model system [170]. I hypothesized that supplementing this vaccine with the Ad5VCA0956 adjuvant would enhance this humoral and T-cell response due to the strong pro-inflammatory innate response it generates. I therefore vaccinated mice by IM injection with varying concentrations of the Ad5-TA vector in combination with the Ad5-VCA0956 vector in equal ratio ranging from 1*106 to 5x109 viral particles (vp). After two weeks, I assessed TA specific T-cell responses in the spleens of the naive and vaccinated animals using an IFNγ ELISpot assay, utilizing the 15mer peptide (VNGSRYYFDTDTAIA) that has been previously shown to elicit the secretion of IFNγ in splenocytes of mice immunized with the Ad5-TA vector [170]. I found that co-injection of equal amounts of the Ad5-TA and the mutant DGC allele vector Ad5-VCA0956* produced no induction of IFNγ secreting T-cells over that of 113 naïve splenocytes at viral doses of 1x106 and 1x107, but did generate significant IFNγ producing T-cells at 1x108 and 5x109 (Fig. 29, white squares). Although coinjection of 1x106 Ad5-TA with Ad5-VCA0956 did not produce increased IFNγ levels, I observed significantly increased (p < 0.05) IFNγ producing T-cells at a dose of 1x107, as compared to cells derived from the DGC mutant treated control (Fig. 29, black squares). No c-di-GMP was detected in the liver of mice infected with Ad5VCA0956 at the 5x109 dose after 14 days, suggesting that even at high doses, IM administration of Ad5-VCA0956 does not lead to long-lasting c-di-GMP production at distal sites (data not shown). Thus, I conclude that c-di-GMP synthesized by Ad5VCA0956 lowers the effective dose to generate an adaptive response to Ad5-TA. Ad5-VCA0956 does not significantly enhance protection of Ad-TA against C. difficile To determine if the Ad5-VCA0956 construct increases the efficacy of the Ad5TA vaccine in vivo, I decided to test the adjuvant-vaccine combination in a cefoperazone-treated mouse C. difficile challenge model [173, 174]. I first sought to determine the optimum dose of C. difficile to administer to effectively challenge the mice. Naïve mice were administered cefoperazone (0.5 mg/mL in drinking water) for 10 days, followed by a two day rest period. Mice were then administered varying doses ranging from 5*103 to 5*106 of hyperinfectious C. difficile CD2015 (NAP-1, 027 ribotype [187]) spores by oral gavage. The strain of C. difficile used is a clinically isolated strain from a hospital in Southeast Michigan (unpublished). Weights of the mice were monitored for a four day period. Mice were sacrificed upon 114 1000 800 SFCs pe r 6 10 sp lenocyt es 600 * 400 150 100 50 6 7 8 9 1x 10 1x 10 1x 10 5x 10 N ai ve 0 Ad5-TA + Ad5-VCA0956 Ad5-VCA0956* Figure 29 29. IFNγ IFNγ ELISPOT analysis of mice vaccinated with Ad5Ad5-TA and Ad5 vectors. vectors. Mice were administered IM varying doses of both Ad-TA and either AdVCA0956 (black) or Ad-VCA0956* (grey). After 14 days, splenocytes were ex vivo stimulated with a C. difficile specific peptide and the number of IFNγ secreting splenocytes was determined using ELISPOT. Each point represents an individual animal. Lines indicate the mean of the replicates, and error bars indicate standard error. * indicates statistical significance using a two-way ANOVA test combined with a Bonferroni posttest (P < 0.05). losing >20% of their initial body weight. I found that none of the mice administered the lowest dose lost >20% of their initial body weight (Fig. 30 A and B). However, when the dose was increased, I observed significantly more weight loss. The 115 Perce nt survival A 100 80 60 5* 5* 5* 5* 106 105 104 103 cfu cfu cfu cfu 5* 5* 5* 5* 106 105 104 103 cfu cfu cfu cfu 40 20 0 B Percent Body Weight 0 1 2 3 Time (Days) 4 110 100 90 80 0 1 2 3 Time (Days) 4 Figure 30. 30. C. difficile challenge of BALB/c mice. Mice were challenged with varying doses of C. difficile CD2015 spores. (A) Survival of mice challenged with varying doses of C. difficile. Mice were sacrificed after losing > 20% of their initial body weight. (B) Weight loss of mice challenged with C. difficile. Each point represents the mean of 5 replicates. Error bars indicate standard error. 116 greatest weight loss was observed in the 5*105 dose of C. difficile, resulting in 80% mortality by day 4 of the challenge. To determine how effective my Ad5-VCA0956 adjuvant was at increasing the protective properties of Ad5-TA, I vaccinated mice with 1*107 vp of both the Ad5-TA vaccine antigen and the Ad5-VCA0956 adjuvant vector; this dose was selected as it produced a significant difference in the ELISPOT analysis (Fig. 29). As controls, I also vaccinated mice with equal doses of the Ad5-TA vaccine with the Ad5VCA0956* and Ad5-Null vectors. After 10 weeks, mice were administered cefoperazone for 10 days, followed by a 2-day rest. Mice were then challenged with 5*105 C. difficile spores. I found that both the mortality and weight loss of all mice treated with the Ad5-TA was less than the naïve mice (Fig. 31). However, I did not observe any significant differences between the mice that received the Ad5VCA0956 treatment compared to either the Ad5-VCA0956* or the Ad5-Null treatments. And while the average weight loss at the end of the challenge was lowest in the Ad5-VCA0956 treated mice, this difference was not statistically significant from that of the Ad5-VCA0956* or Ad5-Null treated mice. Thus, I conclude from this experiment that in the conditions examined here, the Ad5VCA0956 does not contribute to the effectiveness of the Ad5-TA vaccine. 117 Percent survival A 100 Naive Ad-VCA0956 Ad-VCA0956* Ad-Null 80 60 40 20 0 B Percent Body Weight 0 1 2 3 Time (Days) 4 110 Naive Ad-VCA0956 Ad-VCA0956* Ad-Null 100 90 80 70 0 1 2 3 Time (Days) 4 Figure 31. C. difficile challenge of BALB/c mice vaccinated with Ad5Ad5-TA and Ad5Ad5VCA0956. Mice were treated with equal doses of Ad5-TA and either Ad5-VCA0956 (squares), Ad5-VCA0956* (up-triangles), or Ad5-Null (down-triangles). (A) Survival of mice challenged with C. difficile. Mice were sacrificed after losing > 20% of their initial body weight. (B) Weight loss of mice challenged with C. difficile. Naïve: N=5, Ad5-VCA0956: N=9, Ad5-VCA09568 and Ad5-Null: N=15. Error bars indicate standard error. 118 Discussion With a current demand for novel vaccines that target difficult-to-treat diseases, it is crucial to have adjuvants to pair with these vaccines to optimize efficacy. Currently, there are a limited number of adjuvants available for clinical use, and there is a need for new adjuvants which can enhance the efficacy of vaccines to improve immunological protection [61, 188]. Numerous studies have implicated c-di-GMP as a promising novel adjuvant. Indeed, this second messenger molecule has been shown to stimulate a robust type I interferon response and increase the secretion of numerous cytokines and chemokines to initiate a balanced Th1/Th2 response, as well as stimulate the inflammasome pathway and immune cell activation/recruitment [49-57]. My approach to utilize c-di-GMP as an adjuvant is novel in that it utilizes an adenovirus vector to deliver c-di-GMP producing enzyme DNA into cells, thereby synthesizing the adjuvant in vivo and circumventing the inefficient delivery of IM or IV administration of chemically synthesized c-di-GMP. Adenovirus vectors are promising in that they are costefficient to produce and can efficiently deliver specific antigens or adjuvants into cells for in vivo production, greatly reducing the cost of using c-di-GMP as an adjuvant. I have demonstrated that an adenovirus vector carrying a bacterial DGC is capable of synthesizing c-di-GMP in both human and mouse model systems. Consistent with other reports utilizing c-di-GMP, synthesis of c-di-GMP by virally mediated transduction of the DGC gene in vivo increases the secretion of numerous 119 cytokines and chemokines [50, 52-57]. These induced cytokines and chemokines include signals characteristic of both Th1 (Table 2 and 3, IFNγ, IL-12) and Th2 (Table 2 and 3, IL-4, IL-6) type responses. Additionally, c-di-GMP production from Ad5-VCA0956 enhances activation of the innate immune system by activating TLR signaling (Fig. 28, TLR2, MyD88). It appears however that c-di-GMP synthesized in vivo negatively regulates the expression of inflammasome-dependent pathways in hepatocytes (Fig. 28, IL-1β, IL-18). The significance of this finding is unclear, especially as it has been reported that c-di-GMP activates the NLRP3 inflammasome pathway [51]. Importantly, I observed no signs of poor cell physiology or health in my cell cultures and animal models. Furthermore, my data indicate that the c-di-GMP synthesized by the Ad5-VCA0956 vector is transient, and thus should enhance antigen recognition and response while minimizing any potentially unwanted long term effects associated with administration, such as autoimmune activation [189]. I have also shown that c-di-GMP synthesized in vivo reduces the effective antigen dose of Ad5-TA, a vaccine antigen which targets the toxin of the human pathogen C. difficile. Reducing the dose required to initiate an adaptive immune response is of particular significance as high viral particle doses can lead to global toxicities, endothelial cell activation, and liver damage [171, 177, 190-192]. C-diGMP has been shown to enhance protection against other pathogens including S. aureus, K. pneumoniae, and S. pneumoniae [53-55, 58], indicating that c-di-GMP has broad antigen-adjuvant synergy. Although I have only demonstrated an 120 enhanced adaptive response to C. difficile in vitro, Ad5-VCA0956 can be coupled with other vaccines to enhance antigen recognition efficacy. While c-di-GMP is a bacteria-specific second messenger, the Ad5-VCA0956 stimulated c-di-GMP innate immune response could enhance protection in viral or cancer vaccine systems as well. I believe that the approach of synthesizing c-di-GMP in vivo offers numerous advantages over directly administering c-di-GMP as a mucosal, intramuscular, or intravenous adjuvant; these include lower cost, more efficient adjuvant delivery, and a more robust innate immune response. Furthermore, adjustments can be made to alter and optimize the Ad5-VCA0956 for clinical use. V. cholerae contains 40 predicted DGC alleles within its genome, and it has been shown that ectopic expression of these different DGCs results in different intracellular c-di-GMP concentrations [79]. Hence intracellular expression of other DGCs could produce different amounts of c-di-GMP in eukaryotic cells, increasing the efficiency of this adjuvant. Alternatively, other types of second messengers could be used to stimulate innate immunity. One example would be to express a diadenylate cyclase to synthesize the related bacterial second messenger cyclic di-AMP in vivo. Cyclic di-AMP has similarly been shown to induce a robust innate immune response through STING mediated recognition [193, 194]. Additionally, different promoters could be used in lieu of the CMV promoter to produce localized or temporally controlled c-di-GMP production in the body. 121 Future directions With a current demand for development of vaccines for otherwise difficult to treat infectious diseases, vaccine adjuvant development has been proposed as a promising research avenue. I have shown that an adenovirus vector harboring a DGC gene results in c-di-GMP synthesis in eukaryotic cells, leading to robust stimulation of the innate immune system. In this study, I identified two DGCs that produced c-di-GMP upon transfection into HeLa cells (Fig. 26A). The DGC VCA0848 produced significantly more c-di-GMP upon transfection than the VCA0956 strain. As my data indicate that this DGC is more efficient at synthesizing c-di-GMP in the eukaryotic cell environment, I hypothesize that an Ad5 construct harboring this DGC will produce more intracellular c-di-GMP than the Ad5-VCA0956 vector in vivo. This is significant, as less adenovirus would be required to stimulate innate immunity and thus would increase the effect of c-di-GMP specific innate immune activation compared to Ad5-stimulated innate immunity. Furthermore, reducing the dosage of Ad5 required to initiate an innate immune response would reduce the adverse effects associated with Ad5 treatment [171, 177, 190-192]. I propose creating an Ad5-VCA0848 construct and determining its effect on innate and adaptive immunity as described in Chapter 4. Previous studies demonstrated that administration of purified c-di-GMP leads to an innate immune response that conveys protection against S. pneumonia, K. pneumonia, and S. aureus [53, 54, 58]. As the Ad5-VCA0956 is able to synthesize c-di-GMP in vivo, I hypothesize that the Ad5-VCA0956 produces a more efficient 122 stimulation of innate immunity and thus leads to a more robust response. I propose to test the Ad5-VCA0956 in a S. pneumonia challenge model to see if the vector is capable of protecting animals against acute bacterial infection [58]. As a control, the Ad5-VCA0956* and the Ad5-Null vectors should be included to account for any protection granted by the vector or expressing a bacterial protein in an animal. Additionally, purified c-di-GMP should be administered intranasally as a control to determine if the Ad5-VCA0956 is more effective than c-di-GMP. The Ad5-VCA0956 construct is able to increase the adaptive response of Ad5TA at lower doses (Table 2 and 3, Fig. 29). However, treatment of mice with Ad5VCA0956 in combination with Ad5-TA did not increase protection in a C. difficile mouse challenge model compared with Ad5-VCA0956* or the empty Ad5-Null vector. The dose I utilized in this challenge was the dose that increased IFNγ production of splenocytes compared to the negative controls (Fig. 29). It is possible that this dose is not sufficient to produce other cytokines or chemokines important for enhancing an adaptive immune response. I postulate that the Ad5-VCA0956 vector could increase the protection of Ad5-TA at higher doses. I propose repeating this challenge with different doses of the Ad5-VCA0956 while maintaining the Ad5TA dose. Additionally, this experiment should be repeated with the Ad5-VCA0848 to determine if increased Ad5 mediated c-di-GMP production can improve protection. 123 CHAPTER 6 – Conclusions Conclusions While there are numerous studies describing the role c-di-GMP plays in regulating the physiology of many bacteria, the regulatory inputs to this almost ubiquitously utilized second messenger are currently under-characterized. I have demonstrated that the growth environment effects intracellular c-di-GMP concentrations in V. cholerae, and that these differences can be as great as 20-fold (Fig. 5). Furthermore, I have developed a novel method to differentiate the impact of c-di-GMP synthesis and hydrolysis on overall intracellular c-di-GMP. This method can be applied to other bacterial systems to determine the effects of specific environmental factors on c-di-GMP, the effects of deleting DGCs and PDEs, and differences in c-di-GMP synthesis and hydrolysis between different organisms. The intestinal environment is a tropism for many human pathogens. To efficiently cause disease, pathogens recognize specific host-derived cues and subsequently produce virulence factors. It has been proposed that the intracellular c-di-GMP of V. cholerae is elevated in the marine environment, and reduced in the human host. In contrast, my research suggests that the regulation of intracellular cdi-GMP in V. cholerae in the human host is more complex than previously appreciated. I have demonstrated that bile and bicarbonate inversely modulate intracellular c-di-GMP, where bile increases c-di-GMP approximately 4-fold and bicarbonate quenches this induction (Fig. 8, 18). As c-di-GMP signaling is widely distributed amongst bacteria, it is possible that other enteric bacteria employ similar responses as V. cholerae to bile and bicarbonate in order to adapt to 124 different microenvironments within the small intestine. Thus, this study may provide insights about the role c-di-GMP plays in the biology of both human pathogens and normal members of the human intestinal microbiota. Growth rate regulation of bacteria has been a focus of study for the last 50 years, and yet there are still many questions regarding how cell division is governed. I have demonstrated that intracellular c-di-GMP concentration and growth rate are negatively correlated in V. cholerae (Fig. 20, 21, 22). I have also shown that rRNA is decreased upon ectopic DGC expression (Fig. 25). Determining the mechanism governing the c-di-GMP mediated growth rate inhibition is important, as this phenotype could have large implications in core bacterial processes. Understanding the relationship between growth rate regulation and c-diGMP is also informative for drug discovery studies targeting compounds that effect c-di-GMP synthesis and hydrolysis, as these compounds may have unanticipated effects on bacterial growth rate. As new and re-emerging infectious diseases continue to take a toll on our population and burden our healthcare system, vaccine and adjuvant development offer promising solutions to combat this problem. I have shown that a DGC delivered into the cytoplasm of a eukaryotic cell can synthesize c-di-GMP and activate innate immunity by inducing the production of 15 cytokines and chemokines (Fig. 26, 27; Table 2, 3). Furthermore, I have provided evidence that using adenovirus to deliver a DGC gene can be utilized as a vaccine adjuvant to enhance an adaptive immune response (Fig. 29). As c-di-GMP initiates a potent 125 innate response, our novel adenovirus construct is a promising candidate for clinical applications to enhance the efficacy of existing vaccine antigens. Additionally, the Ad5-VCA0956 vector is capable of efficiently producing high levels of c-di-GMP in eukaryotic cells, and thus can be used as a tool elucidate the different c-di-GMP response mechanisms of various eukaryotic cell types. There are still many questions regarding regulatory inputs to c-di-GMP signaling networks, as well as what behaviors c-di-GMP regulates in response to these inputs. Discerning how bacteria interact with and perceive their changing environments is paramount to understanding the ecology and behaviors of these organisms, as well as developing new clinical applications to protect ourselves against devastating human pathogens. I have demonstrated here that the role of cdi-GMP as a second messenger is to translate environmental signals into physiological changes that allow V. cholerae to adapt and persist in various niches. As c-di-GMP is ubiquitously utilized amongst bacteria, we are only beginning to appreciate the role this molecule plays in enabling bacteria to transition between different lifestyles, the physiological changes associated with different levels of intracellular c-di-GMP, and the subsequent impact on the pathogenesis of bacteria. 126 APPENDICES APPENDICES 127 Strain Genotype Escherichia coli Reference Invitrogen DH10B Escherichia coli [103] S17 Vibrio cholerae WT C6706 Vibrio cholerae ∆vpsL Vibrio cholerae ∆hapR Vibrio cholerae ∆luxO Vibrio cholerae ∆cdgA Vibrio cholerae ∆toxR Vibrio cholerae ∆toxT Vibrio cholerae [78] VC0934 (vpsL) [72] knockout mutant hapR knockout [67] mutant VC1021 (luxO) [68] knockout mutant VCA0074 knockout Chapter 2 mutant VC0984 (toxR) Chapter 3 knockout mutant VC0838 (toxT) Chapter 3 knockout mutant VC1372 knockout Chapter 3 ∆VC1372 mutant Vibrio cholerae VC1376 knockout Chapter 3 ∆VC1376 mutant Vibrio cholerae VC1067 knockout Chapter 3 ∆VC1067 mutant VC1372, VC1376 Vibrio cholerae double knockout Chapter 3 ∆VC1372 ∆VC1376 mutant Vibrio cholerae VC1067, VC1372, ∆VC1372 ∆VC1376 VC1376 triple Chapter 3 ∆VC1067 knockout mutant Vibrio cholerae VC1295 knockout Chapter 3 ∆VC1295 mutant VC1295, VC1067, Vibrio cholerae VC1372, VC1376 ∆VC1295 ∆VC1372 Chapter 3 quadruple knockout ∆VC1376 ∆VC1067 mutant Clostridium difficile NAP-1, Ribotype Unpublished CD2015 027 Table A1. Bacteria strain list. All bacteria strains used in this dissertation, and their references. 128 Plasmid Name Description Reference pTL17 pKD3 pEVS141 pVCA0956 pCMW121 pBBRlux 6:C9-lux vpsT-lux Inducible FLP recombinase vector Contains cat flanked by FRT sites Backbone for DGC expression plasmids, vector control IPTG inducible VCA0956 GGDEF expression vector IPTG inducible VC1086 EAL expression vector Backbone for luminescence reporter plasmids VC1673 – c-d-GMP inducible lux reporter vector VC2647 - c-d-GMP inducible lux reporter vector qrgB expression plasmid, pMMB67eh backbone (Pursley, B.R.) IPTG inducible qrgB expression vector IPTG inducible qrgB* active site mutant expression vector IPTG inducible VC1153 (tfoX) expression plasmid (Chan, C.) IPTG inducible VC0900 GGDEF expression vector IPTG inducible VC1029 GGDEF expression vector IPTG inducible VC1067 GGDEF expression vector IPTG inducible VC1104 GGDEF expression vector IPTG inducible VC1185 GGDEF expression vector IPTG inducible VC1216 GGDEF expression vector IPTG inducible VC1353 GGDEF expression vector IPTG inducible VC1367 GGDEF expression vector IPTG inducible VC1370 GGDEF expression vector IPTG inducible VC1372 GGDEF expression vector IPTG inducible VC1376 GGDEF expression vector IPTG inducible VC1593 GGDEF expression vector IPTG inducible VC1599 GGDEF expression vector IPTG inducible VC2224 GGDEF expression vector IPTG inducible VC2285 GGDEF expression vector IPTG inducible VC2454 GGDEF expression vector IPTG inducible VC2697 GGDEF expression vector IPTG inducible VCA0049 GGDEF expression vector IPTG inducible VCA0074 GGDEF expression vector IPTG inducible VCA0165 GGDEF expression vector IPTG inducible VCA0217 GGDEF expression vector IPTG inducible VCA0557 GGDEF expression vector IPTG inducible VCA0560 GGDEF expression vector IPTG inducible VCA0697 GGDEF expression vector IPTG inducible VCA0848 GGDEF expression vector IPTG inducible VCA0939 GGDEF expression vector [86] [84] [195] [79] [72] [69] [107] [196] pBRP2 pCMW75 pCMW98 pANDA2 pVC0900 pVC1029 pVC1067 pVC1104 pVC1185 pVC1216 pVC1353 pVC1367 pVC1370 pVC1372 pVC1376 pVC1593 pVC1599 pVC2224 pVC2285 pVC2454 pVC2697 pVCA0049 pVCA0074 pVCA0165 pVCA0217 VCA0557 pVCA0560 pVCA0697 pVCA0848 pVCA0939 Chapter 3 [72] [72] Chapter 3 Table A2. A2. Plasmid list. All plasmids used in this dissertation, and their references. 129 [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] pVCA0960 pVCA0965 pVCA1082 pVC0072 pVC0130 pVC0398 pVC0653 pVC0658 pVC0703 pVC1934 pVC2750 pVCA0080 pVCA0785 pVC1652 pVCA1083 pBK34 pBK35 pBK36 pBK37 pBK38 pBK39 pBK40 pBK41 pBK42 pBK43 pBK45 pBK46 pBK47 pBK48 pCMW122 pCMW123 pCMW124 Table A2 (cont’d) cont’d) IPTG inducible VCA0960 GGDEF expression vector IPTG inducible VCA0965 GGDEF expression vector IPTG inducible VCA1082 GGDEF expression vector IPTG inducible VC0072 GGDEF/EAL expression vector IPTG inducible VC0130 GGDEF/EAL expression vector IPTG inducible VC0398 GGDEF/EAL expression vector IPTG inducible VC0653 GGDEF/EAL expression vector IPTG inducible VC0658 GGDEF/EAL expression vector IPTG inducible VC0703 GGDEF/EAL expression vector IPTG inducible VC1934 GGDEF/EAL expression vector IPTG inducible VC2750 GGDEF/EAL expression vector IPTG inducible VCA0080 GGDEF/EAL expression vector IPTG inducible VCA0785 GGDEF/EAL expression vector IPTG inducible VC1652 EAL expression vector IPTG inducible VCA1083 EAL expression vector IPTG inducible VC1087 HD-GYP expression vector IPTG inducible VC1295 HD-GYP expression vector IPTG inducible VC1348 HD-GYP expression vector IPTG inducible VC2340 HD-GYP expression vector IPTG inducible VC2497 HD-GYP expression vector IPTG inducible VCA0210 HD-GYP expression vector IPTG inducible VCA0681 HD-GYP expression vector IPTG inducible VCA0895 HD-GYP expression vector IPTG inducible VCA0931 HD-GYP expression vector IPTG inducible VC0137 EAL expression vector IPTG inducible VC1592 EAL expression vector IPTG inducible VC1851 EAL expression vector IPTG inducible VCA0101 EAL expression vector IPTG inducible VCA0536 EAL expression vector IPTG inducible VC1211 EAL expression vector IPTG inducible VC1641 EAL expression vector IPTG inducible VC1710 EAL expression vector 130 [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] [79] Chapter 3 Chapter 3 Chapter 3 Chapter 3 Chapter 3 Chapter 3 Chapter 3 Chapter 3 Chapter 3 Chapter 3 Chapter 3 Chapter 3 Chapter 3 Chapter 3 [72] [72] [72] pVC1372* pVC1376* pVC1067* pBK51 pCMW27 pCMW28 pCMW33 pCMW34 pCWM35 pCMW36 pCMW37 pCMW38 pCMW39 pCMW40 pCMW41 pCMW42 pCMW47 pCMW48 pCMW49 pCMW50 pCMW51 pCMW52 pCMW53 pCMW54 pCMW55 pCMW56 pCMW57 pCMW58 pCMW59 pCMW60 pCMW61 pCMW62 pCMW63 pCMW64 pCMW65 pCMW66 pCMW67 pCMW68 pCMW69 Table A2 (cont’d) IPTG inducible VC1372* active site mutant expression vector IPTG inducible VC1376* active site mutant expression vector IPTG inducible VC1067* active site mutant expression vector toxT knockout vector, pKAS32 backbone VC0072-gfp transcriptional fusion VC0653-gfp transcriptional fusion VC0130-gfp transcriptional fusion VC0900-gfp transcriptional fusion VC1029-gfp transcriptional fusion VC1104-gfp transcriptional fusion VC1185-gfp transcriptional fusion VC1216-gfp transcriptional fusion VC1353-gfp transcriptional fusion VC1367-gfp transcriptional fusion VC1372-gfp transcriptional fusion VCA0074-gfp transcriptional fusion VC1370-gfp transcriptional fusion VC1376-gfp transcriptional fusion VC1593-gfp transcriptional fusion VC1599-gfp transcriptional fusion VC1934-gfp transcriptional fusion VC2224-gfp transcriptional fusion VC2285-gfp transcriptional fusion VC2370-gfp transcriptional fusion VC2454-gfp transcriptional fusion VC2697-gfp transcriptional fusion VC2750-gfp transcriptional fusion VCA0049-gfp transcriptional fusion VCA0080-gfp transcriptional fusion VCA0165-gfp transcriptional fusion VCA0217-gfp transcriptional fusion VCA0557-gfp transcriptional fusion VCA0560-gfp transcriptional fusion VCA0697-gfp transcriptional fusion VCA0785-gfp transcriptional fusion VCA0848-gfp transcriptional fusion VCA0939-gfp transcriptional fusion VCA0956-gfp transcriptional fusion VCA0960-gfp transcriptional fusion 131 Chapter 3 Chapter 3 [79] Chapter 3 [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] pCMW70 pCMW71 pCMW72 pCMW73 pCMW78 pCMW79 pCMW80 pCMW81 pCMW82 pCMW83 pCMW84 pCMW85 pCMW86 pCMW87 pCMW88 pCMW89 pCMW90 pBBRluxVC1295 pBBRLuxVC1348 pBBRluxVC2340 pBBRluxVC2497 pBBRluxVCA0210 pBBRluxVCA0681 pBBRluxVCA0895 pBBRluxVCA0931 pShuttle-CMV pShuttleCMV- qrgB Table A2 (cont’d) VCA0965-gfp transcriptional fusion VC0658-gfp transcriptional fusion VC1067-gfp transcriptional fusion VC0703-gfp transcriptional fusion VCA1082-gfp transcriptional fusion VC0398-gfp transcriptional fusion VC0137-gfp transcriptional fusion VC0515-gfp transcriptional fusion VC1086-gfp transcriptional fusion VC1211-gfp transcriptional fusion VC1592-gfp transcriptional fusion VC1641-gfp transcriptional fusion VC1652-gfp transcriptional fusion VC1710-gfp transcriptional fusion VC1851-gfp transcriptional fusion VCA0101-gfp transcriptional fusion VCA0536-gfp transcriptional fusion [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] [72] VC1295-lux transcriptional fusion [66] VC1348-lux transcriptional fusion [66] VC2340-lux transcriptional fusion [66] VC2497-lux transcriptional fusion Chapter 3 VCA0210-lux transcriptional fusion [66] VCA0681-lux transcriptional fusion [66] VCA0895-lux transcriptional fusion [66] VCA0931-lux transcriptional fusion [66] Backbone for DGC eukaryote expression vectors [165] qrgB expression vector under CMV control Chapter 5 pShuttleCMV- qrgB* active site mutant expression vector under qrgB* CMV control pShuttleCMVVCA0848 pShuttleCMVVCA0848* Chapter 5 VCA0848 expression vector under CMV control Chapter 5 VCA0848* active site mutant expression vector under CMV control Chapter 5 132 Table A2 (cont’d) pShuttleCMVVCA0956 pShuttleCMVVCA0956* VCA0956 expression vector under CMV control Chapter 5 VCA0956* active site mutant expression vector under CMV control Chapter 5 133 Primer Name Description Sequence VCA0074KO1 VCA0074KO2 VCA0074KO3 VCA0074KO4 342-us.2ecoR1 342-ds.2bamH1 VC1087fwecoR1 VC1087rvbamh1 VC1295fwecoR1 VC1295rvbamh1 VC1348fwecoR1 VC1348rvbamh1 VC2340fwecoR1 VC2340rvbamh1 VC2497fwecoR1 VC2497rvbamh1 VCA0210fwecoR1 VCA0210rvbamh1 VCA0681fwecoR1 VCA0681rvbamh1 For ∆cdgA strain For ∆cdgA strain For ∆cdgA strain For ∆cdgA strain TCGATGTGTCTGATCTGCGTATCCG CGTTGGTAT GAAGCAGCTCCAGCCTACACGGGG CAAAGTTTACCAT TAAGGAGGATATTCATATGGGGCTT CTTATGAATCAAAAT GGTTGAGAAGTAGAACACAATCGAT ATCCACACG ATAGAATTCCTCTGCGCTCCTCATA TTCT ATAGGATCCGACAGAGAAAGTTAAG AAAC ATAGAATTCAGGAGCTAAGGAAGCT AAAATGCAAATGGATAACAT ATAGGATCCTTATCCTATATAAACG GGGATCGGC ATAGAATTCAGGAGCTAAGGAAGCT AAATTGACCATTTGGGTTCT ATAGGATCCTTATGCGGCATCTTTA AAGTGTGCT ATAGAATTCAGGAGCTAAGGAAGCT AAAATGGCAACCGCCAATAT ATAGGATCCTCAGCCTGACGCTTGT TGAC ATAGAATTCAGGAGCTAAGGAAGCT AAAATGCAGATGCAACCCAA ATAGGATCCTCAGTGGGTGATTCCC TGGTC ATAGAATTCAGGAGCTAAGGAAGCT AAAGTGGCAAGCATTAAAAT ATAGGATCCTTACTCTTCGCTGTCA AAGAAGTATG ATAGAATTCAGGAGCTAAGGAAGCT AAATTGAAGTGGTTTAAATA ATAGGATCCCTAGTCAGGCAGCGAA GCAC ATAGAATTCAGGAGCTAAGGAAGCT AAAATGAGATGGTCAGAAAT ATAGGATCCTTATGCCCATTCGGTT GGGCTTT For pBRP2 For pBRP2 For pBK34 For pBK34 For pBK35 For pBK35 For pBK36 For pBK36 For pBK37 For pBK37 For pBK38 For pBK38 For pBK39 For pBK39 For pBK40 For pBK40 Restriction Site EcoRI BamHI EcoRI BamHI EcoRI BamHI EcoRI BamHI EcoRI BamHI EcoRI BamHI EcoRI BamHI EcoRI BamHI Table A3. Primer list. All primers used in this dissertation and their sequence, and any restriction sites included for cloning applications. 134 VCA0895fwmfeI VCA0895rvbglII VCA0931fwecoR1 VCA0931rvbamh1 VC0137fwecoR1 VC0137rvbamH1 VC0515fwecoR1 VC0515rvbamH1 VC1592fwecoR1 VC1592rvbamH1 VC1851fwecoR1 VC1851rvbamH1 VCA0101fwecoR1 VCA0101rvbglII VCA0536fwecoR1 VCA0536rvbamH1 CMW464 For pANDA2 CMW465 For pANDA2 For pBK41 For pBK41 For pBK42 For pBK42 For pBK43 For pBK43 For pBK44 For pBK44 For pBK45 For pBK45 For pBK46 For pBK46 For pBK47 For pBK47 For pBK48 For pBK48 For pVC1372* For VC1376-ASM pVC1376* For VC1067-ASM pVC1067* For pBBRlux2497-fw-SacI 2497 VC1372-ASM Table A3 (cont’d) ATACAATTGAGGAGCTAAGGAAGCT AAATTGCAACCTAGCCAAGA ATAAGATCTCTAGTAATCTTCCCGT AAATACGCC ATAGAATTCAGGAGCTAAGGAAGCT AAAATGAGTGTTGCACAAAA ATAGGATCCTTAGGTCTTATCTGAA GAGTGTGGTA ATAGAATTCAGGAGCTAAGGAAGCT AAATTGGTTCGATGTTTATG ATAGGATCCCTAAATTAATCGATTG ATGTCCTGGC ATAGAATTCAGGAGCTAAGGAAGCT AAAATGGATAGTTTCATAGC ATAGGATCCCTATTTAACTTCTTTC ATTAAAGTTCGAC ATAGAATTCAGGAGCTAAGGAAGCT AAAATGAATCACATCCACCC ATAGGATCCCTATAACGCAGACATA CGTGGAG ATAGAATTCAGGAGCTAAGGAAGCT AAATTGAAGTACTCGTATGT ATAGGATCCTTAGGTTTGTACACCA AGAAGATCG ATAGAATTCAGGAGCTAAGGAAGCT AAAATGGCTAAGGAAAAGAC ATAAGATCTTTATATTTTGCTTGTG GCAAAGTTTAATG ATAGAATTCAGGAGCTAAGGAAGCT AAAATGGCGTGCGTCAAAAG ATAGGATCCTTAGTTAGCATTTGCT ACCTTAGCG GAATTCGGGGAACGTGATTAAAGG GGATCCACGCTGCTGACAACTTTCT AAC TCGCGAACTCTTCAGCGGCATATCG AGCGACC ATATTTTTGCCCGTTTCGCCGCTGA TGAGTTTGCCATCTT GGGCGCAGCTATCGGATTGCTGCG G GAGCTC TCTTGGTTACTTGAAAGAGGCC 135 Mfe1 BglII EcoRI BamHI EcoRI BamHI EcoRI BamHI EcoRI BamHI EcoRI BamHI EcoRI BglII EcoRI BamHI EcoRI BamHI SacI 2497-rvBamHI VC1372KO-1 VC1372KO-2 VC1372KO-3 VC1372KO-4 VC1376KO-1 VC1376KO-2 VC1376KO-3 VC1376KO-4 VC1067KO-1 VC1067KO-2 VC1067KO-3 VC1067KO-4 VC1295KO-1 VC1295KO-2 VC1295KO-3 VC1295KO-4 toxRKO-1 toxRKO-2 toxRKO-3 toxRKO-4 For pBBRlux2497 For ∆VC1372 strain For ∆VC1372 strain For ∆VC1372 strain For ∆VC1372 strain For ∆VC1376 strain For ∆VC1376 strain For ∆VC1376 strain For ∆VC1376 strain For ∆VC1067 strain For ∆VC1067 strain For ∆VC1067 strain For ∆VC1067 strain For ∆VC1295 strain For ∆VC1295 strain For ∆VC1295 strain For ∆VC1295 strain For ∆toxR strain For ∆toxR strain For ∆toxR strain For ∆toxR strain Table A3 (cont’d) GGATCC AATCTTAATTACATCTGAGTGAATA ACC CGTCTTGTTGATATTGATGATTGCG TGCGACCAGT GAAGCAGCTCCAGCCTACACGCTTG GTCGTCCTGAATGCACAAA TAAGGAGGATATTCATATGGCTCAG ATTGTTTATAGGATGGGAGC CATTTACAGGTGGAAGACAAAACTC TACTCGCCAC GCGCTTGGCCAGTCACTTGACCGGT TAG GAAGCAGCTCCAGCCTACACGAAAA ATCATGATGAAG TAAGGAGGATATTCATATGAATCGA ATTATGCCTTCAA CCAACGCGGATGGCACGATTGATCC GC GGCCATTTTGGCGATAGCGCCGTAA GTCG GAAGCAGCTCCAGCCTACACAGAC GTTGAATAATGGTATG TAAGGAGGATATTCATATGTCAAAC TGGTGAAATGC CCCTTCCCAACTAAGCAGCCATCGT GGC GCCAGCGGCATTGAACACACC GAAGCAGCTCCAGCCTACAC TTGCAGCACACTTTAAAGATGC TAAGGAGGATATTCATATG TAAGTTACGGCTAGTTAAGTAT GCTCCTGTGAAAACAGGAACTGATC CTAAAGTTCGACATCCCCTTGC GAAGCAGCTCCAGCCTACACGATCT TGCTATGCAAAATAGACACA TAAGGAGGATATTCATATGCTAATG TCCCAGTATCTCCCTG GTTGAGCGGCAGATCGTTTGAG 136 BamHI toxTpkas1pBK51 toxTpkas2pBK51 toxTpkas3pBK51 toxTpkas4pBK51 qrgB-CMVfw qrgB-CMV-rv VCA0848CMV-fw VCA0848CMV-rv VCA0848ASM VCA0956CMV-fw VCA0956CMV-rv VCA0956ASM For ∆toxT strain For ∆toxT strain For ∆toxT strain For ∆toxT strain For pShuttleCMV -qrgB For pShuttleCMV -qrgB For pShuttleCMV -VCA0848 For pShuttleCMV -VCA0848 For pShuttleCMV -VCA0848* For pShuttleCMV -VCA0956 For pShuttleCMV -VCA0956 For pShuttleCMV -VCA0956* Table A3 (cont’d) ATAGAGCTCCCTCATATCAAAGTTT ATGAAGGGAC ATAGCGGCCGCTGCGTTCTACTCTG AAGATATATAAATAA ATAGCGGCCGCATGGAATACGTTTA CTTGATCCTATTTTC ATAGGTACCCAAAGAGTCTAGCGTT TCAGAAATAG SacI NotI NotI KpnI ATACTCGAGTTAAGAAACCTTTTGG ATTCTTAAGTTG KpnI ATACTCGAGTTAAGAAACCTTTTGG ATTCTTAAGTTG XhoI ATAGGTACCCCACCATGAATGACAA AGTGCT KpnI ATACTCGAGTTAGAAAAGTTCAACG TCATCAGAA XhoI GTCTTCTCAACTATTTCGCTTTGCT GCTGAAGAGTTCGTGATTATTTTTT ATAGGTACCCCACCGTGATGACAAC TGAAGATTTCA KpnI ATACTCGAGTTAGAGCGGCATGACT CGAT XhoI TGACAGCTTATCGTTATGCCGCTGA AGAGTTTGCACTGAT 137 REFERENCES 138 REFERENCES 1. Ross, P., et al., Regulation of cellulose synthesis in Acetobacter xylinum by cyclic diguanylic acid. Nature, 1987. 325(6101): p. 279-281. 325 2. Seshasayee, A.S.N., G.M. Fraser, and N.M. Luscombe, Comparative genomics of cyclic-di-GMP signaling in bacteria: post-translational regulation and catalytic activity. Nucleic Acids Research, 2010. 38(18): p. 5970-5981. 38 3. Ausmees, N., et al., Genetic data indicate that proteins containing the GGDEF domain possess diguanylate cyclase activity. Fems Microbiology Letters, 2001. 204(1): p. 163-167. 204 4. Ryjenkov, D.A., et al., Cyclic diguanylate is a ubiquitous signaling molecule in bacteria: Insights into biochemistry of the GGDEF protein domain. Journal of Bacteriology, 2005. 187(5): p. 1792-1798. 187 5. Galperin, M.Y., Bacterial signal transduction network in a genomic perspective. Environmental Microbiology, 2004. 6(6): p. 552-567. 6. Ryan, R.P., et al., Cyclic di-GMP signaling in bacteria: Recent advances and new puzzles. Journal of Bacteriology, 2006. 188(24): p. 8327-8334. 88 7. Schmidt, A.J., D.A. Ryjenkov, and M. Gomelsky, The ubiquitous protein domain EAL is a cyclic diguanylate-specific phosphodiesterase: Enzymatically active and inactive EAL domains. Journal of Bacteriology, 2005. 187(14): p. 4774-4781. 187 8. Ryan, R.P., et al., Cell-cell signaling in Xanthomonas campestris involves an HD-GYP domain protein that functions in cyclic di-GMP turnover. Proceedings of the National Academy of Sciences of the United States of America, 2006. 103(17): p. 6712-6717. 103 9. Romling, U., M.Y. Galperin, and M. Gomelsky, Cyclic di-GMP: the First 25 Years of a Universal Bacterial Second Messenger. Microbiology and Molecular Biology Reviews, 2013. 77(1): p. 1-52. 77 139 10. Christen, B., et al., Allosteric control of cyclic di-GMP signaling. Journal of Biological Chemistry, 2006. 281(42): p. 32015-32024. 281 11. Simm, R., et al., GGDEF and EAL domains inversely regulate cyclic di-GMP levels and transition from sessility to motility. Molecular Microbiology, 2004. 53(4): p. 1123-1134. 53 12. Fang, X. and M. Gomelsky, A post-translational, c-di-GMP-dependent mechanism regulating flagellar motility. Molecular Microbiology, 2010. 76(5): 76 p. 1295-1305. 13. Boehm, A., et al., Second messenger-mediated adjustment of bacterial swimming velocity. Cell, 2010. 141(1): p. 107-116. 141 14. Cheng, K.J., R.P. McCowan, and J.W. Costerton, Adherent epithelial bacteria in ruminants and their roles in digestive-tract function. American Journal of Clinical Nutrition, 1979. 32(1): p. 139-148. 32 15. Costerton, J.W., P.S. Stewart, and E.P. Greenberg, Bacterial biofilms: A common cause of persistent infections. Science, 1999. 284(5418): p. 1318284 1322. 16. Tischler, A.D. and A. Camilli, Cyclic diguanylate (c-di-GMP) regulates Vibrio cholerae biofilm formation. Molecular Microbiology, 2004. 53(3): p. 857-869. 53 17. Galperin, M.Y., A.N. Nikolskaya, and E.V. Koonin, Novel domains of the prokaryotic two-component signal transduction systems. Fems Microbiology Letters, 2001. 203(1): p. 11-21. 203 18. Zähringer, F., et al., Structure and signaling mechanism of a zinc-sensory diguanylate cyclase. Structure, 2013: p. In press. 19. Bernier, S.P., et al., Modulation of Pseudomonas aeruginosa surface- associated group behaviors by individual amino acids through c-di-GMP signaling. Research in Microbiology, 2011. 162(7): p. 680-688. 162 140 20. Carlson, H.K., R.E. Vance, and M.A. Marletta, H-NOX regulation of c-di- GMP metabolism and biofilm formation in Legionella pneumophila. Molecular Microbiology, 2010. 77(4): p. 930-942. 77 21. Karatan, E., T.R. Duncan, and P.I. Watnick, NspS, a predicted polyamine sensor, mediates activation of Vibrio cholerae biofilm formation by norspermidine. Journal of Bacteriology, 2005. 187(21): 187 p. 7434-7443. 22. Wan, X.H., et al., Globins synthesize the second messenger bis-(3 '-5 ')-cyclic diguanosine monophosphate in bacteria. Journal of Molecular Biology, 2009. 388(2): p. 262-270. 388 23. Tuckerman, J.R., et al., An oxygen-sensing diguanylate cyclase and phosphodiesterase couple for c-di-GMP control. Biochemistry, 2009. 48(41): 48 p. 9764-9774. 24. Kanazawa, T., et al., Biochemical and physiological characterization of a BLUF protein-EAL protein complex involved in blue light-dependent degradation of cyclic diguanylate in the purple bacterium Rhodopseudomonas palustris. Biochemistry, 2010. 49(50): p. 10647-10655. 49 25. Agostoni, M., et al., Occurrence of Cyclic di-GMP-Modulating Output Domains in Cyanobacteria: an Illuminating Perspective. mBio, 2013. 4(4). 26. Ryan, R.P., et al., Cell-cell signal-dependent dynamic interactions between HD-GYP and GGDEF domain proteins mediate virulence in Xanthomonas campestris. Proceedings of the National Academy of Sciences of the United States of America, 2010. 107(13): p. 5989-5994. 107 27. Trimble, M.J. and L.L. McCarter, Bis-(3 '-5 ')-cyclic dimeric GMP-linked quorum sensing controls swarming in Vibrio parahaemolyticus. Proceedings of the National Academy of Sciences of the United States of America, 2011. 108(44): p. 18079-18084. 108 28. Huq, A., et al., Ecological relationships between Vibrio cholerae and planktonic crustacean copepods. Applied and Environmental Microbiology, 1983. 45(1): p. 275-283. 45 141 29. Tamplin, M.L., et al., Attachment of Vibrio cholerae serogroup-O1 to zooplankton and phytoplankton of Bangladesh waters. Applied and Environmental Microbiology, 1990. 56(6): p. 1977-1980. 56 30. Halpern, M., et al., Chironomid egg masses as a natural reservoir of Vibrio cholerae non-O1 and non-O139 in freshwater habitats. Microbial Ecology, 2004. 47(4): 47 p. 341-349. 31. Faruque, S.M., M.J. Albert, and J.J. Mekalanos, Epidemiology, genetics, and ecology of toxigenic Vibrio cholerae. Microbiology and Molecular Biology Reviews, 1998. 62(4): p. 1301-+. 62 32. DiRita, V.J., Co-ordinate expression of virulence genes by ToxR in Vibrio cholerae. Molecular Microbiology, 1992. 6(4): p. 451-458. 33. Bina, J., et al., ToxR regulon of Vibrio cholerae and its expression in vibrios shed by cholera patients. Proceedings of the National Academy of Sciences, 2003. 100(5): p. 2801-2806. 100 34. Tischler, A.D. and A. Camilli, Cyclic diguanylate regulates Vibrio cholerae virulence gene expression. Infection and Immunity, 2005. 73(9): p. 5873-5882. 73 35. Tamayo, R., et al., Role of cyclic di-GMP during El tor biotype Vibrio cholerae infection: Characterization of the in vivo-induced cyclic di-GMP phosphodiesterase CdpA. Infection and Immunity, 2008. 76(4): p. 1617-1627. 76 36. Martinez-Wilson, H.F., et al., The Vibrio cholerae hybrid sensor kinase VieS contributes to motility and biofilm regulation by altering the cyclic diguanylate level. Journal of Bacteriology, 2008. 190(19): p. 6439-6447. 190 37. Dawson, P.A., Bile Secretion and the Enterohepatic Circulation, in Sleisenger and Fordtran's Gastrointestinal and Liver Disease (9th), M. Feldman, L.S. Friedman, and L.J. Brandt, Editors. 2010, W.B. Saunders: Philadelphia, PA. p. 1075-1088. 38. O'Connor, C.J. and R.G. Wallace, Physicochemical behavior of bile-salts. Advances in Colloid and Interface Science, 1985. 22(1): p. 1-111. 22 142 39. Begley, M., C.G.M. Gahan, and C. Hill, The interaction between bacteria and bile. Fems Microbiology Reviews, 2005. 29(4): p. 625-651. 29 40. Bernstein, C., et al., Bile salt activation of stress response promoters in Escherichia coli. Current Microbiology, 1999. 39(2): p. 68-72. 39 41. Schubert, R. and K.H. Schmidt, Structural changes in vesicle membranes and mixed micelles of various lipid compositions after binding of different bile salts. Biochemistry, 1988. 27(24): p. 8787-8794. 27 42. Cabral, D.J., et al., Transbilayer movement of bile-acids in model membranes. Biochemistry, 1987. 26(7): p. 1801-1804. 26 43. Provenzano, D., et al., The virulence regulatory protein ToxR mediates enhanced bile resistance in Vibrio cholerae and other pathogenic Vibrio species. Infection and Immunity, 2000. 68(3): p. 1491-1497. 68 44. Colmer, J.A., J.A. Fralick, and A.N. Hamood, Isolation and characterization of a putative multidrug resistance pump from Vibrio cholerae. Molecular Microbiology, 1998. 27(1): p. 63-72. 27 45. Bina, J.E. and J.J. Mekalanos, Vibrio cholerae tolC is required for bile resistance and colonization. Infection and Immunity, 2001. 69(7): p. 468169 4685. 46. Bina, X.R., et al., Vibrio cholerae RND family efflux systems are required for antimicrobial resistance, optimal virulence factor production, and colonization of the infant mouse small intestine. Infection and Immunity, 2008. 76(8): p. 3595-3605. 76 47. Hung, D.T., et al., Bile acids stimulate biofilm formation in Vibrio cholerae. Molecular Microbiology, 2006. 59(1): p. 193-201. 59 48. McWhirter, S.M., et al., A host type I interferon response is induced by cytosolic sensing of the bacterial second messenger cyclic-di-GMP. Journal of Experimental Medicine, 2009. 206(9): p. 1899-1911. 206 143 49. Sauer, J.D., et al., The N-Ethyl-N-Nitrosourea-Induced Goldenticket Mouse Mutant Reveals an Essential Function of Sting in the In Vivo Interferon Response to Listeria monocytogenes and Cyclic Dinucleotides. Infection and Immunity, 2011. 79(2): p. 688-694. 79 50. Ebensen, T., et al., The bacterial second messenger cyclic diGMP exhibits potent adjuvant properties. Vaccine, 2007. 25(8): 25 p. 1464-1469. 51. Abdul-Sater, A.A., et al., Cyclic-di-GMP and cyclic-di-AMP activate the NLRP3 inflammasome. EMBO reports, 2013. 14(10): p. 900-906. 14 52. Ebensen, T., et al., The bacterial second messenger cdiGMP exhibits promising activity as a mucosal adjuvant. Clinical and Vaccine Immunology, 2007. 14(8): 14 p. 952-958. 53. Karaolis, D.K.R., et al., Bacterial c-di-GMP is an immunostimulatory molecule. Journal of Immunology, 2007. 178(4): p. 2171-2181. 178 54. Karaolis, D.K.R., et al., Cyclic Di-GMP stimulates protective innate immunity in bacterial pneumonia. Infection and Immunity, 2007. 75(10): p. 75 4942-4950. 55. Yan, H.B., et al., 3 ',5 '-Cyclic diguanylic acid elicits mucosal immunity against bacterial infection. Biochemical and Biophysical Research Communications, 2009. 387(3): p. 581-584. 387 56. Gray, P.M., et al., Evidence for cyclic diguanylate as a vaccine adjuvant with novel immunostimulatory activities. Cellular Immunology, 2012. 278(1-2): p. 278 113-119. 57. Blaauboer, S.M., V.D. Gabrielle, and L. Jin, MPYS/STING-Mediated TNF-α, Not Type I IFN, Is Essential for the Mucosal Adjuvant Activity of (3'-5')Cyclic-Di-Guanosine-Monophosphate In Vivo. Journal of Immunology, 2014. 192(1): p. 492-502. 192 58. Ogunniyi, A.D., et al., C-di-GMP is an effective immunomodulator and vaccine adjuvant against pneumococcal infection. Vaccine, 2008. 26(36): p. 26 4676-4685. 144 59. Chen, W.X., R. KuoLee, and H.B. Yan, The potential of 3 ',5 '-cyclic diguanylic acid (c-di-GMP) as an effective vaccine adjuvant. Vaccine, 2010. 28(18): p. 28 3080-3085. 60. Mosca, F., et al., Molecular and cellular signatures of human vaccine adjuvants. Proceedings of the National Academy of Sciences, 2008. 105(30): p. 105 10501-10506. 61. Coffman, R.L., A. Sher, and R.A. Seder, Vaccine Adjuvants: Putting Innate Immunity to Work. Immunity, 2010. 33(4): p. 492-503. 33 62. Gupta, R.K., Aluminum compounds as vaccine adjuvants. Advanced Drug Delivery Reviews, 1998. 32(3): p. 155-172. 32 63. Meibom, K.L., et al., Chitin induces natural competence in Vibrio cholerae. Science, 2005. 310(5755): p. 1824-1827. 310 64. Atkinson, S., et al., A hierarchical quorum-sensing system in Yersinia pseudotuberculosis is involved in the regulation of motility and clumping. Molecular Microbiology, 1999. 33(6): p. 1267-1277. 33 65. Hammer, B.K. and B.L. Bassler, Quorum sensing controls biofilm formation in Vibrio cholerae. Molecular Microbiology, 2003. 50(1): p. 101-114. 50 66. Hammer, B.K. and B.L. Bassler, Distinct sensory pathways in Vibrio cholerae El Tor and Classical biotypes modulate cyclic dimeric GMP levels to control biofilm formation. Journal of Bacteriology, 2009. 191(1): p. 169-177. 191 67. Miller, M.B., et al., Parallel quorum sensing systems converge to regulate virulence in Vibrio cholerae. Cell, 2002. 110(3): p. 303-314. 110 68. Zhu, J., et al., Quorum-sensing regulators control virulence gene expression in Vibrio cholerae. Proceedings of the National Academy of Sciences of the United States of America, 2002. 99(5): p. 3129-3134. 99 69. Lenz, D.H., et al., The Small RNA Chaperone Hfq and Multiple Small RNAs Control Quorum Sensing in Vibrio harveyi and Vibrio cholerae. Cell, 2004. 118(1): p. 69-82. 118 145 70. Hammer, B.K. and B.L. Bassler, Regulatory small RNAs circumvent the conventional quorum sensing pathway in pandemic Vibrio cholerae. Proceedings of the National Academy of Sciences of the United States of America, 2007. 104(27): p. 11145-11149. 104 71. Freeman, J.A. and B.L. Bassler, A genetic analysis of the function of LuxO, a two-component response regulator involved in quorum sensing in Vibrio harveyi. Molecular Microbiology, 1999. 31(2): p. 665-677. 31 72. Waters, C.M., et al., Quorum sensing controls biofilm formation in Vibrio cholerae through modulation of cyclic di-GMP levels and repression of vpsT. Journal of Bacteriology, 2008. 190(7): p. 2527-2536. 190 73. Srivastava, D. and C.M. Waters, A tangled web: Regulatory connections between quorum sensing and cyclic di-GMP. Journal of Bacteriology, 2012. 194(17): p. 4485-4493. 194 74. Yildiz, F.H., et al., Molecular analysis of rugosity in a Vibrio cholerae O1 El Tor phase variant. Molecular Microbiology, 2004. 53(2): p. 497-515. 53 75. Zhu, J. and J.J. Mekalanos, Quorum sensing-dependent biofilms enhance colonization in Vibrio cholerae. Developmental Cell, 2003. 5(4): p. 647-656. 76. Lim, B., et al., Cyclic-diGMP signal transduction systems in Vibrio cholerae: modulation of rugosity and biofilm formation. Molecular Microbiology, 2006. 60(2): p. 331-348. 60 77. Koestler, B.J. and C.M. Waters, Exploring Environmental Control of Cyclic di-GMP Signaling in Vibrio cholerae by Using the Ex Vivo Lysate Cyclic diGMP Assay (TELCA). Applied and Environmental Microbiology, 2013. 79(17): 79 p. 5233-41. 78. Thelin, K.H. and R.K. Taylor, Toxin-coregulated pilus, but not mannose- sensitive hemagglutinin, is required for colonization by Vibrio cholerae O1 El Tor biotype and O139 strains. Infection and Immunity, 1996. 64(7): p. 2853-6. 64 146 79. Massie, J.P., et al., Quantification of high-specificity cyclic diguanylate signaling. Proceedings of the National Academy of Sciences of the United States of America, 2012. 109(31): p. 12746-12751. 109 80. Iwanaga, M. and K. Yamamoto, New medium for the production of cholera toxin by Vibrio cholerae 01 biotype El Tor. Journal of Clinical Microbiology, 1985. 22(3): 22 p. 405-408. 81. Clark, D.J. and O. Maaløe, DNA replication and the division cycle in Escherichia coli. Journal of Molecular Biology, 1967. 23(1): p. 99-112. 23 82. Sambrook, J., Russell, D.W., Molecular Cloning - A Laboratory Manual. 3rd ed. 2001, Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press. 83. Svitil, A.L., et al., Chitin degradation proteins produced by the marine bacterium Vibrio harveyi growing on different forms of chitin. Applied and Environmental Microbiology, 1997. 63(2): p. 408-413. 63 84. Datsenko, K.A. and B.L. Wanner, One-step inactivation of chromosomal genes in Escherichia coli K-12 using PCR products. Proceedings of the National Academy of Sciences of the United States of America, 2000. 97(12): 97 p. 6640-6645. 85. Marvig, R.L. and M. Blokesch, Natural transformation of Vibrio cholerae as a tool - Optimizing the procedure. Bmc Microbiology, 2010. 10. 10 86. Long, T., et al., Quantifying the integration of quorum-sensing signals with single-cell resolution. Plos Biology, 2009. 7(3): p. 640-649. 87. Sambanthamoorthy, K., et al., Identification of Small Molecules That Antagonize Diguanylate Cyclase Enzymes To Inhibit Biofilm Formation. Antimicrobial Agents and Chemotherapy, 2012. 56(10): p. 5202-5211. 56 88. Bennett, B.D., et al., Absolute metabolite concentrations and implied enzyme active site occupancy in Escherichia coli. Nature Chemical Biology, 2009. 5(8): p. 593-599. 147 89. Bobrov, A.G., et al., Systematic analysis of cyclic di-GMP signalling enzymes and their role in biofilm formation and virulence in Yersinia pestis. Molecular Microbiology, 2011. 79(2): p. 533-551. 79 90. Bellows, L.E., et al., Hfq-dependent, co-ordinate control of cyclic diguanylate synthesis and catabolism in the plague pathogen Yersinia pestis. Molecular Microbiology, 2012. 86(3): 86 p. 661-674. 91. Newell, P.D., et al., Systematic Analysis of Diguanylate Cyclases That Promote Biofilm Formation by Pseudomonas fluorescens Pf0-1. Journal of Bacteriology, 2011. 193(18): p. 4685-4698. 193 92. Christen, M., et al., Asymmetrical Distribution of the Second Messenger c-diGMP upon Bacterial Cell Division. Science, 2010. 328(5983): 328 p. 1295-1297. 93. Romling, U., M. Gomelsky, and M.Y. Galperin, C-di-GMP: the dawning of a novel bacterial signaling system. Molecular Microbiology, 2005. 57(3): p. 62957 639. 94. Beyhan, S., L.S. Odell, and F.H. Yildiz, Identification and characterization of cyclic diguanylate signaling systems controlling rugosity in Vibrio cholerae. Journal of Bacteriology, 2008. 190(22): p. 7392-7405. 190 95. Fong, J.C.N. and F.H. Yildiz, Interplay between cyclic AMP-Cyclic AMP receptor protein and cyclic di-GMP signaling in Vibrio cholerae biofilm formation. Journal of Bacteriology, 2008. 190(20): p. 6646-6659. 190 96. Mueller, R.S., et al., Indole acts as an extracellular cue regulating gene expression in Vibrio cholerae. Journal of Bacteriology, 2009. 191(11): p. 3504191 3516. 97. Shikuma, N.J., J.C.N. Fong, and F.H. Yildiz, Cellular levels and binding of c- di-GMP control subcellular localization and activity of the Vibrio cholerae transcriptional regulator VpsT. Plos Pathogens, 2012. 8(5). 98. Kulasakara, H., et al., Analysis of Pseudomonas aeruginosa diguanylate cyclases and phosphodiesterases reveals a role for bis-(3 '-5 ')-cyclic-GMP in 148 virulence. Proceedings of the National Academy of Sciences of the United States of America, 2006. 103(8): p. 2839-44. 103 99. Hecht, G.B. and A. Newton, Identification of a novel response regulator required for the swarmer-to-stalked-cell transition in Caulobacter crescentus. Journal of Bacteriology, 1995. 177(21): p. 6223-6229. 177 100. Aldridge, P., et al., Role of the GGDEF regulator PleD in polar development of Caulobacter crescentus. Molecular Microbiology, 2003. 47(6): p. 1695-1708. 47 101. Edmunds, A.C., et al., Cyclic di-GMP Modulates the Disease Progression of Erwinia amylovora. Journal of Bacteriology, 2013. 195(10): p. 2155-65. 195 102. Beyhan, S., et al., Regulation of rugosity and biofilm formation in Vibrio cholerae: Comparison of VpsT and VpsR regulons and epistasis analysis of vpsT, vpsR, and hapR. Journal of Bacteriology, 2007. 189(2): p. 388-402. 189 103. de Lorenzo, V. and K.N. Timmis, Analysis and construction of stable phenotypes in Gram-negative bacteria with Tn5-derived and Tn10-derived minitransposons. Bacterial Pathogenesis, Pt A, 1994. 235: 235 p. 386-405. 104. Graham, D.Y. and M.S. Osato, H. pylori in the pathogenesis of duodenal ulcer: Interaction between duodenal acid load, bile, and H. pylori. American Journal of Gastroenterology, 2000. 95(1): p. 87-91. 95 105. Northfield, T.C. and I. McColl, Postprandial concentrations of free and conjugated bile acids down the length of the normal human small intestine. Gut, 1973. 14(7): p. 513-518. 14 106. Sambrook, J., E. F. Fritsch, and T. Maniatis., Molecular Cloning - A Laboratory Manual, 2nd Edition. 2 ed. 1989, Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press. 107. Srivastava, D., R.C. Harris, and C.M. Waters, Integration of cyclic di-GMP and quorum sensing in the control of vpsT and aphA in Vibrio cholerae. Journal of Bacteriology, 2011. 193(22): p. 6331-6341. 193 149 108. Skorupski, K. and R.K. Taylor, Positive selection vectors for allelic exchange. Gene, 1996. 169(1): p. 47-52. 169 109. Fürste, J.P., et al., Molecular cloning of the plasmid RP4 primase region in a multi-host-range tacP expression vector. Gene, 1986. 48(1): p. 119-131. 48 110. Hofmann, A.F., The continuing importance of bile acids in liver and intestinal disease. Archives of Internal Medicine, 1999. 159(22): 159 p. 2647-2658. 111. Claros, M.G. and G. Vonheijne, TOPPRED-II - an improved software for membrane-protein structure predictions. Computer Applications in the Biosciences, 1994. 10(6): p. 685-686. 10 112. Neron, B., et al., Mobyle: a new full web bioinformatics framework. Bioinformatics, 2009. 25(22): 25 p. 3005-3011. 113. Schultz, J., et al., SMART: a web-based tool for the study of genetically mobile domains. Nucleic Acids Research, 2000. 28(1): p. 231-234. 28 114. Goss, T.J., et al., ToxR recognizes a direct repeat element in the toxT, ompU, ompT and ctxA promoters of Vibrio cholerae to regulate transcription. Infection and Immunity, 2013. 81(3): p. 884-95. 81 115. Abuaita, B.H. and J.H. Withey, Termination of Vibrio cholerae virulence gene expression is mediated by proteolysis of the major virulence activator, ToxT. Molecular Microbiology, 2011. 81(6): p. 1640-1653. 81 116. Provenzano, D. and K.E. Klose, Altered expression of the ToxR-regulated porins OmpU and OmpT diminishes Vibrio cholerae bile resistance, virulence factor expression, and intestinal colonization. Proceedings of the National Academy of Sciences of the United States of America, 2000. 97(18): p. 1022097 10224. 117. Hung, D.T. and J.J. Mekalanos, Bile acids induce cholera toxin expression in Vibrio cholerae in a ToxT-independent manner. Proceedings of the National Academy of Sciences of the United States of America, 2005. 102(8): p. 3028102 3033. 150 118. Seidler, U. and M. Sjöblom, Gastroduodenal Bicarbonate Secretion, in Physiology of the Gastrointestinal Tract, L.R. Johnson, Editor. 2012, Elsevier: San Diego. p. 1311-1339. 119. Hogan, D.L., M.A. Ainsworth, and J.I. Isenberg, Gastroduodenal bicarbonate secretion. Alimentary Pharmacology & Therapeutics, 1994. 8(5): p. 475-488. 120. Abuaita, B.H. and J.H. Withey, Bicarbonate induces Vibrio cholerae virulence gene expression by enhancing ToxT activity. Infection and Immunity, 2009. 77(9): p. 4111-4120. 77 121. Schuhmacher, D.A. and K.E. Klose, Environmental signals modulate ToxTdependent virulence factor expression in Vibrio cholerae. Journal of Bacteriology, 1999. 181(5): 181 p. 1508-1514. 122. Yang, M.H., et al., Bile salt-induced intermolecular disulfide bond formation activates Vibrio cholerae virulence. Proceedings of the National Academy of Sciences of the United States of America, 2013. 110(6): p. 2348-2353. 110 123. Hunter, J.L., et al., The Vibrio cholerae diguanylate cyclase VCA0965 has an AGDEF active site and synthesizes cyclic di-GMP. Bmc Microbiology, 2014. 14(1): p. 22. 14 124. Gu, J.Y., et al., Genomic and systems evolution in Vibrionaceae species. Bmc Genomics, 2009. 10: 10 p. 13. 125. Begley, M., C.G.M. Gahan, and C. Hill, Bile stress response in Listeria monocytogenes LO28: Adaptation, cross-protection, and identification of genetic loci involved in bile resistance. Applied and Environmental Microbiology, 2002. 68(12): p. 6005-6012. 68 126. Tamayo, R., B. Patimalla, and A. Camilli, Growth in a Biofilm Induces a Hyperinfectious Phenotype in Vibrio cholerae. Infection and Immunity, 2010. 78(8): p. 3560-3569. 78 127. Quigley, E.M.M. and L.A. Turnberg, pH of the microclimate lining human gastric and duodenal mucosa in vivo. Gastroenterology, 1987. 92(6): p. 187692 1884. 151 128. Prouty, A.M. and J.S. Gunn, Salmonella enterica serovar typhimurium invasion is repressed in the presence of bile. Infection and Immunity, 2000. 68(12): p. 6763-6769. 68 129. Krukonis, E.S. and V.J. DiRita, From motility to virulence: sensing and responding to environmental signals in Vibrio cholerae. Current Opinion in Microbiology, 2003. 6(2): p. 186-190. 130. Cybulski, L.E., et al., Membrane Thickness Cue for Cold Sensing in a Bacterium. Current Biology, 2010. 20(17): p. 1539-1544. 20 131. Mansilla, M.C., et al., Control of membrane lipid fluidity by molecular thermosensors. Journal of Bacteriology, 2004. 186(20): p. 6681-6688. 186 132. Kingston, A.W., et al., A sigma(W)-dependent stress response in Bacillus subtilis that reduces membrane fluidity. Molecular Microbiology, 2011. 81(1): 81 p. 69-79. 133. Taranto, M.P., et al., Bile salts and cholesterol induce changes in the lipid cell membrane of Lactobacillus reuteri. Journal of Applied Microbiology, 2003. 95(1): p. 86-91. 95 134. Choi, K.-H., R.J. Heath, and C.O. Rock, β-Ketoacyl-Acyl Carrier Protein Synthase III (FabH) Is a Determining Factor in Branched-Chain Fatty Acid Biosynthesis. Journal of Bacteriology, 2000. 182(2): p. 365-370. 182 135. Cerda-Maira, F.A., C.S. Ringelberg, and R.K. Taylor, The Bile Response Repressor BreR Regulates Expression of the Vibrio cholerae breAB Efflux System Operon. Journal of Bacteriology, 2008. 190(22): p. 7441-7452. 190 136. Wibbenmeyer, J.A., et al., Vibrio cholerae OmpU and OmpT porins are differentially affected by bile. Infection and Immunity, 2002. 70(1): p. 12170 126. 137. Oude Elferink, R.P., et al., Regulation of biliary lipid secretion by mdr2 Pglycoprotein in the mouse. The Journal of Clinical Investigation, 1995. 95(1): 95 p. 31-38. 152 138. Klappenbach, J.A., J.M. Dunbar, and T.M. Schmidt, rRNA Operon Copy Number Reflects Ecological Strategies of Bacteria. Applied and Environmental Microbiology, 2000. 66(4): p. 1328-1333. 66 139. Miura, A., et al., Growth-rate-dependent regulation of ribosome synthesis in E. coli: Expression of the lacZ and galK genes fused to ribosomal promoters. Cell, 1981. 25(3): 25 p. 773-782. 140. Sarmientos, P. and M. Cashel, Carbon starvation and growth rate-dependent regulation of the Escherichia coli ribosomal RNA promoters: differential control of dual promoters. Proceedings of the National Academy of Sciences, 1983. 80(22): p. 7010-7013. 80 141. Nomura, M., R. Gourse, and G. Baughman, Regulation of the Synthesis of Ribosomes and Ribosomal Components. Annual Review of Biochemistry, 1984. 53(1): p. 75-117. 53 142. Gourse, R.L., H.A. de Boer, and M. Nomura, DNA determinants of rRNA synthesis in E. coli: Growth rate dependent regulation, feedback inhibition, upstream activation, antitermination. Cell, 1986. 44(1): p. 197-205. 44 143. Condon, C., C. Squires, and C.L. Squires, Control of ribosomal-RNA Transcription in Escherichia-coli. Microbiological Reviews, 1995. 59(4): p. 59 623-&. 144. Gourse, R.L., et al., Rrna transcription and growth rate-dependent regulation of ribosome synthesis in Escherichia coli. Annual Review of Microbiology, 1996. 50: 50 p. 645-677. 145. Schneider, D.A., W. Ross, and R.L. Gourse, Control of rRNA expression in Escherichia coli. Current Opinion in Microbiology, 2003. 6(2): p. 151-156. 146. Jin, D.J., C. Cagliero, and Y.N. Zhou, Growth rate regulation in Escherichia coli. Fems Microbiology Reviews, 2012. 36(2): p. 269-287. 36 147. Svitil, A.L., M. Cashel, and J.W. Zyskind, Guanosine tetraphosphate inhibits protein synthesis in vivo. A possible protective mechanism for starvation 153 stress in Escherichia coli. Journal of Biological Chemistry, 1993. 268(4): p. 268 2307-11. 148. Traxler, M.F., et al., The global, ppGpp-mediated stringent response to amino acid starvation in Escherichia coli. Molecular Microbiology, 2008. 68(5): p. 68 1128-1148. 149. Potrykus, K. and M. Cashel, (p)ppGpp: Still Magical? Annual Review of Microbiology, 2008. 62: 62 p. 35-51. 150. Srivatsan, A. and J.D. Wang, Control of bacterial transcription, translation and replication by (p)ppGpp. Current Opinion in Microbiology, 2008. 11(2): p. 11 100-105. 151. Kriel, A., et al., Direct Regulation of GTP Homeostasis by (p)ppGpp: A Critical Component of Viability and Stress Resistance. Molecular Cell, 2012. 48(2): p. 231-241. 48 152. Kahm, M., et al., grofit: Fitting Biological Growth Curves with R. Journal of Statistical Software, 2010. 33(7): p. 1-21. 33 153. R Development Core Team, R: A language and environment for statistical computing. 2008, R Foundation for Statistical Computing Vienna, Austria. 154. Stokke, C., T. Waldminghaus, and K. Skarstad, Replication patterns and organization of replication forks in Vibrio cholerae. Microbiology, 2011. 157(3): p. 695-708. 157 155. Merritt, J.H., et al., Specific Control of Pseudomonas aeruginosa SurfaceAssociated Behaviors by Two c-di-GMP Diguanylate Cyclases. mBio, 2010. 1(4). 156. Kader, A., et al., Hierarchical involvement of various GGDEF domain proteins in rdar morphotype development of Salmonella enterica serovar Typhimurium. Molecular Microbiology, 2006. 60(3): p. 602-616. 60 154 157. Stevenson, B.S. and T.M. Schmidt, Life History Implications of rRNA Gene Copy Number in Escherichia coli. Applied and Environmental Microbiology, 2004. 70(11): p. 6670-6677. 70 158. Ross, W., et al., The Magic Spot: A ppGpp Binding Site on E. coli RNA Polymerase Responsible for Regulation of Transcription Initiation. Molecular Cell, 2013. 50(3): 50 p. 420-429. 159. Krásný, L. and R. Gourse, An alternative strategy for bacterial ribosome synthesis: Bacillus subtilis rRNA transcription regulation. The EMBO Journal, 2004. 23(22): p. 4473-83. 23 160. Anderson, J.D., et al., Role of bacterial growth rates in the epidemiology and pathogenesis of urinary infections in women. Journal of Clinical Microbiology, 1979. 10(6): p. 766-771. 10 161. Klumpp, S., Z.G. Zhang, and T. Hwa, Growth Rate-Dependent Global Effects on Gene Expression in Bacteria. Cell, 2009. 139(7): p. 1366-1375. 139 162. Hartman, Z.C., D.M. Appledorn, and A. Amalfitano, Adenovirus vector induced innate immune responses: Impact upon efficacy and toxicity in gene therapy and vaccine applications. Virus Research, 2008. 132(1-2): p. 1-14. 132 163. Aldhamen, Y.A., S.S. Seregin, and A. Amalfitano, Immune recognition of gene transfer vectors: focus on adenovirus as a paradigm. Frontiers in immunology, 2011. 2(40): p. 1-12. 164. Fukazawa, T., et al., Adenovirus-mediated cancer gene therapy and virotherapy (Review). International Journal of Molecular Medicine, 2010. 25(1): p. 3-10. 25 165. Seregin, S.S., et al., Use of DAF-Displaying Adenovirus Vectors Reduces Induction of Transgene- and Vector-Specific Adaptive Immune Responses in Mice. Human Gene Therapy, 2010. 22(9): p. 1083-1094. 22 166. Seregin, S.S., et al., CR1/2 is an important suppressor of Adenovirus-induced innate immune responses and is required for induction of neutralizing antibodies. Gene Therapy, 2009. 16(10): p. 1245-1259. 16 155 167. Seregin, S.S., et al., Adenovirus capsid-display of the retro-oriented human complement inhibitor DAF reduces Ad vector-triggered immune responses in vitro and in vivo. Blood, 2010. 116(10): p. 1669-1677. 116 168. Amalfitano, A., et al., Production and Characterization of Improved Adenovirus Vectors with the E1, E2b, and E3 Genes Deleted. Journal of Virology, 1998. 72(2): 72 p. 926-933. 169. Morgan, J., P. Ng, and F. Graham, Construction of First-Generation Adenoviral Vectors, in Gene Therapy Protocols. 2002, Springer New York. p. 389-414. 170. Seregin, S.S., et al., Adenovirus-based vaccination against Clostridium difficile toxin A allows for rapid humoral immunity and complete protection from toxin A lethal challenge in mice. Vaccine, 2012. 30(8): p. 1492-1501. 30 171. Seregin, S.S., et al., Transient Pretreatment With Glucocorticoid Ablates Innate Toxicity of Systemically Delivered Adenoviral Vectors Without Reducing Efficacy. Molecular Therapy, 2009. 17(4): p. 685-696. 17 172. Appledorn, D.M., et al., Wild-type adenoviruses from groups A-F evoke unique innate immune responses, of which HAd3 and SAd23 are partially complement dependent. Gene Therapy, 2008. 15(12): p. 885-901. 15 173. Theriot, C.M., et al., Cefoperazone-treated mice as an experimental platform to assess differential virulence of Clostridium difficile strains. Gut Microbes, 2011. 2(6): p. 326-334. 174. Reeves, A.E., et al., The interplay between microbiome dynamics and pathogen dynamics in a murine model of Clostridium difficile Infection. Gut Microbes, 2011. 2(3): p. 145-158. 175. Nakhamchik, A., C. Wilde, and D.A. Rowe-Magnus, Cyclic-di-GMP regulates extracellular polysaccharide production, biofilm formation, and rugose colony development by Vibrio vulnificus. Applied and Environmental Microbiology, 2008. 74(13): p. 4199-4209. 74 156 176. Thormann, K.M., et al., Control of formation and cellular detachment from Shewanella oneidensis MR-1 biofilms by cyclic di-GMP. Journal of Bacteriology, 2006. 188(7): p. 2681-2691. 188 177. Everett, R.S., et al., Liver Toxicities typically induced by first-generation adenoviral vectors can be reduced by use of E1, E2b-deleted adenoviral vectors. Human Gene Therapy, 2003. 14(18): 14 p. 1715-1726. 178. Nakamura, T., K. Sato, and H. Hamada, Reduction of natural adenovirus tropism to the liver by both ablation of fiber-coxsackievirus and adenovirus receptor interaction and use of replaceable short fiber. Journal of Virology, 2003. 77(4): p. 2512-2521. 77 179. Tanaka, Y. and Z.J. Chen, STING Specifies IRF3 Phosphorylation by TBK1 in the Cytosolic DNA Signaling Pathway. Science Signaling, 2012. 5(214): p. ra20-. 180. de Almeida, L.A., et al., MyD88 and STING Signaling Pathways Are Required for IRF3-Mediated IFN- Induction in Response to Brucella abortus Infection. PLoS ONE, 2011. 6(8): p. e23135. 181. Lucado, J., C. Gould, and A. Elixhauser, Clostridium difficile Infections (CDI) in Hospital Stays, 2009, in HCUP Statistical Brief. 2012, Agency for Healthcare Research and Quality: Rockville, MD. 182. Morris, A.M., et al., Clostridium difficile colitis: An increasingly aggressive iatrogenic disease? Archives of Surgery, 2002. 137(10): p. 1096-1100. 137 183. Redelings, M.D., F. Sorvillo, and L. Mascola, Increase in Clostridium difficile–related mortality rates, United States, 1999–2004. 2007, Emerg Infect Dis. 184. Kyne, L., et al., Health Care Costs and Mortality Associated with Nosocomial Diarrhea Due to Clostridium difficile. Clinical Infectious Diseases, 2002. 34(3): p. 346-353. 34 157 185. Dubberke, E.R. and A.I. Wertheimer, Review of Current Literature on the Economic Burden of Clostridium difficile Infection. Infection Control and Hospital Epidemiology, 2009. 30(1): p. 57-66. 186. Aslam, S., R.J. Hamill, and D.M. Musher, Treatment of Clostridium difficileassociated disease: old therapies and new strategies. The Lancet Infectious Diseases, 2005. 5(9): p. 549-557. 187. Cartman, S.T., et al., The emergence of 'hypervirulence' in Clostridium difficile. International Journal of Medical Microbiology, 2010. 300(6): p. 387300 395. 188. Reed, S.G., et al., New horizons in adjuvants for vaccine development. Trends in Immunology, 2009. 30(1): 30 p. 23-32. 189. Zandman-Goddard, G. and Y. Shoenfeld, Infections and SLE. Autoimmunity, 2005. 38(7): p. 473-485. 38 190. Wolins, N., et al., Intravenous administration of replication-incompetent adenovirus to rhesus monkeys induces thrombocytopenia by increasing in vivo platelet clearance. British Journal of Haematology, 2003. 123(5): p. 903123 905. 191. Appledorn, D.M., et al., Complex interactions with several arms of the complement system dictate innate and humoral immunity to adenoviral vectors. Gene Therapy, 2008. 15(24): p. 1606-1617. 15 192. Schiedner, G., S. Hertel, and S. Kochanek, Efficient transformation of primary human amniocytes by E1 functions of Ad5: Generation of new cell lines for adenoviral vector production. Human Gene Therapy, 2000. 11(15): p. 11 2105-2116. 193. Barker, J.R., et al., STING-Dependent Recognition of Cyclic di-AMP Mediates Type I Interferon Responses during Chlamydia trachomatis Infection. mBio, 2013. 4(3). 158 194. Woodward, J.J., A.T. Iavarone, and D.A. Portnoy, c-di-AMP Secreted by Intracellular Listeria monocytogenes Activates a Host Type I Interferon Response. Science, 2010. 328(5986): p. 1703-1705. 328 195. Dunn, A.K., et al., New rfp- and pES213-derived tools for analyzing symbiotic Vibrio fischeri reveal patterns of infection and lux expression in situ. Applied and Environmental Microbiology, 2006. 72(1): 72 p. 802-810. 196. Lenz, D.H., et al., CsrA and three redundant small RNAs regulate quorum sensing in Vibrio cholerae. Molecular Microbiology, 2005. 58(4): p. 1186-1202. 58 159