Impacts of Mucins on Microbial Physiology and Interactions by Nicole Lynn Kavanaugh B.S. Biology Hofstra University, 2008 Submitted to the Microbiology Graduate Program in partial fulfillment of the requirements for the degree of Doctor of Philosophy at the Massachusetts Institute of Technology June 2015 © 2015 Nicole Kavanaugh. All Rights Reserved. The author hereby grants to MIT permission to reproduce and to distribute publicly paper and electronic copies of this thesis document in whole or in part in any medium now known or hereafter created. Author ……………………………………………………………………………………………... Nicole L. Kavanaugh Microbiology Graduate Program May 22, 2015 Certified by………………………………………………………………………………………... Katharina Ribbeck Assistant Professor of Biological Engineering Thesis Supervisor Accepted by………………………………………………………………………………………... Michael Laub Professor of Biology Director Microbiology Graduate Program 2 Impacts of Mucins on Microbial Physiology and Interactions by Nicole Lynn Kavanaugh Submitted to the Microbiology Graduate Program in partial fulfillment of the requirements for the degree of Doctor of Philosophy at the Massachusetts Institute of Technology ABSTRACT The human body is colonized by trillions of microbes known collectively as the microbiota. Many of these organisms inhabit mucosal surfaces, with most found in the large intestines, but many also dwell in the respiratory and urogenital tracts. Despite the enormous microbial population inhabiting the body, some of which are opportunistic pathogens, most people harbor these organisms without showing any signs of disease. The mucus that covers the wet epithelium and houses the microbiota is a prime candidate for offering protection from pathogens, yet its specific role is poorly understood. The object of this thesis is to explore the effects of mucins, the main gel-forming component of mucus, on microbial behavior. Using an in vitro mucus model consisting of purified mucins, I show that these polymers suppress microbial virulence traits in selected bacterial and fungal opportunistic pathogens and influence the composition of microbial communities. In Chapter 2, I study the impact of mucins on bacterial biofilm formation using the opportunistic pathogen Pseudomonas aeruginosa. I demonstrate that mucins reduce surface attached biofilm formation. However, P. aeruginosa can overcome mucin-induced biofilm suppression if flagellar motility is lost, allowing them to form non-surface attached biofilms that are suspended in mucins. In Chapter 3, I analyze the effects of mucins on the virulence traits of Candida albicans, a fungal opportunistic pathogen. The results show that mucins broadly suppress virulence traits of this organism, such as surface attachment, hyphal formation and biofilm formation, at both the levels of gene expression and phenotype. In Appendix A, I combine P. aeruginosa and C. albicans with mucins to determine the influence of these molecules on interspecies interactions. Whereas P. aeruginosa typically kills C. albicans, mucins protect the fungus from bacterial pathogenicity. Therefore, in addition to influencing microbial virulence, mucins can impact microbial community dynamics. Overall, my work suggests that mucins protect the body from microbes by functioning as a modulator of microbial behavior, coercing certain microbes to downregulate virulence gene and trait expression, thereby influencing microbial effects on the host and within the microbiota. Thesis Supervisor: Katharina Ribbeck Title: Assistant Professor of Biological Engineering 3 4 Acknowledgements First and foremost I would like to thank my thesis mentor Katharina Ribbeck. The passion that Katharina shows for her work is what inspired me to join her lab. She was always there when I needed advice but also gave me generous freedom to pursue my ideas. For this, I will always be grateful. I would like to thank my past and present thesis committee members, Chris Kaiser, Gerry Fink, Tim Lu and Jacquin Niles, as well as Katherine Lemon for joining from the Forsyth Institute. I am very thankful for your time as well as your indispensible advice throughout the years. Thank you to Clarissa Nobile, who has been a tremendously supportive collaborator. I am also very grateful to Alan Grossman for his encouragement when I needed it most. A very special thank you to all the Ribbeckers. Thank you for listening when I wanted to run an idea by you, for always being willing to get coffee or to eat way too much food, and for being great friends. Thank you to Thomas, Nicole B. (AKA “Tall Nicole”), Andrew, Tu, Julia, Leon, Erica, Kate, Wes, Tahoura, and Brad. A special thank you to the original crew, Marina, Oliver, Alex and Regi, an amazing group that helped me get acclimated when I first joined the Ribbeck Lab. I would also like to thank the undergraduate researchers that worked with me, Anna, Emily and Angela. A special thanks to Nicole B, Julia, Erica and Kate for their comments on this thesis. My acknowledgements would not be complete without mentioning the unyielding support I received from my family and friends. Thank you to my friends at the Thirsty Ear Pub and in the Microbiology Program. Not only are you brilliant, you’re incredibly fun. My best friends, Sam, Jackie, Alison, Liz, Sevanne and Danielle, live in NY but have been with me every step of the way. The laughs and support they have given me have been essential to completing my Ph.D. To Reid, thank you for inspiring me to work my hardest, for always listening, and for being so easy to love. To my Aunt Mary Lou who never hesitated to help me when I asked for it, thank you. And to my parents and sisters, Mom, Dad, Kristen and Tara, your constant encouragement means so much to me. It is from you that I get my work ethic, my strength, and my weird sense of humor. I love you all so much and thank you from the bottom of my heart. 5 Table of Contents Abstract 3 Acknowledgments 5 List of Figures 7 List of Tables 9 Chapter 1 Introduction 10 Chapter 2 Mucin biopolymers prevent bacterial aggregation by retaining 33 cells in the free-swimming state Chapter 3 Mucins suppress virulence traits of Candida albicans 58 Appendix A Mucins suppress Pseudomonas aeruginosa virulence toward 83 Candida albicans Appendix B Selected antimicrobial essential oils eradicate Pseudomonas spp. and 91 Staphylococcus aureus biofilms Chapter 4 Conclusions and future directions 6 106 List of Figures Chapter 1 1-1 Schematic depicting the structure of mucins 13 1-2 Mucins protect the epithelium from colitis and bacterial contact 17 1-3 Mechanisms used by pathogenic microbes to overcome the mucus layer 19 1-4 Models for studying the interactions between microbes and mucus 20 1-5 SEM micrographs of natively purified and industrially purified mucins 22 1-6 Model of mucus protection against pathogenic microbes 24 Chapter 2 2-1 Mucins reduce bacterial biofilm formation 37 2-2 P. aeruginosa swimming velocity is unperturbed by mucins 38 2-3 Nonmotile P. aeruginosa Form Flocs in Mucin Environments 39 2-4 The loss of flagellar motility supports floc formation 40 2-5 Selected viscous polymer solutions support the formation of P. aeruginosa flocs 41 2-6 P. aeruginosa cystic fibrosis clinical isolates form flocs in mucins 42 2-7 P. aeruginosa floc formation in mucins is not dependent on alginate 43 2-8 Flocs grown in mucins are antibiotic resistant 45 Chapter 3 3-1 Mucins induce a unique morphological state characterized by suppressed virulence traits 61 3-2 The mucin-induced morphological state is distinct from the opaque state 64 3-3 qPCR of known C. albicans virulence genes comparing gene expression in RPMI with 65 and without mucins 3-4 The effects of osmotic stress and viscosity on hyphal formation 66 3-5 Mucins suppress hyphae formation in YPD + FBS 67 3-6 Mucins suppress hyphal growth from both yeast and hyphal cells 68 3-7 Mucins reduce attachment of C. albicans to polystyrene and mucus-secreting cells 70 3-8 Mucins reduce C. albicans biofilm formation 72 7 Appendix A A-1 Mucins protect wild type C. albicans from P. aeruginosa pathogenicity 85 A-2 Mucins reduce P. aeruginosa attachment to C. albicans 87 Appendix B B-1 Cassia oil kills planktonic bacteria and biofilms with comparable efficiency 94 B-2 Disc diffusion assay identifies essential oils with antimicrobial activity 96 B-3 Activities of selected antibiotics and antimicrobial essential oils against P. aeruginosa 98 PAO1 and P. putida KT2440 B-4 Comparison of two methods of biofilm cultivation for antibiotic and essential oil testing 100 B-5 Susceptibility of S. aureus SC-01 to essential oils 8 101 List of Tables Chapter1 1-1: Human mucin genes and their expression patterns throughout the body 12 Chapter 2 2-1 List of strains and plasmids used in this study 51 Chapter 3 3-1 Analysis of opaque state attributes upon exposure to mucins 63 Appendix B B-1 Minimum inhibitory concentration of colistin and essential oils as determined by the 93 standard microbroth dilution assay B-2 MIC and MBEC of Essential Oil Components Against PAO1 9 102 Chapter 1 Introduction 10 Introduction The mucus found in the human body covers a vast surface area of epithelial cells that are exposed to the environment. It coats the respiratory, digestive, and urogenital tracts, as well as the ocular surface. The mucus barrier is the first line of defense that protects our bodies from environmental threats including toxins, viruses, bacteria and fungi. However, the mucus layer is home to a significant portion of the microbiota, or the native population of microbes that inhabit the body. Despite the enormous microbial population inhabiting the body, many people harbor these organisms without showing any signs of disease. Certain mucosal infections, such as those related to inflammatory bowel disease, are accompanied by a disruption of the mucus, suggesting its importance in keeping infections at bay [1–4]. However, the mechanisms behind mucusmediated protection of the body are not well understood. In this thesis, I explore the capacity of mucins, the main gel-forming molecules of mucus, to act as regulators of microbial virulence. Using the bacterium Pseudomonas aeruginosa and the fungus Candida albicans as model organisms, I determine the effects of mucins on the physiology of these opportunistic pathogens. My work shows that mucins can suppress microbial virulence trait expression, highlighting the importance of mucins as regulators of virulence. Additionally, by combining P. aeruginosa and C. albicans inside a mucin environment, I investigate the role of mucins in mediating interspecies interactions. Mucins stabilize the coexistence of these two organisms, suggesting that these polymers not only impact virulence traits of individual species but can also influence the structure of multispecies microbial communities. Mucins are an integral mucus component Mucus is a mixture of a number of components, including water (~95%), lipids, proteins and antimicrobial peptides. The slimy, jelly-like consistency of mucus is attributed to mucin polymers. Mucins are glycoproteins consisting of a protein backbone rich in serine and threonine residues that are heavily glycosylated, resulting in large (100-10,000kD) molecules. Mucins polymerize via interacting peptide domains and tangle to form a hydrated mucus gel. The human body produces cell-surface associated and secreted mucins. Secreted mucins generate the 11 Mucin Mucin type MUC1 Membrane MUC2 MUC3A MUC3B Secreted Membrane Membrane MUC4 Membrane MUC5AC Secreted MUC5B Secreted MUC6 Secreted MUC7 Secreted MUC8 Secreted MUC9 Secreted MUC11 Membrane MUC12 MUC13 Membrane Membrane MUC15 Membrane MUC16 Membrane MUC17 Membrane MUC19 Secreted MUC20 Membrane MUC21 Membrane Normal expression pattern Epithelial surfaces of the respiratory, female reproductive, and gastrointestinal tracts as well as in the middle ear, salivary, and mammary glands Intestinal and colonic goblet cells Small and large intestine, thymus, liver, lymph nodes, and heart Small and large intestine, thymus, liver, lymph nodes, and heart Epithelial surfaces of the eye, oral cavity, middle ear, lachrymal glands, salivary glands, mammary gland, prostate gland, stomach, colon, lung, trachea, and female reproductive tract. Tracheobronchial goblet cells and in the gastric epithelial cells Salivary, tracheobronchial, and esophageal mucous glands as well as in the pancreatobiliary and endocervical epithelial cells Gastric and duodenal mucous glands, pancreatobiliary, and endocervical epithelial cells Oral cavity epithelial cells, minor salivary gland, and possibly in the respiratory tract, pancreas, and bladder Airway and middle ear epithelial cells and male and female reproductive tracts. Female reproductive tract May represent a differential splice variant of MUC12; Expressed in the colon, stomach, middle ear, and lung epithelium Stomach and colon Gastrointestinal and respiratory tracts Lung, mammary gland, hematopoietic tissues, gonads, and gastrointestinal tract Ocular surface, respiratory tract, and female reproductive tract epithelia Gastrointestinal tract with the highest expression in the duodenum and conjunctival epithelium Mucosal cells of major salivary glands and the epithelial cells from corneal, conjunctival, lacrimal gland, middle ear, and trachea Highly expressed in the kidneys and moderately in the placenta, colon, lung, prostate, and liver Lung, large intestine, thymus, and testis Table 1: Human mucin genes and their expression patterns throughout the body. Adapted from [5] with permission from Springer Science and Business Media. viscoelastic properties of the mucus gel and surface-associated mucins are found in the glycocalyx, a dense layer of glycoproteins and glycolipids directly associated with the top of the epithelial layer. Different mucin types are produced in different body regions; there are roughly 20 human mucin genes (Table 1). 12 Figure 1 Schematic depicting the structure of mucins. A) Mucins contain a both nonglycosylated and glycosylated domains. The nonglycosylated domains are involved in polymerization and mesh formation. The glycosylated domains allow mucins to retain water to form a hydrogel. New sketch based on [6] with permission from Elsevier. B) Core glycan structures 1-4 and examples of mucin glycan structures. Taken from [7] with permission from Cold Spring Harbor Press. 13 Mucins are produced in goblet cells which synthesize, package, and secrete the polymers into the environment. For a more detailed description, I refer the reader to a number of comprehensive reviews [6,8,9]. Briefly, the protein portion of mucins is comprised of several characteristic domains. The central region contains highly repetitive proline serine threonine (PTS) domains that are the site of O-glycosylation (Fig. 1A). The N-terminus of cell-surface mucins contain a transmembrane domain that anchors them to the cell surface. In secreted gel forming mucins, the C- terminus contains cysteine knot domains and the N or both N and C termini contain von Willebrand D (VWD) domains, both of which contribute to mucin oligomerization [10,11]. Secreted mucins are O-glycosylated in the golgi apparatus [12]. O-glycosylation begins with the addition of an N-acetylgalactosamine residue to a threonine or serine hydroxyl group on the mucin peptide backbone. Then, galactose and or N-acetylglucosamine residues are added to form one of four main core structures (Fig. 1B). The core structures can then be further elongated with sugars including galactose, fucose, and sialic acid, to form glycan chains of varying complexities. Sialic acids and sulfates are common terminal moieties of mucin glycans, resulting in a strong negative charge. The composition of mucin glycans is dependent on available glycosyltransferases which vary throughout the body, thus mucins in different body locations exhibit different glycosylation patterns. Next, the mucins are either transported to the membrane or packaged into vesicles for secretion. Once released into the lumen, the mucins quickly absorb large amounts of water and mix with other components found in the environment to form mature mucus. Mucus protects the body from microbial infections The mucus serves many vital purposes for the body. In the stomach, it modulates pH by controlling proton transport [13]. In the lungs, the mucus captures particles and microbes from the environment and is swept away through the coordinated beating of cilia [14]. In the vagina, it varies in thickness throughout the menstrual cycle to regulate the passage of sperm and prevent the ascension of bacteria into the uterus [15]. In the intestines, mucus acts as a selective barrier to allow nutrients to reach the epithelium while preventing microbes from accessing the tissue [12]. While mucus serves specialized roles in each of these regions, it has one common function in all locations: to protect the underlying epithelia from microbial infection. 14 Microbes are in constant contact with the body. Healthy humans teem with microbes that inhabit the skin and the mucus. The gut is the most highly populated mucosal region, where bacterial densities can reach 1011 bacterial cells per gram of feces [16]. Many of these microbes are commensals that are indispensible for human health. For example, the microbiota is partly responsible for digestion of food; nutrients that are released as a result of microbial digestion pass through the mucus and are absorbed by the intestinal epithelium. While pathogens are an obvious danger to the mucosa, even commensal bacteria can cause disease if they colonize host cells[17]. Yet despite the high microbial load supported by the mucus, the mucosal epithelium remains uninfected in healthy people. Mucin interactions with microbes The mechanisms of mucus protection against infections are not well understood. One hypothesis is that mucus acts as a physical barrier to microbes [12]. For example, high viscosity mucus in the cervix and in the intestines has been shown to decrease the motility of certain bacteria [18]. Mucins play a central role in conferring the physical properties of mucus and are therefore important determinants of the efficacy of the barrier. Changes in the expression or glycosylation of mucins can have dramatic effects on the properties of mucus [19]. Additionally, the dense glycans that protrude from mucins often directly interact with microbes, which can have multiple sugar binding proteins expressed on the cell surface [20]. Mucins can provide a structural framework for antimicrobial molecules and substances found within the mucus. They have been shown to bind antimicrobial peptides and may therefore present them to microbes in high local concentrations. For example, the non-gel-forming secreted mucin MUC7 binds to the salivary antimicrobial peptide histatin 1, as does MUC5B which also binds histatins 3 and 5 [21]. In addition to binding these molecules, mucins themselves can act as antimicrobials. Terminal α1,4-linked N-acetylglucosamine found on gastric mucins have antimicrobial affects against Helicobacter pylori via the inhibition of cell wall synthesis [22]. Additionally, MUC7 contains an N-terminal domain with sequence homology to histatin-5, which demonstrates antimicrobial activity against the fungal pathogen Candida albicans [23]. Mucins can also bind directly to microbes. Many bacteria are known to associate with mucins via mucus binding proteins and lectins [24]. For example, the probiotic bacterium 15 Lactobacillus rhamnosus possesses pili that are coated with mucus binding domains [25]. It is hypothesized that these pili immobilize these bacteria inside the mucus layer, separating the microbiota from host cells while allowing them to persist as a commensals [26]. Another possibility is that microbial mucus-binding proteins recognize specific mucin glycans, causing selective microbial colonization in parts of the body that secrete mucins compatible with specific microbial binding proteins. Another potential function of microbial binding to mucins is to serve as decoys for microbial binding sites to epithelial cells. The microbiota is rich with adhesins and glycosidases that can adhere to and degrade host-associated glycans, including mucins [27,28]. Since cellsurface mucins can be liberated from the epithelial layer, one hypothesis is that they release upon binding to microbes, thereby preventing association between host and microbial cells [4]. Additionally, the cell-surface mucin MUC1 reports the attachment of the bacterium P. aeruginosa to the body via signal transduction, indicating that these mucins behave as signaling molecules in addition to decoys [29]. Consequences of dysfunctional mucins The importance of mucus in protecting against microbial infections is demonstrated in disease states and mucus models in which the mucus layer is disrupted. Mouse models lacking certain mucins display significant inflammation of the mucosal epithelium. For example, studies using Muc2 deletion mice show a colitis-like phenotype (Fig. 2A), with inflamed epitheia, bloody diarrhea and death [30]. In Muc2-/- mice, the commensal bacteria directly contact the epithelium (Fig. 2B), causing inflammation and cancer development [31]. Additionally, mucus barrier quality, as measured by penetration by fluorescent beads, is poorer in mouse and human colitis samples (Fig. 2C&D). However, mucus layer thickness in these samples is not correlated with permeability, indicating that mucus quality and not quantity is an important characteristic in inflammatory bowel disease [32]. One explanation for the decrease in mucus barrier effects in these experiments is that the generation of mucins must be accomplished very quickly during periods of inflammation, resulting in insufficient concentrations of mucins or mucins with 16 Figure 2 Mucins protect the epithelium from colitis and bacterial contact. A) Muc2 knockout mice show colitis-like symptoms shortly after birth. Taken from [30] with permission from Elsevier. B) Mouse colon sections of WT and Muc2-/- mice. Epithelium stained blue and bacteria stained red. Scale bar: 100 µm. Taken from [31] Copyright (2008) National Academy of Sciences, USA. C) Control and colitis human colonic biopsy samples inoculated with bacteriasized fluorescent beads on top of the mucus[32]. Mayo 0 represents colitis patients in remission and Mayo 1-3 represents those with active disease. Scale bar: 100 µm. D) Quantification of bead penetration through biopsy samples in C) [32]. “Close to epithelium” is <120µm. C&D) Reproduced from [32] with permission from BMJ Publishing Group Ltd. 17 reduced or abnormal glycosylation patterns that are less effective in preventing disease than those from healthy people [32]. Mucin glycosylation also plays an indispensible role in conferring the protective properties of mucus, as evidenced by human disease pathology and studies of glyco-deficient mouse models. One example illustrating the importance of mucin glycans in human disease is ulcerative colitis (UC), an inflammatory bowel disease which is linked to altered interactions between the immune system and the microbiota [33]. Patients with UC have distinct MUC2 glycosylation patterns compared to mucins from healthy individuals [34]. Specifically, UC patients have a higher abundance of short glycan sequences and fewer long, complex glycans. Interestingly, non-diseased control groups and UC remission groups show similar glycosylation patterns, indicating that modified glycans in UC patients reflect their disease state. Abberant mucin glycosylations in mouse models demonstrate increased disease, and is reviewed extensively in [35]. Briefly, the effects of the loss of four types of core glycans found on the intestinal mucin MUC2, namely core-1, 2, 3 & 4 (Fig. 1B), were analyzed in different studies using glycosyltransferase deficient mice. Mouse models lacking core O-glycans demonstrate spontaneous development of colitis [36], increased bacterial penetration of the mucus [32] and increased intestinal permeability and susceptibility to colitis-causing agents [37, 38]. Enhanced susceptibility to disease in the aforementioned glyco-deficient mice may be a consequence of increased microbial degradation of mucins. Studies of bacterial protease activities against mucins demonstrate that certain mucin glycans, such as ppGalNAc-T3mediated O-glycosylation, prevent proteolytic digestion of mucins by bacterial proteases [39,40]. Additionally, mucin glycan sulfation and sialylation are hypothesized to decrease or inhibit the activity of bacterial glycosidases [41, 42]. Therefore, the role of glycans in regulating the barrier effect of mucus is at least partially due to protection of mucin structure by resisting bacterial degradation. Microbial strategies for overcoming the mucus barrier Certain pathogens have evolved strategies to subvert the mucus barrier to cause infection. These strategies include physical penetration via pH modulation, enzymatic degradation and avoiding 18 Figure 3 Mechanisms utilized by pathogenic microbes to overcome the mucus layer. Helicobacter pylori can swim through the normally viscous stomach mucus by increasing its local pH which reduces mucus viscosity and allows the bacteria to swim through. Other pathogens, such as Vibrio cholera and Pseudomonas aeruginosa, secrete enzymes that degrade mucins. The mucus can be avoided entirely by exploiting M Cells, which have little to no mucus covering. Adapted by permission from Macmillan Publishers Ltd: Nature Reviews Microbiology [12], copyright (2011) the mucus by invading M cells, which lack a mucus coating (Fig. 3, [12]). The canonical example of mucus penetration by a pathogen is H. pylori, which swims through the thick, viscous mucus of the stomach to infect the epithelium below. This bacterium increases its local pH, causing the surrounding mucus microenvironment to become pervious through reduction in viscosity [43]. In addition to raising the local pH, bacteria can degrade mucins to make it more easily penetrable, as is seen with Vibrio cholera, Yersinia enterocolitica, and Pseudomonas aeruginosa [39–42]. To avoid the mucus altogether, pathogens can take advantage of Microfold (M) cells, immune cells that sample the mucosal environment. M cells are found in Peyers 19 Figure 4 Models for studying the interactions between microbes and mucus. patches in the epithelium, which are regions with little or no mucus that allow for the M cells to obtain immunogenic materials from the intestinal lumen [48]. These materials are then delivered to lymphoid tissues, which generate the appropriate immune response. Certain pathogens, such as Salmonella and Shigella, exploit the lack of mucus on M cells to enter the tissue and cause infections [49,50]. Models for studying mucus-microbe interactions A number of model systems for studying mucus are available and vary in their complexity (Fig. 4). The most intricate include mice with mucus irregularities [5] such as mucin deletions [30, 51, 52], mucin overexpression [53], and defective mucin glycosylation patterns [36–38]. These models are particularly useful in the context of microbe-mucus interaction studies since tissues can be excised and analyzed to determine the effects of abnormal mucus on the microbial flora. However, animal models are difficult to maintain and are often insufficient proxies in the context of microbial disease since many microbes vary in their host specificity. To elude the difficulty in interpreting animal model data, in vitro and ex vivo techniques are available, such as mucussecreting tissues and cell lines. Sections of mucosal tissues can be excised from humans and animals and infected with microbes [54–57]. Additionally, cell lines derived from the intestines and lungs that are capable of secreting mucus are available [58, 59]. These cell lines allow for a 20 controlled environment in which secreted mucus can be harvested for further study or inoculated with microbes directly. One drawback of these models is that cell culture medium may not allow for the synthesis of physiologically relevant mucus. Additionally, these cells often have to be coaxed into mass-producing mucus, which may have reduced barrier effects or altered characteristics. An alternative model that is particularly useful in human studies is that of whole mucus. Examples whole human mucus experiments are particularly prevalent in the context of cystic fibrosis, where patients expel copious sputum that can be collected, and saliva which is very easily harvested from healthy and diseased people alike [60–63]. Unfortunately, it is comparatively more difficult to collect mucus from most other body locations, such as the intestines and the cervix, which require invasive techniques to obtain. Another drawback of using whole mucus is that there are many components that comprise mucus, making it difficult to ascertain the specific mechanisms of microbial interactions. Therefore, an even more simplified model can be used to study mucus-microbe interactions: purified mucins. To grasp the basics of microbial interactions with mucus, purified native mucins are a simplified, yet highly informative in vitro system. The glycan structures that coat mucins are responsible for the physical properties of mucus and provide many potential binding sites for microbes. Due to the close interactions between microbes and mucins, there is massive potential to discover strategies employed by these polymers to protect the body against infections. The purification of mucins enables the in vitro study of mucin-related phenomena without the confounding factors of whole mucus. The structure of mucins is crucial for carrying out physiologically relevant experiments, and high-quality purification of mucins that retains post-translational modifications as well as supramolecular structure is paramount. In fact, many industrially purified mucins do not form gels as native mucins do, most likely due to the loss of structural integrity [63, 64] and are therefore not sufficient comparisons to native mucins. Additionally, the structures of gels formed by natively purified mucins and industrially purified mucins vary dramatically (Fig. 5), further highlighting the differences between the two types of molecules. In this thesis, I use purified native porcine gastric, porcine intestinal, and human salivary mucins to study microbial interactions with mucin. When human mucins were not available, such as in in the case of gastric and intestinal mucins, porcine mucins were chosen as a proxy to 21 Figure 5 SEM micrographs of natively purified and industrially purified mucins. Natively purified mucins show the characteristic fine, heterogeneous mesh structure of mucus, whereas the industrially purified gel does not. human mucins due to their high sequence homology [66]. The mucins purified using this method are structurally and functionally superior to industrially purified mucins because they retain the important physical and chemical properties that give native mucus its unique characteristics. Thus, we have the ability to study microbial interactions with mucins in a controlled, highly reproducible, and physiologically relevant manner. Although purified native mucins closely reflect the physical properties of mucus, there are a number of limitations to using these mucins in vitro. Firstly, mucus is made up of numerous components besides mucins, including other proteins, salts, lipids, antimicrobial peptides and commensal bacteria. While mucins are largely responsible for the gel forming properties of mucus, the other components are likely to play a role as well, and may affect microbial behavior. Second, the mucus in the body is constantly being replenished, which is not reflected in many of the experiments described in this thesis. Despite the caveats presented here, the use of purified native mucins is highly informative due to their relative simplicity when compared to other mucus models, their major role in defining the physical properties of mucus, and their high potential for interacting with microbes. Simplified 22 experiments combining microbes and mucins are an important first step toward enhancing the understanding of experiments using more complex conditions. Hypothesis: Mucins protect the body and influence the composition of the microbiota by suppressing microbial virulence The examples of microbial interactions with mucus that I have laid out in this introduction are physical in nature: microbes existing in, binding to, penetrating and degrading mucus. While these interactions are important in understanding the relationship between the microbiota and the mucus barrier, these mechanisms do not explain how the mucus prevents these microbes from causing infection. This thesis aims to explore the influence of mucins on microbial physiology to understand the mechanisms behind the mucin-mediated protection from microbial disease. Mucins are emerging as important regulators of microbial behavior that can influence the expression of virulence traits of microbes. For example, the human cell-surface mucin MUC1 can inhibit surface adhesion of the gastric bacterium Heliobacter pylori [67]. Moreover, the intestinal pathogen Campylobacter jejuni, influences virulence gene expression in the presence of human MUC2 [68]. Other examples include modulation of HIV-1 [69] and influenza [70] infectivity by mucins. However, in-depth analyses of the effects of mucins on virulence processes and microbial interactions are lacking. This thesis aims to increase understanding of the influences of mucins on microbes by performing in depth analyses of bacterial and fungal physiology in the presence of natively purified mucins. Evidence from the literature demonstrates extensive physical interactions between microbes and mucins. Therefore, I propose that mucins have direct impacts on the expression of microbial virulence traits, which functions in tandem with the barrier properties of mucus to reduce the propensity of pathogens to cause disease (Fig. 6). Since the human body contains an enormous pool of potential candidate microbes to study, I began by selecting two well-known model organisms, the bacterium Pseudomonas aeruginosa and the fungus Candida albicans, to observe their expression of virulence traits in the presence of mucins. These two organisms were chosen because both P. aeruginosa and C. albicans can asymptomatically reside in healthy people [71–76] but are formidable pathogens in the context of disease. Specifically, P. aeruginosa forms robust biofilms that are implicated in 23 Figure 6 Model of mucus protection against pathogenic microbes. (1) The canonical view of mucus is that it functions as a physical barrier to infection. (2) I propose a second protective mechanism by which the mucus influences the pathogenic state of microbes: coaxing them toward commensalism, as opposed to pathogenicity, via the regulation of virulence gene and trait expression. wound, eye, and ear infections and perhaps is most well known for infecting the lung mucus of cystic fibrosis patients. In Appendix B, I show that plant essential oils are more successful in killing P. aeruginosa biofilms than classic antibiotics. However, it is difficult to deliver essential oils into the body and there are potential biocompatibility issues, making mucins a more desirable candidate to study in the context of mucosal surfaces. C. albicans also forms biofilms, as well as superficial mucosal infections, such as oral thrush and vaginal yeast infections, and systemic disease. Another reason to study these microbes is that they are typically found near mucosal surfaces, making them prime candidates for studying in the context of mucins. Finally, these two organisms have well documented interactions with each other [77–79] and are therefore logical to study both separately and together inside mucins. Understanding the influence of mucins on the virulence traits and interactions of these microbes may provide insight into how they remain in the body without causing disease and may uncover strategies to prevent them from emerging as pathogens. In the following chapters, I will introduce these pathogens in more depth and then describe the effects of mucins on them when grown individually and in coculture. 24 In Chapter 2, I will detail my contributions to a project studying the influences of mucins on P. aeruginosa biofilm formation [80]. Together with my colleagues, I discovered that mucins inhibit surface biofilm formation of this pathogen by maintaining cells in a motile, dispersed state. However, P. aeruginosa can overcome mucin-mediated biofilm inhibition by forming suspended biofilm-like flocs that resemble those found in the lungs of cystic fibrosis patients. These flocs are formed not by the wild type, but by nonmotile flagellar mutants, a trait that is common to cystic fibrosis clinical isolates. Indeed, clinical isolates form flocs in our experiments, even those that are motile, indicating that other factors besides motility play a role in floc formation. In Chapter 3, I will detail work aimed at elucidating the influence of native mucins on C. albicans virulence gene and trait expression [81]. Specifically, I monitored hyphal formation, surface attachment and biofilm formation in the presence of mucins and found that mucins suppress them all. Additionally, mucins suppress the expression of a number of virulence-associated genes, indicating that mucins directly influence microbial physiology. In Appendix A, the two organisms are combined together to determine the effects of mucins on microbial interactions. Typically, P. aeruginosa is virulent toward C. albicans, in part by forming biofilms on fungal hyphae. Since mucins suppress both P. aeruginosa biofilm formation and C. albicans hyphal formation, I hypothesized that mucins reduce C. albicans killing by P. aeruginosa. Indeed, mucins extend the viability of C. albicans in coculture, suggesting that mucin-mediated suppression of virulence trait expression influences microbial community dynamics. Overall, the results presented in this thesis suggest that mucins play an important role in protecting the body from opportunistic pathogens by suppressing the expression of virulence traits, including aggregation, surface adhesion and biofilm formation. This effect also impacts microbial community dynamics by suppressing pathogenesis between microbes. I suggest that mucins act as more than a physical barrier to infection by modulating microbial virulence. In addition to increasing the understanding of the role of mucins as protectors of the body, this work demonstrates the potential value of mucins as a strategy for preventing and overcoming microbial biofouling and infections. 25 References 1. Corfield A P. Mucins in the gastrointestinal tract in health and disease. Front Biosci. 2001;6: d1321. doi:10.2741/Corfield 2. Corfield AP, Myerscough N, Gough M, Brockhausen I, Schauer R, Paraskeva C. Glycosylation patterns of mucins in colonic disease. Biochem Soc Trans. 1995;23: 840– 845. 3. Derrien M, van Passel MW, van de Bovenkamp JH, Schipper RG, de Vos WM, Dekker J. Mucin-bacterial interactions in the human oral cavity and digestive tract. Gut Microbes. 2010;1: 254–268. doi:10.4161/gmic.1.4.12778 4. Linden SK, Sutton P, Karlsson NG, Korolik V, McGuckin MA. Mucins in the mucosal barrier to infection. Mucosal Immunol. 2008;1: 183–197. doi:10.1038/mi.2008.5 5. Joshi S, Kumar S, Bafna S, Rachagani S, Wagner K-U, Jain M, et al. Genetically engineered mucin mouse models for inflammation and cancer. Cancer Metastasis Rev. 2015; 1–17. doi:10.1007/s10555-015-9549-1 6. Bansil R, Turner BS. Mucin structure, aggregation, physiological functions and biomedical applications. Curr Opin Colloid Interface Sci. 2006;11: 164–170. doi:10.1016/j.cocis.2005.11.001 7. Brockhausen I, Schachter H, Stanley P. O-GalNAc Glycans. In: Varki A, Cummings RD, Esko JD, Freeze HH, Stanley P, Bertozzi CR, et al., editors. Essentials of Glycobiology. 2nd ed. Cold Spring Harbor (NY): Cold Spring Harbor Laboratory Press; 2009. Available: http://www.ncbi.nlm.nih.gov/books/NBK1896/ 8. Hattrup CL, Gendler SJ. Structure and Function of the Cell Surface (Tethered) Mucins. Annu Rev Physiol. 2008;70: 431–457. doi:10.1146/annurev.physiol.70.113006.100659 9. Thornton DJ, Rousseau K, McGuckin MA. Structure and function of the polymeric mucins in airways mucus. Annu Rev Physiol. 2008;70: 459–486. doi:Doi 10.1146/Annurev.Physiol.70.113006.100702 10. Perez-Vilar J, Hill RL. The Structure and Assembly of Secreted Mucins. J Biol Chem. 1999;274: 31751–31754. doi:10.1074/jbc.274.45.31751 11. Lang T, Hansson GC, Samuelsson T. Gel-forming mucins appeared early in metazoan evolution. Proc Natl Acad Sci U S A. 2007;104: 16209–16214. doi:10.1073/pnas.0705984104 12. McGuckin MA, Lindén SK, Sutton P, Florin TH. Mucin dynamics and enteric pathogens. Nat Rev Microbiol. 2011;9: 265–278. doi:10.1038/nrmicro2538 13. Li L, Lieleg O, Jang S, Ribbeck K, Han J. A microfluidic in vitro system for the quantitative study of the stomach mucus barrier function. Lab Chip. 2012;12: 4071–4079. doi:10.1039/C2LC40161D 26 14. Sleigh MA, Blake JR, Liron N. The Propulsion of Mucus by Cilia. Am Rev Respir Dis. 1988;137: 726–741. doi:10.1164/ajrccm/137.3.726 15. Gipson IK, Moccia R, Spurr-Michaud S, Argüeso P, Gargiulo AR, Hill JA, et al. The Amount of MUC5B Mucin in Cervical Mucus Peaks at Midcycle. J Clin Endocrinol Metab. 2001;86: 594–600. doi:10.1210/jcem.86.2.7174 16. Franks AH, Harmsen HJM, Raangs GC, Jansen GJ, Schut F, Welling GW. Variations of Bacterial Populations in Human Feces Measured by Fluorescent In Situ Hybridization with Group-Specific 16S rRNA-Targeted Oligonucleotide Probes. Appl Environ Microbiol. 1998;64: 3336–3345. 17. Brook I. Microbiology and Management of Abdominal Infections. Dig Dis Sci. 2008;53: 2585–2591. doi:10.1007/s10620-007-0194-6 18. Cone RA. Barrier properties of mucus. Adv Drug Deliv Rev. 2009;61: 75–85. doi:10.1016/j.addr.2008.09.008 19. Lai SK, Wang YY, Wirtz D, Hanes J. Micro- and macrorheology of mucus. Adv Drug Deliv Rev. 2009;61: 86–100. 20. Esko JD, Sharon N. Microbial Lectins: Hemagglutinins, Adhesins, and Toxins. 2009; Available: http://www.ncbi.nlm.nih.gov/books/NBK1907/ 21. Iontcheva I, Oppenheim FG, Troxler RF. Human salivary mucin MG1 selectively forms heterotypic complexes with amylase, proline-rich proteins, statherin, and histatins. J Dent Res. 1997;76: 734–743. 22. Kawakubo M, Ito Y, Okimura Y, Kobayashi M, Sakura K, Kasama S, et al. Natural antibiotic function of a human gastric mucin against Helicobacter pylori infection. Science. 2004;305: 1003–1006. doi:10.1126/science.1099250 23. Gururaja TL, Levine JH, Tran DT, Naganagowda GA, Ramalingam K, Ramasubbu N, et al. Candidacidal activity prompted by N-terminus histatin-like domain of human salivary mucin (MUC7). Biochim Biophys Acta BBA - Protein Struct Mol Enzymol. 1999;1431: 107–119. doi:10.1016/S0167-4838(99)00034-5 24. Huang JY, Lee SM, Mazmanian SK. The human commensal Bacteroides fragilis binds intestinal mucin. Anaerobe. 2011;17: 137–141. doi:10.1016/j.anaerobe.2011.05.017 25. Kankainen M, Paulin L, Tynkkynen S, Ossowski I von, Reunanen J, Partanen P, et al. Comparative genomic analysis of Lactobacillus rhamnosus GG reveals pili containing a human- mucus binding protein. Proc Natl Acad Sci. 2009;106: 17193–17198. doi:10.1073/pnas.0908876106 26. Etzold S, Juge N. Structural insights into bacterial recognition of intestinal mucins. Curr Opin Struct Biol. 2014;28: 23–31. doi:10.1016/j.sbi.2014.07.002 27 27. Hoskins LC, Boulding ET, Gerken TA, Harouny VR, Kriaris MS. Mucin Glycoprotein Degradation by Mucin Oligosaccharide-degrading Strains of Human Faecal Bacteria. Characterisation of Saccharide Cleavage Products and their Potential Role in Nutritional Support of Larger Faecal Bacterial Populations. Microb Ecol Health Dis. 1992;5: 193– 207. doi:10.3109/08910609209141586 28. Corfield AP, Wagner SA, Clamp JR, Kriaris MS, Hoskins LC. Mucin degradation in the human colon: production of sialidase, sialate O-acetylesterase, N-acetylneuraminate lyase, arylesterase, and glycosulfatase activities by strains of fecal bacteria. Infect Immun. 1992;60: 3971–3978. 29. Lillehoj EP, Kim H, Chun EY, Kim KC. Pseudomonas aeruginosa stimulates phosphorylation of the airway epithelial membrane glycoprotein Muc1 and activates MAP kinase. Am J Physiol - Lung Cell Mol Physiol. 2004;287: L809–L815. doi:10.1152/ajplung.00385.2003 30. Van der Sluis M, De Koning BAE, De Bruijn ACJM, Velcich A, Meijerink JPP, Van Goudoever JB, et al. Muc2-Deficient Mice Spontaneously Develop Colitis, Indicating That MUC2 Is Critical for Colonic Protection. Gastroenterology. 2006;131: 117–129. doi:10.1053/j.gastro.2006.04.020 31. Johansson MEV, Phillipson M, Petersson J, Velcich A, Holm L, Hansson GC. The inner of the two Muc2 mucin-dependent mucus layers in colon is devoid of bacteria. Proc Natl Acad Sci U S A. 2008;105: 15064–15069. doi:10.1073/pnas.0803124105 32. Johansson MEV, Gustafsson JK, Holmén-Larsson J, Jabbar KS, Xia L, Xu H, et al. Bacteria penetrate the normally impenetrable inner colon mucus layer in both murine colitis models and patients with ulcerative colitis. Gut. 2013; doi:10.1136/gutjnl-2012-303207 33. Morgan XC, Tickle TL, Sokol H, Gevers D, Devaney KL, Ward DV, et al. Dysfunction of the intestinal microbiome in inflammatory bowel disease and treatment. Genome Biol. 2012;13: R79. doi:10.1186/gb-2012-13-9-r79 34. Larsson JMH, Karlsson H, Crespo JG, Johansson MEV, Eklund L, Sjövall H, et al. Altered O-glycosylation profile of MUC2 mucin occurs in active ulcerative colitis and is associated with increased inflammation. Inflamm Bowel Dis. 2011;17: 2299–2307. doi:10.1002/ibd.21625 35. Bergstrom KSB, Xia L. Mucin-type O-glycans and their roles in intestinal homeostasis. Glycobiology. 2013;23: 1026–1037. doi:10.1093/glycob/cwt045 36. Fu J, Wei B, Wen T, Johansson MEV, Liu X, Bradford E, et al. Loss of intestinal core 1– derived O-glycans causes spontaneous colitis in mice. J Clin Invest. 2011;121: 1657– 1666. doi:10.1172/JCI45538 37. Stone EL, Ismail MN, Lee SH, Luu Y, Ramirez K, Haslam SM, et al. Glycosyltransferase Function in Core 2-Type Protein O Glycosylation. Mol Cell Biol. 2009;29: 3770–3782. doi:10.1128/MCB.00204-09 28 38. An G, Wei B, Xia B, McDaniel JM, Ju T, Cummings RD, et al. Increased susceptibility to colitis and colorectal tumors in mice lacking core 3–derived O-glycans. J Exp Med. 2007;204: 1417–1429. doi:10.1084/jem.20061929 39. Post S van der, Subramani DB, Bäckström M, Johansson MEV, Vester-Christensen MB, Mandel U, et al. Site-specific O-Glycosylation on the MUC2 Mucin Protein Inhibits Cleavage by the Porphyromonas gingivalis Secreted Cysteine Protease (RgpB). J Biol Chem. 2013;288: 14636–14646. doi:10.1074/jbc.M113.459479 40. Subramani DB, Johansson MEV, Dahlén G, Hansson GC. Lactobacillus and Bifidobacterium species do not secrete protease that cleaves the MUC2 mucin which organizes the colon mucus. Benef Microbes. 2010;1: 343–350. doi:10.3920/BM2010.0039 41. Nieuw Amerongen AV, Bolscher JG, Bloemena E, Veerman EC. Sulfomucins in the human body. Biol Chem. 1998;379: 1–18. 42. Tobisawa Y, Imai Y, Fukuda M, Kawashima H. Sulfation of Colonic Mucins by NAcetylglucosamine 6-O-Sulfotransferase-2 and Its Protective Function in Experimental Colitis in Mice. J Biol Chem. 2010;285: 6750–6760. doi:10.1074/jbc.M109.067082 43. Celli JP, Turner BS, Afdhal NH, Keates S, Ghiran I, Kelly CP, et al. Helicobacter pylori moves through mucus by reducing mucin viscoelasticity. Proc Natl Acad Sci. 2009;106: 14321–14326. doi:10.1073/pnas.0903438106 44. Crowther RS, Roomi NW, Fahim RE, Forstner JF. Vibrio cholerae metalloproteinase degrades intestinal mucin and facilitates enterotoxin-induced secretion from rat intestine. Biochim Biophys Acta. 1987;924: 393–402. 45. Mantle M, Rombough C. Growth in and breakdown of purified rabbit small intestinal mucin by Yersinia enterocolitica. Infect Immun. 1993;61: 4131–4138. 46. Slomiany BL, Slomiany A. Mechanism of Helicobacter pylori pathogenesis: focus on mucus. J Clin Gastroenterol. 1992;14 Suppl 1: S114–121. 47. Aristoteli LP, Willcox MDP. Mucin Degradation Mechanisms by Distinct Pseudomonas aeruginosa Isolates In Vitro. Infect Immun. 2003;71: 5565–5575. doi:10.1128/IAI.71.10.5565-5575.2003 48. Kucharzik T, Lügering N, Rautenberg K, Lügering A, Schmidt MA, Stoll R, et al. Role of M cells in intestinal barrier function. Ann N Y Acad Sci. 2000;915: 171–183. 49. Jones BD, Ghori N, Falkow S. Salmonella typhimurium initiates murine infection by penetrating and destroying the specialized epithelial M cells of the Peyer’s patches. J Exp Med. 1994;180: 15–23. doi:10.1084/jem.180.1.15 50. Acute inflammation causes epithelial invasion and mucosal destruction in experimental shigellosis. J Exp Med. 1994;180: 1307–1319. 29 51. Spicer AP, Rowse GJ, Lidner TK, Gendler SJ. Delayed Mammary Tumor Progression in Muc-1 Null Mice. J Biol Chem. 1995;270: 30093–30101. 52. Velcich A, Yang W, Heyer J, Fragale A, Nicholas C, Viani S, et al. Colorectal Cancer in Mice Genetically Deficient in the Mucin Muc2. Science. 2002;295: 1726–1729. doi:10.1126/science.1069094 53. Ehre C, Worthington EN, Liesman RM, Grubb BR, Barbier D, O’Neal WK, et al. Overexpressing mouse model demonstrates the protective role of Muc5ac in the lungs. Proc Natl Acad Sci U S A. 2012;109: 16528–16533. doi:10.1073/pnas.1206552109 54. Southern P, Horbul J, Maher D, Davis DA. C. albicans Colonization of Human Mucosal Surfaces. PLoS ONE. 2008;3: e2067. doi:10.1371/journal.pone.0002067 55. Harriott MM, Lilly EA, Rodriguez TE, Fidel PL, Noverr MC. Candida albicans forms biofilms on the vaginal mucosa. Microbiol-Sgm. 2010; doi:10.1099/mic.0.039354-0 56. Gustafsson JK, Ermund A, Johansson MEV, Schütte A, Hansson GC, Sjövall H. An ex vivo method for studying mucus formation, properties, and thickness in human colonic biopsies and mouse small and large intestinal explants. Am J Physiol Gastrointest Liver Physiol. 2012;302: G430–438. doi:10.1152/ajpgi.00405.2011 57. Gomez-Trevino M, Boureau H, Karjalainen T, Bourlioux P. Clostridium dificile adherence to mucus: Results of an in vivo and ex vivo assay. Microb Ecol Health Dis. 1996;9. Available: http://www.microbecolhealthdis.net/index.php/mehd/article/view/8381 58. Lesuffleur T, Barbat A, Dussaulx E, Zweibaum A. Growth adaptation to methotrexate of HT-29 human colon carcinoma cells is associated with their ability to differentiate into columnar absorptive and mucus-secreting cells. Cancer Res. 1990;50: 6334–6343. 59. Choe MM, Tomei AA, Swartz MA. Physiological 3D tissue model of the airway wall and mucosa. Nat Protoc. 2006;1: 357–362. doi:10.1038/nprot.2006.54 60. Palmer KL, Mashburn LM, Singh PK, Whiteley M. Cystic fibrosis sputum supports growth and cues key aspects of Pseudomonas aeruginosa physiology. J Bacteriol. 2005;187: 5267–5277. doi:10.1128/JB.187.15.5267-5277.2005 61. Singh PK, Schaefer AL, Parsek MR, Moninger TO, Welsh MJ, Greenberg EP. Quorumsensing signals indicate that cystic fibrosis lungs are infected with bacterial biofilms. Nature. 2000;407: 762–764. doi:10.1038/35037627 62. Samaranayake LP, Hughes A, Weetman DA, MacFarlane TW. Growth and acid production of Candida species in human saliva supplemented with glucose. J Oral Pathol Med. 1986;15: 251–254. doi:10.1111/j.1600-0714.1986.tb00617.x 63. Mellersh A, Clark A, Hafiz S. Inhibition of Neisseria gonorrhoeae by normal human saliva. Br J Vener Dis. 1979;55: 20–23. doi:10.1136/sti.55.1.20 30 64. Crater JS, Carrier RL. Barrier Properties of Gastrointestinal Mucus to Nanoparticle Transport. Macromol Biosci. 2010;10: 1473–1483. doi:10.1002/mabi.201000137 65. Kocevar-Nared J, Kristl J, Smid-Korbar J. Comparative rheological investigation of crude gastric mucin and natural gastric mucus. Biomaterials. 1997;18: 677–81. 66. Turner BS, Bhaskar KR, Hadzopoulou-Cladaras M, LaMont JT. Cysteine-rich regions of pig gastric mucin contain von Willebrand factor and cystine knot domains at the carboxyl terminal. Biochim Biophys Acta BBA - Gene Struct Expr. 1999;1447: 77–92. doi:10.1016/S0167-4781(99)00099-8 67. Lindén SK, Sheng YH, Every AL, Miles KM, Skoog EC, Florin THJ, et al. MUC1 Limits Helicobacter pylori Infection both by Steric Hindrance and by Acting as a Releasable Decoy. PLoS Pathog. 2009;5: e1000617. doi:10.1371/journal.ppat.1000617 68. Tu QV, McGuckin MA, Mendz GL. Campylobacter jejuni response to human mucin MUC2: modulation of colonization and pathogenicity determinants. J Med Microbiol. 2008;57: 795–802. doi:10.1099/jmm.0.47752-0 69. Bergey EJ, Cho MI, Blumberg BM, Hammarskjöld ML, Rekosh D, Epstein LG, et al. Interaction of HIV-1 and human salivary mucins. J Acquir Immune Defic Syndr. 1994;7: 995–1002. 70. Couceiro JN, Paulson JC, Baum LG. Influenza virus strains selectively recognize sialyloligosaccharides on human respiratory epithelium; the role of the host cell in selection of hemagglutinin receptor specificity. Virus Res. 1993;29: 155–165. 71. Lister PD, Wolter DJ, Hanson ND. Antibacterial-resistant Pseudomonas aeruginosa: Clinical impact and complex regulation of chromosomally encoded resistance mechanisms. Clin Microbiol Rev. 2009;22: 582–610. doi:10.1128/CMR.00040-09 72. Kerckhoffs APM, Ben-Amor K, Samsom M, Rest ME van der, Vogel J de, Knol J, et al. Molecular analysis of faecal and duodenal samples reveals significantly higher prevalence and numbers of Pseudomonas aeruginosa in irritable bowel syndrome. J Med Microbiol. 2011;60: 236–245. doi:10.1099/jmm.0.022848-0 73. Estepa V, Rojo-Bezares B, Torres C, Sáenz Y. Faecal carriage of Pseudomonas aeruginosa in healthy humans: antimicrobial susceptibility and global genetic lineages. FEMS Microbiol Ecol. 2014;89: 15–19. 74. Drell T, Lillsaar T, Tummeleht L, Simm J, Aaspõllu A, Väin E, et al. Characterization of the maginal micro- and mycobiome in asymptomatic reproductive-age Estonian women. PLoS ONE. 2013;8: e54379. doi:10.1371/journal.pone.0054379 75. Ghannoum MA, Jurevic RJ, Mukherjee PK, Cui F, Sikaroodi M, Naqvi A, et al. Characterization of the oral fungal microbiome (mycobiome) in healthy individuals. PLoS Pathog. 2010;6: e1000713. doi:10.1371/journal.ppat.1000713 31 76. Scanlan PD, Marchesi JR. Micro-eukaryotic diversity of the human distal gut microbiota: qualitative assessment using culture-dependent and -independent analysis of faeces. ISME J. 2008;2: 1183–1193. doi:10.1038/ismej.2008.76 77. Hogan DA, Kolter R. Pseudomonas-Candida interactions: An ecological role for virulence factors. Science. 2002;296: 2229–2232. doi:10.1126/science.1070784 78. Brand A, Barnes JD, Mackenzie KS, Odds FC, Gow NA. Cell wall glycans and soluble factors determine the interactions between the hyphae of Candida albicans and Pseudomonas aeruginosa. Fems Microbiol Lett. 2008;287: 48–55. doi:10.1111/j.15746968.2008.01301.x 79. Trejo-Hernández A, Andrade-Domínguez A, Hernández M, Encarnación S. Interspecies competition triggers virulence and mutability in Candida albicans–Pseudomonas aeruginosa mixed biofilms. ISME J. 2014; doi:10.1038/ismej.2014.53 80. Caldara M, Friedlander RS, Kavanaugh NL, Aizenberg J, Foster KR, Ribbeck K. Mucin Biopolymers Prevent Bacterial Aggregation by Retaining Cells in the Free-Swimming State. Curr Biol CB. 2012; doi:10.1016/j.cub.2012.10.028 81. Kavanaugh NL, Zhang AQ, Nobile CJ, Johnson AD, Ribbeck K. Mucins Suppress Virulence Traits of Candida albicans. mBio. 2014;5: e01911–14. doi:10.1128/mBio.01911-14 32 Chapter 2 Mucin biopolymers prevent bacterial aggregation by retaining cells in the free-swimming state Parts of the work in this chapter were published in: Caldara M, Friedlander RS, Kavanaugh NL, Aizenberg J, Foster KR, Ribbeck K. 2012. Mucin Biopolymers Prevent Bacterial Aggregation by Retaining Cells in the FreeSwimming State. Curr. Biol. Reproduced with permission from Elsevier 33 Abstract Many species of bacteria form surface-attached communities known as biofilms. Surrounded in secreted polymers, these aggregates are difficult to prevent and eradicate, posing problems for medicine and industry [1, 2]. Humans play host to hundreds of trillions of microbes that live adjacent to our epithelia and are typically able to prevent harmful colonization. Mucus, the hydrogel overlying all wet epithelia in the body, can prevent bacterial contact with the underlying tissue. The digestive tract, for example, is lined by a firmly adherent mucus layer that is typically devoid of bacteria, followed by a second, loosely adherent layer that contains numerous bacteria [3]. Here, we investigate the role of mucus as a principle arena for hostmicrobe interactions. Using defined in vitro assays, we found that mucin biopolymers, the main functional constituents of mucus, promote the motility of planktonic bacteria, and prevent their adhesion to underlying surfaces. The deletion of motility genes, however, allows Pseudomonas aeruginosa to overcome the dispersive effects of mucus and form suspended antibiotic-resistant flocs, which mirror the immotile natural isolates found in the cystic fibrosis lung mucus [4,5]. Mucus may offer new strategies to target bacterial virulence, such as the design of anti-biofilm coatings for implants. Introduction The formation of dense aggregates of cells is key to many bacterial phenotypes, including those involved in virulence and antibiotic resistance. These communities provide robustness and protection from environmental threats such as phagocytosis and antimicrobial substances [6, 7]. The intensive study of surface-attached aggregates, or biofilms, has revealed the roles of key genes involved in motility, attachment and extracellular polymer secretion [8]. In addition to the well-studied surface-bound biofilms, many species can also form dense aggregates in suspension, termed microcolonies or flocs, under certain conditions. The respiratory tract of cystic fibrosis (CF) patients is a focus for research because the thick static mucus found in the CF lung leads to chronic colonization by Pseudomonas aeruginosa, which is a leading cause of morbidity and mortality in CF patients [9, 10]. Successful colonization by P. aeruginosa results in the formation of flocs (>100µm in diameter) embedded in the thick mucus layer [11]. In this environment, the bacteria undergo extensive adaptation and produce virulence factors that altogether assist bacterial infection and inhibit the host immune defense [12–16]. 34 Despite the importance of mucus colonization for bacterial phenotypes and virulence [11, 17], little is understood regarding the mechanisms that enable efficient growth in mucus, their native environment. Several observations argue that we cannot generalize from typical in vitro biofilm studies to the mucosa. For example, clinical P. aeruginosa isolates often overexpress alginate, an extracellular matrix polysaccharide that is not required for surface colonization [18]. Additionally, P. aeruginosa frequently switches to the non-motile state during adaptation in mucus, while motility appears to promote surface attachment during biofilm formation in vitro [19]. Key to the structural and rheological properties of mucus are mucins—large, densely glycosylated polymers. We hypothesize that when arranged as a 3D network, pure mucin polymers can affect P. aeruginosa colonization in the mucosa. Here, we analyze the effects of mucins on colonization by P. aeruginosa with a simplified mucus model where purified native mucins are presented in solution. Until now, our understanding of the effects of mucins on P. aeruginosa has been limited by two key methodological points. First, a direct test of the effects of mucins on P. aeruginosa aggregation has focused on surface attachment, which is not required for colony formation. Second, mucins are typically purified using procedures that degrade and denature the mucin polymers. As we demonstrate below, this results in critical differences from the native mucus of the lung, where mucin polymers entangle to form a flexible and highly-organized 3D network. With our system we show that mucins can effectively suppress bacterial colonization, both on an immersed solid surface, and in solution. We furthermore show that bacteria can overcome mucin inhibition and form large flocs upon the loss of flagellar motility. Last, we show that non-motile, mucus-suspended flocs have superior antibiotic resistance properties than their motile wild-type counterparts. We conclude that the utilization of a 3D model that includes native mucins is critical to advance our understanding of mucosal colonization. Results Mucins reduce biofilm formation of P. aeruginosa To begin to dissect mucin-bacterial interactions, we developed an in vitro assay that uses defined concentrations of native mucins. As a source of mucins we purified native porcine gastric mucus to obtain an extract composed predominantly of MUC5AC, which is one of the major gel- 35 forming components in the lungs and stomach [20]. The use of natively purified mucins is decisive for the utility of this assay, as commercially available mucins are processed and have lost the ability to form viscoelastic hydrogels, as are generated by the native polymers [21, 22]. The second critical feature for this assay is the presentation of mucins in solution, as they exist in the secreted lung mucus, instead of depositing them onto a surface. This detail is important as the surface deposition of mucins is likely to adsorb functional groups, thereby partially dehydrating and altering the biochemical activity of the polymer. First, we tested the effect of mucins on the ability of bacteria to form biofilms on an immersed surface. A plastic microcentrifuge tube was inoculated with culture medium that contained physiological concentrations of mucins [23]. Using the motile, opportunistic pathogen Pseudomonas aeruginosa, we quantified firm attachment by placing exponential-phase cells into the mucin solution, allowing biofilm formation to proceed, and quantifying the biofilm and planktonic populations using the metabolic stain MTT. At 6 h, a time at which biofilms have begun to form, approximately 90% of P. aeruginosa cells remained planktonic in the presence of mucins, compared with 50-60% in tryptone broth (TB) alone or TB plus PEG or dextran (Fig. 1A). The total amount of MTT signal between the different conditions illustrates no major differences in growth, with a small, yet significant, increase in the signal from the mucinexposed population (Fig. 1B). Thus, the results in Fig. 1A are not a result of killing effects by the mucins. This suggests that mucins suppress biofilm formation, and instead promote a planktonic lifestyle. Mucin gels maintain or augment bacterial swimming It is tempting to speculate that bacteria failed to access the underlying surface because they were trapped within the mucin network. If this is true, we should expect to see a measurable decrease of motility within the mucin hydrogel. To test if motion was hindered in the presence of mucins, we tracked the movements of P. aeruginosa cells that carried a deletion in the flagellar hook gene (flgE), and were thus deficient in self-propulsion. These cells demonstrated a significant decrease in diffusivity (p<0.001) in mucin environments, from 2.4 ± 0.2 × 10-9 cm2/s to 1.0 ± 0.1 × 10-9 cm2/s (n ≥ 96 cells), reflecting a higher apparent viscosity of mucin-containing gels, and suggesting that geometric hindrance was present. However, the wild-type cells remained highly motile in the presence of the mucins. The distribution of velocities of swimming cells in mucins 36 Figure 1 Mucins reduce bacterial biofilm formation. PAO1 wild-type bacteria were grown in polypropylene tubes containing TB with or without 1% (w/v) PEG, dextran, or mucin. After 6 h, the amount of planktonic versus biofilm cells was quantified using MTT staining. A) Percentage of the biofilm and planktonic populations relative to the whole. Asterisks represent p-value <0.001. Error bars represent standard deviation of three replicates. B) Total MTT signal generated from the combined adherent and planktonic populations. Asterisk represents p-value <0.05. Error bars represent standard deviation of three replicates. is similar to that in liquid medium, despite the differences in apparent viscosity (Fig. 2). This effect was apparent when we compared cells in Pseudomonas minimal medium (PMM) as well as in tryptone broth (TB) with or without mucins. Immotile P. aeurginosa cells can form suspended flocs in mucin gels If mucins can prevent surface colonization by maintaining cellular motility, we speculated that the loss of cellular motility may be advantageous for colonizing mucus environments. This line of inquiry may have direct physiological relevance, as isolates of P. aeruginosa from cystic fibrosis (CF) mucus are often non-motile [5]. Since immotile P. aeruginosa are known to have reduced surface attachment and biofilm formation, we looked beyond surface adhesion in the presence of mucin and observed the bacteria in the volume of the mucin gel after 20 h of incubation. The wild-type cells remained largely as individual cells or small, suspended colonies of up to 20 µm2 (this corresponds roughly to clusters of 10-20 cells) distributed throughout the 37 Figure 2 P. aeruginosa swimming velocity is unperturbed by mucins. Boxplots depicting swimming velocities of P. aeruginosa in various conditions. Cells were grown in the media indicated at 50%-strength. Velocities were obtained from particle tracking analyses of 20-s swimming videos obtained at 20 frames per second. Data collected by R. Friedlander. volume of the mucin medium (Fig. 3A). However, when observing the nonmotile flagellar mutant PAO1 āflgE, we noticed a striking difference compared to the behavior of wild-type cells. The flagella mutant formed large aggregated flocs of up to 250 µm2 (Fig. 3A). These differences are not likely due to variations in cellular populations in the mucin medium, as PAO1 displayed similar growth rates in the presence and absence of mucins (Fig. 3B). A similar behavior was found for two additional flagella mutants, āflgK, which lack a hook filament junction protein, and āfliD, which lack an adhesive protein at the tip of the flagellar filament but not for āpilB which lack pilus-mediated adhesion and twitching motility (Fig. 3A). The ability of cells to form suspended flocs was inversely correlated with their ability to form surface biofilms in mucin-free environments (Fig. 3C&D). For example, wild-type and āpilB cells formed substantial surface biofilms, but failed to form large suspended flocs in the presence of mucins. Conversely, the various flagellar mutants formed large flocs, but had reduced surface biofilms in the absence of mucins. Complementing the flgE deletion in PAO1 āflgE restored swimming motility and diminished the capacity of the bacteria to form flocs in mucin, indicating that it is indeed the lack of flagella that caused the formation of flocs (Fig. 3E&F). We hypothesized that loss of flagellar motility (rather than other properties of flagella, such as adhesion) was the dominant contributor to the observed aggregation. To test this, we measured mucin-dependent flocculation by a PA14 strain that carries a fully assembled flagellum, but is paralyzed due to deletions in all four stators in the motor complex (ΔmotABΔmotCD). This mutant formed substantially larger flocs (up to 60 µm2) than the wild 38 Figure 3 Nonmotile P. aeruginosa Form Flocs in Mucin Environments. A) Floc formation of PAO1 wild type, flagellar mutants (ΔflgE, ΔflgK, ΔfliD), a pili mutant (ΔpilB), and double flagella and pili mutant (ΔflgEΔpilB) in PMM with 1% mucins after 20 hr of incubation. Scale bar is 20 µm. B) Duplication times of the mutant panel in mucins. C) Box plots quantifying floc size of wild type, flagella, and pili, mutants indicated in µm2 after 20 hr of growth in 1% mucin. D) Surface-attached biofilm formation of the panel of flagellar and pili mutants. Data are presented as percent biofilm formation relative to wild type. Error bars represent standard devation of three replicates. E) Complementation of ΔflgE restores motility and F) suppresses floc formation. Data in (A-D) collected by M. Caldara. 39 Figure 4 The loss of flagellar motility supports floc formation. P. aeruginosa PA14 without a flagellum (ΔflgK) and with an immotile flagellum (ΔmotABCD) were inoculated into a 1% mucin solution in PMM and incubated for 20h before imaging (B-D) and quantification (A). Data collected by M. Caldara. type (Fig. 4), but the structures were smaller than those formed by the ΔflgK strain. Both a loss of motility and loss of the flagella itself, therefore, appear to contribute to mucus colonization. Notably, floc formation did not occur in PEG, dextran or industrially purified mucins but did occur in locust bean gum and methylcellulose, two plant-based carbohydrate polymers (Fig. 5). Therefore, not all viscous polymer solutions support the formation of flocs but it is possible to mimic the effects found in mucins with certain viscous polymer solutions, suggesting that the physical properties of mucins play a key role in floc formation. P. aeruginosa cystic fibrosis clinical isolates form flocs in mucins The loss of flagellar motility allows P. aeruginosa to form flocs in our in vitro mucin gels that resemble those found in the lungs of cystic fibrosis patients and is also a common characteristic of cystic fibrosis clinical isolates. Another common feature of clinical isolates is the overexpression of extracellular polymeric substances, such as the carbohydrates alginate, psl and 40 Figure 5 Selected viscous polymer solutions support the formation of P. aeruginosa flocs. GFP fluorescent P. aeruginosa PAO1 ΔflgE was inoculated into 0.5% viscous polymer solutions as indicated on each panel and imaged. pel. Alginate overexpression, or mucoidy, is particularly common amongst these isolates. We hypothesized that the mucin gels used in our experiments are a comparable environment to the cystic fibrosis lung and that clinical isolates form flocs. To test this hypothesis, we inoculated mucins with four clinical isolates: two that overexpress alginate (FRD1 & 224) and two that overexpress Psl and Pel (19660 & CF127). The cells were visualized fluorescently by introducing PBBR1, a plasmid that constitutively expresses GFP using the pLac promoter. All of the clinical isolates except for 19660 produced large flocs whereas wild-type PAO1 remained largely individual or formed only small clusters (Fig. 6A). This suggests that elevated secretion of extracellular polymers strongly promotes the formation of flocs in mucin-environments. The phenotypes of the suspended colonies resemble those in the cystic fibrosis lung, suggesting that our mucin-system may present a useful laboratory model for dissecting mechanisms of in vivo colonization in complex lung mucus. To validate that the difference in mucus-borne phenotypes is directly due to increased levels of extracellular matrix, and not a consequence of reduced motility we conducted a standard motility plate assay [19]. All of the strains except for ΔflgE and FRD1 were motile (Fig. 6B). CF224 and CF127 displayed motility in the plate assay, yet were able to form large flocs in the presence of mucins. This result indicates immotility is not a requirement for floc formation, but we cannot rule out the possibility that the contact with mucins induces loss of motility. 41 Figure 6 P. aeruginosa cystic fibrosis clinical isolates form flocs in mucins. A) Fluorescently labeled cystic fibrosis isolates were grown in 0.5% mucins in PMM for 20h and imaged to observe floc formation. All strains except 19660 formed flocs. B) Motility assay plates indicate that all but one isolates are motile. P. aeruginosa floc formation is influenced by the alginate Flagellar loss allows bacteria to effectively colonize mucus. Additionally, motile clinical isolates that overexpress biofilm extracellular matrix polymers display floc formation inside mucin gels. However, in addition to extracellular matrix overexpression, these isolates may have other genetic mutations that result in floc formation. Therefore, we tested a panel of isogenic alginate mutants with varying motility states inside mucins to determine the effects of extracellular matrix on floc formation. Alginate was chosen because it plays only a minor role in biofilm formation [24] but is overexpressed in colonies adapted to growth in CF lung mucus [28, 29]. The mutant panel includes an alginate and flagellum deletion mutant strain (ΔflgEΔalgD), an alginate overexpressing strain (PDO300, a PAO1 strain with a mucoid mucA mutation [27]) and an overexpressing, nonmotile strain (PDO300 ΔflgE). The singular ΔalgD mutant was not included in this panel because it does not form flocs, presumably due to wild type levels of motility. We found that the deletion of alginate in the ΔflgEΔalgD strain significantly reduced floc size in both 0.5% and 1% mucins when compared to ΔflgE alone (Fig. 7), suggesting that alginate or its pathway plays a role in floc formation. Interestingly, the motile PDO300 formed small flocs in 0.5% but not 1% mucins. When the ΔflgE mutation was introduced into the 42 Figure 7 P. aeruginosa floc formation in mucins is not dependent on alginate. A) Fluorescence micrographs of an alginate mutant panel, both deletions and overexpressors, grown in 0.5% and 1% mucins. B) Floc size measurement for each mutant in 0.5% and 1% mucins. Averages are for three replicates; Each replicate used five pictures. Asterisks represent p < 0.05. C) CFU counts demonstrating similar growth patterns of the mutants, therefore any differences depicted in (A&B) are not due to growth defects. Error bars represent standard deviation of three replicates. 43 alginate overexpressing strain PDO300, floc formation was restored, although the flocs in 0.5% mucins were significantly smaller than those formed by ΔflgE alone. The restoration of floc formation in the alginate overexpressing mutant upon flagellar deletion suggests that the loss of motility is more important for floc formation than the overexpression of alginate. Perhaps at lower mucin concentrations, P. aeruginosa can overcome the dispersive effects of mucins by expressing alginate, but at higher mucin concentrations the mucins overcome bacterial colonization strategies. The decreased floc size in PDO300 ΔflgE is interesting because many cystic fibrosis isolates overexpress alginate, yet it appears to hinder the formation of colonies as large as those formed by ΔflgE. Perhaps colony size is less important for bacterial survival in the body than increased expression of alginate. Alginate overexpression likely provides enhanced resistance to antibiotics, a phenotype commonly observed in CF pathology [28] and is therefore highly selected for in the CF lung. P. aeruginosa flocs that emerge in mucin gels are antibiotic resistant Last, we asked whether floc formation can provide bacteria with a selective advantage. By analogy with biofilms, we hypothesized that the immotile cellular aggregates that emerge in the presence of mucins also have a higher resistance toward antibiotics. We grew wild-type and non-motile āflgE cells in mucin media for 20 h, and then subjected both strains to 20 µg/mL of two clinically relevant antibiotics that differ in their mode of action (Fig. 8). This experiment revealed two points: first, both wild-type and ΔflgE bacteria were systematically more resistant to colistin in the presence of mucins as compared to liquid culture without mucins. This suggests that the mucins themselves have the capacity to reduce the efficacy of colistin, regardless of whether cells are planktonic (wild type) or form flocs (ΔflgE). Second, it appeared that the floc-forming ΔflgE cells were more resistant to both antibiotics in the mucin medium than the motile wild-type cells. To test for this possibility we determined the percent survival of the bacteria in either condition, by normalizing to the cell numbers in the untreated samples in liquid and mucin. Inside the mucin medium, the non-motile flagella mutants were on average 14 times more resistant to colistin (Fig. 8B) and approximately 6 times more resistant to ofloxacin (Fig. 8C) than wild-type cells, both of which are statistically significant differences. We conclude that the aggregates that emerge upon loss of motility indeed have an increased resistance compared to motile wild-type cells, possibly due to the 44 Figure 8 Flocs grown in mucins are antibiotic resistant. A) PAO1 Wild type and ΔflgE colony counts in medium with and without mucins after exposure to the antibiotics colistin and ofloxacin. B & C) Data from A) depicted as % survival relative to untreated cells. Asterisks represent different p-value thresholds: (**) if p < 0.01 and (***) if p < 0.001. D) Colony counts of a panel of PAO1 alginate mutants grown in mucins and exposed to colistin. E) The data in D) depicted as percent survival relative to untreated cells. Stars represent significant difference (p <0.05) from the wild type. Error bars in all graphs represent SD of three replicates. presence of an altered composition or quantity of extracellular matrix components, or due to a protective effect of increased cell density [29]. To determine the role of alginate on the resistance of the flocs, we challenged a panel of isogenic alginate mutants grown in mucins with 20 µg/mL colistin. This experiment was carried out in 1% mucins, a condition in which only nonmotile strains form flocs (8 D & E). We found that only the wild type was susceptible to colistin. Both nonmotile, floc forming strains tested, ΔflgEΔalgD and PDO300ΔflgE were resistant to colistin. The resistance of ΔflgEΔalgD illustrates that alginate is not not an important component in determining 45 resistance to colistin in the flocs. Interestingly, the motile PDO300 alginate overexpressing strain was resistant despite the absence of floc formation, suggesting that alginate is sufficient but not necessary for conferring protection from colistin. Discussion Here we have found that animals provide a candidate solution to inhibit biofilm formation, namely mucin polymers. Critically, our results demonstrate that mucins can limit bacterial biofilm formation without killing or trapping bacteria, which will help to limit selective pressure for resistance. Indeed, our only evidence for a resistance phenotype comes in the form of non-motile cells, which are likely to be strongly limited in other modes of virulence [5,30]. Our observations of motility and reduced adhesion in mucin media are similar to findings for Campylobacter jejuni in mouse intestinal crypts. In a previous study, extracted epithelial scrapings from C. jejuni-colonized gnotobiotic mice demonstrated a lack of adhesion and unhindered motility within the crypts [31]. Similar to this, a recent study showed that when supplemented in agar plates, mucins appear to increase motility of P. aeruginosa [32]. At first sight these and our findings contrast with reports on surfaceimmobilized mucins, which arrest [33,34] and can cause large aggregate formation of P. aeruginosa cells [35]. However, these findings can be reconciled if one considers that the effects of mucins on motility may depend on their native three dimensional structure and hence biophysical properties such as viscoelasticity and lubricity, which are preserved in native mucus and presumably inside agar gels, but not when adsorbed to a two-dimensional surface [32]. The gel-forming mucin MUC2 has an ordered repeating ring structure [36], and we speculate that also other gel-forming mucins, such as the MUC5AC used in our experiments, display three dimensional features that affect their interactions with bacteria. Indeed, Berg and Turner have observed that certain structured viscous solutions allow increased velocities of motile bacteria by providing a rigid framework for generating propulsive forces [37]. We anticipate that studying mucins in their native three-dimensional form will reveal valuable novel information about bacterial behavior that cannot be captured by collapsed mucin monolayers. Indeed, we found that cystic fibrosis clinical isolates form flocs inside mucin gels that resemble those found in the lungs of cystic fibrosis patients and that floc-forming bacteria are resistant to antibiotics. Therefore, we suggest that using purified 46 mucins allows for the development of informative in vitro assays for studying the microbial behavior in mucus. Experimental Procedures Mucin purification The source for purification of native MUC5AC was pig stomachs, which secrete MUC5AC, homologous to the human glycoprotein [38]. Porcine gastric mucins were purified as described previously, with the omission of the CsCl density gradient centrifugation [39]. Mass spectrometry analysis was used to determine the composition of the mucin preparation as described previously [40]. Briefly, the analysis was performed at the Harvard Microchemistry and Proteomics Analysis Facility by microcapillary reverse-phase HPLC nanoelectrospray tandem mass spectrometry on a Thermo LTQ-Orbitrap mass spectrometer. The spectra were analyzed using the algorithm Sequest [41]. The analysis showed that MUC5AC was the predominant mucin present in our purified extract, which also contained MUC2, MUC5B and MUC6 as well as other proteins including histones, actin and albumin. In addition, its quality was tested by rheology as described in [21,39], which confirmed that the isolated mucins displayed viscoelastic properties similar to native mucus. Strains and growth conditions All strains, plasmids, and their sources are listed below. Pseudomonas aeruginosa PAO1 was the wild type in this study. P. aeruginosa from the PA14 background was used for the motility mutants presented in Fig.4. The following media were used: lysogeny broth (LB), tryptone broth (TB; 10% w/v tryptone), Pseudomonas minimal medium (PMM; 2.5 mM Na-succinate, 1.2 mM MgSO4, 35 mM K2HPO4, 22 mM KH2PO4, 0.8 mM (NH4)2SO4, E. coli minimal medium (M63 salts) supplemented with 0.2% (w/v) glucose and 0.5% (w/v) casamino acids (M63+). Unless specified otherwise, the standard growth medium for P. aeruginosa was 1% mucin (w/v) in PMM. Mucins were dissolved in the medium by gentle shaking overnight at 4°C. Constructions of deletion mutants and GFP-labeled strains Deletions in P. aeruginosa strains were obtained using Splicing Over Extension (SOE)-PCR, as described previously[42], and confirmed by PCR. In addition, the inability of the strains ΔflgE, 47 ΔflgK, ΔfliD to swim and swarm, and the incapacity of ΔpilB to twitch, was tested as described previously[19]. To facilitate microscopy, we expressed GFP constitutively in each strain. GFP expressing strains were created as described previously [43]. Strains from the PA14 background used in Fig. 4 that express GFP were grown in the presence of carbenicillin (250 µg/ml) as described in [44]. Deletion of fliC was confirmed by PCR and motility agar assay. Generation of complementation plasmids A complementation plasmid containing the flgE gene was created using the plasmid pMQ80[45]. Briefly, the GFP gene was removed from the plasmid via restriction enzyme digestion with ecoRI and hindIII. The flgE gene was amplified from P. aeruginosa PAO1 genomic DNA with primers containing ecoRI and hindIII restriction sites upstream and downstream of flgE respectively. After digestion, the flgE gene was inserted into the plasmid using T4 DNA ligase. For control plasmids, pMQ80 was treated with Klenow polymerase and blunt end-ligated without an insert. The plasmids were transformed[46] into E. coli DH5α cells and plated on LB agar containing 30 µg/mL gentamicin selective media. Resulting colonies were inoculated into overnight cultures and the plasmids were extracted using the GenElute™ Plasmid Miniprep Kit (Sigma Aldrich). The plasmids were transformed into PAO1 strains plated on LB agar plates containing 50 µg/mL gentamicin. Successful transformants were again transformed with the plasmid pSMC21[44], a constitutive GFP expressing plasmid, to facilitate observation of floc formation (note: the fluorescent strains used in the rest of the study are gentamicin resistant and could not be used with pMQ80 which relies on gentamicin selection). Transformants were selected on LB agar containing 400 µg/mL carbenicillin and 50 µg/mL gentamicin. Floc formation in mucin was performed as described in the main text, with the addition of antibiotics (carbenicillin for GFP expression and gentamicin for complementation plasmids) and 100 mM arabinose to induce pMQ80 complementation vector expression. Quantification of biofilm formation in mucin gels Freshly growing cells at an OD600 of 0.01 were inoculated in polypropylene PCR tubes and incubated at 37°C in TB or in TB containing 0.5% (w/v) mucins. After 6 h the planktonic cells were removed for quantification, and the adherent cells in the tubes were washed 2 times with PBS to remove non-adherent cells. Planktonic and adherent cells were stained with 5 mg/ml 48 MTT (3-(4,5-Dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide) for 2 h at 37°C, and subsequently destained with 20% sodium dodecyl sulfate in 50% dimethylformamide (adjusted to pH = 4.7) overnight at 37°C. The resulting solutions were quantified using a plate reader (OD595). Particle tracking For measurement of cell velocities, bacteria were grown to exponential phase as described above, stained with Syto9 live cell stain by adding Syto9 1:1000 into the culture, and incubated for 10 minutes at room temperature. The stained cells were diluted 1:10 into a 50% strength solution of growth medium (as indicated in figure) or growth medium supplemented with mucin, dextran or PEG. These solutions were mixed and dispensed into chambers for visualization. Videos of cells were taken on an inverted fluorescent microscope at 20 frames per second to obtain trajectories (see SI for additional details). The trajectories obtained were processed using Matlab to determine velocities and diffusivities. Diffusivities were based upon mean squared displacement values for a range of lag times. Trajectories were also examined visually to ensure accuracy. Floc size measurements To observe cells growing in mucins, cells from exponential growth phase were added to a 0.5% or 1% mucin gel so that the final concentration was 50-100 cells /µl-1. The same protocol was used for observing cells grown in other polymeric solutions, specifically, PEG, dextran, methylcellulose and locust bean gum. The mix was placed in a 96-well glass bottom plate (MatTek) and incubated at 37°C for 20 h. Images were taken immediately after incubation using an Axiovert 200M (Zeiss). Floc sizes were quantified in two different ways, by measuring the size of each floc and by measuring the area of the flocs in each frame of view. In both cases, the software ImageJ [47] was used. We first subtracted the background using the rolling ball algorithm. The image was then thresholded using an iterative procedure based on the isodata algorithm. We defined a group of cells as a floc when it was composed of at least 2-3 cells. Analysis with Minitab 16 (Minitab Inc.) showed that the data were not normally distributed. We therefore plotted the data using box-plots to provide an unbiased overview of the distribution of 49 the floc sizes for each condition. Four pictures from three independent experiments were analyzed for each strain. Relative biofilm formation P. aeruginosa Biofilms were grown statically at 37°C on the air-liquid interface of 96-well plates as described in [48]. The plates were inoculated with tryptone Broth (TB; 1% Tryptone, 0.5% NaCl w/v) containing cells at an optical density at 600 nm (OD600) of 0.0025. After 24 hrs, the biofilms were stained using 1% crystal violet and destained using 33% acetic acid. The absorbance of the resulting solutions was read with a plate reader at 595 nm. Motility Assay M63 motility plates supplemented with 0.2% glucose, 1 mM MgSO4 and 0.5% casamino acids were created as described previously[49]. To induce expression from the complementation plasmid, 100 mM arabinose was added to the plates. Wooden inoculation sticks were dipped into overnight cultures of the strains being tested and used to stab the center of the motility plates. The plates were incubated overnight (16 h) at 30°C. Statistical Analysis Values are reported in the text as value ± SEM. For statistical comparisons between groups with approximately normal distributions, the Student’s two-tailed t-test was used. Error bars in figures are either standard deviation or SEM, as indicated in the legends. Antibiotic treatment To determine the antibiotic resistance of flocs grown in mucin-media, cells were grown in PMM with 1% (w/v) mucin. After 20 h, the number of cells was determined by counting CFU; this number was used as the reference number prior to treatment. The antibiotics ofloxacin and colistin were added to the cultures at final concentrations of 20 µg/ml, and the cultures were grown at 37°C for 3 h. After treatment, the number of survivors was estimated by measuring the CFU. Each experiment was carried out in triplicate. To determine the resistance of cells grown in the absence of mucins, an exponential phase culture was adjusted to contain the same number of cells as had grown in 1% mucin in 20 h, and challenged with antibiotics as described above. 50 Acknowledgements This work was supported by the Cystic Fibrosis Foundation CFF grant number RIBBEC08I0 and MIT startup funds to KR. KRF is supported by European Research Council grant 242670. RSF is supported through the National Science Foundation Graduate Research Fellowship Program. We thank D.J. Wozniak for the EPS deletion strains, B. Berwin for providing the P. aeruginosa PA14 strains, W. Kim for the labeled conjugating strain, G.A. O’Toole for the complementation vector, and the lab of Roberto Kolter for the E. coli strain ZK2686. Table 1 List of strains and plasmids used in this study. Strains and Description plasmids Reference or source E.coli F- endA1 recA1 galE15 galK16 nupG rpsL DH10B ΔlacX74 Φ80lacZΔM15 araD139 Δ(ara,leu)7697 mcrA Δ(mrr-hsdRMS- Invitrogen mcrBC) λKmr, thi-1, thr, leu, tonA, lacY, supE, SM10 λpir recA::RP4-2-Tc::Mu, pir+ pUX-BF13 [43] (Apr- Tn7 helper) S17-1 λpir Tpr Smr recA, thi, pro, hsdR-M+RP4: 2Tc:Mu: Km Tn7 λpir Mark Silby P. aeruginosa PAO1 wild type, clinical isolate [50] ΔflgE PAO1-ΔflgE This study ΔfliD PAO1-ΔfliD This study ΔflgK PAO1-ΔflgK This study ΔpilB PAO1-ΔpilB This study ΔflgE ΔpilB PAO1-ΔflgE ΔpilB This study ΔalgD PAO1-ΔalgD 51 [51] DJ Wozniak ΔalgD ΔflgE PAO1-ΔalgD ΔflgE PDO300 PAO1-mucA22 (Alginate overexpressor) PDO300 ΔflgE PAO1-mucA22 ΔflgE This study PA14 wild type, clinical isolate + pSMC21 [44] PA14- ΔflgK PA14-ΔflgK + pSMC21 [44] PA14- ΔmotAB ΔmotCD + pSMC21 [44] Cystic Fibrosis Clinical Isolate; Alginate M. Franklin overexpressing [52] Cystic Fibrosis Clinical Isolate; Alginate R. Kolter & overexpressing K. Foster Cystic Fibrosis Clinical Isolate; Psl & Pel M. Parsek overexpressing [53] Cystic Fibrosis Clinical Isolate; Psl & Pel M. Parsek overexpressing [53] PA14- ΔmotAB ΔmotCD FRD1 CF224 19660 CF127 This study [27]DJ Wozniak Plasmids pMQ30 pBKminiTn7Gm/Cm-gfp pEX18Gm + CENURA, Gmr, allelic replacement vector Gmr, Cmr, transposon delivery plasmid [54] [43] Apr, Kanr, Carbr, plasmid containing GFP pSMC21 under the control of Ptac constitutive [44] promoter pBBR1_MCS_GFP Constituitive GFP expression from pBBR1(MCS5)-Plac-gfp pMQ80 Complementation plasmid, GmR pMQflgE Complementation plasmid containing flgE 52 [55,56] [45] GA O’Toole This study References 1. Donlan RM. Biofilm formation: a clinically relevant microbiological process. Clin Infect Dis. 2001;33: 1387–1392. doi:10.1086/322972 2. Petrova OE, Sauer K. Sticky situations - Key components that control bacterial surface attachment. J Bacteriol. 2012; doi:10.1128/JB.00003-12 3. Johansson MEV, Phillipson M, Petersson J, Velcich A, Holm L, Hansson GC. The inner of the two Muc2 mucin-dependent mucus layers in colon is devoid of bacteria. Proc Natl Acad Sci USA. 2008;105: 15064–15069. doi:10.1073/pnas.0803124105 4. Singh PK, Schaefer AL, Parsek MR, Moninger TO, Welsh MJ, Greenberg EP. Quorumsensing signals indicate that cystic fibrosis lungs are infected with bacterial biofilms. Nature. 2000;407: 762–764. doi:10.1038/35037627 5. Mahenthiralingam E, Campbell ME, Speert DP. Nonmotility and phagocytic resistance of Pseudomonas aeruginosa isolates from chronically colonized patients with cystic fibrosis. Infect Immun. 1994;62: 596–605. 6. Leid JG, Shirtliff ME, Costerton JW, Stoodley P. Human leukocytes adhere to, penetrate, and respond to Staphylococcus aureus biofilms. Infect Immun. 2002;70: 6339–45. 7. Davies D. Understanding biofilm resistance to antibacterial agents. Nat Rev Drug Discov. 2003;2: 114–22. doi:10.1038/nrd1008 nrd1008 [pii] 8. Jefferson KK. What drives bacteria to produce a biofilm? FEMS Microbiology Letters. 2004;236: 163–173. doi:10.1111/j.1574-6968.2004.tb09643.x 9. Campodonico VL, Gadjeva M, Paradis-Bleau C, Uluer A, Pier GB. Airway epithelial control of Pseudomonas aeruginosa infection in cystic fibrosis. Trends Mol Med. 2008;14: 120–33. doi:S1471-4914(08)00027-0 [pii] 10.1016/j.molmed.2008.01.002 10. Lyczak JB, Cannon CL, Pier GB. Establishment of Pseudomonas aeruginosa infection: lessons from a versatile opportunist. Microbes Infect. 2000;2: 1051–60. doi:S12864579(00)01259-4 [pii] 11. Worlitzsch D, Tarran R, Ulrich M, Schwab U, Cekici A, Meyer KC, et al. Effects of reduced mucus oxygen concentration in airway Pseudomonas infections of cystic fibrosis patients. J Clin Invest. 2002;109: 317–25. doi:10.1172/JCI13870 12. De Vos D, De Chial M, Cochez C, Jansen S, Tummler B, Meyer JM, et al. Study of pyoverdine type and production by Pseudomonas aeruginosa isolated from cystic fibrosis patients: prevalence of type II pyoverdine isolates and accumulation of pyoverdine-negative mutations. Arch Microbiol. 2001;175: 384–8. 53 13. Mena A, Smith EE, Burns JL, Speert DP, Moskowitz SM, Perez JL, et al. Genetic adaptation of Pseudomonas aeruginosa to the airways of cystic fibrosis patients is catalyzed by hypermutation. J Bacteriol. 2008;190: 7910–7. doi:JB.01147-08 [pii] 10.1128/JB.01147-08 14. Oliver A, Canton R, Campo P, Baquero F, Blazquez J. High frequency of hypermutable Pseudomonas aeruginosa in cystic fibrosis lung infection. Science. 2000;288: 1251–4. doi:8507 15. Schaber JA, Carty NL, McDonald NA, Graham ED, Cheluvappa R, Griswold JA, et al. Analysis of quorum sensing-deficient clinical isolates of Pseudomonas aeruginosa. J Med Microbiol. 2004;53: 841–53. 16. Scharfman A, Arora SK, Delmotte P, Van Brussel E, Mazurier J, Ramphal R, et al. Recognition of Lewis x derivatives present on mucins by flagellar components of Pseudomonas aeruginosa. Infect Immun. 2001;69: 5243–8. 17. Boucher RC. Airway surface dehydration in cystic fibrosis: pathogenesis and therapy. Annu Rev Med. 2007;58: 157–70. doi:10.1146/annurev.med.58.071905.105316 18. Govan JR, Deretic V. Microbial pathogenesis in cystic fibrosis: mucoid Pseudomonas aeruginosa and Burkholderia cepacia. Microbiol Rev. 1996;60: 539–574. 19. O’Toole GA, Kolter R. Flagellar and twitching motility are necessary for Pseudomonas aeruginosa biofilm development. Molecular Microbiology. 1998;30: 295–304. doi:10.1046/j.1365-2958.1998.01062.x 20. Schade C, Flemström G, Holm L. Hydrogen ion concentration in the mucus layer on top of acid-stimulated and -inhibited rat gastric mucosa. Gastroenterology. 1994;107: 180– 188. 21. Kocevar-Nared J, Kristl J, Smid-Korbar J. Comparative rheological investigation of crude gastric mucin and natural gastric mucus. Biomaterials. 1997;18: 677–81. 22. Crater JS, Carrier RL. Barrier Properties of Gastrointestinal Mucus to Nanoparticle Transport. Macromolecular Bioscience. 2010;10: 1473–1483. doi:10.1002/mabi.201000137 23. Kirkham S, Sheehan JK, Knight D, Richardson PS, Thornton DJ. Heterogeneity of airways mucus: variations in the amounts and glycoforms of the major oligomeric mucins MUC5AC and MUC5B. Biochem J. 2002;361: 537–546. 24. Wozniak DJ, Wyckoff TJ, Starkey M, Keyser R, Azadi P, O’Toole GA, et al. Alginate is not a significant component of the extracellular polysaccharide matrix of PA14 and PAO1 Pseudomonas aeruginosa biofilms. Proc Natl Acad Sci U S A. 2003;100: 7907– 12. doi:10.1073/pnas.1231792100 1231792100 [pii] 54 25. Hentzer M, Teitzel GM, Balzer GJ, Heydorn A, Molin S, Givskov M, et al. Alginate overproduction affects Pseudomonas aeruginosa biofilm structure and function. J Bacteriol. 2001;183: 5395–401. 26. Stapper AP, Narasimhan G, Ohman DE, Barakat J, Hentzer M, Molin S, et al. Alginate production affects Pseudomonas aeruginosa biofilm development and architecture, but is not essential for biofilm formation. J Med Microbiol. 2004;53: 679–90. 27. Mathee K, Ciofu O, Sternberg C, Lindum PW, Campbell JIA, Jensen P, et al. Mucoid conversion of Pseudomonas aeruginosa by hydrogen peroxide: a mechanism for virulence activation in the cystic fibrosis lung. Microbiology. 1999;145: 1349–1357. doi:10.1099/13500872-145-6-1349 28. Moreau-Marquis S, Stanton BA, O’Toole GA. Pseudomonas aeruginosa biofilm formation in the cystic fibrosis airway. Pul Pharm. 2008;21: 595–599. doi:10.1016/j.pupt.2007.12.001 29. Connell JL, Wessel AK, Parsek MR, Ellington AD, Whiteley M, Shear JB. Probing Prokaryotic Social Behaviors with Bacterial “Lobster Traps.” mBio. 2010;1. doi:10.1128/mBio.00202-10 30. Josenhans C, Suerbaum S. The role of motility as a virulence factor in bacteria. Int J Med Microbiol. 2002;291: 605–614. 31. Lee A, O’Rourke JL, Barrington PJ, Trust TJ. Mucus colonization as a determinant of pathogenicity in intestinal infection by Campylobacter jejuni: a mouse cecal model. Infect Immun. 1986;51: 536–546. 32. Yeung ATY, Parayno A, Hancock REW. Mucin Promotes Rapid Surface Motility in Pseudomonas Aeruginosa. mBio. 2012;3. doi:10.1128/mBio.00073-12 33. Vishwanath S, Ramphal R. Adherence of Pseudomonas aeruginosa to human tracheobronchial mucin. Infect Immun. 1984;45: 197–202. 34. Arora SK, Ritchings BW, Almira EC, Lory S, Ramphal R. The Pseudomonas aeruginosa flagellar cap protein, FliD, is responsible for mucin adhesion. Infect Immun. 1998;66: 1000–7. 35. Landry RM, An D, Hupp JT, Singh PK, Parsek MR. Mucin–Pseudomonas aeruginosa interactions promote biofilm formation and antibiotic resistance. Molecular Microbiology. 2006;59: 142–151. doi:10.1111/j.1365-2958.2005.04941.x 36. Ambort D, Johansson MEV, Gustafsson JK, Nilsson HE, Ermund A, Johansson BR, et al. Calcium and pH-dependent packing and release of the gel-forming MUC2 mucin. Proc Natl Acad Sci USA. 2012;109: 5645–5650. doi:10.1073/pnas.1120269109 37. Berg HC, Turner L. Movement of microorganisms in viscous environments. Nature. 1979;278: 349–351. 55 38. Turner B, Bansil R, Afdhal NH. Expression of cysteine-rich C-terminal domains of Pig Gastric Mucin in Pichia pastoris. FASEB J. 2007;21. 39. Celli J, Gregor B, Turner B, Afdhal NH, Bansil R, Erramilli S. Viscoelastic properties and dynamics of porcine gastric mucin. Biomacromolecules. 2005;6: 1329–33. 40. Lieleg O, Lieleg C, Bloom J, Buck CB, Ribbeck K. Mucin Biopolymers As BroadSpectrum Antiviral Agents. Biomacromolecules. 2012; doi:10.1021/bm3001292 41. Yates JR, Eng JK, McCormack AL, Schieltz D. Method to Correlate Tandem Mass Spectra of Modified Peptides to Amino Acid Sequences in the Protein Database. Anal Chem. 1995;67: 1426–1436. doi:10.1021/ac00104a020 42. Schweizer HP. Allelic exchange in Pseudomonas aeruginosa using novel ColE1-type vectors and a family of cassettes containing a portable oriT and the counter-selectable Bacillus subtilis sacB marker. Mol Microbiol. 1992;6: 1195–204. 43. Lambertsen L, Sternberg C, Molin S. Mini-Tn7 transposons for site-specific tagging of bacteria with fluorescent proteins. Environ Microbiol. 2004;6: 726–32. 44. Amiel E, Lovewell RR, O’Toole GA, Hogan DA, Berwin B. Pseudomonas aeruginosa evasion of phagocytosis is mediated by loss of swimming motility and is independent of flagellum expression. Infect Immun. 2010;78: 2937–45. doi:IAI.00144-10 [pii] 10.1128/IAI.00144-10 45. Shanks RMQ, Caiazza NC, Hinsa SM, Toutain CM, O’Toole GA. Saccharomyces cerevisiae-based molecular tool kit for manipulation of genes from gram-negative bacteria. Appl Environ Microbiol. 2006;72: 5027–5036. doi:10.1128/AEM.00682-06 46. Choi K-H, Kumar A, Schweizer HP. A 10-min method for preparation of highly electrocompetent Pseudomonas aeruginosa cells: application for DNA fragment transfer between chromosomes and plasmid transformation. J Microbiol Methods. 2006;64: 391– 397. doi:10.1016/j.mimet.2005.06.001 47. Abramoff MD, Magelhaes PJ, Ram SJ. Image Processing with ImageJ. Biophotonics International. 2004;11: 36–42. 48. O’Toole GA. Microtiter dish biofilm formation assay. J Vis Exp. 2011; doi:10.3791/2437 49. O’Toole GA, Kolter R. Initiation of biofilm formation in Pseudomonas fluorescens WCS365 proceeds via multiple, convergent signalling pathways: a genetic analysis. Mol Microbiol. 1998;28: 449–61. 50. Holloway BW. Genetic recombination in Pseudomonas aeruginosa. J Gen Microbiol. 1955;13: 572–81. 56 51. Whitchurch CB, Erova TE, Emery JA, Sargent JL, Harris JM, Semmler ABT, et al. Phosphorylation of the Pseudomonas aeruginosa response regulator AlgR Is essential for type IV fimbria-mediated twitching motility. J Bacteriol. 2002;184: 4544–4554. doi:10.1128/JB.184.16.4544-4554.2002 52. Ohman DE, Chakrabarty AM. Genetic mapping of chromosomal determinants for the production of the exopolysaccharide alginate in a Pseudomonas aeruginosa cystic fibrosis isolate. Infect Immun. 1981;33: 142–148. 53. Wolfgang MC, Kulasekara BR, Liang X, Boyd D, Wu K, Yang Q, et al. Conservation of genome content and virulence determinants among clinical and environmental isolates of Pseudomonas aeruginosa. Proc Natl Acad Sci USA. 2003;100: 8484–8489. doi:10.1073/pnas.0832438100 54. Hoang TT, Karkhoff-Schweizer RR, Kutchma AJ, Schweizer HP. A broad-host-range Flp-FRT recombination system for site-specific excision of chromosomally-located DNA sequences: application for isolation of unmarked Pseudomonas aeruginosa mutants. Gene. 1998;212: 77–86. 55. Kovach ME, Elzer PH, Hill DS, Robertson GT, Farris MA, Roop RM 2nd, et al. Four new derivatives of the broad-host-range cloning vector pBBR1MCS, carrying different antibiotic-resistance cassettes. Gene. 1995;166: 175–176. 56. Billings N, Ramirez Millan M, Caldara M, Rusconi R, Tarasova Y, Stocker R, et al. The extracellular matrix component Psl provides fast-acting antibiotic defense in Pseudomonas aeruginosa Biofilms. PLoS Pathog. 2013;9. doi:10.1371/journal.ppat.1003526 57 Chapter 3 Mucins suppress virulence traits of Candida albicans Work in this chapter was published in: Kavanaugh NL, Zhang AQ, Nobile CJ, Johnson AD, Ribbeck K. 2014. Mucins Suppress Virulence Traits of Candida albicans. mBio 5:e01911–14. 58 Abstract Candida albicans is the most prevalent fungal pathogen of humans, causing a variety of diseases ranging from superficial mucosal infections to deep-seated systemic invasions. Mucus, the gel that coats all wet epithelial surfaces, accommodates C. albicans as part of the normal microbiota where C. albicans resides asymptomatically in healthy humans. Through a series of in vitro experiments combined with gene expression analysis, we show that mucin biopolymers, the main gel-forming constituents of mucus, induce a new oval-shaped morphology in C. albicans in which a range of genes related to adhesion, filamentation, and biofilm formation are downregulated. We also show that corresponding traits are suppressed, rendering C. albicans incapable of forming biofilms on a range of different synthetic surfaces and human epithelial cells. Our data suggests that mucins can manipulate C. albicans physiology and we hypothesize that they are key environmental signals for retaining C. albicans in the host-compatible, commensal state. Introduction Candida albicans is an important opportunistic fungal pathogen in humans that can cause superficial infections, such as vaginitis in women or thrush in babies and HIV patients, and systemic, often fatal, disease in more advanced cases [1]. C. albicans possesses a range of virulence traits, including adherence, filamentation, and secretion of proteases [2]. At the heart of many infections is the formation of surface-associated C. albicans communities, also termed biofilms, which can form on mucosal epithelial surfaces and on implanted medical devices, such as catheters and heart valves. Biofilms show increased resistance to both the immune system and to antifungal treatment [3]. Despite the ability of C. albicans to cause disease, the healthy human body accommodates it as part of the microbiota [4–6]. How the body tolerates the continued presence of potentially virulent C. albicans is largely unknown. Mucus is the slimy coating found on all wet epithelia in the human body, including the eyes, airways, and the gastrointestinal and female genitourinary tracts, and many host-microbe interactions take place in this context. Its major gel-forming components, the mucin glycopolymers, are emerging as important regulators of microbial virulence. For example, the human cell-surface mucin MUC1 can inhibit surface adhesion of the gastric bacterium Heliobacter pylori [7]. Moreover, the secreted human mucin MUC5AC can prevent 59 Pseudomonas aeruginosa surface attachment and biofilm formation by promoting a dispersive state of bacteria [8]. Other examples include mucus-mediated clearance of the bacterium Streptococcus pneumoniae [9] and modulation of HIV-1 [10] and influenza [11] infectivity by mucins. These observations indicate that mucin biopolymers help prevent bacterial and viral infections by regulating cellular processes related to virulence. Fungal pathogens are only distantly related to bacteria and viruses, and little is known about their interactions with mucins. Here, we investigate the role of mucins as potential regulators of C. albicans virulence. Using a combination of gene expression analysis, mucus-secreting cell lines and defined in vitro assays with natively purified mucins, we show that exposure to mucins induces a new ovalshaped morphological state in C. albicans, in which various virulence traits are down-regulated, including surface adhesion, the morphological transition to the filamentous state, and biofilm formation. Our results indicate that mucins are key contributors to host defense against C. albicans, and may offer new strategies to target fungal virulence, such as the design of antifungal treatments or coatings for implants. Results Mucins regulate C. albicans physiology To determine the effects of mucins on C. albicans, the strain SC5314 was cultured in RPMI medium with and without pig gastric mucins (MUC5AC); RPMI favors growth in the filamentous state. Mucins were supplemented in the medium to create a 3D environment, as is found in the native mucus barrier [8]. In the absence of mucins, the cells formed extensive hyphae which clumped together into flocs (Fig. 1A). In contrast, mucin-exposed cells predominantly formed short chains resembling pseudohyphae (Fig. 1A) or unicellular, ellipsoidal cells that are distinct from round, yeast-form cells. An analysis of the growth rate of C. albicans in mucins shows that cells continue to grow over time, and even show an enhanced increase in optical density due to the homogenous suspension of individual cells as opposed to the filamentous flocs formed in RPMI medium alone (Fig. 1B&C). Importantly, the effect of MUC5AC on C. albicans physiology is not limited to this type of mucin but was also observed in two other mucins: MUC2 from pig intestinal mucus and MUC5B from human saliva (Fig. 1D). This indicates that mucins have a general effect on C. albicans that likely extends across all mucosal surfaces. Due to the abundant availability of 60 Figure 1: Mucins induce a unique morphological state characterized by suppressed virulence traits A) Phase contrast images of C. albicans in different morphological states. B) Growth curve of C. albicans +/- mucins in YPD at 30°C. In this case, the growth rates are roughly the same. C) Growth curve of C. albicans +/- mucin in RPMI at 37°C. Here, the cells grown in mucins increase in optical density faster than those grown without mucins due to decreased hyphae formation and flocking of cells in mucins. D) Phase contrast images of C. albicans after growth in the presence or absence of the following mucins: pig gastric (MUC5AC), pig intestinal (MUC2) or human salivary (MUC5B). 61 MUC5AC, the remainder of the experiments in this study was performed using this mucin. At first sight the mucin-induced morphology resembles opaque cells, which are the mating competent form of C. albicans [12] that have reduced virulence in systemic infection models [13]. Since the ability of C. albicans to switch to the opaque state is controlled by genes at the Mating-Type Like (MTL) locus, such that only MTLa or MTLα strains can undergo the transition to the opaque form, we first tested if mucins induce our starting MTLa/α heterozygous strains to become homozygous MTLa or MTLα mating-competent cells. Using PCR with primers specific to the MTL locus, we found that mucin-exposed cells remained heterozygous at the MTL locus (n = 100). To further probe their identity, the mucin-exposed cells were assayed for three features that are indicative of opaque cells: the abilities to form opaque colonies on agar plates [14], mate in the presence of the opposite mating type [15,16], and form mating protrusions in the presence of mating pheromone [17]. Our data show that the mucin-exposed cells were incapable of these three traits: they did not form opaque colonies, were not mating competent, and did not form mating protrusions in the presence of mating α-pheromone (Table 1). Moreover, exposure of wor1 Δ/Δ cells (which cannot form opaque cells) to mucins also induced the oval-shaped morphology (Fig. 2), suggesting that the mucin-dependent morphology can develop independently of the master regulator of opaque status. As one further control, we repeated the white-opaque switching, quantitative mating and pheromone response assays using mating competent MTLa or MTLα cells and found that mucins do not affect these opaque processes in mating competent strains (Table 1). Taken together, these results indicate that exposure to mucins suppresses the formation of hyphae while inducing the formation of a novel phenotype that superficially resembles the opaque cell type, which is distinct in its physiological responses. Mucins down-regulate virulence-associated genes in C. albicans To better characterize the mucin-induced morphological change, we carried out transcriptional profiling experiments. A wild-type C. albicans strain (SC5314) was grown for 8 hrs in RPMI at 37°C in the presence and absence of 0.5% mucins. We performed quantitative PCR (qPCR; Fig. 3) to measure the expression levels of selected virulence genes including those associated with adhesion [18], biofilm formation [19], and secreted proteinases [20]. TAF145, a general transcription factor TFIID subunit [21], did not change expression in the presence of mucins and 62 White to Opaque Switch Assay Strain Cell Type Treatment Switching Frequency n SN425 a/α RPMI + mucin 0% 450 SN425 a/α RPMI 0% 520 RBY717 a/a white RPMI + mucin 6.80% 850 RBY717 a/a white RPMI 5.90% 1200 Response to Alpha Factor Strain SN425 SN425 Cell Type Treatment a/α RPMI + mucin + α-factor RPMI + αfactor RPMI + mucin + α-factor RPMI + αfactor RPMI + mucin + α-factor RPMI + αfactor a/α RBY731 a/a opaque RBY731 a/a opaque RBY717 a/a white RBY717 a/a white Cells with Projections n 0% 200 0% 220 33.80% 330 41.80% 280 0.50% 200 1.00% 200 Quantitative Mating Assay Strains Crossed SN87 X SN95 SN87 X SN95 RBY1177 X RBY1180 RBY1177 X RBY1180 Cell Type Treatment a/α X a/α a/α X a/α a/a opaque X α/α opaque a/a opaque X α/α opaque RPMI + mucin RPMI RPMI + mucin Mating Frequency 0% 0% 2.0 % RPMI 7.9% Table 1 Analysis of opaque state attributes upon exposure to mucins. C. albicans that is either homozygous or heterozygous at the mating locus were assayed for their ability to switch from white to opaque, respond to alpha factor and mate after exposure to mucins. 63 Figure 2 The mucin-induced morphological state is distinct from the opaque state. Phase contrast images comparing the WT mucin-induced morphology to wor1Δ/Δ, which is locked in the white phase. was used as a reference for calibration. This experiment shows that 7 of the 16 tested genes (indicated by asterisks, Fig. 3) were down-regulated in the presence of mucins by more than 1.5 fold (p <0.05) as determined by a two-tailed, unpaired t-test comparing the ΔCT values of samples with and without mucins. Mucins suppress C. albicans transition to the filamentous state To understand in more detail the effect of mucins on C. albicans physiology, we monitored the C. albicans strain HGFP3, which expresses GFP from the hyphal-specific HWP1 promoter, during growth in RPMI with or without 0.5% natively-purified mucins. To test whether the 64 Figure 3 qPCR of known C. albicans virulence genes comparing gene expression in RPMI with and without mucins. RNA was extracted from 6 independent biological replicates. Error bars represent SEM. Asterisks represent different p-value thresholds: (*) if p < 0.05, (**) if p < 0.01, and (***) if p < 0.001. effects are simply a response to an increased viscosity or osmotic stress, we also subjected the cells to 0.5% industrially purified mucins (Sigma-Aldrich), 0.5% methylcellulose, or 1M sorbitol. Industrially purified mucins are proteolytically processed and as a result have lost the gel-forming capacity characteristic of native mucins [22, 23]. The polysaccharide methylcellulose is commonly used to mimic the viscosity of a mucus environment [24, 25]. Sorbitol at high concentrations can induce osmotic stress; testing this condition is informative because the osmotic stress pathway is linked to hyphae formation [26]. Our data show that in the “no polymer control”, as well as in the presence of methylcellulose or 1M sorbitol, the majority of cells formed hyphae, as indicated both by the elongated structures and the presence of GFP fluorescence (Fig. 4). In contrast, in the presence of native mucins, hyphae formation was nearly completely suppressed. Industrial mucins also inhibited hyphae formation, although not as effectively as the native mucins; a proportion of cells remained in the hyphal form. These results suggest that specific biochemical attributes present in the mucins suppress hyphal formation. Natively purified mucins are obtained by a relatively mild purification procedure to preserve their structure, and it is formally possible that contaminants contribute to the regulation of the 65 Figure 4 The effects of osmotic stress and viscosity on hyphal formation A) Fluorescent micrographs overlaid onto phase contrast images of C. albicans grown in RPMI with or without 0.5% mucin (with and without CsCl purification), 0.5% methylcellulose, 1M sorbitol (an osmotic stress inducer) or 0.5% PEG. The strain used here, HGFP3 fluoresces only when true hyphae are formed. C. albicans phenotype. To exclude this possibility, we also tested mucins purified with CsCl, a more stringent condition which removes the majority of associated proteins and lipids from the mucins. We observed the same suppression of hyphal formation as in the presence of natively purified mucins (Fig. 4), indicating that this effect is due to the mucins, and not to any associated factors. The ability of mucins to suppress filamentation was also studied at the level of gene expression. Using qPCR we analyzed the expression levels of the hyphal-specific genes ALS3, ECE1, and HWP1 in four media conditions: medium without polymers, and medium with native mucins, industrial mucins, and methylcellulose. For all three genes, native mucins mediated the strongest down-regulation (Fig. 6B). Of note is that this mucin-induced down-regulation of hyphal-specific genes was also observed in an alternative culture medium, YPD + FBS (Fig. 5), indicating that the effects of mucins are not dependent on a specific medium condition. We next tested the effects of mucins on pre-formed hyphae. Hyphae, formed in the absence of mucins, were inoculated into RPMI with and without 0.5% mucins or 0.5% 66 Figure 5 Mucins suppress hyphae formation in YPD + FBS. A) Fluorescent micrographs overlaid onto phase contrast images of C. albicans grown in YPD + FBS, an alternative hyphaeinducing medium to RPMI, with or without 0.5% mucin. B) qPCR comparing the expression of three hyphal-specific genes with and without mucins in YPD. methylcellulose. Our data show that newly formed cells, which bud off from the hyphae, were predominantly in the yeast-form after exposure to mucins (Fig. 6A, right panels). For comparison, hyphae in the absence of mucins continued to produce hyphae. Hyphal cells inoculated into medium containing methylcellulose also continued to produce more hyphal cells. These results were verified by qPCR, which showed that native mucins have the strongest capacity to suppress hyphal formation (Fig. 6C). We note that the levels of expression of the gene HWP1 showed a lower degree of suppression (yet significant; p < 0.001) when hyphae were added as the starting culture in comparison to when yeast were added (-1.3 fold change in expression with hyphae inoculum vs -15.9 fold change for yeast). Thus, in addition to inhibiting the growth of hyphae in yeast-form cells, we can also conclude that the hyphal-suppressive capacity of mucins is strong enough to downregulate the responsible pathway in existing hyphae. 67 Figure 6 Mucins suppress hyphal growth from both yeast and hyphal cells. Fluorescent microscopy images overlaid on phase contrast images (A) and quantitative PCR of hyphalspecific genes expression from yeast (B) or hyphae (C) after incubation for 8 hours in RPMI , 0.5% Methylcellulose, 0.5% native mucins or 0.5% industrially purified mucins. The strain used, HGFP3, expresses GFP only when cells form true hyphae. RNA was extracted from 6 independent biological replicates. Error bars represent SEM. Asterisks represent different pvalue thresholds: (*) if p < 0.05, (**) if p < 0.01, and (***) if p < 0.001. 68 Mucins suppress surface adhesion of C. albicans A critical, early step of infection by C. albicans is its attachment to a solid surface. In Fig. 3 and Fig. 6 we observed that genes involved in cell adhesion (ALS1 and ALS3) were downregulated in the presence of mucins. Moreover, we know that mucins may physically trap certain particles and cells, thereby preventing their association with an underlying surface. Hence, we tested whether media containing gel-forming mucins also decreases the surface attachment of C. albicans. We analyzed the ability of C. albicans to colonize two different surfaces in the presence of mucins, abiotic polystyrene, which is often used in the context of biofilm formation assays, and human epithelial mucus-secreting cells. For adhesion to polystyrene, a suspension of yeast cells was inoculated into polystyrene 96-well plates containing RPMI without or with 0.5% native mucins, industrially purified mucins and methylcellulose. Fig. 7A&B show that native mucins significantly reduced cell attachment to polystyrene. This effect was detectable as early as 30 min and became stronger over the course of the hour (unpaired two-tailed t-test; p <0.05). Methylcellulose was similarly effective in suppressing cell surface attachment (Fig. 7A&B). For comparison, industrially purified mucins provided no significant protection from surface attachment. Native mucins and methylcellulose both increase the viscosity of the medium, hence this parameter could be responsible for the observed anti-adhesion effect. To test this, we subjected the yeast to medium containing 0.5% of polyethylene glycol (PEG), which is often used as an antifouling coating [27]. Our data show that PEG was not able to decrease surface attachment to the same degree as mucins and methylcelluose (Fig. 7C), suggesting that an increased viscosity alone is not sufficient to protect a surface from C. albicans attachment. We next infected human mucus secreting cells with C. albicans to determine the effects of mucus on attachment to a living surface. C. albicans typically resides in the intestine as a commensal, but this environment can also be the starting point for systemic dissemination [28]. A simple in vitro model for the study of C. albicans interactions with mucus-coated epithelial cells of the intestines makes use of HT29-MTX cells derived from human colorectal adenocarcinoma cells, which secrete native-like mucus when growing in culture [29]. The mucus layer can be removed from the secreting cells with N-acetylcysteine, allowing for a comparison of cell adhesion with and without native mucus. To determine the effects of mucus on the attachment of C. albicans to human mucus-secreting epithelial cells, a suspension of C. albicans was exposed to cells that were lined with an intact mucus layer, or to cells from which the mucus 69 Figure 7 Mucins reduce attachment of C. albicans to polystyrene and mucus-secreting human cells. (A) Fluorescent microscopy images of polystyrene 96-well plates after incubation with C. albicans in different media conditions (RPMI alone, RPMI + 0.5% methylcellulose, 0.5% industrially purified mucin or 0.5% native mucin). Time points were taken every 15 minutes after removal of non-adherent cells. (B) Quantification of attachment to the polystyrene plates. Error bars represent standard deviation of 3 replicates. (C) Quantification of attachment to polystyrene in the presence of viscous polymer solutions. (D) Fluorescence and phase contrast microscopy images of C. albicans SC5314 stained with calcofluor white after 2 hours of incubation with HT29-MTX human mucus-secreting cells. (E) Quantification of C. albicans attachment to HT29-MTX cells. Error bars represent standard deviation of 3 replicates. had been removed. After 2 hours the unbound cells were removed by washing the epithelial cells with phosphate buffered saline. Fig. 7D&E show that significantly more C. albicans cells had attached to the unprotected epithelial cells compared to cells that had been shielded with a layer 70 of mucus. This suggests that the mucus lining of epithelial cells provides efficient protection from C. albicans attachment. Mucins decrease C. albicans biofilm formation C. albicans readily forms surface-attached biofilms on both abiotic and biotic surfaces. Because biofilm formation relies on attachment, filamentation and cell-cell interactions [30, 31], we tested if mucins inhibit biofilm formation by C. albicans. We used a standard biofilm assay in which a polystyrene surface is immersed in a yeast culture (in RPMI) to allow attachment of the cells, and after 90 min, washed to remove non-adherent cells. Then the surface is submerged in fresh, cell-free medium to follow the development of the initially attached cells into a biofilm over 8, 24, and 48 h. We performed this assay in the presence of native mucins and for comparison, with methylcellulose and industrial mucins. To evaluate the effect of the individual polymers on C. albicans surface attachment we removed the supernatant after indicated time points and analyzed the resulting biofilms (Fig. 8). Surface attachment was visibly reduced in the presence of native mucins, and even more effectively prevented in the presence of methylcellulose (Fig. 8A). Confocal xy and xz sections of the biofilms show that both surface coverage (xy images) and thickness of the biofilm are reduced (xz scan) in the presence of mucins (Fig. 8C). Moreover, a lower percentage of cells in the mucin-treated biofilms form hyphae within the thin biofilms, confirming the previous observation from Fig. 4 and Fig. 6. Finally, quantification of the biofilm biomass shows a significant reduction in the presence of mucins (Fig. 8D). These data suggests that both mucins and methylcellulose have the capacity to suppress, or destabilize, the attachment of C. albicans to an underlying polystyrene surface over a relatively long period of time. In Fig. 1B&C we showed that the cells continue to grow strongly in the presence of mucins, indicating that the reduced number of attached cells is not due to toxicity of the polymers. If cell proliferation is not suppressed by the polymers and yet, a reduction of biomass is observed on the surface, one would expect that a significant proportion of the population has been shifted to the supernatant. This could occur if cells are released from the biofilm and continue to propagate in the planktonic phase. Indeed, Fig. 8E shows that the supernatant from biofilms in the presence of mucins or methylcellulose contained higher numbers of cells than the supernatants from biofilms growing without mucins. This experiment also reveals another 71 Figure 8 Mucins reduce C. albicans biofilm formation. Biofilms were grown using strain HGFP3 that produces GFP upon transcription from a hyphal-specific promoter. A) Macroscopic view of biofilms grown in the presence and absence of mucins or methylcellulose. B) Fluorescent microscopy images overlaid on phase contrast images of C. albicans found in biofilm supernatants. C) Confocal images of biofilms grown in the presence and absence of mucins. D&E) Quantification of biofilm biomass (D) and C. albicans found in biofilm supernatants (E). 72 important point: in the presence of mucins, the non-attached cells are largely devoid of filaments, and therefore presumably in a less invasive form (Fig. 8B). In contrast, the cells in the supernatant of methylcellulose cultures were all highly filamentous. Together these data suggest that both mucins and methylcellulose can suppress the formation and maturation of biofilms by reducing attachment of the cells to an underlying surface, possibly by preventing stable maintenance of cells within the biofilm. However, mucins have a distinct effect: they render nearly the entire population devoid of filaments. Discussion This work shows that the mucin MUC5AC, which is expressed in the stomach and in the lungs, can induce the downregulation of several virulence traits in C. albicans, both at the level of gene expression and phenotype. These include the suppression of both filamentous growth and the formation of surface attached biofilms. Studies with other types of mucins [32, 33], suggest that the ability of mucins to manage microbial virulence may be a general mechanism that is present on all mucosal surfaces as part of the innate mucosal immune system. How might mucins prevent the transition to the hyphal form? Mucins, but not the other tested polymers, were capable of blocking this transition, suggesting that specific, mucinassociated glycans might be involved in this process. Consistent with this idea, glucose, maltose, and galactose in solution can all influence the formation of hyphae in C. albicans [34]. The identity of the mucin glycan moieties that are recognized by C. albicans, as well as the receptors and pathways in the yeast that are affected by these sugars, are currently unknown; their identification may suggest new valuable strategies for preventing or recovering C. albicans infections of mucosal surfaces. Preventing biofilm formation on materials exposed to living organisms presents a vexing engineering challenge; the results obtained for mucus hydrogels may provide some interesting new strategies. Our experiments show that mucins and methylcellulose are both effective in suppressing C. albicans surface attachment. Both polymers appear to reduce the ability for initial surface attachment; moreover, they render newly formed cells less capable of stably integrating into an emerging biofilm. How these polymers work to suppress surface attachment, and even whether they function by the same mechanisms, are open questions. Despite their superficial resemblance in surface protection, mucins and methylcellulose have different effects on the 73 surrounding C. albicans cell population. With methylcellulose, the vast majority of cells remain in the filamentous form, while a cell population exposed to mucins remains largely devoid of filaments. There is good experimental evidence that the ability to form filaments is required for C. albicans virulence [35]. Therefore, we have shown that the native mucins in the body are capable of providing a dual mechanism for virulence control: supressing the yeast-hyphae transition and surface attachment. The ability of mucins to suppress virulence traits is not specific for C. albicans but appears to apply also toward a range of other microorganisms, including the bacteria Helicobacter pylori and Pseudomonas aeruginosa [7,8] and also certain viruses such as HIV and influenza [10,11]. Understanding the mechanism of the mucin-Candida host-pathogen interaction could direct treatment strategies for regulating the healthy microbiota and also shed light on the molecular origin of increased susceptibility to microbial disease. Materials and Methods C. albicans Strains and Media Strains were maintained on YPD agar (2% Bacto peptone, 2% glucose, 1% yeast extract, 2% agar) and grown at 30°C. Single colonies were inoculated into YPD broth and grown with shaking overnight at 30°C prior to each experiment. The experiments were performed using RPMI 1640 (Gibco 31800-089) buffered with 165mM MOPS and supplemented with 0.2% NaHCO3 and 2% glucose or YPD + 10% Fetal Bovine Serum (FBS). 0.5% methylcellulose (Sigma, 15 cP) was prepared from a 5% stock solution by dilution in RPMI. Type II mucin from porcine stomach (Sigma) was dialyzed in a Spectra/Por Float-A-Lyzer G2 dialysis tube with a 100 kDa molecular weight cutoff, followed by lyophilization. Industrial PGM and native PGM were dissolved in RPMI and gently vortexed at 4°C overnight. The C. albicans yeast strains used in this study are SC5314 and HGFP3. Strain HGFP3 [36] was constructed by inserting the GFP gene next to the promoter of HWP1, a gene encoding a hyphal cell wall protein, and was provided by E. Mylonakis (Massachusetts General Hospital, Boston, MA, USA) with permission of P. Sundstrom (Stabb et al., 2003). 74 Mucin Purification The mucins were natively purified to preserve their properties, as opposed to industriallypurified mucins which do not form gels in solution [23,22]. Porcine gastric mucins (PGM) were purified from fresh pig stomachs as previously described [37]. Briefly, the mucus layer was isolated from pig stomachs and solubilized in sodium chloride buffer containing protease inhibitors to prevent mucin degradation, and sodium azide to prevent bacterial proliferation. Insoluble components were removed via centrifugation, and the mucins were isolated using gel filtration chromatography on a sepharose column (CL2B). The mucins were then concentrated and lyophilized. As a control to ensure that there were no contaminants in the mucin preparation, CsCl gradient centrifugation prepared mucins (as described in [38,39]) were compared to those prepared without this step. Growth curves 1 µL of overnight culture was added to 100 µL of medium (RPMI or YPD) in a 96-well plate. The wells contained plain medium or medium supplemented with 0.5% native mucin, Sigma PGM or methyl cellulose (Sigma Aldrich, cat # M7140). The plates were incubated at 30°C with shaking and the optical density at 650nm was read once every hour. The contents of each well were pipetted up and down before each reading was taken to ensure homogeneous distribution of cells. After 8 hours of growth, the cells were stained with 10 µg/mL calcofluor white for 5 minutes, aliquoted onto a microscope slide and imaged using a Zeiss Observer Z1 inverted fluorescence microscope with a Zeiss EC Plan-Neofluar® 20X objective lens (ex365/em445). Extraction of RNA and cDNA synthesis 1 mL of RPMI or 0.5% PGM in RPMI in a culture tube was inoculated with 10 µL of an overnight culture of strain SC5314 and incubated at 37°C and 180 rpm for 8 h. RNA was extracted using the Epicentre® MasterPure™ Yeast RNA Purification Kit and treated with Sigma-Aldrich® AMPD1 Amplification Grade DNase I. 500 ng of RNA per sample was used to generate cDNA using the Invitrogen™ Superscript® III First-Strand Synthesis System. cDNA samples were stored at -80°C until use. 75 Quantitative PCR All primers were obtained from Sigma and analyzed for efficiency before use in experiments. Efficiency was calculated by performing qPCR with serial 1:10 dilutions of genomic DNA. BioRad iQ™ SYBR® Green Supermix was used for qPCR reactions. Experiments were performed in a Roche LightCycler® 480 II machine with the following run protocol: (1) 95°C for 3 min, (2) 40 cycles of 95°C for 10 sec, 58°C for 30 sec, and 72°C for 30 sec. Crossing threshold (CT) values were obtained and used for analysis. Fold changes were calculated using the ΔΔCT method in comparison to the reference gene TAF145. White to Opaque Switch Assays White-to-opaque switch assays were performed on synthetic dextrose plates as previously described [40] to determine the white-to-opaque switch frequency for the wild-type a/a white strain RBY717 [41] after growth in liquid RPMI medium in the presence and absence of 0.5% purified mucins at 25ºC for 24h. Data is displayed in Table S1. The switching frequency is the percentage of colonies that displayed opaque sectors or opaque colonies. The “n” is the total number of colonies counted. Response to Alpha Factor The response of C. albicans wild-type white and opaque strains (RBY717a/a and RNY731a/a, respectively [41] in liquid RPMI medium in the presence of 0.5% purified mucin + α-factor, and without mucin + α-factor were assayed as described previously [17] after a 24h exposure period at 25ºC. α-factor in 10% DMSO was added at a final concentration of 10 µg/ml. Cells were scored for formation of elongated projections 24 h after α-factor was added to the cells. This experiment was also performed at 37ºC (data not shown), and mucin had no effect on the observed number of elongated projections in that condition as well. Data is displayed in Table S1. The “n” is the total number of cells counted, and the ratio of cells with projections was calculated as a percentage. Quantitative Mating Assay Quantitative mating assays were performed as previously described [42] in the presence and absence of 0.5% purified mucin at 25ºC for 5 days. The wild-type MTL heterozygote strains 76 SN87 and SN95 and opaque strains RBY1177a/a and RBY1180α/α containing different selectable markers were crossed for this assay. Data is displayed in Table S1. The ratio of cells with mating products was calculated as a percentage. Filamentation Assay 100 µL each of RPMI, 0.5% methylcellulose, 0.5% Industrial PGM, and 0.5% native PGM were inoculated with the strain HGFP3 as yeast-form cells or hyphae in a 96-well plate. 1 µL of an overnight culture was used as a source for yeast-form cells. For hyphae, an overnight culture was diluted 1:100 into YPD + 10% FBS, which stimulates the transition to hyphae, and grown to OD600 = 0.5. The hyphae were spun down, resuspended to OD600 = 5 and 10 µL was added to the aforementioned conditions. The cells were incubated at 37°C with 180 rpm shaking for 8 h. Adherent cells were scraped off of the surface and samples were pipetted vigorously to break up aggregates. 15 µL of each sample was placed on a microscope slide for visualization. Slides were imaged with a Zeiss Observer Z1 inverted fluorescence microscope with a Zeiss PlanApochromat 20X objective lens under phase contrast and FITC (ex475/em530). Polystyrene Attachment Assay Polystyrene 96-well plates were inoculated with 100 µL of RPMI, 0.5% methylcellulose, 0.5% Industrial PGM, and 0.5% native PGM containing yeast-form cells from the strain SC5314. The plates were incubated statically at 37°C. Every 15 minutes, a time point was taken by washing the wells with 200 µL of PBS twice followed by the addition of 100 µL of PBS. After 1 hour, 1 µL of 1mg/mL calcofluor white solution was added to each well. The samples were imaged as previously mentioned using a 10X objective (ex365/em445). The experiment was performed in triplicate with 5 pictures taken of each well. The images were analyzed in ImageJ as follows: each image was converted to 8-bit and the contrast was enhanced (0.4% saturated pixels). The image was then thresholded to create a binary image. The image was then analyzed using the “Analyze Particles” tool to measure the surface area covered by cells. The surface area measurements of the 15 images for each condition and timepoint were averaged. 77 Attachment to Human mucus-secreting colorectal cells (HT29-MTX) HT29-MTX Mucus-secreting cells reliably secrete a thick, homogeneous layer of mucus as soon as 7 days post confluency. The cells were grown in a 24-well plate. 2-weeks post confluency, the cells were treated with 10mM N-acetylcysteine (NAC), which cleaves disulfide bonds between mucins [43], for 30 min to remove the adherent mucus layer or with PBS as a control. For infection, C. albicans strain SC5314 was diluted from an overnight culture into DMEM to OD600 = 0.5. 500 µL of C. albicans was added on top of the HT29-MTX cells and incubated statically at 37°C for 2 hours. After 2 hours, the medium was removed from the wells, which were subsequently washed twice with 500 µL of PBS. The remaining C. albicans cells were stained with calcofluor white and analyzed using a Zeiss Observer Z1 inverted fluorescence microscope and a plate reader (ex355/em460). The HT29-MTX cells were derived from HT-29 cells (ATCC HTB-38) as described in [29]. HT-29 cells were derived from an anonymous donor. Biofilm formation assay and Visualization Biofilms were grown on either of two surfaces: in a 96-well plate (for macroscopic views and quantification) or on 8mm-diameter-silicone circles (for confocal imaging). Before the experiment, the silicone circles were washed with water and autoclaved. The surfaces were incubated in adult bovine serum overnight with shaking at 37°C. The next day, the surfaces were washed in PBS and submerged in a C. albicans cell suspension of OD600 = 0.5 in RPMI with or without 0.5% mucins. The samples were incubated at 37°C with shaking at 180 rpm for 90 minutes to facilitate attachment of yeast cells to the surface. Nonadherent cells were washed away with PBS and the samples were subsequently submerged in fresh RPMI with or without 0.5% mucins, 0.5% industrial mucins, or 0.5% methylcellulose. The biofilms were allowed to grow with shaking (180 rpm) at 37°C, for the indicated amount of time. For biomass quantification, the biofilms were stained with 20 µg/mL calcofluor white in PBS for 10 minutes and analyzed in a Spectra Max M3 plate reader (ex355/em460). Biofilm supernatant quantification was performed in the same plate reader (absorbance at 600nm). For confocal imaging, the biofilms were submerged in 6 mL of 20 µg/mL calcofluor white and stained for 10 minutes. The biofilms were imaged using a photo scanner or a Zeiss LSM 700 Upright Confocal. Planktonic cells were placed on microscope slides and imaged using a Zeiss widefield fluorescent microscope. 78 Acknowledgments We thank Dr. Paula Sundstrum and Dr. Eleftherios Mylonakis for strain HGFP3, Dr. Suzanne Noble for the wor1Δ/Δ strain SN1064, Dr. Richard Bennett for strains RBY717a/a, RBY731a/a, RBY1177 and RBY1180, and Dr. Thécla Lesuffleur for the HT29-MTX cell line. We also thank Dr. Bradley Turner for performing CsCl gradient centrifugation. We are grateful to Dr. Gerry Fink and Dr. Dawn Thompson for helpful comments and advice. This work was supported by CEHS Pilot Project Grant # P30-ES002109 (K.R. & N.L.K.), the MIT/NIGMS Biotechnology Training Program Grant # 5T32GM008334-24 (N.L.K.), Burroughs Wellcome Fund 2012 Collaborative Research Travel Grant (C.J.N.), National Institutes of Health grant K99AI100896 (C.J.N.) and National Institutes of Health grant R01 AI083311 (A.D.J.). References 1. Sudbery P, Gow N, Berman J. The distinct morphogenic states of Candida albicans. Trends Microbiol. 2004;12: 317–324. doi:10.1016/j.tim.2004.05.008 2. Calderone RA, Fonzi WA. Virulence factors of Candida albicans. Trends in Microbiology. 2001;9: 327–335. doi:10.1016/S0966-842X(01)02094-7 3. Donlan RM, Costerton JW. Biofilms: Survival mechanisms of clinically relevant microorganisms. Clin Microbiol Rev. 2002;15: 167–193. doi:10.1128/CMR.15.2.167193.2002 4. Drell T, Lillsaar T, Tummeleht L, Simm J, Aaspõllu A, Väin E, et al. Characterization of the vaginal micro- and mycobiome in asymptomatic reproductive-age Estonian women. PLoS ONE. 2013;8: e54379. doi:10.1371/journal.pone.0054379 5. Ghannoum MA, Jurevic RJ, Mukherjee PK, Cui F, Sikaroodi M, Naqvi A, et al. Characterization of the oral fungal microbiome (mycobiome) in healthy individuals. PLoS Pathog. 2010;6: e1000713. doi:10.1371/journal.ppat.1000713 6. Scanlan PD, Marchesi JR. Micro-eukaryotic diversity of the human distal gut microbiota: qualitative assessment using culture-dependent and -independent analysis of faeces. ISME J. 2008;2: 1183–1193. doi:10.1038/ismej.2008.76 7. Lindén SK, Sheng YH, Every AL, Miles KM, Skoog EC, Florin THJ, et al. MUC1 Limits Helicobacter pylori Infection both by steric hindrance and by acting as a releasable decoy. PLoS Pathog. 2009;5: e1000617. doi:10.1371/journal.ppat.1000617 8. Caldara M, Friedlander RS, Kavanaugh NL, Aizenberg J, Foster KR, Ribbeck K. Mucin biopolymers prevent bacterial aggregation by retaining cells in the free-swimming state. Curr Biol. 2012; doi:10.1016/j.cub.2012.10.028 79 9. Nelson AL, Roche AM, Gould JM, Chim K, Ratner AJ, Weiser JN. Capsule enhances Pneumococcal colonization by limiting mucus-mediated clearance. Infect Immun. 2007;75: 83–90. doi:10.1128/IAI.01475-06 10. Bergey EJ, Cho MI, Blumberg BM, Hammarskjöld ML, Rekosh D, Epstein LG, et al. Interaction of HIV-1 and human salivary mucins. J Acquir Immune Defic Syndr. 1994;7: 995–1002. 11. Couceiro JN, Paulson JC, Baum LG. Influenza virus strains selectively recognize sialyloligosaccharides on human respiratory epithelium; the role of the host cell in selection of hemagglutinin receptor specificity. Virus Res. 1993;29: 155–165. 12. Slutsky B, Staebell M, Anderson J, Risen L, Pfaller M, Soll DR. “White-opaque transition”: a second high-frequency switching system in Candida albicans. J Bacteriol. 1987;169: 189–197. 13. Kvaal C, Lachke SA, Srikantha T, Daniels K, McCoy J, Soll DR. Misexpression of the Opaque-Phase-Specific Gene PEP1 (SAP1) in the white phase of Candida albicans confers increased virulence in a mouse model of cutaneous infection. Infect Immun. 1999;67: 6652–6662. 14. Anderson JM, Soll DR. Unique phenotype of opaque cells in the white-opaque transition of Candida albicans. J Bacteriol. 1987;169: 5579–5588. 15. Xie J, Tao L, Nobile CJ, Tong Y, Guan G, Sun Y, et al. White-opaque switching in natural MTLa/α isolates of Candida albicans: Evolutionary implications for roles in host adaptation, pathogenesis, and sex. PLoS Biol. 2013;11: e1001525. doi:10.1371/journal.pbio.1001525 16. Miller MG, Johnson AD. White-Opaque switching in Candida albicans is controlled by mating-type Locus homeodomain proteins and allows efficient mating. Cell. 2002;110: 293–302. doi:10.1016/S0092-8674(02)00837-1 17. Bennett RJ, Uhl MA, Miller MG, Johnson AD. Identification and characterization of a Candida albicans mating pheromone. Mol Cell Biol. 2003;23: 8189–8201. doi:10.1128/MCB.23.22.8189-8201.2003 18. Hoyer LL, Green CB, Oh SH, Zhao X. Discovering the secrets of the Candida albicans agglutinin-like sequence (ALS) gene family--a sticky pursuit. Med Mycol. 2008;46: 1–15. doi:10.1080/13693780701435317 19. Nobile CJ, Fox EP, Nett JE, Sorrells TR, Mitrovich QM, Hernday AD, et al. A recently evolved transcriptional network controls biofilm development in Candida albicans. Cell. 2012;148: 126–138. doi:10.1016/j.cell.2011.10.048 20. Naglik JR, Challacombe SJ, Hube B. Candida albicans secreted aspartyl proteinases in virulence and pathogenesis. Microbiol Mol Biol Rev. 2003;67: 400–428. doi:10.1128/MMBR.67.3.400-428.2003 80 21. Gangloff Y-G, Romier C, Thuault S, Werten S, Davidson I. The histone fold is a key structural motif of transcription factor TFIID. Trends in Biochemical Sciences. 2001;26: 250–257. doi:10.1016/S0968-0004(00)01741-2 22. Crater JS, Carrier RL. Barrier properties of gastrointestinal mucus to nanoparticle transport. Macromolecular Bioscience. 2010;10: 1473–1483. doi:10.1002/mabi.201000137 23. Kocevar-Nared J, Kristl J, Smid-Korbar J. Comparative rheological investigation of crude gastric mucin and natural gastric mucus. Biomaterials. 1997;18: 677–81. 24. Worku ML, Sidebotham RL, Baron JH, Misiewicz JJ, Logan RP, Keshavarz T, et al. Motility of Helicobacter pylori in a viscous environment. Eur J Gastroenterol Hepatol. 1999;11: 1143–1150. 25. Smith DJ, Gaffney EA, Gadêlha H, Kapur N, Kirkman-Brown JC. Bend propagation in the flagella of migrating human sperm, and its modulation by viscosity. Cell Motility and the Cytoskeleton. 2009;66: 220–236. doi:10.1002/cm.20345 26. Alonso-Monge R, Navarro-García F, Molero G, Diez-Orejas R, Gustin M, Pla J, et al. Role of the mitogen-activated protein kinase Hog1p in morphogenesis and virulence of Candida albicans. J Bacteriol. 1999;181: 3058–3068. 27. Banerjee I, Pangule RC, Kane RS. Antifouling coatings: Recent developments in the design of surfaces that prevent fouling by proteins, bacteria, and marine organisms. Advanced Materials. 2011;23: 690–718. doi:10.1002/adma.201001215 28. Nucci M, Anaissie E. Revisiting the source of candidemia: skin or gut? Clin Infect Dis. 2001;33: 1959–1967. doi:10.1086/323759 29. Lesuffleur T, Porchet N, Aubert JP, Swallow D, Gum JR, Kim YS, et al. Differential expression of the human mucin genes MUC1 to MUC5 in relation to growth and differentiation of different mucus-secreting HT-29 cell subpopulations. J Cell Sci. 1993;106: 771–783. 30. Baillie GS, Douglas LJ. Role of dimorphism in the development of Candida albicans biofilms. J Med Microbiol. 1999;48: 671–679. doi:10.1099/00222615-48-7-671 31. Ramage G, VandeWalle K, López-Ribot JL, Wickes BL. The filamentation pathway controlled by the Efg1 regulator protein is required for normal biofilm formation and development in Candida albicans. FEMS Microbiol Lett. 2002;214: 95–100. 32. Ogasawara A, Komaki N, Akai H, Hori K, Watanabe H, Watanabe T, et al. Hyphal formation of Candida albicans is inhibited by salivary mucin. Biol Pharm Bull. 2007;30: 284–286. 33. Butrus SI, Klotz SA. Blocking Candida adherence to contact lenses. Curr Eye Res. 1986;5: 745–750. 81 34. Maidan MM, Thevelein JM, Van Dijck P. Carbon source induced yeast-to-hypha transition in Candida albicans is dependent on the presence of amino acids and on the G-proteincoupled receptor Gpr1. Biochem Soc Trans. 2005;33: 291–293. doi:10.1042/BST0330291 35. Saville SP, Lazzell AL, Monteagudo C, Lopez-Ribot JL. Engineered control of cell morphology in vivo reveals distinct roles for yeast and filamentous forms of Candida albicans during infection. Eukaryotic Cell. 2003;2: 1053–1060. doi:10.1128/EC.2.5.10531060.2003 36. Staab JF, Bahn Y-S, Sundstrom P. Integrative, multifunctional plasmids for hypha-specific or constitutive expression of green fluorescent protein in Candida albicans. Microbiology (Reading, Engl). 2003;149: 2977–2986. 37. Lieleg O, Lieleg C, Bloom J, Buck CB, Ribbeck K. Mucin biopolymers as broad-spectrum antiviral agents. Biomacromolecules. 2012;13: 1724–1732. doi:10.1021/bm3001292 38. Smith BF, LaMont JT. Hydrophobic binding properties of bovine gallbladder mucin. Journal of Biological Chemistry. 1984;259: 12170–12177. 39. Gong D, Turner B, Bhaskar KR, LaMont JT. Lipid binding to gastric mucin: Protective effect against oxygen radicals. American Journal of Physiology - Gastrointestinal and Liver Physiology. 1990;259: G681–G686. 40. Zordan RE, Miller MG, Galgoczy DJ, Tuch BB, Johnson AD. Interlocking transcriptional feedback loops control white-opaque switching in Candida albicans. PLoS Biol. 2007;5: e256. doi:10.1371/journal.pbio.0050256 41. Bennett RJ, Johnson AD. The role of nutrient regulation and the Gpa2 protein in the mating pheromone response of C. albicans. Mol Microbiol. 2006;62: 100–119. doi:10.1111/j.13652958.2006.05367.x 42. Bennett RJ, Miller MG, Chua PR, Maxon ME, Johnson AD. Nuclear fusion occurs during mating in Candida albicans and is dependent on the KAR3 gene. Mol Microbiol. 2005;55: 1046–1059. doi:10.1111/j.1365-2958.2005.04466.x 43. Sheffner AL. The reduction in vitro in viscosity of mucoprotein solutions by a new mucolytic agent, N-Acetyl-L-Cysteine. Annals of the New York Academy of Sciences. 1963;106: 298–310. doi:10.1111/j.1749-6632.1963.tb16647.x 82 Appendix A Mucins suppress Pseudomonas aeruginosa virulence toward Candida albicans 83 Background The opportunistic fungal pathogen Candida albicans and the bacterium Pseudomonas aeruginosa are often co-localized in both healthy and diseased humans. As a result of their coexistence, these two microbes have evolved a pathogenic relationship in which P. aeruginosa forms biofilms on C. albicans hyphae and secretes small molecules, such as phospholipases and redox active phenazines, that result in fungal cell death [1, 2]. Although these microbes typically exist near the mucosa, the impact of mucus on this relationship is unclear. Previous research from our lab shows that mucins, the main gel-forming components of mucus, suppress virulence traits of certain microbes, including C. albicans and P. aeruginosa [3, 4]. Therefore, we hypothesize that mucins likely impact antagonistic interactions between microbes. Through coculturing C. albicans and P. aeruginosa in the presence and absence of mucins, we explored the effects of mucins on microbial interactions. We found that C. albicans is protected from P. aeruginosa virulence in the presence of mucins. The protective effects of mucins were lost toward a constitutively hyphal C. albicans mutant, suggesting that mucin-mediated hyphal suppression may account for the protective effects. Additionally, we showed that mucins reduce physical contact between C. albicans and P. aeruginosa, leading to a decrease in bacterial biofilm formation on the fungus. We propose that through modulation of microbial virulence, mucins can influence microbial population dynamics and are likely important factors controlling microbial ecology in more complex communities, such as the microbiota. Results We began by coculturing P. aeruginosa PA14 and C. albicans SC5314 in the presence and absence of porcine gastric mucins mucins (PGM) to determine the effect of the polymers on C. albicans survival. PGM were chosen due to their enriched MUC5AC content, which is homologous to human gastric and lung mucins [5]. As a control, C. albicans was grown as a monoculture with and without mucins to identify any influences on growth; no differences were observed (Fig. 1A). Next, we added P. aeruginosa PA14 to the experiment to create C. albicans - P. aeruginosa cocultures. As expected, cocultures grown in the absence of mucins yielded a reduction in colony forming units (CFUs) as soon as 24 hours after addition of the bacteria, with complete eradication by 48 hours (Fig. 1A). Interestingly, the addition of mucins delayed 84 Figure 1 Mucins protect wild-type C. albicans from P. aeruginosa pathogenicity. Viability of C. albicans SC5314 wild type (A) or constitutive hyphal mutant tup1Δ/Δ (B) with and without pig gastric mucins in monoculture or coculture with P. aeruginosa. (C) Morphology of wild-type and tup1Δ/Δ C. albicans grown with and without mucins. (D) Viability of wild-type C. albicans with and without 0.5% methylcellulose in monoculture or coculture with P. aeruginosa. 85 C. albicans eradication by 24hrs, indicating that mucins play protective role against P. aeruginosa virulence. One possible explanation for reduced killing of C. albicans in the presence of mucins is that mucin-induced suppression of hyphae formation protects the yeast from P. aeruginosa virulence. Previous research demonstrates that P. aeruginosa selectively attached to and kills C. albicans hyphae, not yeast [1, 2]. To explore this possibility, we tested a constitutively hyphal strain, tup1Δ/Δ, in the survival assay. This strain forms hyphae within mucins, in contrast to the wild type in which hyphae formation is suppressed (Fig. 1C). When cocultured with PA14, tup1Δ/Δ viability decreased at the same rate regardless of the presence of mucins (Fig. 1B), suggesting that mucin-induced suppression of hyphae formation in the wild type confers the protection from P. aeruginosa observed in Fig. 1A. Additionally, these results suggest that mucin protection is conferred by influences on C. albicans and not P. aeruginosa because the bacterium causes tup1Δ/Δ cell death in both the presence and absence of mucins. We next tested 0.5% methylcellulose in the survival assay. Methylcellulose is a polysaccharide that forms viscous solutions that can be used to mimic mucus. It is highly effective at preventing C. albicans surface attachment, but does not inhibit hyphal formation [3]. This experiment was designed to determine if viscous polymer solutions in general confer protection, or if suppression of hyphal formation is necessary. During coculture, there appears to be a small protective effect of methylcellulose at 24 hrs, but by 48 hrs all cells are dead regardless of the presence or absence of the polymer. This suggests that mucin-mediated suppression of hyphal formation is more important in conferring protection than the viscosity of mucins. One possibility to explain mucin-mediated protection of the wild type is that P. aeruginosa selectively attaches to and kills hyphae, which are suppressed in the wild-type, mucin-exposed culture. Another possibility is that a non-attachment mediated form of virulence is responsible for the selective killing of hyphae, such as secreted factors. To distinguish between these possibilities, we imaged wild type or tup1Δ/Δ C. albicans with or without mucins, 24, 48 and 72hrs post coculture using scanning electron microscopy. Interestingly, the wild-type, mucin-exposed C. albicans had fewer attached P. aeruginosa at all time points compared to cells grown in the absence of mucins (Fig. 2A). To explore this further, the constitutive hyphal tup1Δ/Δ mutant, which was killed with equal efficiency with and without mucins in the survival 86 Figure 2 Mucins reduce P. aeruginosa attachment to C. albicans. Scanning electron microscopy images of wild-type (A) and tup1Δ/Δ (B) C. albicans after coculture with P. aeruginosa PA14 for the indicated period of time. 87 assay, was analyzed using SEM in the presence and absence of mucins. Interestingly, tup1Δ/Δ showed a slight reduction in attachment at 24 and 48 hrs in the presence of mucins, but was fully colonized by P. aeruginosa by 72hrs (Fig. 2B). Since the mutant is constitutively hyphal, we suggest that the presence of mucins and not morphology is the determining factor in bacterial attachment. Based on the data from the survival assay, tup1Δ/Δ is killed at the same rate in the presence or absence of mucins, yet attachment is less prevalent in the presence of mucins. Therefore, we suggest that killing is not completely dependent on attachment and is partially dependent on secreted factors. This theory is supported by previous research that shows cell-free spent coculture medium has killing effects against C. albicans [2]. One possibility is that hyphae are more sensitive to P. aeruginosa secreted factors and the constitutive presence of hyphae in the tup1Δ/Δ mutant allows for increased killing by P. aeruginosa despite the anti-attachment effects of mucins. This also explains the eventual killing of wild-type C. albicans by 72hrs, despite reduced bacterial attachment. Discussion In summary, we showed that mucins had significant impacts on the interactions between C. albicans and P. aeruginosa. Mucins protected wild-type C. albicans from P. aeruginosa virulence in coculture, but did not affect a constitutive hyphal mutant, suggesting that morphology plays a role in determining the protective effects of mucins. However, we found that mucins suppress bacterial attachment to both the wild type and tup1Δ/Δ, indicting that secreted bacterial factors are likely more effective in killing C. albicans than attachment. In this model, hyphae are sensitive to P. aeruginosa derived secreted factors whereas mucin-exposed wild-type cells are resistant. While this system suggests a role for mucins as regulators of microbe-microbe dynamics, it raises the question of how widespread the influences of mucins are. Presumably, a reduction in microbe-microbe attachment and virulence could have far reaching consequences on the microbiota. We hypothesize that mucins play large roles in shaping microbial communities and may be an important determinant of microbiota composition. Since mucins allow for the coexistence of C. albicans and P. aeruginosa and thereby increase microbial diversity, they 88 could allow for the stabilization of the microbiota and preventing certain species from dominating the population. Experimental Procedures Strains and Growth Conditions The strains used in this study are P. aeruginosa PA14 and C. albicans SC5314 and tup1Δ/Δ (courtesy of the Alexander Johnson, UCSF). C. albicans strains were streaked on YPD agar (2% Bacto peptone, 2% glucose, 1% yeast extract, 2% agar) from glycerol stocks and grown at 30°C. Single colonies were inoculated into YPD broth and grown with shaking overnight at 30°C prior to each experiment. P. aeruginosa was inoculated into LB broth from glycerol stocks and incubated overnight with shaking at 37°C. Experiments were carried out using RPMI (165mM MOPS, 2% glucose) for C. albicans growth and spent LB (SLB) for coculture. SLB was obtained by allowing PA14 to grow to OD600 = 1.6, centrifuging the culture and filtering the supernatant using a 0.2 µm syringe filter. Coculture survival 1 µL of an SC5314 overnight culture was inoculated into 100 µL of RPMI in a 96-well plate (Mattek), with or without 0.5% natively purified pig gastric mucins or methylcellulose (15cp, Sigma Aldrich) and grown for 4 hours with shaking at 37°C. Concurrently, 2mL of LB was inoculated with 40 µL PA14 and grown for 4 hours with shaking at 37°C. RPMI was then removed from C. albicans and replaced with 200 µL SLB. P. aeruginosa was added to C. albicans to a final OD600 = 0.25. A control without P. aeruginosa was included. At 0h, 24h, 48h and 72h, the contents of the wells were homogenized and a 20µL aliquot was serially diluted in PBS. 50 µL of the dilutions were plated on YPD agar + 30 µg/mL gentamicin and 60 µg/mL tetracycline (to select for C. albicans) and Cetrimide agar (to select for P. aeruginosa) and incubated overnight at 30°C and 37°C respectively. Colonies were enumerated after incubation. Scanning Electron Microscopy After coculture for the specified period of time, the cultures were filtered onto 2µm polycarbonate membranes (Millipore) and fixed with 2% glutaraldehyde. The samples were equilibrated in 0.1M sodium cacodylate buffer and stained with 1% OsO4. Next, the samples 89 were dehydrated in a series of ethanol baths, critical point dried and sputter coated before imaging with a JEOL 5600 Scanning Electron Microscope. References 1. Hogan DA, Kolter R. Pseudomonas-Candida Interactions: An Ecological Role for Virulence Factors. Science. 2002;296: 2229–2232. doi:10.1126/science.1070784 2. Brand A, Barnes JD, Mackenzie KS, Odds FC, Gow NA. Cell wall glycans and soluble factors determine the interactions between the hyphae of Candida albicans and Pseudomonas aeruginosa. Fems Microbiol Lett. 2008;287: 48–55. doi:10.1111/j.15746968.2008.01301.x 3. Kavanaugh NL, Zhang AQ, Nobile CJ, Johnson AD, Ribbeck K. Mucins Suppress Virulence Traits of Candida albicans. mBio. 2014;5: e01911–14. doi:10.1128/mBio.0191114 4. Caldara M, Friedlander RS, Kavanaugh NL, Aizenberg J, Foster KR, Ribbeck K. Mucin Biopolymers Prevent Bacterial Aggregation by Retaining Cells in the Free-Swimming State. Curr Biol CB. 2012; doi:10.1016/j.cub.2012.10.028 5. Turner BS, Bhaskar KR, Hadzopoulou-Cladaras M, Specian RD, LaMont JT. Isolation and characterization of cDNA clones encoding pig gastric mucin. Biochem J. 1995;308 ( Pt 1): 89–96. 90 Appendix B Selected antimicrobial essential oils eradicate Pseudomonas spp. and Staphylococcus aureus biofilms Work in this chapter was published in: Kavanaugh NL, Ribbeck K. 2012. Selected Antimicrobial Essential Oils Eradicate Pseudomonas Spp. and Staphylococcus Aureus Biofilms. Appl. Environ. Microbiol. 78:4057–4061. 91 Abstract One major challenge posed by cells within biofilms is their higher resistance to antibiotics than their free-living counterparts. Here, we show that selected antimicrobial essential oils can eradicate biofilms with higher efficiency than certain important antibiotics, making them interesting candidates for the treatment of biofilms. Main Text Microbial biofilms pose a challenge in clinical and industrial settings where the need for sterility is paramount. Bacteria within biofilms are more resistant to antibiotics and disinfectants than individual cells in suspension [1,2]. Several mechanisms can account for the increased antibiotic resistance in biofilms, including the physical barrier formed by exopolymeric substances [3], a proportion of dormant bacteria that are inert toward antibiotics [4], and resistance genes that are uniquely expressed in biofilms [5–8]. Together, these bacterial measures against antibiotics are urging the discovery of novel strategies that will effectively kill bacterial biofilms. Plant essential oils have been used for hundreds of years as natural medicines to combat a multitude of pathogens, including bacteria, fungi and viruses [9]. Several essential oils confer antimicrobial activity by damaging the cell wall and membrane, leading to cell lysis, leakage of cell contents, and inhibition of proton motive force [10]. In addition, there is evidence that they effectively kill bacteria without promoting the acquisition of resistance [11,12]. Finally, many essential oils are relatively easy to obtain, have low mammalian toxicity, and degrade quickly in water and soil, making them relatively environmentally friendly [13]. Here, we probe the ability of selected essential oils to kill biofilms formed by Pseudomonas aeruginosa (PAO1), Pseudomonas putida (KT2440) and Staphylococcus aureus SC-01. P. aeruginosa is a gram-negative bacterium found in the soil, water, and in animals, but is also an opportunistic pathogen in humans. It can infect the pulmonary and urinary tracts, wounds, and burns, and cause devastating medical complications by forming biofilms on medical devices, such as catheters. The biofilms formed by P. aeruginosa allow this pathogen to evade treatment with antibiotics and cause persistent, sometimes deadly, infections. The closely related P. putida can also form biofilms, but is not a pathogen. In rare cases, P. putida can cause infections in immunocompromised individuals. Usually, P. putida is found in the environment, especially in soil, freshwater, and on the roots of plants. The gram-positive S. aureus can exist 92 MIC MIC P. aeruginosa PAO1 P. putida KT2440 Colistin 3.0 µg ml-1 Not tested Cassia 0.2% (v/v) 0.2% Clove > 5% > 5% Lavender > 5% > 5% Peru Balsam 2.5% 2.5% Red Thyme > 5% 2.1 ± 0.4% Tea Tree 5% 2.5% Table 1 Minimum inhibitory concentration of colistin and essential oils as determined by the standard microbroth dilution assay. The data here represent the average minimum inhibitory concentrations of the antibiotic colistin and various essential oils. Each experiment was performed in triplicate. The highest concentration of each essential oil tested was 5% (v/v). Any oil that did not show antimicrobial activity in the range tested is listed as “ > 5%.” Standard error is reported, unless the results for all three trials were identical. both as a commensal and as a pathogen. As a pathogen, this bacterium is responsible for a broad range of maladies, from superficial skin infections to serious systemic infections. Treatment of S. aureus is complicated by antibiotic resistance, which is especially problematic in multidrug resistant strains such as methicillin-resistant S. aureus (MRSA). Essential extracts from the bark of plants in the genus Cinnamomum have antibacterial activity toward a range of different microbes, including P. aeruginosa [14–16]. Importantly, the effect of Cinnamomum extract on P. aeruginosa was described against individual bacteria in solution. We then asked if this potent antimicrobial would also be effective against this bacterium within a biofilm. To address this question, P. aeruginosa biofilms were grown on the air-liquid interface of a microscope slide, which was halfway submerged in Mueller Hinton Broth (MHB) containing PAO1 at an OD600 = 0.0025. After 24 hours of growth at room temperature, biofilms were 93 Figure 1 Cassia oil kills planktonic bacteria and biofilms with comparable efficiency. Cells were exposed to colistin or cassia oil for 2 h and then stained with a LIVE/DEAD stain to determine viabililty. Live cells are labeled in green (SYT09), and dead cells are labeled in red (propidium iodide). Shown here is one representative of three experiments. washed with H2O and then challenged with cation-adjusted MHB containing 0.2% or 0.1%(v/v) cassia oil (Cinnamomum aromaticum, 100% pure from Aura Cacia) or 3µg mL-1 colistin. In a separate assay, the CLSI microbroth dilution method modified with a 2-hour challenge period [17], 0.2% (v/v) cassia oil and 3µg/mL colistin were determined to be the lowest concentration of each chemical required to eradicate P. aeruginosa in solution (Table 1). In the case of cassia, 0.1% (v/v) Tween80 was added to mix the oil with the medium [18]. At this concentration, Tween80 did not affect the growth or viability of planktonic cells or cells in a biofilm (data not 94 shown). After 2 hours, the treated biofilms were rinsed with H2O, stained with LIVE/DEAD® BacLight™ (Invitrogen), and imaged by wide field fluorescence microscopy. BacLight™ uses a combination of two nucleic acid dyes: SYTO9, a membrane-permeable green dye that labels both viable and dead cells, and propidium iodide, a membrane-impermeable red dye that only labels membrane-compromised cells and eliminates the green SYTO9 signal. Planktonic cells (final OD600 = 0.25) were challenged with the same concentration of cassia or colistin used against the biofilms for 2 hours, and then placed into a glass-bottom 96-well plate for imaging. Our results show that the minimal inhibitory concentration (MIC) of colistin (3µg mL-1) needed to eradicate planktonic cells is not effective against cells within a biofilm, since a large fraction of the cells remains stained in green (Fig. 1, top right panel). In contrast, the MIC of cassia oil against planktonic cells (0.2%, Table 1) is also sufficient to kill the vast majority of P. aeruginosa cells within a biofilm (Fig. 1, middle panels), suggesting that these cells are not protected from cassia oil. A slightly lower concentration of the essential oil (0.1%) neither kills bacteria in solution, nor in biofilms (bottom panels). Are other antimicrobial essential oils similarly effective as cassia oil in killing Pseudomonas biofilms? To address this question, we screened for oils that can kill P. aeruginosa PAO1 in a disc diffusion assay using MHB agar according to the Clinical Laboratory and Standards Institute standard protocol [19]. The essential oils were supplied by Aura Cacia and New Directions Aromatics and described as 100% pure. 20µL of each oil was spotted undiluted onto filter paper discs created from 3-layers of Whatman filter paper (190µm). Our data revealed the following oils as effective in killing P. aeruginosa: Cassia, Clove (Syzygium aromaticum), Peru Balsam (Myroxylon balsamum), Red Thyme (Thymus vulgaris) and Tea Tree (Melaleuca alternifolia) (Fig. 2). To account for the possibility that the oils penetrate into the agar to different degrees, resulting in what falsely appears to be a reduced antimicrobial effect, any oil that produced a visible zone of inhibition was considered for subsequent experiments. In the next step, we explored whether the oils that were active in the disc diffusion assay are also effective in killing biofilms. To address this point, we determined two parameters for individual oils: the minimal inhibitory concentration (MIC) required to kill planktonic cells and the minimal biofilm eradication concentration (MBEC). Biofilms were grown on a MBEC™ device (Innovotech Inc., Edmonton, Canada), a modified microtiter plate that contains 96 polystyrene pegs attached to the lid [1]. The pegs were immersed in MHB containing 106 cells 95 Figure 2 Disc diffusion assay identifies essential oils with antimicrobial activity. Antibiotics at a concentration of 20 mg ml−1 (A) and pure essential oils (B) were tested against P. aeruginosa PAO1. The substances that produced a zone of inhibition were further analyzed; lavender oil served as a negative control. mL-1 while shaking at 37°C or 30°C for P. aeruginosa PAO1 and P. putida KT2440, respectively. After 24 hours, the biofilms that had grown on the pegs were rinsed and subjected to a 1:1 serial dilution of antibiotics and essential oils in cation-adjusted MHB as indicated in Fig. 3; the medium used to dilute the oils was supplemented with 0.1% Tween80. The volume of the challenge medium was 200 µl, and the highest concentration of antibiotics and essential oils tested was 100µg mL-1 and 5%, respectively. Ampicillin and lavender served as negative controls for antibiotics and essential oils, respectively. After 2 hours of incubation the pegs were washed, immersed in 150µL fresh MHB, and sonicated for 10 minutes in a Branson 2510 sonicator (40kHz) to release and dissociate the peg-associated biofilms. The average number of cells on each peg was determined by breaking the pegs off of the lid and sonicating them individually in microcentrifuge tubes containing 200µL of PBS. The resulting solution was 96 serially diluted and plated onto MHB agar plates to determine colony-forming units. The CFU counts revealed that the average numbers of cells per peg were 3 × 107 for PAO1 and 4 × 106 for KT2440. To obtain the MIC, the same number of planktonic cells was added per well to challenge with antibiotics or essential oils. After 2 hours, 20µL from each well was added to fresh MHB. After overnight incubation, the lowest concentration of each chemical that prevented survival the biofilm and planktonic cells was determined. The experiments were performed in triplicate and the average MIC or MBEC values were determined. Fig. 3a and 3b show the MIC and MBEC of each chemical tested. Only one antibiotic, ofloxacin, was able to eradicate planktonic and biofilm bacteria with almost equal efficiency. The other antibiotics, colistin and gentamicin, were not effective in killing biofilms, even at concentrations 10-fold higher than the MIC. In contrast, the essential oils cassia and Peru balsam were effective against biofilms and planktonic bacteria at nearly equal concentrations. This observation confirms the result in Fig. 1 for cassia oil, where little difference between the MIC and MBEC was observed. Interestingly, red thyme oil is effective against biofilms at a concentration of ~2%, but is unable to kill planktonic cells at any of the concentrations tested. This suggests that thyme oil is more effective against biofilms than it is against bacteria in solution. To determine statistical significance, we performed a one-sample t-test to compare the biofilm population to the mean of the planktonic population. Since the planktonic population for red thyme was not killed by the highest concentration tested, we used the maximum value (5%) as the population mean to see if significance could be detected at this level. Indeed, the difference between the planktonic and biofilm populations for both PAO1 and KT2440 were significant (p <0.05), indicating that red thyme oil is more effective against biofilms than planktonic cells. We conclude that the essential oils tested here can act against biofilms more effectively than the tested antibiotics. To test for potential strain-specific effects of the essential oils we assessed their effect on a close relative, P. putida (KT2440) (Fig. 3c). Our data illustrate that P. putida is more sensitive than P. aeruginosa to clove, red thyme and tea tree oils (Fig. 3). This effect is especially evident for clove oil, which does not eliminate PAO1 but is potent against both KT2440 biofilms and planktonic cells (Fig. 3). Using one-sample t-tests to compare the P. putida and P. aeruginosa data for clove oil (assuming that the mean for the P. aeruginosa samples is 5%), we found the effect of clove oil on P. putida is significantly different from that on P. aeruginosa (P < 0.05). 97 Figure 3 Activities of selected antibiotics and antimicrobial essential oils against P. aeruginosa PAO1 (A, B) and P. putida KT2440 (C). The MIC and MBEC of various substances were determined by challenging bacteria that were planktonic or within biofilms, respectively. Asterisks represent data that extend beyond the plot range, indicating that no killing was observed at the tested concentrations. Each experiment was performed in triplicate, and the error bars represent standard error. 98 Additionally, red thyme is effective against planktonic bacteria at 5%, the highest concentration tested and tea tree oil is effective against biofilm cells. We are unable to calculate statistical significance in these cases due to the samples surviving at the highest concentration tested. The differences between the two Pseudomonas species indicate species-specific activity of the oils and suggest that specific mechanisms of resistance to the oils may be at work. For example, since certain essential oils appear to work on the cell wall or cell membrane, it is possible that the composition of these cellular components is key to determining susceptibility to essential oils. The species-specific activity of the oils suggests that tailored combinations to target a range of different microbes may be effective against multi-species biofilms. One similarity between the P. aeruginosa and P. putida data is that red thyme oil is more effective against biofilm cells than their planktonic counterparts. The same is true for tea tree oil against P. putida, which is ineffective at concentrations of 5% or less against planktonic bacteria, but is effective against biofilm bacteria at a concentration of ~4%. In these cases, being inside a biofilm turns into a disadvantage to the bacteria, as it renders them more susceptible to the activity of this particular essential oil. It is possible that the extracellular matrix of the biofilm adsorbs the active components and increases their local concentration. Another possibility is that the cell membrane or cell wall in biofilm cells is different than planktonic cells due to differential gene expression in the two cell types. It should be noted that the data obtained using the MBEC device is reproducible with a different assay where biofilms are grown in the wells of a 96-well microtiter plate [20] instead of on polystyrene pegs (Fig. 4). In this assay, the protocol is the same as that for the MBEC device, with the difference that the plates are not shaken during incubation, allowing the biofilms to grow on the sides of the wells at the air-liquid interface. Additionally, the biofilms are formed in TB (1% tryptone, 0.5% NaCl) as opposed to MHB. In Fig. 4, we compare the susceptibility of biofilms that were grown with both approaches. The biofilms grown with the two different methods contain comparable number of cells (2x107 CFU/well vs. 3 x 107 CFU/peg). The MBEC values obtained for each method were the same for all substances except ofloxacin. It is unclear why cells grown on the MBEC device are more susceptible to ofloxacin than cells grown in 96well plates, especially considering that the other substances tested do not show significant variation in efficacy. However, it is possible that the difference in the medium or structure of the biofilms caused the increased susceptibility of biofilms grown on the MBEC device to ofloxacin. 99 Conc. of antibiotic (μg/mL) A 100 Conc. of essential oil (% v/v) * * * * * 90 80 70 60 Wells 50 MBEC Device 40 30 20 10 0 B * * Colistin Gentamicin Ofloxacin * * 5 Ampicilin * * * * 4 3 2 1 0 Cassia Clove Peru Balsam Red Thyme Tea Tree Lavender Figure 4 Comparison of two methods of biofilm cultivation for antibiotic and essential oil testing. P. aeruginosa biofilms were grown either on the sides of wells in a 96-well plate or on the pegs of an MBEC device, and their sensitivities toward antibiotics and essential oils were determined. After testing two closely related gram-negative bacteria, we studied the effect of essential oils against the gram-positive bacterium S. aureus (Fig. 5). Our goal was to determine if the oils discriminate between gram-positive and gram-negative bacteria. The strain used in this study, SC-01, is a biofilm-forming, oxacillin- and methicillin-resistant clinical isolate [21]. Certain essential oils, such as tea tree, thyme, and peppermint, are effective against planktonic [18,22,23] and biofilm [24,25] MRSA. However, essential oils from cassia, red thyme, or clove have not 100 Figure 5 Susceptibility of S. aureus SC-01 to essential oils. (A) A disc diffusion assay reveals that SC-01 is sensitive to various essential oils. (B) The MIC and MBEC of essential oils were determined by challenging planktonic cells and biofilms, respectively. Asterisks represent data that extend beyond the plot range, indicating that no killing was observed at the tested concentrations. Each experiment was performed in triplicate, and the error bars represent errors. been tested against MRSA biofilms of any strain. Moreover, the strain used in this study (SC-01) has not been challenged with essential oils in previous work. First we performed a disc diffusion assay to determine if the same oils that are effective against Pseudomonas work against S. aureus. Indeed, all of the oils tested, including lavender, showed a zone of inhibition (Fig. 5A). 101 MIC Cinnamaldehyde 0.1% (v/v) MBEC 0.2% Eugenol >5% 3.3 ± 0.8% Linalool > 5% > 5% Table 2 MIC and MBEC of Essential Oil Components Against PAO1 Next, we tested the essential oils against biofilms formed on the pegs of the MBEC device. The protocol is the same as described above, and the average number of cells per peg was 1.5 x 105. The results show that the biofilms were killed by the same or similar concentrations of cassia, Peru balsam, and red thyme oils as were effective against P. aeruginosa (Fig. 5B). Notably, this strain is resistant to methicillin yet is killed effectively by four essential oils tested in this assay. After determining that essential oils are effective against biofilms, we tested individual components of the essential oils for antimicrobial efficacy. We assessed the molecules cinnamaldehyde, eugenol and linalool (from cassia, clove and lavender oils, respectively) [26,27] for their effect against P. aeruginosa planktonic and biofilm cells. All three components were obtained from Sigma Aldrich. The protocol used in this assay is identical to that used for testing whole essential oils, including the use of Tween80 in the medium to suspend the components, which have a low solubility in water. Table 2 summarizes the data, which indicate that cinnamaldehyde is as effective as the complex cassia oil. Additionally, whereas clove oil is not effective in killing P. aeruginosa biofilms in 5% v/v solutions, its ingredient eugenol is effective at 3.3%. The finding that single essential oil components are effective at eradicating bacterial biofilms is promising, as it may allow the dissection of their mechanisms of action, as well as inspire the molecular design of new antimicrobial components. In summary, we demonstrate here that the essential oils cassia, Peru balsam and red thyme are more effective in eradicating Pseudomonas and S. aureus biofilms than selected important antibiotics, making them interesting candidates for the treatment of biofilms. Important future goals include identifying further active antimicrobial components within the oils, as well as the molecular mechanisms by which these components so effectively breach the 102 biofilm barrier. In this study, we only sampled a small number of different oils, but a plethora of other oils is available in nature, bearing an enormous potential for the discovery of alternative treatments to antibiotics. Acknowledgements We thank Marina Caldara for discussion of intermediate results and for valuable comments on the manuscript. This work was supported by the MIT startup funds for Katharina Ribbeck. References 1. Ceri H, Olson ME, Stremick C, Read RR, Morck D, Buret A. The Calgary Biofilm Device: new technology for rapid determination of antibiotic susceptibilities of bacterial biofilms. J Clin Microbiol. 1999;37: 1771–1776. 2. Stewart PS, William Costerton J. Antibiotic resistance of bacteria in biofilms. The Lancet. 2001;358: 135–138. doi:doi: DOI: 10.1016/S0140-6736(01)05321-1 3. Lynch AS, Robertson GT. Bacterial and fungal biofilm infections. Rev Med. 2008;59: 415– 428. doi:10.1146/annurev.med.59.110106.132000 4. Mah TF, O’Toole GA. Mechanisms of biofilm resistance to antimicrobial agents. Trends Microbiol. 2001;9: 34–39. 5. Mulcahy H, Charron-Mazenod L, Lewenza S. Extracellular DNA chelates cations and induces antibiotic sesistance in Pseudomonas aeruginosa biofilms. PLoS Pathog. 2008;4: e1000213. doi:10.1371/journal.ppat.1000213 6. Nguyen D, Joshi-Datar A, Lepine F, Bauerle E, Olakanmi O, Beer K, et al. Active starvation responses mediate antibiotic tolerance in biofilms and nutrient-limited bacteria. Science. 2011;334: 982 –986. doi:10.1126/science.1211037 7. O’Toole GA, Mah TF, Pitts B, Pellock B, Walker GC, Stewart PS. A genetic basis for Pseudomonas aeruginosa biofilm antibiotic resistance. Nature. 2003;426: 306–310. doi:10.1038/nature02122 8. Zhang L, Mah T-F. Involvement of a novel efflux system in biofilm-specific resistance to antibiotics. Journal of Bacteriology. 2008;190: 4447 –4452. doi:10.1128/JB.01655-07 9. Hammer K, Carson C, Riley T. Antimicrobial activity of essential oils and other plant extracts. Journal of Applied Microbiology. 1999;86: 985–990. 10. Burt S. Essential oils: their antibacterial properties and potential applications in foods--a review. Int J Food Microbiol. 2004;94: 223–253. doi:10.1016/j.ijfoodmicro.2004.03.022 103 11. Ali S, Khan A, Ahmed I, Musaddiq M, Ahmed K, Polasa H, et al. Antimicrobial activities of Eugenol and Cinnamaldehyde against the human gastric pathogen Helicobacter pylori. Annals of clinical microbiology and antimicrobials. 2005;4: 20. 12. Ohno T, Kita M, Yamaoka Y, Imamura S, Yamamoto T, Mitsufuji S, et al. Antimicrobial activity of essential oils against Helicobacter pylori. Helicobacter. 2003;8: 207–215. 13. Isman M.B. Plant essential oils for pest and disease management. Crop Protection. 2000;19: 603–608. doi:10.1016/S0261-2194(00)00079-X 14. Bouhdid S, Abrini J, Amensour M, Zhiri A, Espuny MJ, Manresa A. Functional and ultrastructural changes in Pseudomonas aeruginosa and Staphylococcus aureus cells induced by Cinnamomum verum essential oil. J Appl Micro. 2010;109: 1139–1149. doi:10.1111/j.1365-2672.2010.04740.x 15. Nuryastuti T, van der Mei HC, Busscher HJ, Iravati S, Aman AT, Krom BP. Effect of Cinnamon Oil on icaA Expression and Biofilm Formation by Staphylococcus epidermidis. Appl Env Microbiol. 2009;75: 6850–6855. doi:10.1128/aem.00875-09 16. Prabuseenivasan S, Jayakumar M, Ignacimuthu S. In vitro antibacterial activity of some plant essential oils. BMC Complement Altern Med. 2006;6: 39–39. doi:10.1186/14726882-6-39 17. CLSI. Methods for dilution antimicrobial susceptibility tests for bacteria that grow aerobically; approved standard M07-A8, 8th ed. Clinical and Laboratory Standards Institute, Wayne, PA.; 2009. 18. Carson CF, Riley TV. Antimicrobial activity of the essential oil of Melaleuca alternifolia. Lett Appl Microbiol. 1993;16: 49–55. doi:10.1111/j.1472-765X.1993.tb00340.x 19. CLSI. Performance standards for antimicrobial disk susceptibility tests; approved standard 02-A10, 10th ed. Clinical and Laboratory Standards Institute, Wayne, PA.; 2009. 20. Friedman L, Kolter R. Genes involved in matrix formation in Pseudomonas aeruginosa PA14 biofilms. Mol Microbiol. 2004;51: 675–690. 21. Tenover FC, Arbeit R, Archer G, Biddle J, Byrne S, Goering R, et al. Comparison of traditional and molecular methods of typing isolates of Staphylococcus aureus. J Clin Microbiol. 1994;32: 407–415. 22. Nelson R. In-vitro activities of five plant essential oils against methicillin- resistant Staphylococcus aureus and vancomycin-resistant Enterococcus faecium. Journal of Antimicrobial Chemotherapy. 1997;40: 305–306. doi:10.1093/jac/40.2.305 23. Nostro A, Blanco AR, Cannatelli MA, Enea V, Flamini G, Morelli I, et al. Susceptibility of methicillināresistant staphylococci to oregano essential oil, carvacrol and thymol. FEMS Microbiology Letters. 2004;230: 191–195. doi:10.1016/S0378-1097(03)00890-5 104 24. Brady A, Loughlin R, Gilpin D, Kearney P, Tunney M. In vitro activity of tea-tree oil against clinical skin isolates of methicillin-resistant and -sensitive Staphylococcus aureus and coagulase-negative staphylococci growing planktonically and as biofilms. J Med Microbiol. 2006;55: 1375–1380. doi:10.1099/jmm.0.46558-0 25. Jia P, Xue YJ, Duan XJ, Shao SH. Effect of cinnamaldehyde on biofilm formation and sarA expression by methicillin-resistant Staphylococcus aureus. Lett Appl Microbiol. 2011;53: 409–416. doi:10.1111/j.1472-765X.2011.03122.x 26. Lis-Balchin M. Lavender: the genus Lavandula. CRC Press; 2002. 27. Parthasarathy VA, Chempakam B, Zachariah TJ. Chemistry of spices. CABI; 2008. 105 Chapter 4 Conclusions and future directions 106 In this thesis I propose that mucins defend the body from infections by modulating bacterial and fungal virulence and supporting microbial diversity. I found that mucins effectively suppress virulence traits in two opportunistic pathogens: Pseudomonas aeruginosa and Candida albicans. The mucin-mediated effects on virulence not only impact each microbe individually, but also the relationship between the two when grown in coculture. Specifically, mucins reduce P. aeruginosa pathogenicity towards C. albicans, indicating that mucins can affect the microbial composition of cocultures. Based on this work, I hypothesize that mucins have considerable potential to influence human health by modulating bacterial physiology and diversity in the host microbiota. Prior to this thesis, few studies focused on the effects of mucus on microbial physiology; most research concentrated on the opposite effect where microbes impact mucus integrity and expression. While such studies are indispensible for understanding the barrier properties of mucus during mucosal diseases, the influence of the mucus environment on microbial physiology is equally important. Characterizing mucus-mediated effects on microbes allows us to elucidate effective strategies for manipulating microbial physiology, particularly in ways that reduce virulence. In Chapter 2, I showed that mucins suppress biofilm formation of wild type P. aeruginosa. Biofilms are highly problematic due to their mechanical robustness and significant resistance to antibiotics. Plant essential oils are also effective at targeting bacterial biofilms when antibiotics fail, as I show in Appendix B. However essential oils can have toxic effects at high concentrations inside the human body and are difficult to deliver to some parts of the mucosa, such as the lungs and intestines. Therefore, mucins may serve as effective, biocompatible materials to reduce biofilm formation. The mechanisms by which mucins suppress biofilm formation are unknown, but several possibilities exist. Mucins may form surface coatings that prevent bacterial adhesion. Indeed, previous studies show that mucin coatings spontaneously form on a number of surfaces and repel binding of mammalian [1] and bacterial cells [2]. Another possibility is that mucins occlude P. aeruginosa surface proteins that are involved in surface attachment. For example, P. aeruginosa binds to mucins via the flagellum [3–5], which is also involved in adhesion to surfaces [6]. Mucins may therefore block flagellar surface attachment domains either by binding to them directly or by steric hindrance. 107 Although mucins are successful in inhibiting surface associated biofilms, we show that P. aeruginosa can overcome the protective effects of mucins when it loses flagellar motility. The loss of motility is a common characteristic of P. aeruginosa cystic fibrosis isolates, consistent with this being advantageous for the colonization of mucus [7]. In our experiments, flagellar deletion mutants form flocs that resemble biofilms in their structure and demonstrate resistance to antibiotics, much like the colonies formed in the lungs of cystic fibrosis patients. Indeed, clinical cystic fibrosis isolates also formed flocs in mucin solutions. Interestingly, the loss of motility is not a requirement for floc formation, since a number of motile clinical isolates also colonized mucus. Future work includes elucidating the traits of the clinical isolates that allow them to colonize mucus despite their motility. Regardless of the mechanism, the ability to recapitulate clinical phenotypes in vitro presents a model system in which P. aeruginosa clinical isolates can be studied in a physiologically relevant manner. For example, bacteria from cystic fibrosis patients could be inoculated into mucins and challenged with antibiotics to more accurately predict successful treatment options. Although P. aeruginosa is able to overcome the suppressive effects of mucins through the loss of motility, the dispersive effect of these polymers on wild-type bacteria suggests that they have potential to reduce microbial virulence. Indeed, mucin-mediated virulence trait suppression is also observed with the fungus C. albicans. In Chapter 3, I found that mucins suppress the yeast to hyphal transition, surface attachment, biofilm formation, and the expression of a number of virulence-associated genes in C. albicans. The modulation of virulenceassociated gene expression by mucins hints at an exciting potential function of these polymers: to persuade microbes to remain in an avirulent state when they are in close proximity to the epithelium. The suppression of a selection of C. albicans virulence traits by mucins suggest that these polymers are important for protection from disease but a number of open questions remain. First, do mucins impact non-virulent processes? Presumably, the answer is yes. C. albicans is highly responsive to environmental factors such as mechanical stress, sugar concentration and oxygen availability [8–10], therefore high concentrations of mucin glycans and the mucin hydrogel structure likely impact many aspects of C. albicans physiology. Secondly, how do mucins modulate virulence gene and trait expression? The work presented in this thesis does not elucidate the mechanisms by which mucins suppress C. albicans virulence traits, but future work 108 to address this point may include studies of global gene expression. Only a small portion of virulence-related genes were analyzed in this study. Therefore a broader characterization of the impact of mucins on gene expression may suggest the involvement of previously characterized pathways, which would enhance our understanding of mucin-mediated effects on C. albicans physiology. Another method to guide the discovery of mucin-mediated virulence suppression is a deletion mutant screen. For example, a library of characterized C. albicans deletion mutants could be inoculated into mucins and observed for the formation of hyphae in the presence of mucins, which is normally suppressed. Although such a study may provide valuable insights into the mechanisms of mucin-mediated virulence trait suppression, one caveat regarding these types of studies is that C. albicans has redundancies in many of its pathways, including filamentation and adhesion [11–13], making it difficult to find single deletion mutants that behave differently from the wild-type in the presence of mucins. Screening a panel of transcription factor mutants allows for the identification of gene networks involved in the response to mucins, and may alleviate concerns of redundancies if redundant genes have common regulatory networks. Regardless of these caveats, the use of global gene expression analysis or mutant libraries is a first step that can be followed by more focused approaches, such as the analysis of deletion mutants or the study of single proteins. In Appendix A, P. aeruginosa and C. albicans were combined inside mucin environments to determine the impacts of the biopolymers on microbial community dynamics. P. aeruginosa is antagonistic toward C. albicans, typically outcompeting the fungus within 48 hours of coculture. However, the presence of mucins allowed the microbes to coexist for 72 hours. In this system, mucins influence microbial diversity by reducing bacterial virulence toward its fungal counterpart. The coculture experiments suggest that the suppression of microbial pathogenicity induced by mucins is beneficial for microbes by allowing them to thrive in environments that are normally deadly. In this scenario, mucins not only impact microbial virulence, but also the structure of the microbiota, which is emerging as an important determinant of human health. Certain microbiota compositions in humans are linked to different health and disease states, such as obesity [14], inflammatory bowel disease [15] and arthritis [16]. One can imagine that people with mucin defects, such as aberrant glycosylation patterns or insufficient mucin production, may experience dysbiosis that leads to disease. This idea is supported by a study in which a mouse strain deficient in the B4galnt2 glycosyltransferase, 109 which glycosylates intestinal epithelial cell surfaces, shows alterations in the intestinal microbiota composition compared to wild-type mice [17], suggesting that glycans are important determinants of the microbiota composition. An analysis of mucin glycosylation patterns may therefore provide insight into patient health, both to diagnose and to predict ailments. This idea has been put forth in the context of ulcerative colitis, where patients with active disease show altered mucin glycosylation patterns when compared to healthy individuals [18]. Supplementation with healthy mucins may prove to be therapeutic for those suffering from dysbiosis related to insufficient or damaged mucins. The use of mucins as therapeutics has many advantages, one being that they suppress microbial virulence without sacrificing cell viability. The reduction in P. aeruginosa and C. albicans biofilm formation without cell death is particularly important, because antibiofilm methods that are bacteriocidal, e.g. antibiotics or disinfectants, often lead to the development of resistance over time. In the case of mucins, the selection of resistant population by killing sensitive cells does not occur, therefore reducing the risk of evolved resistance. Evolution assays in which microbes are passaged daily into fresh mucins can be employed to test if the inoculum loses its sensitivity to mucins over time. For example, the C. albicans strain HGFP3, which expresses GFP only when true hyphae are formed, could be monitored for fluorescence after each passage to determine if mucins lose their ability to suppress hyphae formation. Mutants would then be sequenced to identify evolved loci and thereby the mechanism used by the fungus to gain resistance. Important to note is that none of the other tested polymers, namely industrial mucins, methylcellulose, or PEG, display all of the effects of mucins. In the case of C. albicans, industrial mucins suppressed hyphal formation but did not disrupt surface attachment. The opposite case was true for methylcellulose, which reduced surface attachment but not hyphal formation. Mucins can confer multiple methods of virulence trait suppression against C. albicans, which suggests that they are highly evolved to tackle many facets of microbial virulence. In the coculture assay, methylcellulose did not extend the viability of C. albicans as was seen with mucins. Perhaps the presence of both peptides and glycans components on mucins, both of which can vary based on gene expression and bodily location, allows for finely tuned influences on microbial behavior that cannot be achieved by the presence of a viscous solution alone. Because the manipulation of mucin glycan structures is very difficult to achieve 110 in a standardized manner, thorough analysis of mucin structure may elucidate important mucin characteristics that influence disease. For example, molecular characterization of glycan structure and the peptide backbone can be performed on mucins from patients that suffer from C. albicans infections and compared to those from healthy people. To strengthen the idea that mucins are general regulators of microbial physiology and community structure, a range of microbes need to be studied in the presence of mucins, including pathogenic and commensal bacteria and fungi. An example from one of my colleagues shows that mucins reduce biofilm formation of the Streptococcus mutans, a bacterium responsible for dental cavity formation [19], suggesting that these polymers have broad effects on different microbial species. Finally, more complex but perhaps more physiologically relevant studies of dual species or multi species communities in mucus are needed to determine the effects of mucins on microbial community structure. The work presented in this thesis indicates that mucins influence microbial virulence traits and interactions, which represents a new perspective on the protective functions of mucus. Mucus is typically considered a physical barrier to infection. Here, I show that mucins have massive potential to reduce microbial virulence and impact the composition of the microbiota. A picture of the far-reaching influences of mucins is just starting to emerge, and our work already suggests that these polymers are widely effective against microbial virulence. This work represents a first step toward understanding the potential of mucins with the hope that one day we can harness their strategies to develop effective antimicrobial therapeutics. 111 References 1. Crouzier T, Jang H, Ahn J, Stocker R, Ribbeck K. Cell patterning with mucin biopolymers. Biomacromolecules. 2013;14: 3010–3016. doi:10.1021/bm400447z 2. Lei Shi RA. Mucin coating on polymeric material surfaces to suppress bacterial adhesion. Colloids Surf B Biointerfaces. 2000;17: 229–239. doi:10.1016/S0927-7765(99)00121-6 3. Sajjan U, Reisman J, Doig P, Irvin RT, Forstner G, Forstner J. Binding of nonmucoid Pseudomonas aeruginosa to normal human intestinal mucin and respiratory mucin from patients with cystic fibrosis. J Clin Invest. 1992;89: 657–665. doi:10.1172/JCI115632 4. Arora SK, Ritchings BW, Almira EC, Lory S, Ramphal R. The Pseudomonas aeruginosa flagellar cap protein, FliD, is responsible for mucin adhesion. Infect Immun. 1998;66: 1000–7. 5. Scharfman A, Arora SK, Delmotte P, Van Brussel E, Mazurier J, Ramphal R, et al. Recognition of Lewis x derivatives present on mucins by flagellar components of Pseudomonas aeruginosa. Infect Immun. 2001;69: 5243–8. 6. Tran VB, Fleiszig SMJ, Evans DJ, Radke CJ. Dynamics of Flagellum- and Pilus-Mediated Association of Pseudomonas aeruginosa with Contact Lens Surfaces. Appl Environ Microbiol. 2011;77: 3644–3652. doi:10.1128/AEM.02656-10 7. Mahenthiralingam E, Campbell ME, Speert DP. Nonmotility and phagocytic resistance of Pseudomonas aeruginosa isolates from chronically colonized patients with cystic fibrosis. Infect Immun. 1994;62: 596–605. 8. Maidan MM, Thevelein JM, Van Dijck P. Carbon source induced yeast-to-hypha transition in Candida albicans is dependent on the presence of amino acids and on the G-proteincoupled receptor Gpr1. Biochem Soc Trans. 2005;33: 291–293. doi:10.1042/BST0330291 9. Kumamoto CA. A contact-activated kinase signals Candida albicans invasive growth and biofilm development. Proc Natl Acad Sci U S A. 2005;102: 5576–5581. doi:10.1073/pnas.0407097102 10. Synnott JM, Guida A, Mulhern-Haughey S, Higgins DG, Butler G. Regulation of the Hypoxic Response in Candida albicans. Eukaryot Cell. 2010;9: 1734–1746. doi:10.1128/EC.00159-10 11. Sheppard DC, Yeaman MR, Welch WH, Phan QT, Fu Y, Ibrahim AS, et al. Functional and Structural Diversity in the Als Protein Family of Candida albicans. J Biol Chem. 2004;279: 30480–30489. doi:10.1074/jbc.M401929200 12. Lu Y, Su C, Solis NV, Filler SG, Liu H. Synergistic Regulation of Hyphal Elongation by Hypoxia, CO2, and Nutrient Conditions Controls the Virulence of Candida albicans. Cell Host Microbe. 2013;14: 499–509. doi:10.1016/j.chom.2013.10.008 112 13. Kadosh D, Lopez-Ribot JL. Candida albicans: Adapting to succeed. Cell Host Microbe. 2013;14: 483–485. doi:10.1016/j.chom.2013.10.016 14. Ley RE, Turnbaugh PJ, Klein S, Gordon JI. Microbial ecology: human gut microbes associated with obesity. Nature. 2006;444: 1022–1023. doi:10.1038/4441022a 15. Sartor RB. Microbial influences in inflammatory bowel diseases. Gastroenterology. 2008;134: 577–594. doi:10.1053/j.gastro.2007.11.059 16. Bordon Y. Autoimmunity: Altered microbiota linked to rheumatoid arthritis. Nat Rev Immunol. 2013;13: 848–848. doi:10.1038/nri3575 17. Staubach F, Künzel S, Baines AC, Yee A, McGee BM, Bäckhed F, et al. Expression of the blood-group-related glycosyltransferase B4galnt2 influences the intestinal microbiota in mice. ISME J. 2012;6: 1345–1355. doi:10.1038/ismej.2011.204 18. Larsson JMH, Karlsson H, Crespo JG, Johansson MEV, Eklund L, Sjövall H, et al. Altered O-glycosylation profile of MUC2 mucin occurs in active ulcerative colitis and is associated with increased inflammation. Inflamm Bowel Dis. 2011;17: 2299–2307. doi:10.1002/ibd.21625 19. Frenkel ES, Ribbeck K. Salivary mucins protect surfaces from colonization by cariogenic bacteria. Appl Environ Microbiol. 2015;81: 332–338. doi:10.1128/AEM.02573-14 113