Temporal and Tissue Specific Changes in Expression of Nutrient Transporters and Host Defense Peptides in Young Broilers during Salmonella and Campylobacter Infections.

Javier Shalin Garcia

Dissertation submitted to the faculty of the Virginia Polytechnic Institute and State University in partial fulfillment of the requirements for the degree of

Doctor of Philosophy In Animal and Poultry Sciences

Eric A. Wong, Committee Chair J. Allen Byrd Rami A. Dalloul Elizabeth R. Gilbert

April 27, 2017 Blacksburg, VA

Keywords: Salmonella, Campylobacter, Host defense peptides, Nutrient transporters.

Copyright 2017, Javier S. Garcia

Temporal and Tissue Specific Changes in Expression of Nutrient

Transporters and Host Defense Peptides in Young Broilers during Salmonella

and Campylobacter Infections.

Academic Abstract

Salmonella and Campylobacter are the leading causes of bacterial foodborne illness in the United

States. Commonly found in the gastrointestinal tract of poultry, Salmonella and Campylobacter may show little to no signs of infection in birds. The objective of this dissertation was to evaluate the influence on mRNA abundance of nutrient transporters and host defense peptides (HDPs) during a Salmonella or a Campylobacter challenge in young commercial broilers. Comparisons were made between non-challenged and challenged (106, 107, or 108 colony forming units of

Salmonella or Campylobacter) broilers on expression of nutrient transporters and host defense peptides in the duodenum, jejunum, ileum and cecum at various days after inoculation. During a

Salmonella challenge, changes in mRNA abundance of nutrient transporters and avian beta- defensins (AvBD) vary by day, tissue and challenge dose. ZnT1 may play an important role during a Salmonella challenge as mRNA abundance of ZnT1 significantly increased (P<0.05) by day 7 in the 108 group compared to the control. Early changes in LEAP2 mRNA abundance were observed in the 106 group than the 107 and 108 groups. However, at a later time point post challenge, a lower abundance of almost all AvBD mRNA (P<0.05) was observed in the lower gastrointestinal tract especially in the 107 and 108 groups compared to the control group, indicating that the pathogen may be influencing intestinal expression of AvBD mRNA. In Campylobacter, analyses revealed that expression of transporter 1 (ZnT1) increased (P<0.05) in the duodenum, ileum and ceca in the 106 group on day 7. An increase (P<0.05) in the expression of avian beta-defensins were observed on day 14 in the ileum and ceca in the 106 group compared to

the control group. Pathogens like Salmonella and Campylobacter may have an influence on the mRNA abundance of nutrient transporters and HDPs. Manipulation of these may ensure the survivability of these pathogens. Through sequestration of nutrients, the pathogen would have the ability to colonize the host and replicate. However, it must evade the host immune system as well. The processing of infected poultry with these pathogens may lead to foodborne illness in humans. Further research is needed to investigate possible methods to counter the influence these pathogens have on host immunity genes.

vi

Temporal and Tissue Specific Changes in Expression of Nutrient

Transporters and Host Defense Peptides in Young Broilers during Salmonella

and Campylobacter Infections

General Audience Abstract

Foodborne pathogens such as Salmonella and Campylobacter are commonly found in the gastrointestinal tract of poultry, causing little to no disease symptoms in poultry. Consumption of uncooked or mishandled meat and eggs from infected poultry could result in foodborne illness in humans. Little is known, however, about the influence of Salmonella and Campylobacter on the intestinal expression of nutrient transporters and immune genes such as host defense peptides in broiler chickens. Nutrient transporters are responsible for the transport of a variety of nutrients across the intestinal lumen to the blood. Host defense peptide are small peptides, which can be effective against invading bacteria, viruses and fungi. Therefore, the effects of Salmonella and

Campylobacter at low, medium, and high challenge doses were determined in broiler chickens. In chickens challenged with Salmonella, changes in the expression of nutrient transporters and host defense peptides were dependent on day, intestinal segment and challenge dose. The expression of the zinc transporter increased in chickens challenged with the highest Salmonella dose. In chickens challenged with Campylobacter, changes in expression of nutrient transporters and host defense peptides were also observed. Expression of the zinc transporter increased in chickens challenged with the lowest Campylobacter dose. Expression of host defense peptides increased in chickens challenged with the lowest Campylobacter challenge dose. These results indicate that cellular zinc levels as well as host defense peptides may play an important role in modulating a

Salmonella and Campylobacter infection.

vii

ACKNOWLEDGEMENTS

Dr. Wong: I appreciate all the support and guidance you have given me throughout my PhD career. Thank you for always looking out for me and for making sure I was on track to graduating. Thank you for always making time to review my data and for all your help with the writing of my dissertation.

Dr. Byrd: Thank you for all your help and guidance through my Masters and PhD programs. I appreciate all your help with the experimental design, trial runs and sample collection for the

Salmonella and Campylobacter projects.

Dr. Dalloul: Thank you for all your help with the Salmonella pre-trials. Thank you for allowing me to stop by your office whenever I had a question or just to chat.

Dr. Gilbert: Thanks for all your support and words of encouragement. I appreciate all your help with the statistical analysis for my data and willingness to answer all my questions. Thank you for always being so positive and kind.

Dr. Wood: Thanks for being an amazing teaching mentor. I am grateful for the learning opportunities you provided me with and for pushing me to excel as an instructor. Thank you providing me with feedback when I taught, for always taking time to answer my questions and for your help with the NACTA dossier. Thank you for believing in me.

Dr. Persia: Thank you for your open door policy. I knew I could always stop by your office to chat about anything. I am grateful for all the advice you have given me about academia, industry and just life in general. Thank you for providing me with opportunities to be involved with the local poultry community.

v Dr. Westfall-Rudd: Thank you for allowing me to be a GTS member. Through the GTS, I was able to develop as an educator and became more comfortable in front of the classroom. Thank you for encouraging me to pursue a career in academia.

Dr. Johnson: Thank you for your help with the NACTA dossier and for always answering my questions.

Melodie and Sparky: I couldn’t have survived grad school without y’all. Thank you for being amazing friends. Melodie, thank you for the adventures and for being my partner in crime.

Sparky, thank you for all the late runs to the store or restaurants and for rescuing me from

Melodie yum yums.

Connie: Thank for being a great friend. I’m glad I had an adventure buddy after Melodie graduated. Thanks for being my happy meal partner and for not taking yourself too seriously.

Nathaniel: Thank you for being an amazing friend. I appreciate all the food and beer adventures we have had.

Haihan: Thank you for everything. I would have not been able to do RNAscope without you. I am grateful to also call you a friend. Thank you for being my lunch buddy and for sharing your culture with me and for inviting me to dinner with your family.

Shengchen: Thank you for teaching me how to do most of the molecular techniques/protocols we do in the lab and for being patient with me when I was first learning how to do everything.

Shuai: Thank you for helping me run the statistical analysis from some of my data and for your willingness to answer all my questions.

Yugo (Coco) Hou: Thank you for being a good friend. Your positive attitude is contagious.

Pat Williams: Thank you for all your help, my projects wouldn’t have been possible without you.

vi

Denise Caldwell: Thanks for all your help with my Masters and PhD projects. Thank you for teaching me about everything I know about working with Salmonella and Campylobacter.

Claudia and Chad Brodkin: Thank you for treating me as one of your own and allowing me to live in your house when I first moved back to Virginia.

Virginia Agricultural Council: Thank you for funding my projects.

George Washington Carver Scholars Program: Thank you Dr. Grayson for your support and for always making me feel welcomed.

Graduate Teaching Scholars Program and Cohorts: Thank you for exposing me to the field of pedagogy. To my fellow GTS colleagues, thank you for sharing your teaching experiences with me. You have all inspired me to become a better educator.

Mom and Dad: Thank you for all your hard work and sacrifice. Thank you for believing in me and always pushing me to do my best.

Leonardo, Judit, and Elias: Thank you for being amazing siblings. Thank you for checking up on me, for always encouraging me to push forward and always being there when I needed y’all.

vii

TABLE OF CONTENTS

TITLE PAGE

ACADEMIC ABSTRACT i

GENERAL AUDIENCE ABSTRACT iii

ACKNOWLEDGEMENTS v

TABLE OF CONTENTS viii

LIST OF TABLES x

LIST OF FIGURES xii

ABBREVIATIONS xiii

CHAPTER 1: INTRODUCTION 1

CHAPTER 2: LITERATURE REVIEW 2 Gastrointestinal tract of chickens 2 Structures of the small intestine 2 Salmonella 3 Salmonella in the poultry industry 4 Salmonella in commercial and table eggs 5 Salmonella in broilers 7 Campylobacter 8 Campylobacter in eggs 10 Nutrient Transporters 10 Avian Immune System 12 Innate Immune System 13 Host-defense peptides 14 Objectives 16 CHAPTER 3: Temporal- and Tissue-Specific Changes in the Expression of Nutrient Transporters in Young Commercial Broilers Challenged with Various Levels of Salmonella Typhimurium 18 Abstract 18 Introduction 18 Materials and Methods 19 Results 23 Discussion 26

viii

CHAPTER 4: Temporal- and Tissue-Specific Changes in mRNA Abundance of Avian Beta-Defensins and Liver-Expressed Antimicrobial Peptide 2 in Young Broilers Challenged with Salmonella Typhimurium 37 Abstract 37 Introduction 37 Materials and Methods 38 Results 40 Discussion 42 CHAPTER 5: mRNA Expression of Nutrient Transporters, Avian Beta-Defensins and Liver-Expressed Antimicrobial Peptide 2 during a Campylobacter Challenge in One and Two Week Old Broilers 52 Abstract 52 Introduction 52 Materials and Methods 53 Results 58 Discussion 61 CHAPTER 6: In situ Hybridization of AvBD-8 and AvBD-10 in the Ileum of Campylobacter-Challenged Young Broilers 73 Abstract 73 Introduction 73 Materials and Methods 74 Results 75 Discussion 76 CHAPTER 7: EPILOGUE 83

LITERATURE CITED 84

ix

LIST OF TABLES

Table 3.1: Forward and Reverse Primers of Nutrient Transporter Genes 29

Table 3.2: Liver, Spleen and Ceca Colonization 1, 2, 5 and 7 days after Salmonella Challenge 30

Table 3.3: Expression of Nutrient Transporters on Day 1 in Young Broilers

Challenged with Salmonella Typhimurium 31

Table 3.4: Expression of Nutrient Transporters on Day 2 in Young Broilers

Challenged with Salmonella Typhimurium 32

Table 3.5: Expression of Nutrient Transporters on Day 5 in Young Broilers

Challenged with Salmonella Typhimurium 33

Table 3.6: Expression of Nutrient Transporters on Day 7 in Young Broilers

Challenged with Salmonella Typhimurium 34

Table 4.1: Forward and Reverse Primers of Host Defense Peptide Genes 46

Table 4.2: Liver, Spleen and Ceca Colonization 1, 2, 5 and 7 days after

Salmonella Challenge 47

Table 4.3: Expression of Host Defense Peptides in Young Broilers

Challenged with Salmonella Typhimurium on Day 1 48

Table 4.4: Expression of Host Defense Peptides in Young Broilers

Challenged with Salmonella Typhimurium on Day 2 49

Table 4.5: Expression of Host Defense Peptides in Young Broilers

Challenged with Salmonella Typhimurium on Day 5 50

Table 4.6: Expression of Host Defense Peptides in Young Broilers

Challenged with Salmonella Typhimurium on Day 7 51

x

Table 5.1: Forward and Reverse Primers of Nutrient Transporter Genes 64

Table 5.2: Forward and Reverse Primers of Host Defense Peptide Genes 65

Table 5.3: Campylobacter jejuni Liver and Spleen Colonization 7, 14, 21 and

28 Days after Challenge 66

Table 5.4: Day 7 mRNA Abundance of Nutrient and Sugar Transporters in

Campylobacter jejuni Challenged and Non-Challenged Young Broilers 67

Table 5.5: Day 7 mRNA Abundance of Host Defense Peptides in

Campylobacter jejuni Challenged and Non-Challenged Young Broilers 68

Table 5.6: Day 14 mRNA Abundance of Nutrient and Sugar Transporters in

Campylobacter jejuni Challenged and Non-Challenged Young Broilers 69

Table 5.7: Day 14 mRNA Abundance of Host Defense Peptides in

Campylobacter jejuni Challenged and Non-Challenged Young Broilers 70

xi

LIST OF FIGURES

Figure 3.1: Treatment layout. 35

Figure 3.2: Salmonella Cecal Recovery in Young Broilers 1, 2, 3, 4, 5, 6 and 7

Days after Challenge. 36

Figure 5.1: Treatment layout. 71

Figure 5.2: Campylobacter Cecal Recovery in Young Broilers 7, 14, 21, and 28

Days after Challenge. 72

Figure 6.1: Expression profile of AvBD8 and AvBD10 in the ileum of control and

Campylobacter-challenged broilers at Day 7. 78

Figure 6.2: Expression profile of AvBD8 and AvBD10 in the ileum of control and

Campylobacter-challenged broilers at Day 7. 79

Figure 6.3: Expression profile of AvBD8 and AvBD10 in the ileum of control and

Campylobacter-challenged broilers at Day 14. 80

Figure 6.4: Expression profile of AvBD8 and AvBD10 in the ileum of control and

Campylobacter-challenged broilers at Day 14. 81

xii

ABBREVIATIONS

AvBD Avian Beta-Defensins bo,+AT b(o,+)-type Amino Acid Transporter

CATH Cathelicidins

CAT2 Cationic Amino Acid Transporter-2

CDC Centers for Disease Control

CFU Colony Forming Units

EAAT3 Excitatory Amino Acid Transporter 3

FSIS Food Safety Inspection Service

GLUT2 2

HDPs Host Defense Peptides

LAT1 L-type Amino Acid Transporter-1

LEAP2 Liver-expressed antimicrobial peptide 2

NA Nalidixic acid

N/N Novobiocin and Nalidixic acid

NO Novobiocin

RV Rappaport-Vassiliadis Enrichment Broth

SGLT1 Sodium glucose transporter-1

XTL4 Xylose-Lysine-Tergitol 4

ZnT1 Zinc Transporter 1

xiii

Chapter 1

Introduction

In 2012, the commercial broiler industry grossed $24.8 billion in the United States, which was an 8% increase from 2011 (Grant et al., 2016). Interestingly, in the U.S. annual hospital costs associated with Salmonella were $2.71 billion in 2010 (Andino and Hanning, 2015) while hospital costs associated with Campylobacter were $1.7 billion dollars in 2011 (Hoofmann et al.,

2012). Salmonella and Campylobacter are bacteria commonly found in the gastrointestinal tract of poultry. These pathogens may show little to no clinical signs of infection in poultry.

However, when undercooked or mishandled contaminated poultry meat or eggs are consumed by humans, these pathogens may cause foodborne illness.

The literature review of this dissertation will focus on Salmonella and Campylobacter in poultry flocks as well as the function of nutrient transporters and host defense peptides (HDPs) in chickens. Nutrient transporters play an important role in the transport of nutrients from the intestinal lumen through the absorptive enterocytes into the blood stream (Hediger et al., 2004).

Host defense peptides are a group of small cationic peptides, which can be effective against bacteria, viruses, and fungi (Cuperus et al., 2013). Nutrient transporter and HDPs are essential in ensuring the health and well-being of poultry.

Manipulation of nutrient transporters and HDPs could be essential in the survivability of

Salmonella and Campylobacter within poultry. The objective of this dissertation is to investigate the influence that Salmonella and Campylobacter have on the expression of nutrient transporters and HDPs over various time points in broilers.

1 Chapter 2

Literature Review

Gastrointestinal tract of chickens

The small intestine is a major site for digestion and nutrient absorption. The proximal segment is the duodenum, which is also referred to as the duodenal loop, and is the main site of lipid digestion (Sklan et al., 1975). The next segment is known as the jejunum and is generally defined as the area posterior to the duodenal loop up to Meckel’s diverticulum. The jejunum plays an important role as the major site for nutrient digestion and absorption (Riesenfeld et al., 1980;

Noy et al., 1995). The distal segment of the small intestine is the ileum and is defined as the segment from Meckel’s diverticulum to the ileo-cecal junction. At the ileo-cecal junction, a pair of blind-ended sacs known as the ceca branch off. The ceca are the site for microbial fermentation with some nutrient, water, and nitrogenous component absorption abilities (Clench, 1999). The lower gastrointestinal tract of poultry is a major site of colonization for many bacteria especially those of human health concerns such as Salmonella and Campylobacter.

Morphological structures of the small intestine

The mucosa of the small intestine features thousands of macroscopic finger-like projections known as villi, which project into the intestinal lumen increasing the surface area available for the absorption of nutrients (Crawley et al., 2014). Each villus is lined by a simple epithelium consisting of a monolayer of epithelial cells, and include goblet cells, enteroendocrine cells and enterocytes. In mammals, paneth cells are also found; however, paneth cells have yet to be fully described in chickens. These differentiated cells originate from the multipotent stem cells located

2 in the intestinal crypts of Lieberkühn (Loeffler et al., 1993). Newly developed epithelial cells move up along the villus from the bottom of the crypt to the tip of the villus, from where they are extruded after 3-4 days in chickens (Imondi and Bird, 1965; Uni et al., 2000) and 4-5 days in rodents (Mowat and Agace, 2014). Absorptive enterocytes make up about 80% of all small intestinal epithelial cells and are hyperpolarized cells responsible for the uptake of macromolecules. On the apical surface of the enterocytes are rigid, closely placed microvilli. The tips of the microvilli contain large, negatively charged glycoproteins that form a brush border glycocalyx. The glycocalyx provides a highly degradative microenvironment that promotes the digestion and absorption of nutrients (Snoeck et al., 2005).

Salmonella

In the United States, Salmonella is the leading cause of bacterial foodborne illness. In

2010, the annual hospitalization costs associated with Salmonella infections were estimated to be

$2.71 billion for 1.4 million cases (Andino and Hanning, 2015). Salmonella is a Gram-negative, facultative anaerobic, rod-shaped bacterium. It belongs to the Enterobacteriaceae family and is an important pathogen that affects both humans and animals. The genus Salmonella is divided into two species, S. bongori and S. enterica. S. enterica is further divided into six subspecies that contain more than 2,000 known serovars (Tinall et al., 2005). In humans, Salmonella causes salmonellosis, a self-limiting illness that results in diarrhea, fever and abdominal cramps with most persons recovering without treatment. In infants, the elderly and immunocompromised individuals, the illness could be severe and may require hospitalization.

The molecular mechanisms relating to Salmonella colonization, pathogenesis and transmission have been described mainly in rodents with little being known about these

3 mechanisms in chickens (Velge et al., 2012). Infection of the host is initiated by the consumption of contaminated food or water, where the pathogen must travel from the stomach to the intestines.

Once it has reached the lower gastrointestinal tract, it must adhere and enter the intestinal epithelial cells. Through invasion of either the enterocytes or microfold cells at the apical side, the pathogen travels through the intestinal wall via transcytosis to the basolateral side and exits into the intestinal space of the lamina propria (Clark et al., 1994; Muller et al., 2012). Within the lamina propria, the pathogen can be taken up by different phagocytes and is distributed rapidly through the blood stream. The type III secretion system (T3SS-1) plays an essential role in the bacterial invasion of host cells. The pathogen binds to the host cell surface and injects effectors into the host cell. These effectors allow Salmonella to manipulate the cytoskeletal machinery and induce the host cell to engulf the pathogen as a Salmonella-containing vacuole (Velge et al., 2012). Once inside the host cell, the pathogen will use a second type III secretion system (T3SS-2) to transport additional effectors, allowing for bacterial survival and replication (Malik-Kale et al., 2011). However, in chickens, the role that the type III secretion systems play during a Salmonella infection still remains unclear.

Salmonella in the poultry industry

S. Gallinarium and S. Pullorum cause the systemic diseases of poultry, fowl typhoid and pullorum diseases, respectively (Barrow, 2007). These diseases are often accompanied by high levels of morbidity and mortality, which devastated the early days of the poultry industry. The

National Poultry Improvement Plan was designed to reduce the level of S. Gallinarium and S.

Pullorum infections in the United States. By the mid-1970s, both serovars were eliminated from commercial poultry flocks in the U.S. Consequently, with the eradication of S. Gallinarium and

4

S. Pullorum, other serovars such as S. Enteritidis and S. Typhimurium, which are linked to human foodborne illness, were able to fill the ecological niche (Baumer et al., 2000; Foley et al., 2008).

Unlike S. Gallinarium and S. Pullorum serovars, serovars of human health concern show little or no signs of disease in poultry. These serovars do, however, cause inflammatory response leading to the expression of cytokines and chemokines (Withanage et al., 2004; Setta et al., 2012; Matulova et al., 2013; Wigley et al., 2014). This reaction causes an influx of heterophils and monocytic phagocytes to the small intestine, leading to inflammation and damage to the gut. Although there is a considerable immune response, it is relatively short lived, allowing Salmonella to persist in the gastrointestinal tract of the bird.

Salmonella in commercial and table eggs

Eggs have been linked to various food outbreaks, mainly due to the consumption of undercooked and/or raw eggs or cross contamination during cooking with eggs. Salmonella

Enteriditis is the most common serovar linked to egg-related outbreaks, although other serovars such as S. Typhimurium, S. Hadar and S. Heidelberg have also been linked to egg-related outbreaks

(Foley et al., 2011). The contamination of eggs may occur via two routes of transmission: vertical or horizontal. In vertical transmission, Salmonella originates from an infected hen, which transmits the Salmonella into the egg during egg development. In horizontal transmission, the egg becomes infected after it has been laid and is exposed to a contaminated environment (Cox et al.,

2000).

Salmonella-contaminated eggs can be produced by inoculating hens with Salmonella.

Laying hens that were orally challenged with one million colony forming units (cfu) of Salmonella were found to have infected ovarian follicles and oviducts. These hens, however, showed no

5 clinical signs of infection and Salmonella was not detected in the feces (Timoney et al., 1989).

These results suggested that infected hens may be present within flocks and may spread the pathogen without being detected. Keller et al. (1995) found that infection of the lower oviduct was crucial to the production of contaminated eggs. They observed that during egg development, hens with infected ovarian follicles gave rise to infected eggs. As the egg travels down the oviduct, the addition of the albumen, which has antibacterial compounds, may reduce Salmonella levels.

However, serovars such as S. Enteritidis have developed mechanisms to survive the harsh environment found in the albumen (Baron et al., 2016). As the egg enters the lower oviduct, the egg may become re-contaminated with the addition of the inner and outer shell membranes.

Eggs may also be contaminated after oviposition if they become exposed to an environment contaminated with Salmonella. The pathogen can be carried on the shell of the egg or it may also be able to penetrate the eggshell. Several studies have shown that Salmonella has the ability to penetrate through egg structures such as the cuticle, the inner and outer shell membranes; and once inside the egg, it can proliferate within the egg contents (Cox et al., 2000). The quality of an egg may also be a major determinant of bacterial penetration. Egg quality can be determined by objective measures such as conductance, specific gravity and shell strength. Conductance is a measure of eggshell porosity, which could reflect the ability of the eggshell to allow the transfer of water vapor and gases. Specific gravity can be used to determine eggshell thickness since the two are positively correlated and the destruction of the egg is not necessary. Sauter and Patersen

(1974) found that eggs with low specific gravity had a higher chance of being penetrated by

Salmonella. Once the pathogen is able to penetrate the eggshell, it can infect the egg content or infect the developing embryo.

6

The ability of Salmonella to contaminate eggs has made it a threat to both producers and consumers. Infected table eggs could lead to foodborne illnesses if not properly handled or cooked prior to consumption. Infected hatching eggs could not only give rise to chicks infected with

Salmonella but in turn, could lead to the infection of other chicks within the same hatching unit.

Once placed in the grow-out farm, these infected chicks could spread Salmonella to other chicks within the flock. The problem may persist to the processing plant where cross-contamination of carcasses may occur.

Salmonella in broilers

Although eggs are the main source of poultry-related foodborne illness outbreaks, poultry meat has also been linked to cases of salmonellosis (Skarp et al., 2016). Broilers may become infected with Salmonella during any stage of production. At the hatchery, eggs that have been contaminated with Salmonella could lead to cross contamination of other eggs at the hatchery.

Broiler breeder and broiler hatcheries have been found to have a high occurrence of contamination of Salmonella. Salmonella was isolated from 71% of egg fragments, 80% of chick conveyor belts swab samples, and 74% of sample pads placed under newly hatched chicks (Cox et el., 1990; Cox et al., 1991; Bailey et al., 1994). Salmonella-positive eggs may also lead to the contamination of other eggs within the hatching cabinet. When the chicks hatch, Salmonella can be spread throughout the cabinet via fan forced-air. Cason et al. (1994) reported that >80% of chicks above or below the Salmonella-positive eggs tested positive for the strain used to infect the eggs. Once chicks have been infected at the hatchery, the infection could be easily spread within the flock at the grow-out farm.

7

A wide range of environmental and management-based factors such as temperature, humidity, pre- and post- harvest interventions have been reported to affect the appearance and disappearance of Salmonella in commercial poultry flocks. Bailey et al. (2001) reported that higher recovery rates were seen during the fall season. They were also able to recover significant rates of Salmonella in the dirt located at the entrance of poultry houses and in litter samples taken from boot swabs. Controlling Salmonella at the various stages of production could be essential to reducing Salmonella in flocks.

Campylobacter

In 2011, the hospitalization costs due to campylobacteriosis was about $1.7 billion in the

United States. It was the second most common cause of bacterial foodborne illness in the U.S.

(Hoffmann et al., 2012). Campylobacter is a spiral shaped, Gram-negative, microaerophilic bacterium commonly found in the gastrointestinal tract of poultry (Bolton, 2015). Generally, it is regarded as a fragile bacterium that is sensitive to environmental stressors such as heating, drying and exposure to oxygen. However, Campylobacter is able to thrive in the hostile gut environment of poultry, which can contaminate carcasses during slaughter and processing (Humphrey et al.,

2007). This suggests that the bacterium has adaptive responses to different environmental stressors to ensure survival. In 2016, updated performance standards were announced by the United States

Department of Agriculture (USDA) with the intention of reducing Salmonella and Campylobacter in raw chicken parts and in not-ready-to-eat ground chicken and turkey products (FSIS, 2016).

Before these standards were introduced, poultry products were regularly tested for the presence of

Salmonella, but not for the presence of Campylobacter.

8

In humans, a Campylobacter infection may cause diarrhea, cramping, abdominal pain, and fever. In a few cases, the infection may lead to Guillain-Barré syndrome, which may cause paralysis or reactive arthritis (Moore et al., 2005). In order to infect humans, the pathogen must bypass the mechanical and immunological barriers of the gastrointestinal tract. The motility, its corkscrew shape and the relatively short O-sidechain of its lipooligosaccharides, allow it to pass through the mucus layer and adhere to the enterocytes (Young et al., 2007; Epps et al., 2013).

Through use of microtubule polymerization, the pathogen is internalized by the host cell in the form of a Campylobacter-containing vacuole. Interaction between the pathogen and the host epithelial cells induces the production of interleukin (IL)-8, which recruits dendritic cells, macrophages and neutrophils to the site of infection. Although the mechanisms for controlling a

Campylobacter-infection are unclear, an adaptive immune response may be responsible for controlling the infection (Panigrahi et al., 1992; Guerry et al., 2000).

In poultry, birds may show little to no signs of being infected. Interestingly, Byrne et al.

(2007) found that chicken intestinal mucus attenuated the virulence of Campylobacter compared to human intestinal mucus. Campylobacter does not elicit intestinal inflammation in the bird nor does it demonstrate cellular attachment or invasion of the intestines in poultry (Young et al., 2007).

Vaezirad et al. (2016) reported that an increase in IL-6, IL-8 and iNOS transcripts, which are responsible for inducing a pro-inflammatory response, were observed in broiler chickens challenged with Campylobacter. This may indicate that the immune response of the bird may be able to sense the pathogen. However, it is unknown how Campylobacter fails to induce an immune response similar to that observed in humans. In vitro, HDPs have been shown to be very effective against Campylobacter (van Dijk, et al., 2012). Conversely, the pathogen has the ability to use its lipopolysaccharide structures to protect its membrane from HDPs and is also able to suppress local

9 and systemic expression of HDPs at the transcriptional level. Campylobacter has developed mechanisms to help it survive and thrive in the intestines of birds.

Campylobacter in eggs

Similar to Salmonella, Campylobacter can be vertically transmitted from broiler breeders to their eggs. Campylobacter has been detected in the reproductive tracts of both hens and roosters, which may lead to the contamination of eggs during egg development (Buhr et al., 2002; Hiett et al., 2002, 2003; Cox et al., 2005). Campylobacter has a limited ability to penetrate the eggshell and survivability of Campylobacter on the eggshell surface and in the egg content is relatively low

(Sahin et al., 2003; Fonseca et al., 2014). Byrd et al. (2007) tested tray liners from commercial hatcheries and were able to recover low levels of Campylobacter. These infected chicks could be actively shedding Campylobacter, which could infect other chicks during production.

Nutrient transporters

Nutrients obtained from the consumption of feedstuff are absorbed in the small intestine.

Through the process of digestion, nutrients in feedstuffs are broken down from complex structures into simple structure such as small peptides, amino acids and monosaccharides. These molecules are then absorbed by the enterocytes that line the villus of the small intestine. The movement of these essential nutrients are regulated by transporters. Transporters are responsible for allowing small molecules to enter or exit cells and based on their function, can be divided into passive or active mechanisms (Heiger et al., 2004). In passive transport, solutes are able to diffuse across the membranes down an electrochemical gradient. In facilitated diffusion, which is a type of passive transport, through the use of a transport , nutrients move down the concentration gradient

10 of the protein without the use of ATP. In active transport, ions or small molecules move through a membrane against its concentration gradient with the use of ATP.

Nutrient transporters are part of the solute carrier (SLC) family, which is the largest family of membrane transport in both mammalian and avian species. In the , the SLC gene family contains 395 genes that are organized into 52 families (Hediger et al., 2013). SLCs are responsible for regulating the transport of a variety of compounds such as inorganic ions, amino acids, sugars, and fatty acids across cell membranes (Schlessinger et al.,

2010). In the intestinal enterocytes, transporters are located on both the brush border and basolateral membranes.

Previous research conducted in our lab has investigated the influence of Eimeria on the expression of nutrient transporters. The downregulation of the nutrient transporters EAAT3, bo,+AT, CAT2, LAT1, GLUT2, and ZnT1 was observed during an Eimeria infection (Paris et al.,

2013; Su at el., 2014; 2015; Yin et al., 2015). Although Eimeria is an intestinal protozoan, it infects the enterocyte intracellularly, similar to Salmonella. Based on these results, the nutrient transporters below were selected for analysis in the Salmonella and Campylobacter studies of this dissertation.

Brush border membrane transporters are responsible for transporting molecules into or out of the enterocyte from or to the intestinal lumen. Excitatory amino acid transporter 3 (EAAT3,

SLC1A1) is responsible for the transport of anionic amino acids such as glutamate and aspartate into the enterocyte. This transporter is dependent on the inward Na+ electrochemical potential and the outward K+ potential so that glutamate can be transported against the concentration gradient into the cell (Kanai and Hediger, 2004). Glutamate is essential because it serves as the major fuel source for the enterocyte. The transporter bo,+AT (SLC7A9) is responsible for the transport of

11 cystine and other cationic amino acids into the enterocyte and an efflux of neutral amino acids out of the enterocyte (Palacin and Kanai, 2004). The sodium-dependent glucose transporter 1

(SGLT1, SLC5A1) is a uniporter and is the primary facilitator for the uptake of glucose in the small intestine. It pumps in one glucose molecule along with 2 Na+ ions into the enterocyte

(Hediger and Rhoads, 1994).

Basolateral membrane transporters are responsible for transporting molecules into or out of the enterocyte from or to the bloodstream. Cationic amino acid transporter 2 (CAT2, SLC7A2) is a low-affinity, high capacity transporter of cationic amino acids (Verrey et al., 2004). L-type amino acid transporter 1 (LAT1, SLC7A5) is a Na+-independent transporter responsible for the transport of neutral amino acids with large branched or aromatic side chains (Verrey et al., 2004).

Glucose transporter 2 (GLUT2, SLC2A2) is mainly found in the basolateral membrane of the enterocyte, although in some cases it may move to the brush border membrane (Mithieux, 2005).

This transporter is Na+-independent and is responsible for the low affinity transport of glucose, galactose, mannose, and fructose and high-affinity transport of glucosamine (Uldry and Thorens,

2004). Zinc transporter 1 (ZnT1, SLC22A18) is responsible for the efflux of zinc out of the enterocyte (Tako et al., 2005). Zinc is an essential mineral that is required for various cellular functions such as DNA synthesis, enzymatic reactions and gene expression (Gielda et al., 2012).

Nutrient transporters play a crucial role in the influx and efflux of nutrients through the small intestine.

Avian Immune System

Similar to other classes of animals, birds have nonspecific and specific immune responses to potential threats such as bacteria, viruses, parasites and other antigenic materials (Korver, 2006).

12

The two arms of the immune system are the innate (nonspecific) and acquired (specific) immunity.

When a pathogen is recognized, it is initially processed by the innate immune system. If the innate immune system requires further assistance, then the acquired immune system is activated (Erf et al., 2004). This dissertation will primarily focus on the innate immune system, specifically host defense peptides, however, both systems are essential to the health and survival of the host

Innate Immune System

The avian innate immune system is comprised of a variety of defense mechanisms responsible for protecting the host from invading pathogens (Wigley, 2013). It is known as a non- specific, barrier immune function; however, it is effective in combating invading pathogens and is essential in driving the necessary response from the adaptive immune system for fighting off an infection (Kaiser, 2010). The innate immune system is made up of physical barriers such as the skin and mucosal layer as well as molecular and cellular defenses such as macrophages, heterophils, natural killer cells and host defense peptides (Kaiser, 2010). Macrophages originate from stem cells found in the bone marrow and play a significant role in phagocytosis, the production of cytokines and in antigen presentation to T and B cells (Quershi, 2003; Korver, 2006).

Heterophils are polymorphonuclear cells responsible for the phagocytosis of invading pathogens

(Kaiser, 2010). Usually the first cell type to respond to an area of infection, heterophils are considered the main effector cells of the innate immune system. After the invading pathogen has been internalized, the heterophil kills the pathogen with a mechanism known as the respiratory burst. Dendritic cells are responsible for capturing and processing antigens and are able to express a high level of the major histocompatibility complex. Mature dendritic cells can also interact with

13 macrophages to stimulate the release of cytokines and with B cells to induce the production of antibodies (Korver, 2006).

Host-defense peptides

Host-defense peptides, formerly known as antimicrobial peptides, are a diverse group of small peptides found in most classes of life (Leher and Ganz, 2002; Cuperus et al., 2013). HDPs are generally 10-50 amino acids long and are enriched in hydrophobic and cationic amino acid residues. These peptides are part of the innate immune system, which is the first line of defense against a pathogen. The majority of the HDPs are synthesized by the host’s phagocytic and mucosal epithelial cells. Mature HDPs can be active against Gram-negative and Gram-positive bacteria but they can be active against fungi and viruses (Zhang and Sunkara, 2014). Host-defense peptides have various modes of action. However, the most common is killing bacteria through physical electrostatic interactions, which cause disruptions in the bacterial cell membrane. There are three classes of host-defense peptides found in avians: β-defensins, cathelicidins, and liver- expressed antimicrobial peptide 2 (LEAP2). HDPs also have the ability to recruit and activate immune cells, which can be essential in fighting off pathogens (Cuperus et al., 2013).

Avian β-defensins (AvBDs) were the first to be discovered in chickens and currently 14

AvBDs have been described (Evans et al., 1994). Sequencing of all 14 AvBDs revealed that the

N-terminal signal peptide regions are highly conserved. Most of the other residues are very diverse among AvBDs (Zhang and Sunkara, 2014). AvBD genes are made up of four exons, except for

AvBD12 in which the last two exons have fused (Xiao et al., 2004). AvBDs are synthesized as inactive prepropeptides, which consist of a short signal peptide, a propeptide and the mature peptide. Transcriptional analysis of AvBD genes have found that AvBD1, 2, and 4 to 7 are mainly

14 expressed in the bone marrow, while AvBD3, and 8 to 14 are expressed by epithelial cells (Lee et al., 2016). However, AvBDs expressed in either the bone marrow or epithelial cells can be expressed in other tissues. Hong et al. (2002) found tissue-specific expression of AvBDs in broilers with necrotic enteritis, which is caused by Clostridium perfringens. In the intestines,

AvBD8, AvBD10 and AvBD13 were all upregulated during a Clostridium infection in two commercial broiler chicken lines. AvBDs may play an important role in controlling pathogenic infection in the small intestine.

Presently, four avian cathelicidins (CATH) have been discovered in chickens: CATH1,

CATH2, CATH3 and CATHB1 (Zanetti, 2005). CATH1 and CATH2 mRNA are predominantly expressed in the bone marrow. They have also been reported to be expressed in other tissues.

CATH1 mRNA is moderately expressed in the gizzard, small intestine and large intestine. CATH2 mRNA expression has been shown to be high in liver tissue and moderate in the cecal tonsils.

Tissue expression of CATH3 has yet to be determined; however, it is speculated to have similar tissue expression as CATH1 and CATH2. CATHB1 expression has been predominantly found in the bursa of Fabricius (Lee et al., 2016). CATH2 was found to be highly effective against

Campylobacter; however, the bacterium has developed mechanisms to suppress intestinal CATH2 expression (van Dijk et al., 2012). CATH may be effective in controlling bacterial infections; however, some bacteria have developed mechanisms to survive interactions with CATH.

Liver-expressed antimicrobial peptide 2 (LEAP2) was first observed in human blood; however, it was later found to be expressed in avians as well (Krause et al., 2003; Lynn et al.,

2004). In chickens, LEAP2 may be expressed as part of the epithelial innate defense system (Lynn et al., 2004; Townes et al., 2004) and may function to prevent pathogens from interacting with epithelial surfaces and potentially invading tissues. Townes et al. (2004) found that in 4-day-old

15 chicks, which had been orally inoculated with Salmonella, there was a significant upregulation of

LEAP2 mRNA in the small intestine and liver. It was later reported that in vitro, LEAP2 is able to affect the permeability of the bacterial outer membrane, leading to cell death (Townes et al.,

2009). Host defense peptides are essential since they are the first response to pathogens as part of the innate immune system.

In summary, Salmonella and Campylobacter are of major concern to the commercial poultry industry due to the threat they pose to consumers. Infected birds may carry these pathogens without showing any sign of infection, which may lead to spread of these pathogens to other birds within a flock. Nutrient transporters are important in the absorption and circulation of nutrients from the intestinal lumen to the blood. Manipulation of these transporters could be beneficial to the colonization and proliferation of these pathogens. Host defense peptides play an important role as part of the innate immune system. Developing mechanisms to avoid HDPs could be essential for these pathogens. Understanding how these pathogens influence the gut environment of the bird could be beneficial to the development of new prevention interventions.

Objectives

Changes in expression of nutrient transporters and HDPs have been partially investigated in Salmonella- and Campylobacter-infected broiler chickens. In this dissertation, the mRNA abundance of amino acid transporters (bo,+AT, CAT2, EAAT3, and LAT1), sugar transporters

(GLUT2 and SGLT1), a mineral transporter (ZnT1), and immune factors (AvBD1, AvBD6,

AvBD8, AvBD10, AvBD11, AvBD12, AvBD13 and LEAP2) were examined in the duodenum, jejunum, ileum and cecum of Salmonella-and Campylobacter-challenged broiler chickens. The objective of the first experiment (Chapters 3&4) was to examine the influence of Salmonella

16 dosage level on the mRNA abundance of nutrient transporters and HDPs over a seven-day infection period. The objective of the second experiment (Chapters 5&6) was to examine the influence of Campylobacter dosage level on the mRNA abundance of nutrient transporters and

HDPs over a two-week infection period. Samples from this study were also used to localize the expression of AvBD8 and AvBD10 to cells in the ileum of infected and non-infected broilers.

17

Chapter 3

Temporal- and Tissue-Specific Changes in the Expression of Nutrient

Transporters in Young Broilers Challenged with Various Levels of Salmonella

Typhimurium.

Abstract

Salmonella is a zoonotic pathogen commonly found in poultry that is of great concern to human health. Once Salmonella is able to colonize and proliferate in the gastrointestinal tract of birds, it can be actively shed via the feces, which can easily spread within a poultry house. However, little is known about its influence on the gut environment and how it is able to thrive. The objective of this study was to examine the influence of Salmonella at three challenge levels (106, 107, 108 colony forming units (cfu)) on the expression of nutrient transporters in the small intestine of young broilers as measured by qPCR. An upregulation of the nutrient transporters EAAT3, ZnT1, and GLUT2 was seen as early as day 1 in birds infected with 106 cfu of Salmonella Typhimurium.

However, by day 7, an increased expression of ZnT1 was observed in the 108 cfu-challenged group.

This indicates that zinc may play an important role during a Salmonella invasion.

Introduction

In the United States, Salmonella is the leading cause of bacterial foodborne illness. In humans, non-typhoidal Salmonella may cause abdominal pain, fever, and diarrhea (CDC, 2016).

However, in most cases, the illness is self-limiting and medical attention may not be necessary.

Salmonella serovars that cause illness in humans, usually do not cause illness in poultry. In most

18 cases, birds may show little to no signs of being infected with Salmonella. Once Salmonella enters the gastrointestinal tract of poultry, it can colonize the lower portion of the small intestine and lead to shedding of the pathogen via the feces.

Manipulation of nutrient transporters may be essential for the invasion of foodborne pathogens such as Salmonella. Nutrient transporters are located on the epithelial cells that line the villi and are responsible for the influx and efflux of nutrients to the absorptive enterocytes of the small intestine (Hediger et al., 2004). Campylobacter, another foodborne pathogen, has the ability to influence the expression of nutrient transporters in the gut of chickens. Awad et al. (2014) suggested that this pathogen may be able to change the gut environment in order to favor the replication and colonization of Campylobacter. However, little is known about its effects on the absorption of nutrients in chickens during Salmonella colonization. The objective of this study was to examine changes in the mRNA expression of nutrient transporters in the GI tracts of young broiler chicks following a Salmonella challenge.

Materials and Methods

Animals

Two-hundred day-of-hatch Ross 308 cross broiler chicks were obtained from a local commercial hatchery. The paper chick tray liners were taken for Salmonella testing. Upon arrival, chicks were orally challenged with Salmonella and then placed into pens (Figure 3.1). The trial was run in one disinfected, environmentally-controlled room with four 1.8 x 1.4 m floor pens with fresh pine shavings. The chicks were provided with heat lamps to ensure they received adequate heat for the first few days. All floor pens were equipped with nipple drinkers and feed trays, which were monitored on a daily basis. All birds received a non-medicated corn-soybean meal starter

19 diet obtained from the Texas A&M University Poultry Research Center. Diet was formulated to meet or exceed the National Research Council nutrient requirements for poultry (NRC, 1994).

Chicks were provided feed and water ad libitum from time of placement until termination. This study was approved by the Southern Plains Agricultural Research Center Animal Care and Use

Committee and conducted at the Food and Feed Safety Research Laboratory (USDA Agricultural

Research Service, College Station, TX)

Tray Liner Evaluation

Using aseptic techniques, each individual tray liner was placed into a gallon-size bag (S.C.

Johnson & Johnson, Racine, WI) and 150 mL of Buffered Peptone Water (BPW) (Buven et al.,

1984) was added into each bag. Bags were then thoroughly massaged for 60 seconds to ensure there was proper contact between the paper tray-liner and the BPW. The peptone-tray liner samples were then incubated at 37°C for 24 hours. After 24 hours, one mL of liquid was transferred into tubes containing 9 mLs of Rappaport-Vassiliadis enrichment broth (Becton

Dickinson, Franklin Lake, NJ) and incubated for 24 hours at 42°C. Following enrichment, each sample was streaked onto Xylose Lysine Tergitol-4 (XTL-4) (Becton Dickinson) agar with novobiocin (25 µg/mL) and incubated for 24 hours at 37°C. Colonies were analyzed for colony morphology.

Challenge

To make the challenge, a strain of Salmonella Typhimurium (ST) with novobiocin (NO) and nalidixic acid (NA) resistance was used (Corrier et al., 1990). The media which was used to culture the resistant strain contained 25 µg/mL of NO and 20 µg/mL of NA. The challenge

20 inoculum was prepared using an overnight culture, which had been transferred 3 times within three days in tryptic soy broth (Becton Dickinson). The culture was serially diluted in sterile - buffered saline to make three different challenge dosages that were approximately 106, 107, 108 cfu per mL. The optical density of the cell dilution was measured with a spectrophotometer at 625 nm and the number of cells in the inoculum was determined using a standard curve (Byrd et al.,

2001). The viable cell concentrations of the three challenge inoculums were confirmed by colony units on XLT-4 plates. Plates were incubated for 24 hours at 37°C and expressed as log10 to determine cfu/mL. On day of hatch, the control group received 0.5 mL of sterile phosphate- buffered saline by oral gavage and occurred before the challenge groups. Challenge was administered by crop gavage of 0.5 mL to each bird within a challenge group. Challenge was administered with the lowest cfu dosage group occurring first and ending with the highest cfu challenge dosage group.

Tissue collection

Before challenge, six day of hatch chicks were randomly selected, euthanized by cervical dislocation and subjected to necropsy immediately after euthanasia for tissue collection. Chicks were observed daily for 7 days after the challenge. From day 1 to day 7, six chicks from each of the four groups were randomly selected, euthanized by cervical dislocation and subjected to necropsy immediately after euthanasia for tissue collection. To prevent contamination, control group was sampled first followed by the challenged groups 106, 107, and 108 cfu. A cecum from each bird was aseptically removed and cecal contents were aseptically collected, weighed and serially diluted at 1:10, 1:100, 1:1000, and 1: 10,000 in 9 mL Butterfield’s solution tubes and spread onto XLT-4 plates with novobiocin (25 µg/mL) and nalidixic acid (20 µg/mL) (N/N). XLT-

21

4 plates were incubated at 37°C for 24 hours and colonies were counted and analyzed for colony morphology. A piece of tissue from the remaining ceca was taken and placed into RNAlater®

(Qiagen, Germantown, MD) solution, the remaining tissue was then placed into 9 mLs of

Rappaport-Vassiliadis enrichment broth, incubated at 42°C for 24 hours and streaked onto XLT-4

N/N plates. Plates were incubated for an additional 24 hours at 37°C and observed for the presence or absence of Salmonella. The small intestine from each bird was also aseptically removed and a

1.25-2.0 cm whole tissue sample was taken from the duodenum, jejunum and ileum, washed in phosphate buffer solution and immediately placed into collection tubes containing RNAlater® solution (Qiagen). All intestinal samples were frozen at -80°C until being shipped to Virginia

Tech.

RNA Extraction and Qualitative Real-Time PCR

Total RNA was extracted from duodenal, jejunal, ileal and cecal samples from birds (n=5) collected at days 1, 2, 5, and 7 post-challenge, following the Direct-zol RNA Miniprep (Zymo

Research Corp, Irvine, CA) protocol. Days 1, 2, 5, and 7 were selected to examine early and later responses to the Salmonella challenge and were based on results from pilot trials. The RNA concentration for each sample was quantified using a Nanodrop 1000 (Fisher Scientific, Hampton,

NH) and then diluted to 200 ng/µL. The genes analyzed include the neutral (LAT1), Na+- independent (bo,+AT), anionic (EAAT3) and cationic (CAT2) amino acid transporters, the facilitated (GLUT2) and Na+-dependent (SGLT1) sugar transporters and the zinc (ZnT1) transporter. The cDNA was synthesized from 500 ng of total RNA using a High-capacity cDNA reverse transcription kit (Applied Biosystems, Waltham, MA). The cDNA samples were diluted

1:30 for qPCR. Each of the qPCR wells contained 5 µL of Fast SYBR Green Master mix (Applied

22

Biosystems), 1 µL of forward primer (5 nM), 1 µL of reverse primer (5 nM), 2 µL of diethyl pyrocarbonate treated water and 1 µL of cDNA and the reactions were run in a 7500 Fast Real- time PCR instrument (Applied Biosystems). The primers for each of the genes plus beta-actin are listed in Table 3.1. These genes were selected based on previous research done in our lab, which evaluated expression of nutrient transporters during an Eimeria challenge (Paris et al., 2013; Su et al., 2014, 2015; Yin et al., 2015).

Fold change was calculated using the ΔΔCt method (Livak and Schmittgen, 2001).

Expression of beta-actin served as the reference gene to calculate ΔCt. For each individual day, the average ΔCt of the control duodenum, jejunum, ileum and ceca was used as the calibrator to calculate ΔΔCt of corresponding treatment tissues.

Statistical Analysis

Salmonella recovery levels (cfu/g of cecal contents) were analyzed by ANOVA using the

JMP Pro 13 software (SAS Institute, Cary, NC). The model included the main effect of log10 cfu, sorted by day. Significant differences (P<0.05) were further separated using Tukey’s Test.

Changes in gene expression of nutrient transporters were analyzed by ANOVA in the JMP

Pro 13 software (SAS Institute). The model included the main effect of treatment, sorted by tissue and genes. Significant differences (P<0.05) were further separated using Tukey’s Test.

Chi-Square analysis was performed using the JMP Pro 13 software (SAS Institute) to determine significant differences (P<0.05) between groups in Salmonella colonization rate.

23

Results

From day 1 through day 7, all three treatment groups had greater levels of Salmonella

Typhimurium when compared to the control group (Figure 3.2). However, there were no differences between treatment groups during the trial. On days 1, 2, 5 and 7, all collected ceca from all challenged groups tested positive for Salmonella, which indicated colonization by

Salmonella was successful and positive liver and spleen samples indicated a systemic infection

(Table 3.2).

On day 1, there were differences (P<0.05) in mRNA abundance of EAAT3, ZnT1, GLUT2,

CAT2, and LAT1 (Table 3.3). Abundance of EAAT3 mRNA was lower in the 107 group compared to the control in the duodenum and cecum. However, in the jejunum and ileum, mRNA abundance of EAAT3 was lower in the 107 and 108 groups than the 106 group, which was not different from control. Abundance of ZnT1 mRNA was lower in the 107 and 108 groups compared to the 106 group, which was not different from control in the duodenum and ileum. In the jejunum, mRNA abundance was lower in the 107 group compared to the 106 group, which again was not different from control. Abundance of GLUT2 mRNA was lower in the 108 group and the 107 and 108 groups compared to the 106 group in the duodenum and ileum respectively. Similarly, there was no difference between the 106 group and the control. In the jejunum, GLUT2 mRNA abundance was greater in the 106 group than the control and the 107 and 108 groups. For CAT2, there was lower mRNA abundance in the 106 group than the 107 group, which was not different from control. For

LAT1, there was greater mRNA abundance in the 107 group compared to control and the 108 group.

On day 2, there were differences (P<0.05) in mRNA abundance of EAAT3, bo,+AT, ZnT1,

CAT2, LAT1 and SGLT1 (Table 3.4). The mRNA abundance of EAAT3 was lower in the 106

24 group, 107 and 108 groups and all challenged groups compared to control in the jejunum, ileum, and cecum, respectively. For bo,+AT, mRNA abundance was lower in the 106 and 108 groups, 107 and 108 groups, and all challenged groups compared to control in the duodenum, ileum, and cecum, respectively. ZnT1 mRNA abundance was only lower in the 108 group compared to control in the duodenum. There was upregulation of CAT2 mRNA in the 108 group compared to control and the 106 group in the jejunum. For LAT1, there was increase in mRNA abundance in the 108 group compared to the control and the 106 group in the ileum and upregulation in the 107 group compared to control in the cecum. SGLT1 mRNA abundance was downregulated in the 108 group compared to control in the cecum.

On day 5, there were differences (P<0.05) in mRNA abundance of EAAT3, bo,+AT, ZnT1, and CAT2 (Table 3.5). For EAAT3, there was increased mRNA abundance in the 107 and 108 groups compared to the 106 group in the duodenum, which was not different from control. The mRNA abundance of bo,+AT was lower in the 107 group compared to control in the duodenum. In the jejunum, ileum, and cecum, there was lower bo,+AT mRNA abundance for all challenged groups compared to control. For ZnT1, there was greater mRNA abundance for the 107 group compared to the control and 108 group in the jejunum. In the ileum, ZnT1 mRNA abundance was lower in the 108 group compared to the 106 group, which was not different from control. For

CAT2, the mRNA abundance was lower in the 108 group compared to control in the duodenum.

In the jejunum, mRNA abundance of CAT2 was lower in all challenged groups compared to the control and in the ileum, mRNA abundance of CAT2 was lower in the 106 and 108 groups compared to the control. CAT2 mRNA abundance was lower in the 107 and 108 groups compared to the control in the cecum.

25

On day 7, there were differences (P<0.05) in mRNA abundance of EAAT3, bo,+AT, ZnT1,

GLUT2, CAT2, and LAT1 (Table 3.6). EAAT3 mRNA abundance was greater in the 107 group than the control and 108 group in the cecum. Also in the cecum, mRNA abundance of bo,+AT was lower in the 108 group than the 107 group, which was not different from control. In the duodenum,

ZnT1 mRNA abundance was greater in the 107 and 108 groups compared to control and greater in the 107 and 108 groups compared to control and the 106 group in the jejunum. ZnT1 mRNA abundance was greater in the 108 group than the control and the 106 group in both the ileum and the cecum. GLUT2 mRNA abundance was lower in the 107 and 108 groups compared to control in the ileum and lower in the 108 group compared to control in the cecum. CAT2 mRNA abundance was greater in the 108 group than control in the cecum. For LAT1, mRNA abundance was greater in the 106 group compared to control and the 107 and 108 groups in the duodenum. In the ileum and cecum, LAT1 mRNA abundance was greater in the 106 group than the control.

Discussion

Nutrient transporters play an essential role in the absorption of nutrients from the intestinal lumen. Pathogens such as Salmonella may have an influence on the expression of nutrient transporters in order to replicate and colonize the gastrointestinal tract of chickens. In this study, changes in the expression of nutrient transporters were observed. These changes appear to be dependent on challenge dosage, day and tissue.

Changes in mRNA abundance could be divided into two phases post challenge: early (day

1 and day 2) and delayed (day 5 and day 7). Early changes were observed in the nutrient transporter

EAAT3. Abundance of EAAT3 mRNA abundance was dose-dependent. A decrease in EAAT3 mRNA abundance was consistent in the ileum and cecum of the 107 group compared to the control

26 group. The ileum and cecum are the main sites for Salmonella colonization (Meade et al., 2009).

However, little to no changes in EAAT3 mRNA abundance could be observed by day 7. EAAT3 is responsible for the transport of glutamate and aspartate. Glutamate is essential to the enterocyte serving as the main energy source for these cells. Decreasing expression of EAAT3 could essentially shut down the enterocyte, which would make the cell less favorable for Salmonella to invade.

Changes in bo,+AT mRNA abundance were observed on day 2 and day 5. Early changes in bo,+AT mRNA abundance were observed in the 107 and 108 groups, with decreases in mRNA abundance in both groups in comparison to the control. By day 5, bo,+AT mRNA abundance in the jejunum, ileum and cecum were lower in all challenge groups. The transporter bo,+AT is responsible for the transport of cystine and other cationic amino acids into the enterocyte and efflux of neutral amino acids out of the enterocyte (Palacin and Kanai, 2004). Delayed changes in CAT2 and LAT1 were also observed. Abundance of CAT2 mRNA was lower in all tissues of the 108 group in comparison to the control on day 5. Cationic amino acid transporter 2 is a low-affinity, high capacity transporter of cationic amino acids (Verrey et al., 2004). Abundance of LAT1 mRNA was greater in all tissues of the 106 group in comparison to the control on day 7. L-type amino acid transporter 1 is a Na+-independent transporter responsible for the transport of neutral amino acids with large branched or aromatic side chains (Verrey et al., 2004). Manipulation of bo,+AT, CAT2, and LAT1 could provide Salmonella with amino acids needed for bacterial and growth.

Delayed changes in the nutrient transporter GLUT2 were observed on day 5 and day 7.

The decrease in mRNA abundance of GLUT2 was consistently observed in the ileum and cecum of the 108 group. GLUT2 is responsible for the transport of glucose, galactose, mannose, and

27 fructose on the basolateral membrane of the enterocyte. The availability of glucose may be essential to the pathogen. Less of GLUT2 may lead to an increase of glucose inside the enterocyte, which may be beneficial for Salmonella growth.

ZnT1 mRNA abundance was affected throughout the trial period. Interestingly on day 1, changes in ZnT1 mRNA abundance were observed between the challenged groups; however, not with the control group. An increase in ZnT1 mRNA abundance was observed in the 106 group in the duodenum, jejunum, and ileum compared to the 107 and 108 groups. The differences in ZnT1 mRNA abundance could be influenced by the dosage given to each group. By day 7, a delayed change was observed, with an increase of ZnT1 mRNA abundance in the 107 and 108 groups in all tissues. Increased expression of ZnT1 would result in the increased efflux of zinc out of the enterocyte. Zinc is essential in resisting disease, improving wound healing and in helping maintain gut integrity (Zhang et al., 2012). The delayed efflux of zinc by the 107 and 108 groups could indicate a possible immune response to the Salmonella challenge. However, it should be noted that only one zinc transporter was analyzed in this study; zinc transporters on the brush border membrane should be looked into in the future to fully understand the role of zinc during a

Salmonella challenge. Manipulation of nutrient transporters could be essential to the survival of

Salmonella within the gastrointestinal tracts of poultry. Early changes in nutrient transporters could be a response by the enterocyte to limit the invasion of Salmonella. However, the response may be dependent on the challenge dose.

28

Table 3.1: Forward and Reverse Primers of Nutrient Transporter Genes Gene Gene full Location Function Forward/reverse primer name 7, Brush Na+-independent neutral/cystine, cationic amino CAGTAGTGAATTCTCTGAGTGTGAAGCT b0,+AT a member 9 border acid exchanger /GCAATGATTGCCACAACTACCA (SLC7A9) Cationic amino acid TGCTCGCGTTCCCAAGA CAT2 a Basolateral Transports lysine, arginine and histidine transporter-2 /GGCCCACAGTTCACCAACAG (SLC7A2) Excitatory amino acid Brush TGCTGCTTTGGATTCCAGTGT EAAT3 a Transports aspartate, glutamate and cysteine transporter 3 border /AGCAATGACTGTAGTGCAGAAGTAATATATG (SLC1A1) L type amino acid CACACTATGGGCGCATGCT LAT1 a Basolateral Transports hydrophobic amino acids transporter-1 /ATTGTGCCTGGAGGTGTTGGT (SLC7A5) Glucose Transports fructose, mannose, galactose, glucose CTCAGCCAGGTGTACTGTGCTT GLUT2 a transporter-2 Basolateral and glucosamine /CGTCATCCGCTTCAGTCTCA (SLC2A2) Sodium glucose Brush GCCATGGCCAGGGCTTA SGLT1 a Transports low concentrations of D-glucose transporter-1 border /CAATAACCTGATCTGTGCACCAGTA (SLC5A1) Zinc TCCGGGAGTAATGGAAATCTTC ZnT1b transporter-1 Basolateral Efflux of Zn2+ /AATCAGGAACAAACCTATGGGAAA (SLC22A18) GTCCACCGCAAATGCTTCTAA Beta-actin a Beta-actin Cytosol Reference gene /TGCGCATTTATGGGTTTTGTT a Primer sequence designed by Gilbert, et al., 2007 b Primer sequence designed by Paris and Wong, 2010

29

Table 3.2: Liver, Spleen and Ceca Colonization 1, 2, 5 and 7 days after Salmonella Challenge.

Day 1 Day 2 Day 5 Day 7

Salmonella culture- positive/total Salmonella culture- positive/total Salmonella culture- positive/total Salmonella culture- positive/total

Challenge Liver Spleen Ceca Liver Spleen Ceca Liver Spleen Ceca Liver Spleen Ceca

CON 0/6 0/6 0/6 0/6 0/6 0/6 0/6 0/6 0/6 0/6 0/6 0/6

L 3/6 3/6 6/6* 5/6* 3/6 6/6* 6/6* 5/6* 6/6* 6/6* 4/6* 6/6*

M 5/6* 4/6* 6/6* 6/6* 4/6* 6/6* 6/6* 6/6* 6/6* 6/6* 6/6* 6/6*

H 6/6* 2/6 6/6* 6/6* 5/6* 6/6* 6/6* 6/6* 6/6* 6/6* 6/6* 6/6*

*Values are significantly different from control (P <0.05)

30

Table 3.3: Expression of Nutrient Transporters on Day 1 in Young Broilers Challenged with Salmonella Typhimurium Tissue Challenge/Gene EAAT3 bo,+AT ZnT1 GLUT2 CAT2 LAT1 SGLT1 Duodenum CON 1.14A 1.01 1.05AB 1.14AB 1.01 1.26 1.08 L-106 1.24A 0.85 1.42 A 2.72 A 1.22 4.23 1.55 M-107 0.43B 1.06 0.51B 1.27 AB 2.64 4.08 0.93 H-108 0.59AB 0.78 0.61B 0.76 B 0.54 0.99 0.86 P-value 0.01 0.63 0.004 0.02 0.16 0.13 0.18 Pooled SEM 0.16 0.18 0.17 0.40 0.64 1.22 0.23

Jejunum CON 1.05AB 1.00 1.12AB 1.03B 1.13 1.07B 1.07 L-106 1.54A 1.29 1.22A 2.42A 1.35 2.69AB 2.69 M-107 0.53B 0.90 0.43B 0.89B 3.48 3.38A 3.37 H-108 0.67B 0.86 0.54AB 0.69B 1.46 1.53AB 1.53 P-value 0.001 0.45 0.018 0.003 0.08 0.03 0.29 Pooled SEM 0.14 0.20 0.19 0.30 0.66 0.54 0.16

Ileum CON 1.10AB 1.08 1.17AB 1.54AB 1.17 1.36 1.50 L-106 1.53A 1.57 1.76A 4.51A 3.23 1.58 2.76 M-107 0.54B 0.74 0.52B 0.76B 0.32 1.05 2.57 H-108 0.66B 0.97 0.58B 0.39B 0.03 0.95 1.69 P-value 0.002 0.12 0.003 0.01 0.67 0.42 0.15 Pooled SEM 0.16 0.23 0.22 0.80 0.40 0.63 0.27

Cecum CON 1.44A 1.20 1.14 1.01 1.12AB 1.12 1.03 L-106 0.57AB 0.50 0.78 4.84 0.63B 0.63 2.02 M-107 0.24B 0.53 0.78 0.59 2.60A 2.60 3.46 H-108 0.45AB 0.96 0.90 0.44 0.93AB 0.93 1.45 P-value 0.04 0.24 0.58 0.08 0.03 0.08 0.38 Pooled SEM 0.28 0.27 0.20 1.26 0.45 0.64 0.36 A,BMeans with no common superscript within a gene and a tissue differ significantly (P<0.05).

31

Table 3.4: Expression of Nutrient Transporters on Day 2 in Young Broilers Challenged with Salmonella Typhimurium Tissue Challenge/Gene EAAT3 bo,+AT ZnT1 GLUT2 CAT2 LAT1 SGLT1 Duodenum CON 1.02 1.02A 1.06A 1.16 1.15 1.09 1.02 L-106 0.62 0.61B 0.88AB 1.44 0.95 1.89 1.11 M-107 0.74 0.70AB 0.64AB 0.93 1.60 2.83 1.16 H-108 0.82 0.68B 0.50B 0.45 1.37 1.95 1.07 P-value 0.053 0.04 0.01 0.07 0.24 0.29 0.97 Pooled SEM 0.10 0.10 0.11 0.25 0.23 0.61 0.20

Jejunum CON 1.02A 1.01 1.26 1.04 1.06B 1.02 1.08 L-106 0.49B 0.79 1.03 1.09 0.85B 1.79 1.13 M-107 0.59AB 0.78 0.61 0.89 1.62AB 2.95 0.99 H-108 0.83AB 0.83 0.58 0.73 2.17A 2.40 0.92 P-value 0.02 0.25 0.23 0.54 0.02 0.16 0.91 Pooled SEM 0.11 0.10 0.26 0.19 0.29 0.59 0.22

Ileum CON 1.06A 2.51A 1.28 1.12 1.17 1.07B 1.04 L-106 0.57AB 1.99AB 1.09 0.66 0.68 1.35B 1.30 M-107 0.36B 1.51B 0.98 0.91 1.44 2.78AB 1.35 H-108 0.32B 1.47B 0.52 0.42 1.45 3.21A 1.25 P-value 0.02 0.02 0.32 0.08 0.28 0.006 0.88 Pooled SEM 0.16 0.22 0.29 0.19 0.30 0.43 0.29

Cecum CON 1.03A 1.14A 1.04 1.57 1.04 1.03B 1.02A L-106 0.18B 0.24B 1.12 0.68 0.59 1.69AB 0.67AB M-107 0.13B 0.20B 0.80 0.52 1.00 2.94A 0.66AB H-108 0.21B 0.31B 0.58 0.27 0.87 2.00AB 0.48B P-value <.0001 0.01 0.29 0.11 0.46 0.049 0.02 Pooled SEM 0.10 0.18 0.21 0.37 0.21 0.44 0.20 A,B,CMeans with no common superscript within a gene and a tissue differ significantly (P<0.05).

32

Table 3.5: Expression of Nutrient Transporters on Day 5 in Young Broilers Challenged with Salmonella Typhimurium Tissue Challenge/Gene EAAT3 bo,+AT ZnT1 GLUT2 CAT2 LAT1 SGLT1 Duodenum CON 1.15AB 1.11A 1.05 1.06 1.12A 1.05 1.04 L-106 0.70B 0.47AB 1.85 0.64 0.55AB 1.43 1.20 M-107 1.67A 0.39B 1.66 1.22 0.64AB 2.07 1.24 H-108 1.67A 0.67AB 1.18 1.32 0.05B 1.25 1.03 P-value 0.02 0.03 0.04* 0.20 0.004 0.22 0.90 Pooled SEM 0.23 0.16 0.21 0.23 0.17 0.35 0.26

Jejunum CON 1.08 1.06A 1.05C 1.08 1.16A 1.07 1.04 L-106 0.89 0.38B 2.23AB 0.83 0.37B 0.94 1.44 M-107 1.54 0.28B 2.30A 1.32 0.33B 1.06 1.60 H-108 1.57 0.57B 1.27BC 1.63 0.03B 1.11 1.50 P-value 0.13 0.001 0.003 0.09 0.003 0.96 0.30 Pooled SEM 0.23 0.11 0.24 0.21 0.18 0.24 0.22

Ileum CON 1.10 1.07A 1.05AB 1.20A 1.28A 1.03 1.05 L-106 0.71 0.39B 1.60A 0.22B 0.27B 0.90 1.80 M-107 1.14 0.30B 1.37AB 0.38B 0.43AB 0.83 1.68 H-108 1.08 0.41B 0.91B 0.42B 0.05B 1.10 1.43 P-value 0.37 0.001 0.01 0.005 0.01 0.63 0.52 Pooled SEM 0.19 0.12 0.37 0.17 0.23 0.16 0.37

Cecum CON 1.08 1.13A 1.13 1.14A 1.22A 1.04 1.13 L-106 0.45 0.35B 1.01 0.42B 0.41AB 2.48 0.72 M-107 0.51 0.21B 0.89 0.23B 0.28B 1.97 0.83 H-108 0.57 0.22B 0.51 0.20B 0.08B 1.76 0.35 P-value 0.054 0.001 0.12 0.002 0.01 0.09 0.21 Pooled SEM 0.16 0.15 0.18 0.16 0.22 0.37 0.25 A,B,Means with no common superscript within a gene and a tissue differ significantly (P<0.05). *Values could not be further separated using Tukey’s Test.

33

Table 3.6: Expression of Nutrient Transporters on Day 7 in Young Broilers Challenged with Salmonella Typhimurium Tissue Challenge/Gene EAAT3 bo,+AT ZnT1 GLUT2 CAT2 LAT1 SGLT1 Duodenum CON 1.01 1.01 1.03C 1.04 1.06 1.06B 1.09 L-106 1.40 1.60 1.49BC 1.43 2.35 12.19A 1.03 M-107 1.00 1.04 2.82AB 0.90 2.64 2.58B 1.20 H-108 1.39 0.98 3.67A 1.07 1.50 1.34B 1.59 P-value 0.046* 0.10 0.001 0.65 0.62 0.01 0.32 Pooled SEM 0.12 0.19 0.41 0.30 0.94 2.26 0.23

Jejunum CON 1.02 1.07 1.03B 1.02 1.07 1.22 1.04 L-106 1.23 1.59 1.88B 0.85 1.56 6.94 1.20 M-107 1.10 1.34 3.29A 0.84 1.36 4.62 1.20 H-108 1.24 1.01 3.60A 0.63 1.62 1.83 1.80 P-value 0.37 0.08 <0.0001 0.11 0.60 0.11 0.23 Pooled SEM 0.10 0.17 0.30 0.10 0.31 1.71 0.28

Ileum CON 1.07 1.70 1.04B 1.13A 1.04 1.23B 1.09 L-106 0.97 1.73 1.49B 0.76AB 1.28 12.75A 1.14 M-107 0.92 1.47 2.44AB 0.21B 2.32 6.03AB 1.03 H-108 0.81 1.20 3.11A 0.15B 1.71 3.09AB 1.21 P-value 0.87 0.36 0.004 0.01 0.11 0.049 0.81 Pooled SEM 0.22 0.23 0.21 0.20 0.36 2.80 0.16

Cecum CON 1.31B 1.42AB 1.04B 1.25A 1.02B 1.39B 1.30 L-106 1.40AB 1.80AB 1.11B 1.08AB 1.49AB 9.10A 1.43 M-107 2.86A 2.15A 1.67AB 0.44AB 1.16AB 6.39AB 1.81 H-108 0.49B 0.38B 3.06A 0.21B 1.84A 3.05AB 1.71 P-value 0.003 0.02 0.001 0.01 0.02 0.01 0.89 Pooled SEM 0.37 0.36 0.40 0.22 0.18 1.56 0.52 A,B,Means with no common superscript within a gene and a tissue differ significantly (P<0.05). *Values could not be further separated using Tukey’s Test.

34

Figure 3.1: Treatment layout*

6 L=10 H=108

Passageway

Control M=107

*Not Drawn to Scale Door

35

Figure 3.2: Salmonella Cecal Recovery in Young Broilers 1, 2, 3, 4, 5, 6 and 7 Days after Challenge 10

A 9

8 A A A A A A A A A A 7 A A A A A A A A A 6 A

5

Logcfus 10 4

3

2

1 B B B B B B B 0 1 2 3 4 5 6 7 Day

CON L M H

A,BMeans with no common superscript differ significantly CON=Control, L=106 challenge group, M= 107 challenge group, H=108 challenge group

36

Chapter 4

Temporal- and Tissue- Specific Changes in mRNA Abundance of Avian Beta-

Defensins and Liver-Expressed Antimicrobial Peptide 2 in Young Broilers

Challenged with Salmonella Typhimurium.

Abstract

Salmonella is the most common cause of bacterial foodborne illness in the United States.

The pathogen can be actively shed by birds through the feces. However, these infected birds may show little to no signs of a Salmonella infection. Salmonella has been shown to induce an immune response in poultry. Avian Beta-defensins (AvBD) and liver-expressed antimicrobial peptide 2

(LEAP2) are host defense peptides, which can be effective against bacterial infections. The objective of this experiment was to study the effects of a Salmonella challenge on the mRNA abundance of AvBD and LEAP2 over seven days. A host immune response to Salmonella challenge was initially observed. However, the response was dependent on dosage, as greater

LEAP2 mRNA abundance was observed on day 1 in the 106 group when compared to the 107 and

108 groups. At later points after challenge, a lower abundance of AvBD1, 6, 8, 10, and 12 mRNA was observed in the lower gastrointestinal tract especially in the 108 group, indicating that the pathogen may be influencing intestinal expression of AvBD mRNA.

Introduction

Salmonella is a bacterium commonly found in the digestive tract of poultry. In the United

States, it is the leading cause of bacterial foodborne illness (CDC, 2013). This microorganism infects poultry as well as humans via the oral route. It must travel down the gastrointestinal tract

37 and reach the distal ileum and cecum, where it outcompetes the resident microflora and penetrates the mucosal epithelium (Berndt et al., 2007). Salmonella colonization activates an inflammatory response leading to an innate immune response by the host.

Host defense peptides (HDPs) are a component of the innate immune system. HDPs are a diverse group of small peptides that are enriched in hydrophobic and cationic amino acid residues

(van Dijk et al., 2011). In avian species, three class of HDPs, avian beta-defensins (AvBD), cathelicidins (CATH) and liver-expressed antimicrobial peptide (LEAP2) have been described.

Salmonella infected chickens have shown an increased expression of AvBD3, AvBD10, and

AvBD12 six hours post-challenge (Meade et al., 2009). Increased expression of LEAP2 has also been seen in Salmonella infected chickens (Townes et al., 2004). Thus, Salmonella induces a host immune response, leading to an increased expression of certain HDPs.

Previous research has shown that mRNA abundance of HDPs is greater during a

Salmonella challenge; however, the pathogen is able to survive and thrive within the host, which may lead to contamination issues when these birds are processed. However, little research has been done to observe changes in AvBDs and LEAP2 mRNA expression during the gradual colonization incidence of Salmonella. The objective of this study was to examine the influence of a Salmonella challenge on the mRNA abundance of AvBD and LEAP2 in the small intestines of young broilers over a 7 day time period.

Materials and Methods

RNA Extraction and Relative qPCR

Tissue samples used in this study were those described and collected in Chapter 3. Total

RNA was extracted from duodenal, jejunal, ileal and cecal samples from birds (n=5) in each treatment group collected at days 1, 2, 5, and 7 following the Direct-zol RNA Miniprep (Zymo

38

Research Corp, Irvine, CA) protocol. The RNA concentration for each sample was quantified using a Nanodrop 1000 (Fisher Scientific, Hampton, NH) and then diluted to 200 ng/µL. The genes analyzed included LEAP2 and the following avian beta-defensins: AvBD1, AvBD6,

AvBD8, AvBD10, AvBD11, AvBD12, and AvBD13. The cDNA was synthesized from 500 ng of total RNA using a High-capacity cDNA reverse transcription kit (Applied Biosystems,

Waltham, MA). The cDNA samples were diluted 1:30 for qPCR. Each of the qPCR wells contained 5 µL of Fast SYBR Green Master mix (Applied Biosystems), 1 µL of forward primer

(5 nM), 1 µL of reverse primer (5 nM), 2 µL of diethyl pyrocarbonate treated water and 1 µL of cDNA and the reactions were run in a 7500 Fast Real-time PCR instrument (Applied Biosystems).

The primers for each of the genes plus beta-actin are listed in Table 4.1. These genes were selected based on intestinal expression of AvBDs observed by Hong et al. (2012) and previous research done in our lab, which evaluated expression of AvBDs and LEAP2 during an Eimeria challenge

(Paris et al., 2013; Su et al., 2014; 2017) and

Fold change was calculated using the ΔΔCt method (Livak and Schmittgen, 2001).

Expression of beta-actin served as the reference gene to calculate ΔCt. For each individual day, the average ΔCt of the control duodenum, jejunum, ileum and ceca was used as the calibrator to calculate ΔΔCt of the corresponding treatment tissues.

Statistical Analysis

Changes in gene expression of host defense peptides (AvBD and LEAP2) were analyzed by ANOVA using the JMP Pro 13 software (SAS Institute, Cary, NC). The model included the main effects of treatment, sorted by tissue and genes. Significant differences (P<0.05) were further separated using Tukey’s Test.

39

Chi-Square analysis was performed using the JMP Pro 13 software (SAS Institute) to determine significant differences (P<0.05) between groups in Salmonella colonization rate.

Results

From day 1 to day 7, all three challenge groups had greater number of colony forming units

(cfu) of Salmonella Typhimurium when compared to the control group (Chapter 3, Figure 3.2).

On day 1, 50 to 100 percent of collected livers tested positive for Salmonella in the 107 and 108 groups. Only 33 to 67 percent of collected spleens tested positive for Salmonella in the challenged groups. On day 2, 83 to 100 percent of collected livers tested positive for Salmonella in the challenged groups. Of collected spleens, 50 to 63 percent tested positive for Salmonella. On days

5 and 7, all collected livers from all challenged groups tested positive for Salmonella. In collected spleens, 83 to 100 percent tested positive for Salmonella on day 5 and 67 to 100 percent positive for Salmonella on day 7. On days 1, 2, 5 and 7, all collected ceca from all challenged groups tested positive for Salmonella. All livers, spleens and ceca collected from the control group tested negative for Salmonella throughout the entirety of the trial (Table 4.2).

On day 1, there were differences (P<0.05) in mRNA abundance of AvBD-10 and LEAP2

(Table 4.3). In the duodenum, abundance of LEAP2 mRNA was greater in the 106 group compared to control, 107 and 108 groups. In the 108 group, LEAP2 mRNA abundance was lower compared to the control group. In the jejunum, ileum, and cecum there was lower abundance of LEAP2 mRNA in the 107 and 108 groups compared to the 106 group. In the jejunum and ileum, the 106 group was not different from the control group whereas in the cecum, the 106 group was greater than the control. Also in the cecum, AvBD-10 mRNA abundance was lower in the 107 group compared to the control group.

40

On day 2, there were differences (P<0.05) in mRNA abundance of AvBD12 and AvBD13

(Table 4.4). In the jejunum, AvBd12 mRNA abundance was increased in the 108 group compared to the control group. In the duodenum, a difference in abundance of AvBD13 was observed; however, when a Tukey’s test was performed, the groups could not be further separated.

Numerically, a greater abundance of AvBD13 mRNA was observed in both the 106 and 107 groups compared to the control group. In the cecum, the 106 group had a greater abundance of AvBD13 compared to the 108 group but not different from the control group.

On day 5, there were differences (P<0.05) in mRNA abundance of all AvBD (Table 4.5).

For AvBD-1, there was a decrease in mRNA abundance in the 107 and 108 groups compared to the control group in the jejunum, decrease in all challenged groups compared to the control in the ileum and decrease in the 107 and 108 groups compared to the control in the cecum. For AvBD6, there was lower mRNA abundance in the 107 and 108 groups compared to the control group in the jejunum and cecum and lower in all challenged groups compared to the control in the ileum. For

AvBD8, mRNA abundance was downregulated in the ileum of all challenged groups compared to the control group and downregulated in the cecum of the 107 and 108 groups compared to the control. In the duodenum, jejunum, ileum and cecum, AvBD10 mRNA abundance was lower in all challenged groups compared to the control. For AvBD11, mRNA abundance was lower in all challenged groups compared to the control in the duodenum and ileum. A difference in mRNA abundance of AvBD-11 was observed in the jejunum, however, groups could not be further separated using a Tukey’s test. Numerically, a lower abundance of AvBD11 mRNA was observed in all challenged groups compared to the control group. For AvBD12, mRNA abundance was decreased in all challenged groups in the duodenum, ileum and cecum compared to the control. In the jejunum, mRNA abundance was decreased in the 107 group compared to the control group. A

41 difference in mRNA abundance of AvBD13 was observed in the ileum, however, groups could not be further separated using a Tukey’s test. Numerically, a lower abundance of AvBD13 mRNA was observed in all challenged groups compared to the control group. In the cecum, AvBD13 mRNA abundance was lower in the 107 and 108 groups compared to the control. For LEAP2, mRNA abundance in the ileum was lower in all challenged groups compared to the control group.

On day 7, there were differences (P<0.05) in mRNA abundance for AvBD6, 10, 11, 12, 13 and LEAP2 (Table 4.6). For AvBD6, there was decrease in mRNA abundance in the 107 and 108 groups compared to the 106 group in the duodenum, which was not different from the control. In the cecum, mRNA abundance was lower in the 108 group compared to the 106 group, however not different from the control. For AvBD8, a difference in mRNA abundance was observed in the duodenum and jejunum, however, groups could not be further separated using a Tukey’s test. In the cecum, mRNA abundance was greater in the 106 group compared to the 108 group, but the 106 group was not different compared to the control. For AvBD10, mRNA abundance was greater in the 106 group compared to the 107 and 108 groups in the duodenum, but the 106 group was not different from the control group. For AvBD11, AvBD12 and AvBD13, mRNA abundance in the duodenum was increased in the 106 group compared to the control and the 107 and 108 groups. For

LEAP2, a difference in mRNA abundance was observed in the duodenum, however, groups could not be further separated using a Tukey’s test. In the jejunum, mRNA abundance of LEAP2 was greater in the 107 and 108 groups compared to the control.

Discussion

Results from tested ceca indicate that the Salmonella challenge of the challenged groups was successful. Salmonella was also able to successfully colonize the ceca as recovered cfu

42 numbers remained high throughout the trial. The ileum and ceca serve as the primary sites for

Salmonella challenge and colonization (Meade et al., 2009). From day 1 to day 7, a gradual increase in positive livers and spleens were observed, indicating that a systemic Salmonella infection had been established in the challenged groups.

The majority of AvBDs did not show an early change in response to the Salmonella challenge. Meade et al. (2009) observed an increase in AvBD expression six hours post-

Salmonella challenge with an increase in heterophils, monocytes and macrophages at 20 and 48 hours post-challenge. The most common mode of action for AvBDs is bacterial killing; however,

AvBDs may also recruit and activate other immune cells (Cuperus et al., 2013). This could explain why a change in the intestinal AvBD mRNA abundance at ~24 hours post challenge was not observed and the increase in heterophils and macrophages observed by Meade et al. (2009), suggesting AvBD . The Salmonella challenge did affect the abundance of LEAP2 mRNA, which seems to be dependent on challenge dose. In the duodenum and cecum, an increase in LEAP2 mRNA abundance was observed in the 106 group compared to the control, while in the jejunum and ileum, LEAP2 mRNA was only numerically greater. This increase in LEAP2 is similar to the response seen by Townes et al. (2004). LEAP2 has been suggested to affect the permeability of the outer Salmonella membrane, which may lead to cell death (Townes et al., 2009). The increase of LEAP2 may indicate an immune response to the challenge and an attempt to terminate it.

Interestingly, the opposite response is seen in the abundance of LEAP2 mRNA in all tissues of the

108 group. This decrease in LEAP2 mRNA may indicate the pathogen may have been able to suppress intestinal expression of LEAP 2 at a high challenge dose. Very little changes in AvBD or LEAP2 mRNA abundance were observed on day 2. It is a very different response from what is

43 observed on day 1, which may indicate the pathogen was able to evade the response by the host immune system.

A remarkable response to the Salmonella challenge was observed on day 5 as the abundance of all HDPs mRNA were downregulated in various tissues in the Salmonella challenged groups compared to the control. AvBD10 mRNA abundance was lower in all tissues in both the

107 and 108 challenged groups. This may indicate that AvBD10 plays an important role in regulating bacterial infections of the gastrointestinal tract. The Salmonella challenge appears to have a pronounced impact on the mRNA abundance of AvBD1, 6, 8, 10, and 12 in the lower gastrointestinal tract, which is the main site of Salmonella colonization, especially in the 107 and

108 groups. This may indicate that the pathogen might be able to influence the intestinal AvBD expression as part of the immune evasion strategy of the pathogen. This ability has also been observed after Campylobacter infection of chickens (van Dijk et al., 2012). The manipulation of

AvBD might be essential in the systemic spread of Salmonella as all livers and spleens collected from the 107 and 108 groups tested positive for Salmonella. However, by day 7, the downregulation of AvBD that was observed on day 5 was no longer present.

AvBDs and LEAP2 have been shown to be effective peptides against pathogens in in vitro settings. In the current study, we observed abundance of AvBDs and LEAP2 mRNA during a

Salmonella challenge. The host immune system was initially able to respond to the Salmonella challenge. However, the response was dependent on dosage, with greater LEAP2 mRNA abundance in the 106 group compared to the 107 and 108 groups. As the infection progressed, a lower abundance of AvBD mRNA was observed in the lower gastrointestinal tract. AvBDs and

LEAP2 may be effective against pathogen infection; however, pathogens may have developed

44 mechanisms to ensure survival from host HDPs, which leads to successful colonization of the gastrointestinal tract of poultry.

45

Table 4.1: Forward and Reverse Primers of Host Defense Peptide Genes Gene Gene full name Forward/reverse primer.

AvBD11 Avian Beta Defensin 1 GAGTGGCTTCTGTGCATTTCTG /TTGAGCATTTCCCACTGATGAG AvBD61 Avian Beta Defensin 6 GCCCTACTTTTCCAGCCCTATT / GGCCCAGGAATGCAGACA AvBD81 Avian Beta Defensin 8 ATGCGCGTACCTAACAACGA /TGCCCAAAGGCTCTGGTATG AvBD101 Avian Beta Defensin 10 CAGACCCACTTTTCCCTGACA /CCCAGCACGGCAGAAATT AvBD111 Avian Beta Defensin 11 GGTACTGCATCCGTTCCAAAG /GCATGTTCCAAATGCAGCAA AvBD121 Avian Beta Defensin 12 TGTAACCACGACAGGGGATTG /GGGAGTTGGTGACAGAGGTTT AvBD131 Avian Beta Defensin 13 CAGCTGTGCAGGAACAACCA /CAGCTCTCCATGTGGAAGCA LEAP22 Liver-expressed antimicrobial peptide 2 CTCAGCCAGGTGTACTGTGCTT /CGTCATCCGCTTCAGTCTCA Beta-actin3 Beta-actin GTCCACCGCAAATGCTTCTAA /TGCGCATTTATGGGTTTTGTT 1 Primer sequence designed by Su et al., 2017 2Primer sequence designed by Casterlow et al., 2011 3Primer sequence designed by Gilbert et al., 2007

46

Table 4.2: Liver, Spleen and Ceca Colonization 1, 2, 5 and 7 days after Salmonella Challenge.

Day 1 Day 2 Day 5 Day 7

Salmonella culture- positive/total Salmonella culture- positive/total Salmonella culture- positive/total Salmonella culture- positive/total

Challenge Liver Spleen Ceca Liver Spleen Ceca Liver Spleen Ceca Liver Spleen Ceca

CON 0/6 0/6 0/6 0/6 0/6 0/6 0/6 0/6 0/6 0/6 0/6 0/6

L 3/6 3/6 6/6* 5/6* 3/6 6/6* 6/6* 5/6* 6/6* 6/6* 4/6* 6/6*

M 5/6* 4/6* 6/6* 6/6* 4/6* 6/6* 6/6* 6/6* 6/6* 6/6* 6/6* 6/6*

H 6/6* 2/6 6/6* 6/6* 5/6* 6/6* 6/6* 6/6* 6/6* 6/6* 6/6* 6/6*

*Values are significantly different from control (P <0.05)

47

Table 4.3: Expression of Host Defense Peptides in Young Broilers Challenged with Salmonella Typhimurium on Day 1

Tissue Challenge/Gene AvBD1 AvBD6 AvBD8 AvBD10 AvBD11 AvBD12 AvBD13 LEAP2 Duodenum CON 1.12 1.12 1.20 2.51 1.16 1.16 1.18 1.08B L-106 1.87 1.18 1.80 0.53 1.81 1.43 1.10 2.02A M-107 0.88 0.77 1.27 0.34 0.72 0.67 0.62 0.43BC H-108 0.74 0.50 0.77 0.86 0.56 0.70 0.48 0.05C P-value 0.59 0.65 0.66 0.40 0.44 0.40 0.30 <.0001 Pooled SEM 0.62 0.43 0.58 0.97 0.58 0.36 0.30 0.23

Jejunum CON 1.13 1.12 1.25 1.89 1.34 1.17 1.23 1.05AB L-106 2.03 2.20 2.92 1.37 1.50 2.30 1.84 1.30A M-107 1.07 1.06 1.69 0.35 0.68 0.72 0.49 0.45BC H-108 1.11 0.95 1.52 1.10 0.83 1.05 0.87 0.06C P-value 0.19 0.20 0.29 0.41 0.42 0.33 0.39 <.0001 Pooled SEM 0.35 0.44 0.64 0.63 0.39 0.62 0.55 0.15

Ileum CON 1.16 1.09 1.20 1.22 1.07 1.29 1.54 1.05AB L-106 1.86 1.74 2.39 0.45 1.23 2.35 1.36 1.36A M-107 0.81 0.73 1.16 0.18 0.44 0.46 0.55 0.16BC H-108 0.99 1.00 1.60 1.90 1.06 1.02 0.79 0.02C P-value 0.22 0.07 0.13 0.13 0.25 0.23 0.38 0.0001 Pooled SEM 0.36 0.25 0.38 0.53 0.28 0.62 0.45 0.23

Cecum CON 1.22 2.85 1.34 1.27A 1.06 1.07 1.05 1.09B L-106 1.20 2.69 1.17 0.34AB 0.91 1.03 0.98 2.07A M-107 1.17 2.58 1.27 0.28B 0.66 0.69 0.78 0.14C H-108 0.52 1.58 0.98 0.69AB 0.52 0.85 0.75 0.02C P-value 0.54 0.67 0.92 0.04 0.11 0.70 0.65 <0.0001 Pooled SEM 0.39 0.71 0.39 0.23 0.16 0.25 0.20 0.23 A,B,CMeans with no common superscript within a gene and a tissue differ significantly (P<0.05)

48

Table 4.4: Expression of Host Defense Peptides in Young Broilers Challenged with Salmonella Typhimurium on Day 2.

Tissue Challenge/Gene AvBD1 AvBD6 AvBD8 AvBD10 AvBD11 AvBD12 AvBD13 LEAP2 Duodenum CON 1.92 1.25 1.07 1.08 1.03 1.15 1.03 1.01 L-106 1.86 1.26 0.88 0.97 1.02 1.73 1.44 0.93 M-107 3.28 1.31 0.97 1.26 1.41 1.60 1.46 0.93 H-108 0.34 1.42 1.24 1.04 1.21 1.25 0.89 0.67 P-value 0.09 0.26 0.54 0.51 0.33 0.42 0.046* 0.44 Pooled SEM 0.73 0.26 0.18 0.14 0.17 0.28 0.16 0.15

Jejunum CON 1.59 1.08 1.10 1.01 1.05 1.05B 1.08 1.06 L-106 1.65 1.20 0.87 1.57 1.13 2.04AB 0.94 1.05 M-107 2.12 1.61 1.16 1.70 1.24 1.58AB 0.91 0.76 H-108 0.58 2.38 1.91 2.32 1.81 2.16A 1.10 0.96 P-value 0.19 0.25 0.35 0.18 0.29 0.04 0.83 0.86 Pooled SEM 0.48 0.48 0.42 0.4 0.29 0.26 0.18 0.28

Ileum CON 2.15 1.21 1.03 1.10 1.11 1.14 1.06 1.08 L-106 1.03 0.81 0.59 2.01 1.60 1.23 1.27 0.40 M-107 2.02 1.05 0.86 2.10 1.82 1.34 1.69 1.00 H-108 0.42 1.50 1.35 1.91 1.85 1.03 1.34 0.37 P-value 0.07 0.38 0.39 0.24 0.49 0.86 0.51 0.25 Pooled SEM 0.49 0.27 0.31 0.37 0.37 0.27 0.29 0.31

Cecum CON 2.14 2.53 1.06 1.07 1.01 1.09 1.02AB 1.07 L-106 1.02 1.66 0.49 1.01 1.09 1.07 1.14A 0.38 M-107 1.34 2.34 0.57 0.85 1.14 0.85 1.02AB 0.70 H-108 0.17 1.57 0.52 0.58 0.67 0.34 0.44B 0.42 P-value 0.23 0.43 0.06 0.27 0.20 0.06 0.04 0.08 Pooled SEM 0.65 0.48 0.16 0.18 0.16 0.20 0.17 0.19 A,B,Means with no common superscript within a gene and a tissue differ significantly (P<0.05). *Values could not be further separated using Tukey’s Test.

49

Table 4.5: Expression of Host Defense Peptides in Young Broilers Challenged with Salmonella Typhimurium on Day 5.

Tissue Challenge/Gene AvBD1 AvBD6 AvBD8 AvBD10 AvBD11 AvBD12 AvBD13 LEAP2 Duodenum CON 1.27 1.14 1.13 1.13A 2.74A 1.07A 1.23 1.06 L-106 0.49 0.52 0.59 0.35B 0.96B 0.43B 0.45 0.85 M-107 0.61 0.69 0.81 0.31B 0.93B 0.43B 0.37 0.96 H-108 0.42 0.42 0.63 0.27B 0.93B 0.39B 0.50 1.03 P-value 0.13 0.60 0.24 0.001 0.004 0.0051 0.10 0.76 Pooled SEM 0.26 0.18 0.20 0.14 0.34 0.13 0.25 0.15

Jejunum CON 1.26A 1.16A 1.42 1.29A 1.33 1.27A 1.49 1.04 L-106 0.47AB 0.57AB 0.69 0.26B 0.27 0.40AB 0.46 0.93 M-107 0.33B 0.37B 0.45 0.29B 0.27 0.31B 0.38 1.14 H-108 0.33B 0.36B 0.53 0.28B 0.41 0.50AB 0.70 1.00 P-value 0.03 0.02 0.11 0.02 0.04* 0.04 0.23 0.71 Pooled SEM 0.22 0.18 0.20 0.24 0.27 0.23 0.40 0.13

Ileum CON 1.21A 1.31A 1.30A 1.23A 1.13A 1.11A 1.45 1.19A L-106 0.33B 0.31B 0.31B 0.22B 0.28B 0.40B 0.30 0.42B M-107 0.28B 0.30B 0.33B 0.31B 0.33B 0.41B 0.34 0.50B H-108 0.12B 0.12B 0.17B 0.14B 0.14B 0.20B 0.20 0.35B P-value 0.01 0.01 0.01 0.002 0.001 0.003 0.04* 0.01 Pooled SEM 0.19 0.22 0.22 0.18 0.15 0.15 0.31 0.16

Cecum CON 1.18A 1.30A 1.20A 1.24A 1.42 1.09A 1.15A 1.04 L-106 0.57AB 0.56AB 0.46AB 0.26B 0.36 0.46B 0.47AB 0.70 M-107 0.23B 0.29B 0.24B 0.19B 0.24 0.29B 0.21B 0.51 H-108 0.24B 0.16B 0.15B 0.08B 0.10 0.10B 0.13B 0.41 P-value 0.01 0.003 0.004 0.002 0.058 0.001 0.01 0.16 Pooled SEM 0.19 0.19 0.18 0.19 0.34 0.15 0.19 0.20 A,B,Means with no common superscript within a gene and a tissue differ significantly (P<0.05). *Values could not be further separated using Tukey’s Test.

50

Table 4.6: Expression of Host Defense Peptides in Young Broilers Challenged with Salmonella Typhimurium on Day 7.

Tissue Challenge/Gene AvBD1 AvBD6 AvBD8 AvBD10 AvBD11 AvBD12 AvBD13 LEAP2 Duodenum CON 1.05 1.14AB 1.20 1.04AB 1.05B 1.10B 1.07B 1.07 L-106 1.66 1.82A 1.42 1.64A 1.98A 3.92A 2.80A 1.08 M-107 0.77 0.28B 0.17 0.41B 0.53B 0.66B 0.53B 1.85 H-108 0.52 0.32B 0.13 0.43B 0.53B 0.57B 0.50B 1.93 P-value 0.06 0.002 0.03* 0.0001 0.0001 0.0001 0.0001 0.02* Pooled SEM 0.28 0.26 0.35 0.16 0.15 0.21 0.21 0.22

Jejunum CON 1.19 1.16 1.18 1.19 1.20 1.14 1.16 1.04C L-106 2.22 1.43 1.10 1.05 1.39 1.78 1.49 1.24BC M-107 0.93 0.58 0.28 0.62 0.84 1.08 1.07 2.02AB H-108 0.68 0.38 0.11 0.54 0.61 0.68 0.60 2.43A P-value 0.07 0.08 0.047* 0.20 0.43 0.42 0.43 0.001 Pooled SEM 0.42 0.30 0.30 0.24 0.36 0.47 0.35 0.21

Ileum CON 1.05 1.19 1.15 1.14 1.18 1.25 1.16 1.06 L-106 1.67 1.51 1.85 1.32 1.79 3.07 2.35 0.64 M-107 1.00 1.03 0.69 1.10 1.26 1.53 1.31 0.64 H-108 1.25 0.42 0.31 0.83 0.73 0.88 0.80 0.54 P-value 0.56 0.14 0.20 0.72 0.29 0.11 0.20 0.07 Pooled SEM 0.36 0.31 0.50 0.31 0.37 0.63 0.51 0.14

Cecum CON 1.17 1.55AB 1.30AB 1.32 1.24 1.48 1.20 1.16 L-106 1.88 2.55A 1.70A 1.86 2.42 3.04 2.14 0.86 M-107 1.12 1.41AB 0.67AB 1.26 2.01 2.28 2.14 1.49 H-108 1.07 0.87B 0.32B 0.66 0.78 0.89 0.92 0.61 P-value 0.39 0.046 0.03 0.14 0.15 0.08 0.12 0.18 Pooled SEM 0.37 0.38 0.32 0.34 0.52 0.47 0.42 0.28 A,B,CMeans with no common superscript within a gene and a tissue differ significantly (P<0.05). *Values could not be further separated using Tukey’s Test.

51

Chapter 5 mRNA Expression of Nutrient Transporters, Avian Beta-Defensins and Liver- Expressed Antimicrobial Peptide 2 during a Campylobacter Challenge in One and Two Week Old Broilers

Abstract

Campylobacter is a bacterium ubiquitously found in the gastrointestinal tract of poultry.

However, there is limited research on its effect on the mRNA abundance of nutrient transporters, defensins and liver-expressed antimicrobial peptides in Campylobacter infected broilers. The objective of this study was to characterize the effects of Campylobacter on the mRNA abundance of these three factors. Day of hatch commercial broiler chicks were inoculated with one of three

(106, 107, 108) colony-forming unit (cfu) levels of Campylobacter jejuni. Quantitative RT-PCR analyses revealed that expression of zinc transporter 1 (ZnT1) increased (P<0.05) on day 7 in the duodenum, ileum and ceca in birds challenged with 106 cfu of Campylobacter. An increase

(P<0.05) in the expression of all AvBD were observed on day 14 in the ileum and ceca in the 106 group. These results demonstrate the role that avian beta-defensins play during a Campylobacter challenge.

Introduction

Campylobacter is one of the most common causes of bacterial foodborne illness in the

United States (Ghareeb et al., 2012). Commonly found in the gastrointestinal tract of poultry, birds may carry high loads of the bacterium without showing any clinical signs of infection

(Newell et al., 2003; Van Deun et al., 2008). However, colonization by Campylobacter negatively

52 impacts the production of poultry by affecting the well-being of the birds (Humphrey, 2006;

Williams et al., 2013).

The small intestine plays an important role in the absorption of nutrients. Due to the expression of nutrient transporters throughout the entire gastrointestinal tract, the chicken is able to absorb amino acids and sugars through its entirety (Awad et al., 2007). Downregulation of nutrient transporters in the duodenum, jejunum, and ceca has been reported in Campylobacter- infected chickens (Awad et al., 2014). However, the ileum of these birds were not tested.

Avian beta-defensins (AvBDs) and liver-expressed antimicrobial peptide 2 (LEAP2) are two classes of host defense peptides that belong to the innate immune system (Cuperus et al.,

2013). Defensins are cysteine-rich cationic peptides, with the beta-defensins being the sole defensin family found in birds. The expression of host defense peptides seems to be influenced during bacterial infections. Expression of AvBDs was upregulated in the gastrointestinal and reproductive tracts of chickens during Salmonella challenges (Yoshimura et al., 2006; Milona et al., 2007; Akbari et al., 2008). However, challenge may cause vastly variable effects in defensin expression. In Salmonella infected chickens, LEAP2 was found to be upregulated (Townes et al.,

2004). The objective of this study was to evaluate the mRNA expression of nutrient transporters,

AvBD and LEAP2 during a Campylobacter challenge at three challenge levels.

Materials and Methods

Animals

One hundred and fifty-two day-of-hatch Ross 308 cross broiler chicks were obtained from a local commercial hatchery. The paper chick tray liners were taken for Campylobacter testing.

Upon arrival, chicks were orally challenged with Campylobacter and then placed into pens. The trial was run in one disinfected, environmentally controlled room with four 1.8 x 1.4 m floor pens

53 with fresh pine shavings (Figure 5.1). The chicks were provided with heat lamps to ensure they received adequate heat for the first few days. All floor pens were equipped with nipple drinkers and feed trays, which were monitored on a daily basis. All birds received a non-medicated corn- soybean meal starter diet obtained from the Texas A&M University Poultry Research Center. The diet was formulated to meet or exceed the National Research Council nutrient requirements for poultry (NRC, 1994). Chicks were provided feed and water ad libitum from time of placement until termination. This study was approved by the Southern Plains Agricultural Research Center

Animal Care and Use Committee and conducted at the Food and Feed Safety Research Laboratory

(USDA Agricultural Research Service, College Station, TX)

Tray Liner Evaluation

Using aseptic techniques, each individual tray liner was placed into a gallon-size bag and

150 mL of Buffered Peptone Water (BPW) was added into each bag. Bags were then thoroughly massaged for 60 seconds to ensure there was proper contact between the paper tray-liner and the

BPW. One milliliter of tray-liner BPW solution was transferred into 9 mL of 2X Bolton’s solution

(Musgrove et al., 1997). After 24 hours, each solution was streaked onto Campy-Cefex plates

(Stern et al., 1992) and incubated in a microaerophilic environment (85% N2, 10% CO2 and 5%

O2) for 48 hours at 42°C. Colonies were then analyzed for morphology.

Challenge

For the challenge, a strain of Campylobacter jejuni (wild-type) was used (Byrd et al.,

1998). The challenge inoculum was prepared using an overnight culture in Bolton’s broth. The culture was serially diluted in Butterfield’s solution (Sigma Chemical Company, St. Louis, MO)

54 to make three different challenge dosages that were approximately 106, 107, 108 colony-forming units (cfu) per mL. The viable cell concentrations of the three challenge inoculums were confirmed by cfu on campy-cefex plates. Plates were incubated for 48 hours at 42°C in a microaerophilic environment and expressed as log10 to determine cfu/mL. The control group received 0.5 mL of sterile Butterfield’s solution by oral gavage and occurred before the challenge groups. Challenge was administered by crop gavage of 0.5 mL to each bird within a challenge group. Challenge was administered with the 106 cfu challenge dosage group occurring first and ending with the 108 cfu challenge dosage group.

Tissue collection

Before challenge, six day of hatch chicks were randomly selected, euthanized by cervical dislocation and subjected to necropsy immediately after euthanasia for tissue collection. The small intestine from each bird was aseptically removed and a 1.25-2.0 cm whole tissue sample was taken from the duodenum, jejunum, ileum and ceca, washed in phosphate buffer solution and immediately placed into collection tubes containing RNAlater® solution (Qiagen, Germantown,

MD). Chicks were observed daily for 28 days after the challenge. On days 7, 14, 21, and 28, six chicks from each of the four groups were randomly selected, euthanized by cervical dislocation and subjected to necropsy immediately after euthanasia for tissue collection. To prevent contamination, the control group was sampled first followed by the challenged groups (106, 107,

108 cfu). A cecum from each bird was aseptically removed and cecal contents were aseptically collected, weighed and serially diluted 1:10, 1:100, 1:1000, and 1: 10,000 in 9 mL Butterfield’s solution and spread onto Campy-Cefex plates. Campy-Cefex plates were incubated at 42°C for

48 hours in a microaerophilic environment and colonies were counted and analyzed for colony

55 morphology. Suspected colonies were confirmed as Campylobacter spp. by examination of colony morphology and motility on a wet mount slide under phase-contrast microscopy. A piece of tissue from the remaining ceca was taken and placed into RNAlater® solution (Qiagen) and the remaining tissue, along with the liver and spleen were individually placed into 9 mL of Bolton’s broth, incubated at 42°C for 24 hours and streaked onto Campy-Cefex plates. Plates were incubated for an additional 48 hours in a microaerophilic environment at 42°C and observed for presence or absence of Campylobacter. The small intestine from each bird was also aseptically removed and a 1.25-2.0 cm whole tissue sample was taken from the duodenum, jejunum and ileum, washed in phosphate buffer solution and immediately placed into collection tubes containing RNAlater® solution (Qiagen). All intestinal samples were frozen at -80°C until shipped to Virginia Tech.

RNA Extraction and Relative qPCR

Total RNA was extracted from duodenal, jejunal, ileal,and cecal samples from birds (n=5) collected on d7 and d14, following the Direct-zol RNA Miniprep protocol (Zymogen, Irvine, CA).

The RNA concentration for each sample was quantified using a Nanodrop 1000 (Thermo

Scientific, Waltham, MA) and then diluted to 200 ng/µL. The nutrient transporter genes analyzed included the neutral (LAT1), Na+-independent (bo,+AT), anionic (EAAT3), and cationic (CAT2) amino acid transporters, the facilitated (GLUT2) and Na+-dependent (SGLT1) sugar transporters and the zinc (ZnT1) transporter. The host defense peptides analyzed included: LEAP2 and the avian beta-defensins AvBD1, AvBD6, AvBD8, AvBD10, AvBD11, AvBD12, and AvBD13. The cDNA was synthesized from 500 ng of total RNA using a High-capacity cDNA reverse transcription kit (Applied Biosystems, Waltham, MA). The cDNA samples were diluted 1:30 for qPCR. Each of the qPCR wells contained 5 µL of Fast SYBR Green Mastermix (Applied

56

Biosystems), 1 µL of forward primer (5 nM), 1 µL of reverse primer (5 nM), 2 µL of Diethyl pyrocarbonate treated water, and 1 µL of cDNA and the reactions were run in a 7500 Fast Real- time PCR instrument (Applied Biosystems). The primers for each of the genes plus beta-actin are listed in Tables 5.1 and 5.2. These genes were selected based on previous research done in our lab, which evaluated expression of nutrient transporters and AvBD during an Eimeria challenge

(Paris et al., 2013; Su et al., 2014; 2015; 2017; Yin et al., 2015).

Fold change was calculated using the ΔΔCt method (Livak and Schmittgen, 2001). Beta- actin served as the reference gene to calculate ΔCt. For each individual day, the average ΔCt of the control duodenum, jejunum, ileum and ceca were used as the calibrator to calculate ΔΔCt of the corresponding challenge group.

Statistical Analysis

Campylobacter recovery levels (cfu/g of cecal contents) were analyzed by ANOVA using the JMP Pro 13 software, the model included the main effect of log10 cfu, sorted by day.

Significant differences (P<0.05) were further separated using Tukey’s Test.

Changes in gene expression of nutrient transporters and host defense peptides were analyzed by ANOVA using the JMP Pro 13 software. The model included the main effect of treatment, sorted by tissue and genes. Significant differences (P<0.05) were further separated using Tukey’s Test.

Chi-Square analysis was performed using the JMP Pro 13 software to determine significant differences between groups in Campylobacter colonization rate.

57

Results

From day 7 through day 28, all three challenge groups had greater cfu number of C. jejuni compared with the control groups (Figure 5.2). On day 7, the 107 group had a lower cfu number when compared to the 108 group. However, for the remainder of the trial there were no differences in cfu numbers between challenge groups. Control birds were all Campylobacter-free except for one bird on day 28.

Although samples were collected at day 7, 14, 21, and 28, only day 7 and 14 samples were analyzed for gene expression to evaluate early changes during Campylobacter colonization. On day 7, there were differences (P<0.05) in mRNA abundance of six nutrient transporters (Table

5.4). For EAAT3, mRNA abundance in the duodenum was greater in the 106 group compared to the 108 group with the 106 group not different from the control group. The bo,+AT mRNA abundance in the duodenum was greater in the 107 group compared to the control group. For

ZnT1, mRNA abundance in the duodenum and ileum was greater in the 106 and 107 group compared to the control. In the cecum, mRNA abundance was greater in the 106 group compared to the control. GLUT2 mRNA abundance in the ileum was greater in the 107 group compared to the control. In the cecum, GLUT2 mRNA abundance was greater in the 106 group compared to

108 group with the 106 group not different from the control. For LAT1, mRNA abundance was greater in the 108 group compared to the control in the duodenum and the 107 group compared to the control in the jejunum. SGLT1 mRNA abundance was lower in the 106 group compared to the control.

Also on day 7, there were differences (P<0.05) in mRNA abundance of three AvBDs

(Table 5.5). AvBD10 mRNA abundance was greater in the 106 group compared to the control in duodenum, lower in the 107 and 108 groups compared to the control in the ileum, and greater in

58 the 106 group compared to the 108 group in the cecum with the 106 group not different from the control. A difference in mRNA abundance for AvBD11 was observed in the ileum; however, it could not be further separated with a Tukey’s test. Numerically, the mRNA abundance in the 106 and 107 groups was lower compared to the control. For AvBD12, mRNA abundance was lower in the 108 group compared to the control group in the cecum. The AvBD13 mRNA abundance in the cecum was greater in the 106 group compared to the 107 and 108 groups with the 106 group not different from the control. For LEAP2, mRNA abundance decreased in the 107 and 108 groups compared to the control.

On day 14, there were differences in mRNA abundance of six nutrient transporters (Table

5.6). For EAAT3, mRNA abundance was greater in the jejunum of the 106 group compared to the

108 group with the 106 group not different from the control and in the ileum, mRNA abundance was greater in the 106 group compared to the control. The bo,+AT mRNA abundance was greater in the 106 group compared the 107 and 108 groups in the duodenum, with the 106 group not different from the control group. In the jejunum and ileum, there was an increase in bo,+AT mRNA abundance in the 106 group compared to the control, and the 107 and 108 groups. For ZnT1, mRNA abundance in the ileum was greater in the 106 group compared to the control. For GLUT2, mRNA abundance in the jejunum and cecum was greater in the 106 group compared to the 108 group with the 106 group not different to the control. In the ileum, there was an increase in GLUT2 mRNA abundance in the 106 group compared to the control and the 107 and 108 groups. For CAT2, mRNA abundance in the duodenum was greater in the 107 group compared to the control. For SGLT1, there was an increase in mRNA abundance in the 106 and 108 groups compared to the control.

Also on day 14, there were differences in mRNA abundance of all AvBDs and LEAP2

(Table 5.7). AvBD1 mRNA abundance in the duodenum was greater in the 106 and 107 groups

59 compared to the 108 group with the 106 and 107 groups not different from control. In the jejunum, mRNA abundance was lower in the 108 group compared to the control. In the ileum and cecum, mRNA abundance was greater in the 106 group compared to the control group. For AvBD6, there was an increase in mRNA abundance in the 106 compared to the 108 group with the 106 group not different to the control. In the jejunum, mRNA abundance was lower in the 108 group compared to the control. In the ileum and cecum, mRNA abundance was greater in the 106 group compared to the control and the 107 and 108 groups. For AvBD8, there was a decrease in mRNA abundance in the duodenum and jejunum in the 108 group compared to the control. In the ileum and cecum, there was an increase in AvBD8 mRNA abundance in the 106 group compared to the control and the 107 and 108 groups. AvBD10 mRNA abundance in the duodenum was greater in the 106 group compared to the 108 group with the 106 group not different compared to the control. In the ileum,

AvBD10 mRNA abundance was greater in the 106 group compared to the control and the 107 and

108 groups. AvBD10 mRNA abundance in the cecum was increased in the 106 group compared to the control and the 107 group. For AvBD11, mRNA abundance in the duodenum was greater in the 106 group compared to the 108 group with the 106 group not different to the control. In the ileum, AvBD11 mRNA abundance was increased in the 106 group compared to the control and

107 and 108 groups and in the cecum, mRNA abundance was increased in the challenged groups compared to the control. AvBD12 mRNA abundance in the ileum was greater in the 106 group compared to the control and in the cecum, mRNA abundance was greater in the 106 and 108 groups compared to the control. For AvBD13, mRNA abundance was greater in the 106 group compared to the 108 group in the duodenum. In the ileum, AVBD13 mRNA abundance was increased in the

106 group compared to the control and 107 group. AvBD13 mRNA abundance was increased in the cecum of all challenged groups compared to the control group. LEAP2 mRNA abundance in

60 the cecum was greater in the 108 group compared to the 107 group with the 108 group not different from control.

Discussion

There was an increase in the expression of ZnT1 in the 106 group compared to the control in the duodenum, ileum and ceca on day 7 and ileum on day 14. Zinc is a trace mineral that has many important cellular functions, such as enzymatic reactions, DNA synthesis and gene expression (Berg et al., 1996). An increase in mRNA expression of ZnT1 could indicate an increased efflux of zinc from the enterocytes into the blood supply. However, with the analysis limited to one zinc transporter, the influx of zinc into the enterocytes is unknown. Further research into other zinc transporters could help explain the role zinc could play during a Campylobacter challenge.

A delayed change in bo,+AT mRNA abundance was observed on day 14. An increase in bo,+AT mRNA abundance was observed in the jejunum and ileum of the 106 group compared to the control group. The transporter bo,+AT is responsible for the transport of cystine and other cationic amino acids. By increasing bo,+AT expression, there would be an increase of cystine and other cationic amino acids entering the enterocyte. This may be a host response to limit available amino acids in the brush border membrane of the ileum since Campylobacter colonization occurs in the intestinal mucosa. Also in the ileum, mRNA abundance of GLUT2 was increased in the 106 group compared to the control. GLUT2 is responsible for the transport of glucose, galactose and mannose, and fructose out of the enterocyte. Interestingly, SGLT1 mRNA was also increased in the ileum of the 106 group compared to the control, increasing the transport of glucose into the

61 enterocyte. In the ileum of the 106 group, GLUT2 and SGLT1 may be working together to limit the availability of glucose into the intestinal mucosal, which could be used by Campylobacter.

Interestingly, an early decrease in the expression of LEAP2 in the ceca was seen in the 107 and 108 groups on day 7. This decrease in LEAP2 is similar as reported in Eimeria infections, in which the protozoan appears to be influencing the gut environment (Casterlow et al., 2011; Su et al., 2015; 2017). A decrease in AvB12 mRNA was also observed in the cecum of the 108 group compared to the control. Suppression of the intestinal expression of CATH2, another host defense peptide, which has the ability to eliminate Campylobacter, was observed in young broilers infected with Campylobacter (van Dijk et al., 2012). The decrease in LEAP2 expression may indicate a similar evasion strategy by the bacterium to ensure it survives.

The increase in the expression of all observed AvBDs in the ileum of the 106 group on day

14 could explain the increased transport of nutrients. The increase of glucose could also be used to fuel the production of HDPs as an increase of mRNA abundance of all AvBDs is seen in the ileum and cecum of the 106 group. In the ceca, expression of AvBDs varied within the challenged groups, with the 106 group expressing an increase in all AvBDs. Increased expression of AvBDs could be an attempt by the host to prevent the bacteria from spreading systemically.

Campylobacter has been reported to induce a response from the immune system, with the release of pro-inflammatory cytokines (Vaezirad et al., 2016). Avian beta-defensins can induce pro- inflammatory cytokine expression in macrophages, indicating AvBDs may play an important role in immune responses (Semple et al., 2011).

During the duration of the trial, isolation of Campylobacter from the ceca was never below

6 logs cfu within any of the challenge groups. This may indicate that although the host was able

62 to activate the innate immune system in response to the bacterium, it was unsuccessful in eliminating the bacterium from the gastrointestinal tract.

Although not many changes in the expression of nutrient transporters were observed, there were a few differences seen compared to the results reported by Awad et al. (2014). In 14-day- old broilers challenged with 108 cfu of C. jejuni, a decrease in expression of GLUT2, SGLT1 and

EAAT3 was observed. In the current study, we observed increases of GLUT2, SGLT1 and EAAT3 mRNA in the ileum. However, at times, the greater expression seen in these genes were between the challenge groups and not the control group. Age of birds when challenged and the type of birds used could explain why differences were observed between these two studies.

In summary, changes in mRNA expression of nutrient transporters, AvBDs, and LEAP2 were observed. These changes appear to be dependent on the challenge dose, age, and tissue. The increased expression of AvBDs may suggest the role they play in the immune response that has been previously reported to occur during a Campylobacter infection. It is still unclear what role the zinc transporter may play during a Campylobacter infection; however, a similar effect has been seen in Salmonella infections.

63

Table 5.1: Forward and Reverse Primers of Nutrient Transporter Genes Gene Gene full Location Function Forward/reverse primer name Solute carrier family 7, Na+-independent neutral/cystine, cationic amino acid CAGTAGTGAATTCTCTGAGTGTGAAGCT b0,+ATa Brush border member 9 exchanger /GCAATGATTGCCACAACTACCA (SLC7A9) Cationic amino acid TGCTCGCGTTCCCAAGA CAT2a Basolateral Transports lysine, arginine and histidine transporter-2 /GGCCCACAGTTCACCAACAG (SLC7A2) Excitatory amino acid TGCTGCTTTGGATTCCAGTGT EAAT3a Brush border Transports aspartate, glutamate and cysteine transporter 3 /AGCAATGACTGTAGTGCAGAAGTAATATATG (SLC1A1) L type amino acid CACACTATGGGCGCATGCT LAT1a Basolateral Transports hydrophobic amino acids transporter-1 /ATTGTGCCTGGAGGTGTTGGT (SLC7A5) Glucose Transports fructose, mannose, galactose, glucose and CTCAGCCAGGTGTACTGTGCTT GLUT2a transporter-2 Basolateral glucosamine /CGTCATCCGCTTCAGTCTCA (SLC2A2)

Sodium glucose GCCATGGCCAGGGCTTA SGLT1 a Brush border Transports low concentrations of D-glucose transporter-1 /CAATAACCTGATCTGTGCACCAGTA (SLC5A1)

Zinc TCCGGGAGTAATGGAAATCTTC ZnT1b transporter-1 Basolateral Efflux of Zn2+ /AATCAGGAACAAACCTATGGGAAA (SLC22A18) GTCCACCGCAAATGCTTCTAA Beta-actina Beta-actin Cytosol Reference gene /TGCGCATTTATGGGTTTTGTT a Primer sequence designed by Gilbert, et al., 2007 b Primer sequence designed by Paris and Wong, 2013

64

Table 5.2: Forward and Reverse Primers of Host Defense Peptide Genes Gene Gene full name Forward/reverse primer.

AvBD11 Avian Beta Defensin 1 GAGTGGCTTCTGTGCATTTCTG /TTGAGCATTTCCCACTGATGAG AvBD61 Avian Beta Defensin 6 GCCCTACTTTTCCAGCCCTATT /GGCCCAGGAATGCAGACA AvBD81 Avian Beta Defensin 8 ATGCGCGTACCTAACAACGA /TGCCCAAAGGCTCTGGTATG AvBD101 Avian Beta Defensin 10 CAGACCCACTTTTCCCTGACA /CCCAGCACGGCAGAAATT AvBD111 Avian Beta Defensin 11 GGTACTGCATCCGTTCCAAAG /GCATGTTCCAAATGCAGCAA AvBD121 Avian Beta Defensin 12 TGTAACCACGACAGGGGATTG /GGGAGTTGGTGACAGAGGTTT AvBD131 Avian Beta Defensin 13 CAGCTGTGCAGGAACAACCA /CAGCTCTCCATGTGGAAGCA LEAP22 Liver-expressed antimicrobial peptide 2 CTCAGCCAGGTGTACTGTGCTT /CGTCATCCGCTTCAGTCTCA Beta-actin3 Beta-actin GTCCACCGCAAATGCTTCTAA /TGCGCATTTATGGGTTTTGTT 1 Primer sequence designed by Su et al., 2017 2Primer sequence designed by Casterlow et al., 2011 3Primer sequence designed by Gilbert et al., 2007

65

Table 5.3: Campylobacter jejuni Liver and Spleen Colonization 7, 14, 21 and 28 Days after Challenge Day 7 Day 14 Day 21 Day 28

Challenge Campylobacter Campylobacter Campylobacter Campylobacter Campylobacter Campylobacter Campylobacter Campylobacter

culture- culture- culture- culture- culture- culture- culture- culture-

positive/total – positive/total - positive/total - positive/total - positive/total - positive/total - positive/total - positive/total -

Liver Spleen Liver Spleen Liver Spleen Liver Spleen

CON 0/6 0/6 0/6 0/6 0/6 0/6 0/6 0/6

L 0/6 2/6 2/6 2/6 1/6 6/6* 1/6 3/6

M 0/6 0/6 1/6 2/6 3/6 3/6 2/6 2/6

H 0/6 0/6 0/6 2/6 2/6 5/6* 3/6 2/6

*Values are significantly different from control (P <0.05)

66

Table 5.4: Day 7 mRNA Abundance of Nutrient and Sugar Transporters in Campylobacter jejuni Challenged and Non-Challenged Young Broilers. Tissue Treatment EAAT3 bo,+AT ZnT1 GLUT2 CAT2 LAT1 SGLT1 Duodenum CON 1.04AB 1.02B 1.12B 1.13 1.10 1.08B 1.02A L-106 1.52A 1.47AB 4.46A 1.38 1.56 2.78AB 0.56B M-107 0.96AB 1.56A 4.29A 1.25 2.07 2.55AB 0.80AB H-108 0.79B 1.16AB 2.01B 1.21 2.15 3.46A 0.62AB P-value 0.03 0.01 0.001 0.96 0.18 0.05 0.02 Pooled SEM 0.16 0.11 0.55 0.33 0.36 0.56 0.10

Jejunum CON 1.09 1.05 1.16 1.04 1.07 1.09B 1.12 L-106 1.25 1.19 3.14 1.17 1.32 1.74AB 0.73 M-107 1.04 1.17 3.31 1.11 2.39 3.21A 0.63 H-108 1.06 1.14 2.27 0.87 1.93 2.46AB 1.05 P-value 0.80 0.93 0.053 0.57 0.07 0.03 0.08 Pooled SEM 0.17 0.17 0.56 0.16 0.36 0.47 0.15

Ileum CON 1.10 0.80 1.13B 1.07B 1.06 1.10 1.21 L-106 1.72 1.26 5.11A 2.48AB 0.81 1.72 0.76 M-107 1.33 1.46 4.76A 2.83A 1.00 1.61 0.99 H-108 1.09 1.19 3.6AB 1.43AB 0.79 1.33 0.73 P-value 0.39 0.13 0.001 0.01 0.65 0.23 0.56 Pooled SEM 0.29 0.19 0.78 0.38 0.18 0.24 0.27

Cecum CON 1.17 1.02 1.05B 1.07AB 1.04 1.05 1.19 L-106 0.58 0.78 4.07A 2.08A 0.71 2.61 0.49 M-107 0.61 0.70 2.62AB 0.95AB 2.87 2.49 0.69 H-108 0.77 1.34 2.32AB 0.37B 1.26 1.41 1.32 P-value 0.29 0.34 0.01 0.02 0.06 0.13 0.12 Pooled SEM 0.23 0.26 0.52 0.35 0.56 0.52 0.29 A,BMeans with no common superscript within a gene and a tissue differ significantly (P <0.05).

67

Table 5.5: Day 7 mRNA Abundance of Avian Beta-Defensins in Campylobacter jejuni Challenged and Non-Challenged Young Broilers. Tissue Treatment AvBD1 AvBD6 AvBD8 AvBD10 AvBD11 AvBD12 AvBD13 LEAP2 Duodenum CON 1.08 1.13 1.19 1.08B 1.17 1.13 1.22 1.03 L-106 1.19 1.84 1.16 2.03A 1.82 1.54 1.19 0.92 M-107 1.59 1.63 1.05 1.18AB 1.33 0.96 0.66 0.73 H-108 1.25 1.42 0.96 1.64AB 1.72 1.12 0.88 0.44 P-value 0.93 0.80 0.95 0.02 0.59 0.21 0.25 0.16 Pooled SEM 0.56 0.52 0.32 0.22 0.38 0.19 0.22 0.18

Jejunum CON 1.13 1.19 1.20 1.31 1.19 1.14 1.28 1.04 L-106 1.21 0.75 0.94 1.32 1.84 1.41 1.20 0.75 M-107 1.28 0.58 1.01 0.58 0.88 0.64 0.55 0.69 H-108 0.62 0.39 0.60 0.47 1.98 0.59 0.46 0.52 P-value 0.64 0.07 0.51 0.051 0.11 0.40 0.07 0.06 Pooled SEM 0.39 0.21 0.33 0.29 0.52 0.23 0.25 0.13

Ileum CON 1.10 1.40 1.24 1.13A 1.03 1.03 1.14 1.01 L-106 0.80 1.28 0.81 0.59AB 0.64 0.95 0.87 1.36 M-107 0.77 0.77 0.76 0.42B 0.43 0.50 0.44 1.33 H-108 0.66 0.70 0.54 0.49B 1.00 0.68 0.61 0.44 P-value 0.32 0.51 0.24 0.02 0.04* 0.11 0.11 0.07 Pooled SEM 0.17 0.39 0.24 0.15 0.15 0.16 0.20 0.25

Cecum CON 1.05 0.76 1.03 1.10AB 1.30 1.12A 1.11AB 1.18A L-106 1.10 0.84 2.29 1.19A 1.06 0.85AB 1.55A 0.71AB M-107 1.93 1.81 1.56 0.71AB 0.60 0.58AB 0.77B 0.34B H-108 0.66 0.49 0.53 0.40B 0.38 0.39B 0.53B 0.19B P-value 0.30 0.11 0.12 0.02 0.06 0.03 0.01 0.01 Pooled SEM 0.47 0.38 0.50 0.18 0.24 0.16 0.18 0.19 A,BMeans with no common superscript within a gene and a tissue differ significantly (P <0.05). *Values could not be further separated using Tukey’s Test.

68

Table 5.6: Day 14 mRNA Abundance of Nutrient and Sugar Transporters in Campylobacter jejuni Challenged and Non-Challenged Young Broilers. Tissue Treatment EAAT3 bo,+AT ZnT1 GLUT2 CAT2 LAT1 SGLT1 Duodenum CON 1.08 1.03AB 1.04 1.06 1.02B 1.03 1.05 L-106 1.31 1.81A 0.97 1.57 1.36AB 1.43 1.21 M-107 0.95 0.82B 0.93 0.74 1.85A 1.56 1.38 H-108 0.66 0.97B 1.44 0.55 1.59AB 1.65 1.68 P-value 0.07 0.02 0.17 0.07 0.03 0.34 0.33 Pooled SEM 0.16 0.21 0.16 0.26 0.18 0.25 0.24

Jejunum CON 1.05AB 1.04B 1.09 1.03AB 1.01 1.01 1.02 L-106 1.22A 2.07A 1.65 1.46A 0.96 1.15 1.74 M-107 0.89AB 1.17B 1.34 1.07AB 0.97 0.98 1.62 H-108 0.59B 1.16B 1.50 0.65B 1.54 1.27 1.50 P-value 0.03 0.001 0.53 0.004 0.13 0.67 0.24 Pooled SEM 0.14 0.16 0.28 0.13 0.19 0.19 0.25

Ileum CON 1.09B 1.04B 1.07B 1.05B 1.11 1.05 1.08B L-106 2.29A 3.39A 2.42A 2.66A 1.33 1.19 2.71A M-107 1.45AB 1.26B 2.09AB 0.99B 1.26 1.21 2.46AB H-108 1.79AB 1.54B 2.07AB 0.77B 1.51 1.34 2.97A P-value 0.03 <.0001 0.04 0.01 0.53 0.49 0.01 Pooled SEM 0.26 0.25 0.31 0.34 0.19 0.13 0.38

Cecum CON 1.06 1.08 1.11 1.09AB 1.08 1.01 1.13 L-106 1.46 1.43 1.72 1.30A 0.75 1.32 2.34 M-107 1.48 1.16 1.64 0.53AB 0.91 0.94 2.44 H-108 0.99 0.98 2.53 0.48B 0.96 1.22 2.30 P-value 0.24 0.39 0.14 0.02 0.41 0.32 0.24 Pooled SEM 0.2 0.19 0.40 0.20 0.13 0.16 0.50 A,BMeans with no common superscript within a gene and a tissue differ significantly (P <0.05).

69

Table 5.7: Day 14 mRNA Abundance of Avian Beta-Defensins in Campylobacter jejuni Challenged and Non-Challenged Young Broilers. Tissue Treatment AvBD1 AvBD6 AvBD8 AvBD10 AvBD11 AvBD12 AvBD13 LEAP2 Duodenum CON 1.16AB 1.12AB 1.05A 1.05AB 1.02AB 1.03 1.09AB 1.04 L-106 1.29A 1.54A 1.51A 1.64A 1.33A 1.35 1.37A 0.87 M-107 1.30A 1.10AB 1.01A 1.02AB 1.17AB 1.78 1.21AB 0.64 H-108 0.34B 0.37B 0.38B 0.56B 0.49B 0.5 0.55B 1.30 P-value 0.01 0.01 0.0003 0.02 0.04 0.20 0.02 0.25 Pooled SEM 0.19 0.20 0.13 0.20 0.19 0.41 0.17 0.22

Jejunum CON 1.06A 1.09A 1.10A 1.01 1.04 1.06 1.01 1.06 L-106 0.66AB 0.72AB 0.87AB 1.06 0.96 0.75 1.23 1.09 M-107 0.74AB 0.67AB 0.75AB 0.80 0.60 0.90 1.19 1.09 H-108 0.28B 0.32B 0.35B 0.48 0.40 0.32 0.47 1.26 P-value 0.01 0.02 0.02 0.051 0.050 0.09 0.18 0.95 Pooled SEM 0.14 0.15 0.15 0.15 0.17 0.20 0.26 0.26

Ileum CON 1.13B 1.16B 1.17B 1.06B 1.11B 1.09B 1.13B 1.09 L-106 3.18A 3.43A 3.61A 3.82A 3.92A 3.38A 4.11A 2.18 M-107 1.47AB 1.37B 1.39B 1.57B 1.59B 1.55AB 1.36B 2.53 H-108 1.24AB 1.29B 1.49B 1.87B 1.94B 1.99AB 2.21AB 0.89 P-value 0.03 0.001 0.01 0.001 0.001 0.02 0.01 0.08 Pooled SEM 0.50 0.37 0.47 0.41 0.41 0.47 0.56 0.48

Cecum CON 1.16B 1.02B 1.20B 1.08C 1.08C 1.07B 1.09B 1.14AB L-106 4.45A 4.44A 4.91A 7.02A 5.66A 4.88A 5.94A 3.35AB M-107 3.30AB 1.89B 2.28B 3.62BC 4.09AB 3.82AB 3.9A 1.01B H-108 2.20AB 2.01B 2.34B 5.44AB 3.36B 4.18A 4.28A 3.97A P-value 0.01 0.001 0.003 0.0003 0.0003 0.01 0.0002 0.02 Pooled SEM 0.63 0.46 0.59 0.75 0.56 0.75 0.57 0.72 A,B,CMeans with no common superscript within a gene and a tissue differ significantly (P <0.05)

70

Figure 5.1: Treatment layout*

6 L=10 H=108

Passageway

Control M=107

*Not Drawn to Scale Door

71

Figure 5.2: Campylobacter Cecal Recovery in Young Broilers 7, 14, 21, and 28 Days after Challenge 9 A A A A 8 AB A A A A B A 7 A 6

5

4

Log 10 cfus10 Log 3

2 B 1 C B B 0 7 14 21 28 Day

CON L M H

A,B,CMeans with no common superscript differ significantly CON=Control, L=106 challenge group, M= 107 challenge group, H=108 challenge group

72

Chapter 6

In situ Hybridization of AvBD-8 and AvBD-10 in the Ileum of

Campylobacter-Challenged Young Broilers

Abstract

Campylobacter is a bacterium ubiquitous to the gastrointestinal tract of poultry. Avian β- defensins (AvBDs) are small cationic peptides, which are effective against bacteria. In the current study, the location of AvBD8 and AvBD10 mRNA expression in the ileum of Campylobacter- infected and non-infected broiler chickens was examined using the in situ hybridization RNAscope method. At day 7, the expression of AvBD8 and AvBD10 mRNA was detected in cells dispersed throughout the intestinal villus and crypt. At day 14, expression of AvBD8 and AvBD10 mRNA was less intense than day 7. The expression pattern was similar as day 7 with cells expressing

AvBD8 and AvBD10 mRNA dispersed throughout the intestinal villus and crypt. No obvious differences between Campylobacter-infected and non-infected birds was observed in the current study. Because this was a pilot study with a small sample size, additional samples would need to be analyzed to determine the influence Campylobacter may have on location and cell expression of AvBDs in broiler chickens.

Introduction

Campylobacter is a zoonotic bacterium commonly found in the gastrointestinal tract of poultry, and is a major cause of bacterial foodborne illness in the United States (Hoffmann et al.,

2012). Campylobacter induces a pro-inflammatory immune response in birds during an infection

73

(Vaezirad et al., 2016). However, the immune response is self-limiting and the pathogen remains in the lower portion of the gastrointestinal tract, where it can colonize and proliferate.

Avian β-defensins (AvBDs) are a class of host defense peptides (HDPs), which can be effective against Gram-negative and Gram-positive bacteria as well as viruses and fungi (Cuperus et al., 2013). Fourteen AvBDs have been described in chickens (Evans et al., 1995). Cathelicidins, another class of HDPs, have been shown to be effective against Campylobacter in vitro (van Dijk et al., 2012). However, little is known about the cells that express AvBDs in the gastrointestinal tract of chickens. The objective of this study was to identify cells that express AvBD8 and

AvBD10 mRNA in the ileum of chickens.

Materials and Methods

Tissue Processing

Tissue samples used in this study were those described and collected in Chapter 5 where

1- to 1.5-cm2 long ileal samples were stored in RNAlater® for ~2 years (Qiagen, Germantown,

MD). Tissues were placed into phosphate buffered 4% paraformaldehyde for 24 hours. The tissues were then stored in 70% ethanol at 4°C for 24 hours after fixation and then sent to the

Virginia Tech Veterinary Teaching Hospital Histopathology Lab (Blacksburg, VA) for paraffin embedding. The paraffin blocks were cut into 4- to 6-µm sections with a microtome (Microm HM

355S, Thermo-Fisher Scientific, Waltham, MA) and stored at room temperature.

In situ Hybridization Analysis

The in situ hybridization (ISH) technique was performed using the RNAscope kit and method (Advanced Cell Diagnostics Inc., ACD, Newark, CA) described by Wang et al., (2012).

74

Ileal tissues were collected from two different chickens from both control and infected groups on day 7 and day 14. Only two samples were examined because the integrity of the tissues was not optimal because tissues were not processed and stored for the RNAscope technique at the time of collection. A set of probes for chicken AvBD8 and AvBD10 were custom designed by ACD. The ileal samples were processed in accordance to the manufacturer’s directions using the HybEZ oven and RNAscope HD Detection reagent kits. The ileal tissues within one time-point were pre-treated and hybridized as a group to minimize variation. The red chromogen was used for detection of

AvBD8 and AvBD10. After RNAscope processing, the slides were stained with a 50% Gill #2 hematoxylin solution (Sigma Aldrich, St. Louis, MO), rinsed in distilled water and then placed into 0.02% ammonia water. After slides were allowed to air dry, a drop of VectaMount™ permanent mounting medium (Vector Laboratories, Burlingame, CA) was added and a coverslip was applied on top. Images were captured with a Nikon Eclipse 80i microscope and DS-Ri1 digital camera.

Results

Cells expressing AvBD8 and AvBD10 mRNA in the ileum of non-challenged and

Campylobacter-challenged birds at day 7 and day 14 post-challenge were identified by in situ hybridization. At day 7, AvBD-8 and AvBD-10 mRNA were detected in the ileum; however, the expression appeared to be variable between control and challenged birds. Strong expression of both AvBD8 and AvBD10 mRNA was seen in the ileum of the control #1 and challenged #1

(Figure 6.1). AvBD8 and AvBD10 mRNA were located in cells of the intestinal villi and crypts.

Expression of AvBD8 and AvBD10 mRNA in the replicate day 7 samples (control #2 and challenged #2) was lower in comparison to the first sample (Figure 6.2). However, expression of

75

AvBD8 and AvBD10 mRNA was still in cells throughout the intestinal villi and crypts. At day

14, some expression of AvBD8 and AvBD10 mRNA was detected. In Figure 6.3, AvBD8 and

AvBD10 mRNA were detected at the base of the intestinal villi and in the crypt at a lower level.

Expression of AvBD8 and AvBD10 was not detected in the ileum of the challenged bird (#1). In

Figure 6.4, AvBD8 and AvBD10 mRNA was again lowly expressed in the intestinal villi and crypt of control #2. In the challenged bird, expression of AvBD8 and AvBD10 mRNA appeared throughout the intestinal villi and crypt.

DISCUSSION

Avian beta-defensins are an important first line of defense against pathogens as part of the innate immune system. Classified as host defense peptides, AvBDs are synthesized in the host phagocytic and mucosal epithelial cells (Zhang and Sunkara, 2014). They are produced and secreted as precursors and can be processed into mature peptides upon an infection. The objective of this study was to localize the expression of AvBD8 and AvBD10 in non-infected and

Campylobacter-infected broilers. Due to the storage and the age of the ileal tissue samples, intestinal villus and crypt integrity was not ideal for the RNAscope method. However, detection of AvBD8 and AvBD10 showed some success. Variation of expression of AvBD8 and AvBD10 mRNA can be observed between samples within the control or challenge group. This variation could be due to individual variabilities in the immune responses to the pathogen. On day 7, expression of AvBD8 and AvBD10 mRNA appeared to be strong and dispersed within the intestinal villus and crypt. On day 14, the intensity of expression for AvBD8 and AvBD10 appeared to be lower than that seen on day 7; however, the distribution pattern appeared to be similar. Detection of AvBD8 and AvBD10 mRNA in the intestinal tissue was consistent with a

76 previous report showing expression of AvBD8 and AvBD10 by qPCR (Hong et al., 2012). AvBD8 and AvBD10 may play an important role in the intestinal immune response against pathogens.

In mammals, paneth cells, secretory cells found in the epithelium of the small intestines, are the primary producers of HDPs in the gastrointestinal tract (Bevins and Salzman, 2011).

Paneth cells are located at the base of the crypts of Lieberkühn. Although paneth-like cells have been recently described in chickens, little is known about their function (Wang et al., 2016).

Localization of LEAP2 in chickens was found in the enterocytes of the small intestines (Su et al.

2017). Host defense peptides may also be produced by the bone marrow and circulating phagocytic cells (Diamond et al., 2009). In the current study, localization of AvBD8 and AvBD10 mRNA was not limited to the crypts of Lieberkühn and expression was observed throughout the intestinal villus and crypts. This distribution in expression of AvBD8 and AvBD10 may indicate that the production of AvBDs may occur throughout the villus and crypts by various epithelial cells or may be of myeloid origin and circulate in the blood to the intestinal villus.

77

AvBD8

AvBD10 0

Figure 6.1: Expression profile of AvBD8 and AvBD10 in the ileum of control and Campylobacter-challenged broilers at Day 7. Red dots show the location of AvBD8 and AvBD10 mRNA.

78

AvBD8

AvBD10

Figure 6.2: Expression profile of AvBD8 and AvBD10 in the ileum of control and Campylobacter-challenged broilers at Day 7. Red dots show the location of AvBD8 and AvBD10 mRNA.

79

AvBD8

AvBD10

Figure 6.3: Expression profile of AvBD8 and AvBD10 in the ileum of control and Campylobacter-challenged broilers at Day 14. Red dots show the location of AvBD8 and AvBD10 mRNA.

80

AvBD8

AvBD10

Figure 6.4: Expression profile of AvBD8 and AvBD10 in the ileum of control and Campylobacter-challenged broilers at Day 14. Red dots show the location of AvBD8 and AvBD10 mRNA.

81

Chapter 7

Epilogue

Salmonella challenge may have early and late influences on the expression of nutrient transporters and host defense peptides (HDPs) in broiler chickens. Early changes in nutrient transporter mRNA abundance could signify a possible response by the host to limit the pathogen invasion of the enterocytes. The efflux of zinc out of the enterocyte may signify another host response in the use of zinc in a possible immune response to the pathogen. Further research is needed to examine the role that zinc may play during a Salmonella challenge. Salmonella was also shown to have a late influence on the abundance of HDPs in the lower gastrointestinal tract.

This may indicated that early attempts to control Salmonella were unsuccessful and once

Salmonella intracellularly invades the host enterocytes, HDPs become ineffective, leading to the decrease of HDPs production and secretion.

Campylobacter may have late influences on the expression of nutrient transporters and

HDPs in broiler chickens. Zinc may also play a role during the start of a Campylobacter challenge as an increase in ZnT1 mRNA abundance was observed. Changes in nutrient transporter mRNA abundance could signify a possible response by the host to limit the nutrients available on the apical side of the enterocytes. However, results from mRNA abundance of AvBDs may indicate an attempt by the host to induce an immune response to Campylobacter at a low challenge level.

Salmonella and Campylobacter are of great concern to the commercial poultry industry.

Commonly found in the gastrointestinal tract of poultry, these pathogens may exist within a poultry flock without being detected. The results of this dissertation may bring forth possible areas of future research that could provide a better understanding of Salmonella and Campylobacter

82 pathogenesis in poultry and in the development of new possible intervention strategies in controlling these pathogens.

83

LITERATURE CITED

Akbari, M.R., H.R. Haghighi, J.R. Chambers, J. Brisbin, L.R. Reed, and S. Sharif. 2008.

Expression of antimicrobial peptides in cecal tonsils of chickens treated with probiotics

and infected with Salmonella enterica serovar typhimurium. Clin. Vaccine Immunol.

15:1689-1693.

Andino, A., and I. Hanning. 2015. Salmonella enterica: Survival, colonization, and virulence

differences among serovars. Sci. World J. Article ID 520179:1-16.

Awad, W.A., E. Razzazi-Fazeli, J. Böhm, and J. Zentek. Influence of deoxynivalenol on the D-

glucose transport across the isolated epithelium of different intestinal segments of laying

hens. J. Anim. Physiol. Anim. Nutr. 91:175-180.

Awad, W.A., J.R. Aschenbach, K. Ghareeb, B. Khayal, C. Hess, and M. Hess. 2014.

Campylobacter jejuni influences the expression of nutrient transporter genes in the

intestine of chickens. Vet. Microbiol. 172:195-201.

Bailey, J.S., N.A. Cox, and M.E. Berrang. 1994. Hatchery-acquired Salmonellae in broiler chicks.

Poult. Sci. 73:1153-1157.

Bailey, J.S., N.J. Stern, P. Fedorka-Cray, S.E. Craven, N.A. Cox, D.E. Cosby, S. Ladely, and M.T.

Musgrove. 2001. Sources and movement of Salmonella through integrated poultry

operations: A multistate epidemiological investigation. J. Food Prot. 64:1690-1697.

Baron, F., F. Nau, C. Guerin-Dubiard, S. Bonnassie, M. Gautier, S.C. Andrews and S. Jan. 2016.

Egg white versus Salmonella Enteritidis! A harsh medium meets a resilient pathogen. Food

Microbiol. 53:82-93.

Bäumler, A.J., B.M. Hargis, and R.M. Tsolis. 2000. Tracing the origins of Salmonella outbreaks.

Science 287:50-52.

84

Barrow, P.A. 2007. Salmonella infections: immune and non-immune protection with vaccines.

Avian Pathol. 36:1-13.

Bevins, C.L., and N.H. Salzman. 2011. Paneth cells, antimicrobial peptides and maintenance of

intestinal homeostatsis. Nat. Rev. Microbol. 9:356-368.

Bolton, D.J., 2015. Campylobacter virulence and survival factors. Food Micro. 48:99-108

Buhr, R.J., N.A. Cox, N.J. Stern, M.T. Musgrove, J.L. Wilson, and K.L. Hiett. 2002. Recovery of

Campylobacter from segments of the reproductive tract of broiler breeder hens. Avian Dis.

46:919-924

Byrd, J.A., D.E. Corrier, M.E. Hume, R.H. Bailey, L.H. Stanker, and B.M. Hargis. 1998. Effect of

feed withdrawal on Campylobacter in the crops of market-age broiler chickens. Avian Dis.

42:802-806.

Byrd, J., R. H. Bailey, R. Wills, and D. Nisbet. 2007. Recovery of Campylobacter from

commercial broiler hatchery trayliners. Poult. Sci. 86:26-29.

Byrne, C.M., M. Clyne, and B. Bourke. 2007. Campylobacter jejuni adhere to and invade chicken

intestinal epithelial cells in vitro. Microbiology. 153:561-569.

Cason, J.A., N.A. Cox, and J.S. Bailey. 1994. Transmission of Salmonella typhimurium during

hatching of broiler chicks. Avian Dis. 38:583-588.

Casterlow, S., H. Li, E.R. Gilbert, R.A. Dalloul, A.P. McElroy, D.A. Emmerson, and E.A. Wong.

2011. An antimicrobial peptide is downregulated in the small intestine of Eimeria maxima-

infected chickens. Poult. Sci. 90:1212-1219.

Clark, M.A., M.A. Jepson, N.L. Simmons, and B.H. Hirst. 1994. Preferential interactions of

Salmonella typhimurium with mouse Peyer’s patch M cells. Res. Microbiol. 145:543-552.

Clench, M.H. 1999. The avian cecum: Update and motility review. J. Exp. Zool. 283:441-447.

85

Corrier, D.E., A.J. Hinton, R.L. Ziprin, R.C. Beier, and J.R. DeLoach. 1990. Effect of dietary

lactose on cecal pH, bacteriostatic volatile fatty acids and Salmonella typhimurium

colonization of broiler chicks. Avian Dis. 34:617-625.

Cox, N. A., J.S. Bailey, J.M. Mauldin, and L.C. Blankenship. 1990. Presence and impact of

Salmonella contamination in commercial broiler hatcheries. Poult. Sci. 69:1606-1609.

Cox, N.A., J.S. Bailey, J.M. Mauldin, L.C. Blankenship, and J.L. Wilson. 1991. Extent of

Salmonellae contamination in breeder hatcheries. Poult. Sci. 70:416-418.

Cox, N., M. Berrang, and J. Cason. 2000. Salmonella penetration of egg shells and proliferation in

broiler hatching eggs--a review. Poult. Sci. 79:1571-1574.

Cox, N. A., J. S. Bailey, L. J. Richardson, R. J. Buhr, D. E. Cosby, J. L. Wilson, K. L. Hiett, G. R.

Siragusa, and D. V. Bourassa. 2005. Presence of naturally occurring Campylobacter and

Salmonella in the mature and immature ovarian follicles of late-life broiler breeder hens.

Avian Dis. 49:285-287.

Crawley, S.W., M.S. Mooseker, and M.J. Tyska. 2014. Shaping the intestinal brush border. J. Cell

Biol. 207:441-451.

Cuperus, T., M. Coorens, A. van Dijk, and H.P. Haagsman. 2013. Avian host defense peptides.

Dev. Comp. Immunol. 41:352-369.

Diamond, G., B. Beckloff, A. Weinberg, and K.O. Kisich. 2009. The role of antimicrobial peptides

in innate host defense. Curr. Pharm. Des. 15:2377-2392.

Epps, S.V.R., R.B. Harvey, M.E. Hume, T.D. Phillips, R.C. Anderson, and D.J. Nisbet. 2013.

Foodborne Campylobacter: infections, metabolism, pathogenesis and reservoirs. Int. J.

Environ. Res. Public Health. 10:6292-6304.

Erf, G.F. 2004. Cell-mediated immunity in poultry. Poult. Sci. 83:580-590.

86

Evans, E.W., G.G. Beach, J. Wunderlich, and B.G. Harmon. 1994. Isolation of antimicrobial

peptides from avian heterophils. J. Leukocyte Biol. 56:661-665.

Evans, E.W., F.G. Beach, K.M. Moore, M.W. Jackwood, J.R. Glisson, and B.G. Harmon. 1995.

Antimicrobial activity of chicken and turkey heterophil peptides CHP1, CHP2, THP1, and

THP3. Vet. Microbiol. 47:295-303.

Foley, S. L., A. M. Lynne, and R. Nayak. 2008. Salmonella challenges: Prevalence in swine and

poultry and potential pathogenicity of such isolates. J. Anim. Sci. 86:149-162.

Foley, S.L., R. Nayak, I.B. Hanning, T.J. Johnson, J. Han, and S.C. Ricke. 2011. Population

dynamics of Salmonella enterica serotypes in commercial egg and poultry production.

Appl. Envir. Microbiol. 77:4273-4279.

Fonseca, B.B., M.E. Beletti, R.T. de Melo, E.P. Mendonca, L.R. Coelho, P.C. Nalevaiko, and D.A.

Rossi. 2014. Campylobacter jejuni in commercial eggs. Braz. J. Microbiol. 45:76-79.

Fotiadis, D., Y. Kanai, and M. Palacin. 2013. The SLC3 and SLC7 families of amino acid

transporters. Mol. Aspects Med. 34:139-158.

FSIS. 2016. Changes to the Salmonella and Campylobacter Verification Testing Program:

Proposed Performance Standards for Salmonella and Campylobacter in Not-Ready-to-Eat

Comminuted Chicken and Turkey Products and Raw Chicken Parts and Related Agency

Verification Procedures and Other Changes to Agency Sampling. Pages 3940-3950 USDA

ed. FSIS, Federal Register.

Ghareeb, K., W.A. Awad, M. Mohnl, R. Porta, M. Biarnes, J. Bohm, and G. Schatzmayr. 2012.

Evaulating the efficact of an avian-specific probiotic to reduce the colonization of

Campylobacter jejuni in broiler chickens. Poult. Sci. 91:1825-1832.

87

Gielda, L.M. and V.J., DiRita. 2012. Zinc competition among intestinal microbiota. mBio.

3(4):00171-12.

Grant, A., F. Hashem, and S. Parveen. 2016. Salmonella and Campylobacter: Antimicrobial

resistance and bacteriophage control in poultry. Food Microbiol. 53:104-109.

Guerry, P., C.P. Ewing, T.E. Hickey, M.M. Prendergast, and A.P. Moran. 2000. Sialylation of

lipooligosaccharide core affects immunogenicity and serum resistance of Campylobacter

jejuni. Infect. Immun. 68:6656-6662.

Hediger, M.A., and D.B. Rhoads. 1994. Molecular physiology of sodium-glucose .

Physiol. Rev. 74:993-1026.

Hediger, M.A., M.F. Romero, J.B. Peng, A. Rolfs, H. Takanaga, and E.A. Bruford. 2004. The

ABCs of solute carriers: physiological, pathological and therapeutic implications of human

membrane transport proteins. Introduction. Pflugers Arch. 447:465-468.

Hediger, M.A., B. Clemencon, R.E. Burrier, and E.A. Bruford. 2013. The ABCs of membrane

transporters in health and disease (SLC series): Introduction. Mol. Aspects Med. 34:95-

107.

Hiett, K. ., N.A. Cox, R.J. Buhr, and N.J. Stern. 2002. Genotype analyses of Campylobacter

isolated from distinct segments of the reproductive tracts of broiler breeder hens. Curr.

Microbiol. 45:400-404.

Hiett, K.L., G.R. Siragusa, N.A. Cox, R.J. Buhr, M.T. Musgrove, N.J. Stern, and J. L. Wilson.

2003. Genotype analyses of Campylobacter isolated from the gastrointestinal tracts and the

reproductive tracts of broiler breeder roosters. Avian Dis. 47:406-414.

88

Hoffmann, S., M.B. Batz, and J.G. Morris. 2012. Annual cost of illness and quality-adjusted life

year losses in the United States due to 14 foodborne pathogens. J. Food Prot. 75:1292-

1302.

Hong, Y.H., W. Song, S.H. Lee, and H.S. Lillehoj. 2012. Differential gene expression profiles of

β-defensins in the crop, intestine, and spleen using a necrotic enteritis model in 2

commercial broiler chicken lines. Poult. Sci. 91:1081-1088.

Humphrey, T. 2006. Are happy chickens safer chickens? Poultry welfare and disease

susceptibility. Br. Poult. Sci. 47:379-391.

Humphrey, T., S. O’Brien, and M. Madsen. 2007. Campylobacters as zoonotic pathogens: a food

production perspective. Int. J. Food Microbiol. 117: 237-257.

Imondi, A.R., and F.H. Bird. 1966. The turnover of intestinal epithelium in the chick. Poult. Sci.

45:142-147.

Kaiser, P. 2010. Advances in avian immunology-prospects for disease control: a review. Avian

Pathol. 39:309-324.

Kanai, Y., and M.A. Hediger. 2004. The glutamate/neutral amino acid transporter family SLC1:

molecular, physiological and pharmacological aspects. Pflugers Arch. 447:469-479.

Keller, L. ., C.E. Benson, K. Krotec, and R.J. Eckroade. 1995. Salmonella enteritidis colonization

of the reproductive tract and forming and freshly laid eggs of chickens. Infect. Immun.

63:2443-2449.

Kohl, K.D. M.E., Ciminari, J.G. Chediack, J.O. Leafloor, W.H. Karasov, S.R. McWilliams, and

E. Caviedes-Vidal. 2017. Modulation of digestive enzyme activities in the avian digestive

tract in relation to diet composition and quality. J. Comp. Physiol. B. 187:339-351.

89

Korver, D.R. 2006. Overview of immune dynamics of the digestive system. J. Appl. Poult. Res.

15:123-135.

Krause, A., R. Sillard, B. Kleemeier, E. Kluver, E. Maronde, J.R. Conejo-Garcia, W.G.

Forssmann, P. Schulz-Knappe, M.C. Nehls, F. Wattler, S. Wattler, and K. Adermann. 2003.

Isolation and biochemical characterization of LEAP-2, a novel blood peptide expressed in

the liver. Protein Sci. 12:143-152.

Lee, M.O., H. Jang, D. Rengaraj, S.Yang, J.Y. Han, S.J. Lamont and J.E. Womack. 2016. Tissue

expression and antibacterial activity of host defense peptides in chicken. BMC Vet. Res.

12:231.

Lee, S.I., H.J. Jang, H. Jeon, M.O. Lee, J.S. Kim, I. Jeon, and S.J. Byun. 2016. Transcriptional

regulation of cathelicidin gene in chicken bone marrow cells. Poult. Sci. 95:912-919.

Lehrer, R.I., and T. Ganz. 2002. Defensins of vertebrate animals. Curr. Opin. Immunol. 14:96-

102.

Livak, K.J., and T.D. Schmittgen. 2001. Analysis of relative gene expression data using real-time

quantitative PCR and the 2-ΔΔCT method. Methods 25:402-408.

Loeffler, H., A. Birke, D. Winton, and C. Potten. 1993. Somatic mutation, monoclonality and

stochastic models of stem cells organization in the intestinal crypt. J. Theor. Biol. 160:471-

491.

Lynn, D.J., R. Higgs, S. Gaines, J. Tierney, T. James, A.T. Lloyd, M.A. Fares, G. Mulcahy, and

C. O’Farrelly. 2004. Bioinformatic discovery and initial characterisation of nine novel

antimicrobial peptide genes in the chicken. Immunogenetics 56:170-177.

Malik-Kale, P., C.E. Jolly, S. Lathrop, S. Winfree, C. Luterbach, and O. Steele-Mortimer. 2011.

Salmonella- at home in the host cell. Front. Microbiol. 125:1-9.

90

Matulova M., K. Varmuzova, F. Sisak, H. Havlickova, V. Babak, K. Stejskal, Z. Zdrahal, and I.

Rychlik. 2013. Chicken innate immune response to oral infection with Salmonella enterica

serovar Enteritidis. Vet. Res. 44:37-48.

Meade, K.G., F. Narciandi, S. Cahalane, C. Reiman, B. Allan and C. O’Farrelly. 2009.

Comparative in vivo infection models yield insights on early host immune response to

Campylobacter in chickens. Immunogenetics 61:101-110.

Milona, P., C.L. Townes, R.M. Bevan, and J. Hall. 2007. The chicken host peptides, gallinacins 4,

7, and 9 have antimicrobial activity againist Salmonella serovars. Biochem. Biophys. Res.

Commun. 356:476-479.

Mithieux, G., P. Misery, C. Magnan, B. Pillot, A. Gautier-Stein, C. Bernard, F. Rajas, and C.

Zitoun. 2005. Portal sensing of intestinal gluconeogenesis is a mechanistic link in the

diminution of food intake induced by diet protein. Cell Metab. 2: 321-329.

Moore, J.E., D. Corcoran, J.S. Dooley, S. Fanning, B. Lucey, M. Matsuda, D.A. McDowell, F.

Megraud, B.C. Millar, R. O’Mohony, L. O’Riordan, M. O’Rourke, J.R. Rao, P.J. Rooney,

A. Sails, and P. Whyte. 2005. Campylobacter. Vet. Res. 36:351-382.

Mowat, A.M., and W.W. Agace. 2014. Regional specialization within the intestinal immune

system. Nat. Rev. Immunol. 14:667-685.

Muller, A.J., P. Kaiser, K.E. Dittmar, T.C. Weber, S. Haueter, and K. Endt. 2012. Salmonella gut

invasion involves TTSS-2 dependent epithelial traversal, basolateral exit, and uptake by

epithelium-sampling lamina propria phagocytes. Cell Host Microbe 11:19-32.

Musgrove, M.T., N.J. Stern, J.S. Bailey. 1997. A comparsion of enrichment methods for the

recovery of Campylobacter spp. in broiler litter samples. Poult. Sci. 76 (Suppl. 1):114.

(Abstr.)

91

Newell, D.G., and C. Fearnley. 2003. Sources of Campylobacter colonization in broiler chickens.

Appl. Environ. Microbiol. 69:4343-4351.

Noy, Y., and D. Sklan. 1995. Digestion and absorption in the young chick. Poult. Sci. 74:366-373.

Palacin, M., and Y. Kanai. 2004. The ancillary proteins of HATs: SLC3 family of amino acid

transporters. Eur. J. Phys. 447:490-494.

Panigrahi, P., G. Losonsky, L.J. DeTolla, and J.G. Morris Jr. 1992. Human immune response to

Campylobacter jejuni proteins expressed in vivo. Infect. Immun. 60:4938-4944.

Paris, N.E., and E.A. Wong. 2013. Expression of digestive enzymes and nutrient transporters in

the intestine of Eimeria maxima-infected chickens. Poult. Sci. 92:1331-1335.

Quershi, M.A. 2003. Avian macrophage and immune response: An overview. Poult. Sci. 82:691-

698.

Riesenfeld, G., D. Sklan, A. Bar, U. Eisner, and S. Hurwitz. 1980. Glucose absorption and starch

digestion in the intestine of the chicken. J. Nutr. 110:117-121.

Sahin O., T.Y. Morishita, and Q. Zhang. 2003. Detection and survival of Campylobacter in

chicken eggs. J. Appl. Microbiol. 95:1070-1079.

Sauter, E. A., and C. F. Petersen. 1974. The effect of egg shell quality on penetration by various

Salmonellae. Poult. Sci. 53:2159-2162.

Semple, F., H. MacPherson, S. Webb, S.L. Cox, L.J. Mallin, C. Tyrell, G.R. Grimes, C.A. Semple,

M.A. Nix, G.L. Millhauser, and J.R. Dorin. 2011. Human beta-defensin 3 affects the

activity of pro-inflammatory pathways associated with MyD88 and TRIF. Eur. J. Immunol.

41:3291-3300.

92

Setta, A., P.A. Barrow, P. Kaiser, and M.A. Jones. 2012. Immune dynamics following infection

of avian macrophages and epithelial cells with typhoidal and non-typhoidal Salmonella

enterica serovars; bacterial invasion and persistence, nitric oxide and oxygen production,

differential host gene expression, NF-κB signaling and cell cytotoxicity. Vet. Immunol.

Immunopathol. 146:212-224.

Schlessinger, A., P. Matsson, J.E. Shima, U. Peiper, S.W. Yee, L. Kelly, L. Apeltsin, R.M. Stroud,

T.E. Ferrin, K.M. Giacomini, and A. Sali. 2010. Comparison of human solute carriers.

Protein Sci. 19:412-428.

Skarp, C.P.A., M.L. Hänninen, and H.I.K. Rautelin. 2016. Campylobacteriosis: the role of poultry

meat. Clin. Microbiol. Infect. 22:103-109.

Sklan, D., S. Hurwitz, P. Budowski, and I. Ascarelli. 1975. Fat digestion and absorption in chicks

fed raw or heated soybean meal. J. Nutr. 105:57-63.

Snoeck, V., B. Goddeeris, and E. Cox. 2005. The role of enterocytes in the intestinal barrier

function and antigen intake. Micro. Infect. 7:997-1004.

Stern, N.J., B. Wojton, and K. Kwaitek. 1992. A differential-selective medium and dry ice-

generated atmosphere for recovery of Campylobacter jejuni. J. Food Prot. 55:514-517.

Su, S., K.B. Miska, R.H. Fetterer, M.C. Jenkins, and E.A. Wong. 2014. Expression of digestive

enzymes and nutrient transporters in Eimeria acervulina-challenged layers and broilers.

Poult. Sci. 93:1217-1226.

Su, S., K.B. Miska, R.H. Fetterer, M.C. Jenkins, and E.A. Wong. 2015. Expression of digestive

enzymes and nutrient transporters in Eimeria-challenged broilers. Exp. Parasitol. 150:13-

21.

93

Su, S., D.M. Dwyer, K.B. Miska, R.H. Fetterer, M.C. Jenkins, and E.A. Wong. 2017. Expression

of host defense peptides in the intestine of Eimeria-challenged chickens. Poult. Sci.

http://dx.doi.org/10.3382/ps/pew468.

Tako, E., P.R. Ferket, and Z. Uni. 2005. Changes in chicken intestinal zinc exporter mRNA

expression and small intestinal functionality following intra-amniotic zinc-methionine

administration. J. Nutr. Biochem. 16:339-346.

Tindall, B.J., P.A.D. Grimont, G.M. Garrity and J. Euzeby. 2005. Nomenclature and taxonomy of

the genus Salmonella. Int. J. Syst. Evol. Microbiol. 55:521-524.

Townes, C.L., G. Michailidis, C.J. Nile, and J. Hall. 2004. Induction of cationic chicken liver-

expressed antimicrobial peptide 2 in response to Salmonella enterica infection. Infect.

Immun. 12:6987-6993.

Townes, C.L., G. Michailidis, and J. Hall. 2009. The interaction of the antimicrobial peptide

cLEAP-2 and the bacterial membrane. Biochem. Biophys. Res. Comm. 387:500-503.

Timoney, J., H. Shivaprasad, R. Baker, and B. Rowe. 1989. Egg transmission after infection of

hens with Salmonella enteritidis phage type 4. Vet. Rec. 125:600-601.

Uldry, M. and B. Thorens. 2004. The SLC2 family of facilitated hexose and polyol transporters.

Pflugers Arch. 447:480-489.

Uni, Z., A. Geyra, H. Ben-Hur, and D. Sklan. 2000. Small intestinal development in the young

chick: Crypt formation and enterocyte proliferation and migration. Br. Poult. Sci. 41:544-

551.

Vaezirad, M.M., A.M. Keestra-Gounder, M.R. de Zoete, M.G. Koene, J.A. Wagenaar and J.P.M

van Putten. 2016. Invasive behavior of Campylobacter jejuni in immunosuppressed

chicken. Virulence 8:248-260.

94

Van Duen, K., F. Pasmans, R. Ducatelle, B. Flahou, K. Vissenberg, A. Martel, W. Van den Broeck,

F. Van Immerseel, and F. Haesebrouck. 2008. Colonization stragery of Campylobacter

jejuni results in persistent infection of the chicken gut. Vet. Microbiol. 130:285-297. van Dijk, A., E.J.A. Veldhuizen, and H.P. Haagsman. 2008. Avian defensins. Vet. Immunol.

Immunopathol. 124:1-18. van Dijk, A.J., E.M. Molhoek, F.J. Bikker, P.L. Yu, E.J. Velduizen, and H.P. Haagsman. 2011.

Avian cathelicidins: paradigms for the development of anti-infectives. Vet. Microbiol.

153:27-36. van Dijk, A., M. Herrebout, M.H.G. Tersteeg-Zijderveld, J.L.M. Tjeerdsma-van Bokhoven, N.

Bleumink-Pluym, A.J.M. Jansman, and E.J.A. Veldhuizen. 2012. Campylobacter jejuni is

highly susceptible to killing by chicken host defense peptide cathelicidin-2 and suppresses

intestinal cathelicidin-2 expression in young broilers. Vet. Microbiol. 160: 347-354.

Velge, P., A. Wiedemann, M. Rosselin, N. Abed, Z. Boumart, A.M. Chausse, O. Grepinet, F.

Namdari, S.M. Roche, A. Rossignol, and I. Virlogeux-Payant. 2012. Multiplicity of

Salmonella entry mechanisms, a new paradigm for Salmonella pathogenesis. Microbiol.

Open. 1:243-258.

Verrey, F., E. I. Closs, C. A. Wagner, M. Palacin, H. Endou, and Y. Kanai. 2004. CATs and HATs:

the SLC7 family of amino acid transporters. Pflugers Arch. l. 447:532-542.

Wagner, C. A., F. Lang, and S. Broer. 2001. Function and structure of heterodimeric amino acid

transporters. Am. J. Cell Physiol. 281:C1077-1093.

Wang, L., J. Li, J. Li Jr., R.X. Li, C.F. Lv, S. Li, Y.L. Mi, and C.Q. Zhang. 2016. Identification of

the Paneth cells in chicken small intestine. Poult. Sci. 95:1631-1635.

95

Wigley, P. 2013. Immunity of bacterial infection in the chicken. Dev. Comp. Immunol. 41:413-

417.

Wigley, P. 2014. Salmonella enterica in the chicken: How it has helped our understanding of

immunology in a non-biomedical model species. Front. Immunol. 5:482-489.

Williams, L.K., L.C. Sait, E.K. Trantham, T.A. Cogan, and T.J. Humphrey. 2013. Campylobacter

infection has different outcomes in fast and slow growing broiler chickens. Avian Dis.

57:238-241.

Withanage, G.S., P. Kaiser, P. Wigley, C. Powers, P. Mastroeni, H. Brooks, P. Barrow, A. Smith,

D. Maskell, and I. McConnell. 2004. Rapid expression of chemokines and

proinflammatory cytokines in newly hatched chickens infected with Salmonella enterica

serovar Typhimurium. Infect. Immun. 72: 2152-2159.

Xiao, Y., A.L. Hughes, J. Ando, Y. Matsuda, J.F. Cheng, D. Skinner-Noble, and G. Zhang. 2004.

A genome-wide screen identifies a single β-defensin gene cluster in the chicken:

Implications for the origin and evolution of mammalian defensins. BMC Genomics 13:56-

67.

Yin, H., L.H. Sumners, R.A. Dalloul, K.B. Miska, R.H. Fetterer, M.C. Jenkins, Q. Zhu, and E.A.

Wong. 2015. Changes in expression of an antimicrobial peptide, digestive enzymes, and

nutrient transporters in the intestine of E. praecox-infected chickens. Poult. Sci. 94:1521-

1526.

Yoshimura, Y., H. Ohashi, K. Subedi, M. Nishibori, and N. Isobe. Effects of age, egg-laying

activity, and Salmonella-inoculation on the expressions of gallinacin mRNA in the vagina

of the hen oviduct. J. Reprod. Dev. 52:211-218.

96

Young, K.T., L.M. Davis and V.J. DiRita. 2007. Campylobacter jejuni: molecular biology and

pathogenesis. Nature 5:665-679.

Zanetti, M. 2005. The role of cathelicidins in the innate host defenses of mammals. Curr. Issues.

Mol. Biol. 7:179-196.

Zhang, G. and L.T. Sunkara. 2014. Avian antimicrobial host defense peptides: From biology to

therapeutic applications. Pharmaceuticals 7:220-247.

97