USING RNA-SEQ TO CHARACTERIZE THE

BIOLOGICAL BASIS OF VARIATION

IN FEED EFFICIENCY

IN BROILER CHICKENS

by

Nan Zhou

A thesis submitted to the Faculty of the University of Delaware in partial fulfillment of the requirements for the degree of Master of Science in Animal Science

Spring 2015

© 2015 Nan Zhou All Rights Reserved

ProQuest Number: 1596914

All rights reserved

INFORMATION TO ALL USERS The quality of this reproduction is dependent upon the quality of the copy submitted.

In the unlikely event that the author did not send a complete manuscript and there are missing pages, these will be noted. Also, if material had to be removed, a note will indicate the deletion.

ProQuest 1596914

Published by ProQuest LLC (2015). Copyright of the Dissertation is held by the Author.

All rights reserved. This work is protected against unauthorized copying under Title 17, United States Code Microform Edition © ProQuest LLC.

ProQuest LLC. 789 East Eisenhower Parkway P.O. Box 1346 Ann Arbor, MI 48106 - 1346

USING RNA-SEQ TO CHARACTERIZE THE

BIOLOGICAL BASIS OF VARIATION

IN FEED EFFICIENCY

IN BROILER CHICKENS

by

Nan Zhou

Approved: ______Behnam Abasht, Ph.D. Professor in charge of thesis on behalf of the Advisory Committee

Approved: ______Limin Kung, Jr., Ph.D. Chair of the Department of Animal and Food Sciences

Approved: ______Mark W. Rieger, Ph.D. Dean of the College of Agriculture and Natural Resources

Approved: ______James G. Richards, Ph.D. Vice Provost for Graduate and Professional Education ACKNOWLEDGMENTS

The work was partly funded by Delaware Bioscience Center for Advanced Technology and Heritage Breeders, LLC. Thank to my collaborators, Behnam Abasht and William R Lee, for their great contributions on this work. The work can’t be completed without you. I would like to thank my committee members for their support. First, I want to thank Behnam Abasht for providing me many good opportunities and teaching me so many things. You brought me to the U.S. and gave me a very warm “home” for the last two and a half years. You made me to be a better person. Thank you to Larry Cogburn, Walter Bottje and William R Lee for serving on my thesis committee and helping me throughout my degree. Your support has been greatly appreciated over the last two years. Thank you to my lab mates, Zhu Zhuo, Marie Mutryn and Weixuan Fu, for helping me with my study and live. Thank you all my friends for helping me and making me not alone during the last two and a half years. I want to give a special thank to my parents for their support and unconditional love. You are always my greatest treasure.

iii TABLE OF CONTENTS

LIST OF TABLES ...... vii LIST OF FIGURES ...... viii ABSTRACT ...... ix

Chapter

1 BACKGROUND ...... 1

1.1 The Definition of Feed Efficiency ...... 1 1.2 The Significance of Improving Feed Efficiency ...... 2 1.3 Factors Affecting Feed Efficiency in Chickens ...... 3

1.3.1 External Factors ...... 3 1.3.2 Internal Factors ...... 4

1.4 Previous Studies in Chicken Feed Efficiency ...... 5

1.4.1 Mitochondrial Function and Feed Efficiency ...... 5 1.4.2 Global Expression Study of Feed Efficiency ...... 7

1.5 Introduction to RNA-seq Technique ...... 8 1.6 Introduction to Skeletal Muscle ...... 10

1.6.1 Skeletal Muscle Metabolism ...... 10 1.6.2 Skeletal Muscle Fiber Types ...... 11

1.7 Adverse Effects of Intense Genetic Selection ...... 12 1.8 Project Introduction ...... 13

REFERENCES ...... 15

2 MESSENGER RNA SEQUENCING AND PATHWAY ANALYSIS PROVIDE NOVEL INSIGHTS INTO THE BIOLOGICAL BASIS OF CHICKENS’ FEED EFFICIENCY ...... 20

2.1 Background ...... 20 2.2 Methods ...... 22

iv 2.2.1 Animals and Sample Collection ...... 22 2.2.2 RNA Isolation ...... 23 2.2.3 RNA-seq Library Preparation and Sequencing ...... 24 2.2.4 Mapping Reads to the Chicken Reference Genome ...... 24 2.2.5 Differential Expression Analysis ...... 25 2.2.6 Nanostring nCounter® Assay ...... 25 2.2.7 Ingenuity Pathway Analysis ...... 26

2.3 Results and Discussion ...... 26

2.3.1 Phenotype Measurements ...... 26 2.3.2 Transcriptional Profile of Chicken Breast Muscle ...... 27 2.3.3 Gene Differential Expression Analysis ...... 29 2.3.4 Confirmation of RNA-seq Data ...... 31 2.3.5 Overview of IPA Analysis ...... 33 2.3.6 Increased Muscle Growth and Remodeling in High-FE Chickens...... 37 2.3.7 Growth Hormone (GH) and IGFs/PI3K/AKT Signaling Pathway Over-represented in the Differentially Expressed …...... 43 2.3.8 Inflammatory Response in the Breast Muscle of High-FE Broilers… ...... 48 2.3.9 Free Radical Scavenging Enriched in the Differentially Expressed Genes between High- and Low-FE Broilers ...... 52 2.3.10 Transcriptional Regulation of Hypoxia-inducible Factor-1α (HIF1α)...... 58

2.4 Conclusions ...... 61

REFERENCES ...... 64

3 CHARACTERIZATION OF METABOLIC DIFFERENCES IN THE BREAST MUSCLE OF HIGH AND LOW FEED EFFICIENCY CHICKENS USING RNA-SEQ ...... 79

3.1 Introduction ...... 79 3.2 Methods ...... 81

3.2.1 Animals and RNA-seq Analysis ...... 81 3.2.2 Ingenuity Pathway Analysis ...... 82 3.2.3 Ultimate pH (pHu) Measurement and Analysis ...... 82 3.2.4 Determination of Breast Muscle Glycogen Content ...... 83 3.2.5 Quantitative Reverse Transcription-PCR (qRT-PCR) ...... 83

v 3.3 Results and Discussion ...... 85

3.3.1 Carbohydrate Metabolism ...... 85

3.3.1.1 Glycerol-3-Phosphate Shuttle ...... 89 3.3.1.2 Transcriptional Control of Carbohydrate Metabolism . 92

3.3.2 Glycogen Depletion in Breast Muscle of the HFE Chickens ...... 93 3.3.3 Transcriptional Regulation of Lipid Metabolism ...... 96

3.3.3.1 Cholesterol Biosynthesis and Reverse Cholesterol Transport… ...... 97 3.3.3.2 Lipid Catabolism ...... 99 3.3.3.3 Lipogenesis ...... 101

3.3.4 Metabolism ...... 102

3.3.4.1 The Regulation of Amino Acid Metabolism ...... 103 3.3.4.2 Glutamine Metabolism ...... 106

3.3.5 Metabolism and Inflammatory Response ...... 107

3.3.5.1 Effects of Inflammatory Response on Metabolic Regulation...... 107 3.3.5.2 Lipid-derived Inflammatory Mediators ...... 108

3.4 Conclusions ...... 110

REFERENCES ...... 113

vi LIST OF TABLES

Table 2.1 Statistics of the Measurements from High- and Low-Feed Efficiency (FE) Chickens ...... 27

Table 2.2 Top 10 Up-Regulated and Down-Regulated Genes in High-FE Chickens ...... 34

Table 2.3 Top Biological Functions and Pathways Enriched by Differentially Expressed Genes between High- and Low-FE Chickens...... 35

Table 2.4 Top Functions Enriched by Genes Up-Regulated in High- or Low-FE Broilers ...... 35

Table 2.5 Activated or Inhibited Upstream Transcription Regulators Predicted by Ingenuity® Pathway Analysis (IPA) Software...... 36

Table 2.6 Differentially Expressed Genes from Matrix Metalloproteinases (Mmps) Family ...... 40

Table 2.7 Over-represented Pathways Involved in Immune Response ...... 50

Table 3.1 Metabolism-associated DE Genes ...... 87

vii LIST OF FIGURES

Figure 2.1 The Number of Properly Mapped, Improperly Mapped and Unmapped Reads is Shown for Each Sample...... 29

Figure 2.2 Volcano Plot Showing Differentially Expressed Genes between High- and Low-Fe Chickens...... 30

Figure 2.3 Comparison of Gene Expression between High- and Low-FE Broiler Chickens...... 31

Figure 2.4 Correlation of Log2 Fold-Change between RNA-Seq and Nanostring for Significantly Differentially Expressed Target Genes...... 32

Figure 2.5 The Correlation of Log2 Fold-Change between RNA-Seq and Nanostring Increased with Gene Expression Levels...... 33

Figure 2.6 Upstream Regulators JunB and MEF2C...... 42

Figure 2.7 Growth Hormone Signaling Pathway Analysis Using Ingenuity Molecule Activity Predictor (MAP)...... 46

Figure 2.8 IGFs/PI3K/AKT Signaling Pathway Analysis Using Ingenuity Molecule Activity Predictor (MAP)...... 47

Figure 2.9 Upstream Regulator JUN and FOS...... 51

Figure 2.10 NRF2-mediated Oxidative Stress Response...... 57

Figure 2.11 Upstream Regulator HIF1α and Its Target Genes...... 60

Figure 3.1 UDP-N-acetyl-D-glucosamine biosynthesis and Glycerol-3-Phosphate Shuttle...... 91

Figure 3.2 Upstream Regulator NR4A3 and Its Target Genes...... 93

Figure 3.3 Upstream Regulator SREBF2, ATF4 and Sp1...... 105

Figure 3.4 Proposed Network between Immune and Metabolic Systems in Breast Muscles of HFE chickens...... 110

viii ABSTRACT

Feed cost plays a key role in the total expenses of modern broiler chicken production. Especially with feed prices increasing constantly in recent years, improving feed efficiency has become a concern for poultry producers. Although huge progress has been made on the optimization of feed efficiency in broiler chickens over the past 50 years, a significant portion of feed energy and nutrients is still wasted because of poor efficiency of nutrient utilization. A further consequence is that an excess of manure is produced, causing environmental concerns in the regions with intense poultry production. Therefore, from both economical and environmental standpoints, efficient use of feed is vital for sustainable poultry production. Different selection criteria have been used for improving feed efficiency in broiler chickens. However, the biological basis of the variation in chicken feed efficiency is not well understood. As a result, some undesirable side effects would appear with the cumulative genetic selection for high feed efficiency. A more profound understanding of this highly complex trait is needed to develop more efficient selection strategies and to foresee potential long-term issues that may arise by selection for high feed efficiency. The objective of this project is to characterize the molecular basis of the variation in chicken feed efficiency, with a long-term goal of sustainably improving feed efficiency in poultry production. Using high-throughput RNA sequencing, we aimed to identify the gene expression differences in breast muscle tissues of broiler chickens with extremely high and low feed efficiencies. Total RNA was isolated from breast muscle samples harvested from 10 high and 13 low feed efficiency chickens at

ix 7 weeks of age. Using Truseq Stranded RNA Prep kit (Illumina), each sample was converted to a uniquely indexed cDNA library, and the resulting cDNA libraries were pooled and sequenced on an Illumina Hiseq 2000 sequencer. An average of 34 million paired-end reads (75 bp) were produced from each sample, 80% of which were properly mapped to the reference genome (Ensembl Galgal4). We analyzed the sequence data using bioinformatics tools Tophat and Cufflinks and detected 1059 genes differentially expressed (FDR < 0.05) in the breast muscle between the high and low feed efficiency chickens. Gene function analysis revealed that genes involved in muscle remodeling, inflammatory response and free radical scavenging were mostly up-regulated in the high feed efficiency birds. Additionally, growth hormone and IGFs/PI3K/AKT signaling pathways were enriched in differentially expressed genes, which might contribute to the high breast muscle yield in high feed efficiency birds and partly explain the FE advantage of these chickens.

x Chapter 1

BACKGROUND

Chicken meat is one of the most important and rapidly growing food in the world. [1]. To meet the strong demand for chicken meat, U.S. broiler production has grown fast from 1990 to 2000 [2]. However, the production growth slowed down in the last decade mainly due to the decline in broiler productivity growth [1]. In addition, the constant increase in feed ingredient price raised the production costs substantially in the last five years, consequently, bringing a huge challenge for the poultry industry [3]. To keep the broiler production growth and maintain the production cost steady, improving feed efficiency has become increasingly urgent for the poultry industry.

1.1 The Definition of Feed Efficiency

Feed efficiency is a trait defining the animal’s ability to convert feed into body weight. Unlike body weight gain or feed intake, feed efficiency cannot be measured directly. There are two common methods to assess this trait.

1) Feed conversion ratio (FCR). FCR is defined as the amount of feed consumed per unit of weight gain [4]. It is a ratio of feed intake and body weight gain over a defined period.

Feed conversion ratio (FCR)= feed intake / body weight gain (1) The less the FCR is, the more efficient the birds are. Because of its moderate heritability and simple calculation, FCR has been successfully applied in the poultry

1 breeding program [5]. However, FCR cannot reflect the true feed efficiency of chickens as it blends with the growth rate. 2) Residual feed intake (RFI). Koch et al. (1963) first proposed the concept of RFI, which is defined as the feed intake that can not be explained by metabolic body weight and body weight gain. The linear regression model is:

RFI = FI – [a + b1 × BW0.75 + b2 × BWG] (2) where FI represents the feed intake of each bird, BW0.75 is the metabolic weight, BWG is the body weight gain, a is the intercept and b1 and b2 are the partial regression coefficients of FI on BW0.75 and BWG [5]. RFI shows moderate to high heritability, ranging from 0.42 to 0.62 for laying hens [6] and 0.2 to 0.4 for broiler chickens [7]. One of the advantages of RFI is its independence from the production traits. Thus, selection based on RFI can reduce the feed consumption without influencing the body weight and production performance

[8]. There is a high genetic correlation between RFI and FCR, rg = 0.66 in Angus cattle [9].

1.2 The Significance of Improving Feed Efficiency

Feed is the largest single cost in raising poultry, accounting for up to 70% of the total costs of poultry production. Especially in recent years, due to the multiple usages of corn, not only as a feed grain but also in biofuel production, feed ingredient prices increased to an unprecedentedly high level, approaching $320 per ton in 2013. As a result of continual growth of feed ingredient prices, the total costs of poultry production increased almost 100% from September 2006 to 2008 [3]. To maintain the profitability while feed price continued increasing, poultry producers have paid more attention to optimizing feed efficiency and reducing feed costs. Feed efficiency is a

2 key trait in poultry production because small improvement in this trait produce substantial savings in feed costs. One study found that a 1% improvement in feed conversion can reduce corn requirements by 5.3 million bushels for the poultry industry in U.S. [3]. Therefore, seeking ways to improve feed efficiency is essential for a sustainable poultry industry. Over the past four or five decades, tremendous progress has been made in feed efficiency in broiler chickens. In a study comparing the growth performance between 1957 Athens-Canadian Randombred strain and 2001 Ross 308 strain of broilers, the chickens in 2001 only need one-third the rearing time and feed that birds needed in 1957 to gain the same weight. Genetic, nutrition and management changes in poultry production are the major contributors for this substantial enhancement. Among these factors, it was determined that genetic selection was responsible for approximately 85 to 90% of the improvement in feed efficiency, whereas changes in nutrition brought about 10 to 15% of the growth [10]. Management modifications, like optimizing temperature and lighting, have also contributed to the improvement in feed efficiency [11].

1.3 Factors Affecting Feed Efficiency in Chickens

1.3.1 External Factors

Feed efficiency is a complex trait influenced by a variety of environmental factors. Nutrition is the largest factor affecting feed efficiency in chickens. Despite meeting all other nutrient requirements, birds fed a low diet have a lower growth rate and higher FCR compared with birds fed with a control diet [12, 13]. Furthermore, environmental temperature plays a critical role in modulating poultry

3 productivity. Since the chicken is endotherm [14], a large quantity of energy must be consumed to maintain a relatively constant body temperature constant even when chickens experience either cold or hot temperatures. Therefore, a moderate environment temperature is essential for the efficient usage of feed in poultry [15]. In addition, other factors including flock size, stocking density, lighting, and noise also have impacts on chicken feed efficiency [16].

1.3.2 Internal Factors

Feed intake is an important trait in poultry production that is closely associated with feed efficiency. As in mammals, the hypothalamic regions as well as the specific neuroendocrine regulation are major modulators controlling feed intake in chickens. There are two main body systems responsible for the regulation of feed intake in mammals [17,18]. One is short-term system, also called peripheral satiety system, that regulates the meal-to-meal feed intake [18]. In response to intestinal nutrient contents and the presence of food in the gastrointestinal tract, the gut, liver and pancreas release some regulatory peptides to control gut motility and secretion. Some of these peptides, like ghrelin and Cholecystokinin (CCK), can serve as satiety signals transmitted through the brainstem to the hypothalamus. By activating the neural pathways or releasing hormonal signals into blood, these satiety signals are able to mediate the feed intake in a short-term [18,19]. The other system modulating feed intake is the long-term regulation of energy balance, which maintains the homeostasis of body energy via activating neural and neuroendocrine pathways in hypothatlamus [19]. Two kinds of regulation, anabolic and catabolic control, constitute the long-term regulation. Anabolic regulation

4 increases energy intake and storage, whereas catabolic modulation decreases energy intake and storage [20].

1.4 Previous Studies in Chicken Feed Efficiency

1.4.1 Mitochondrial Function and Feed Efficiency

Numerous studies revealed a relation between mitochondrial function and feed efficiency in broiler chickens. Bottje et al. (2002) disclosed a divergence in muscle mitochondrial function between low feed efficiency (LFE) and high feed efficiency (HFE) broiler chickens. They found that mitochondria isolated from breast and leg muscles of LFE birds exhibited less efficient coupling of electron transport than those from HFE birds. Because higher levels of reactive oxygen species (ROS) were also examined in the mitochondria from muscles of LFE chickens, they hypothesized that the inefficient coupling of electron transport was associated with increased electron leak in electron transport chain (ETC). Their further studies proved their assumption. Site-specific defects were identified in the mitochondrial complex I and III of the LFE birds, which were proposed to be the potential areas where the electron leak happens. The activities of ETC complex , determined by the rate of oxidation, were also found to be lower in the mitochondria obtained from muscles of LFE chickens. Combined with previous studies, they postulated that the less activity of mitochondrial complexes was a potential result of ROS damage [21,22]. Similar results were obtained in the subsequent studies using samples from chickens intestine and liver [21,23,24]. Moreover, the oxidized glutathione (GSSG) to GSH ratio was found to be greater in the duodenal mitochondria from the LFE chickens. Because the GSSG to

5 GSH ratio is a critical indicator of oxidative stress in cells, they conjectured that there was a mild oxidative stress in the mitochondria of LFE birds [25]. In addition, the protein oxidation was also predicted to be elevated in the LFE chickens as a greater level of protein carbonyl was observed in the mitochondria from muscle, liver, intestine and lymphocytes of these broilers [26]. They also found that ubiquitine was higher in breast muscle of LFE chickens than HFE birds. Since ubiquitin is a polypeptide conjugate with oxidized to facilitate degradation, they concluded that the protein turnover was increased in breast muscle of LFE broilers, which was likely caused by oxidative stress and resulted in the inefficiency of chickens [21]. To develop biomarkers applied in selection for high feed efficiencies, some studies were aimed to identify proteins that are closely associated with chicken feed efficiency. As components of respiratory chain complexes, the expressions of ETC proteins were analyzed in the LFE and HFE birds. Two mitochondrial proteins, cytochrome oxidase (COX) II of complex IV and cytochrome c1 of complex III, were observed differentially expressed between the HFE and LFE phenotypes within a tissue. However, the role of these two proteins were still questionable because their expression divergence was not consistent between different tissues [21, 26-28] . Several extramitochondrial proteins have also shown differentially expressed between HFE and LFE birds, and further studies are required to determine whether these proteins can be developed to be biomarkers. In addition to protein expression differences, researchers also looked into the specific gene expression differences between chickens with high and low feed efficiencies. Ojano-Dirain et al. (2007) measured the mRNA expression levels of six genes in breast muscle and duodenum of the HFE and LFE chickens. Compared with

6 HFE birds, the mRNA levels of avian uncoupling protein (avUCP) were greater in the breast muscle of LFE chickens [29]. Uncoupling proteins (UCP) are proton transporters in the mitochondrial inner membrane. They transport protons from the intermembrane space back to the inner membrane, as a result, consuming the proton gradient and lowering ATP synthesis [30]. Since it has been demonstrated that reversible glutathionylation of UCP2 and UCP3 could negatively regulate ROS production [30,31], Bottje et al. (2009) speculated that the up-regulated UCP in the LFE birds was used to attenuate the higher levels of ROS produced in the LFE mitochondria [21]. Another major mitochondrial carrier protein with uncoupling property, avian adenine nucleotide translocase (avANT), has shown no significant difference in mRNA expression levels between HFE and LFE broilers.. The mRNA expression of cytochrome oxidase III (COX III) was found lower in the LFE chickens, which was likely attributed to more oxidative damages in these birds.

1.4.2 Global Gene Expression Study of Feed Efficiency To uncover the biological mechanisms underlying the variation in chicken feed efficiencies, a global gene expression study was conducted in breast muscle of HFE and LFE broiler chickens using a chicken 44K oligo microarray. It’s found that most of the genes up-regulated in the HFE chickens were related to anabolic metabolism, whereas genes up-regulated in the LFE chickens were mostly associated with muscle fiber development, muscle function, cytoskeletal organization and stress response [32]. Moreover, Bottje et al. (2012) provided more biological insights into the cellular basis of feed efficiency in broilers. After filtering the differentially expressed genes, they identified 27 genes as focus genes for in-depth interpretation. Among these focus genes, 14 genes were up-regulated and 13 genes were down-regulated in the breast

7 muscle of the HFE birds. All up-regulated focus genes were found to be associated with important signal transduction and in sensing of cell energy status as well as regulating energy production in the cell. JNK pathway and protein kinase A (PKA) cascade mechanisms were over-represented with nine of the up-regulated focus genes mapped to these pathways. Furthermore, some up-regulated genes were shown to be associated with the mitochondrial complex I of the ETC, consistent with higher activities of complex I in the HFE broiler phenotype. In contrast, genes down- regulated in the HFE chickens were involved in cytoskeletal organization (e.g., actin- myosin filaments), fatty acid oxidation, growth factors and antioxidant protection [33].

1.5 Introduction to RNA-seq Technique Transcriptome is a collection of whole RNA molecules in the cell, including mRNA, tRNA, rRNA and non-coding RNA [34]. Since transcriptome variation exists between different individuals, studying the transcriptome differences between individuals can provide more insights into the biological mechanisms behind different phenotypes [35]. RNA-seq, also called whole transcriptome sequencing, is an advanced technique developed based on the next-generation sequencing (NGS) technology. NGS is a high-throughput DNA sequencing technology that identifies the bases of a small fragment of DNA from signals emitted from the re-synthesized complementary bases in a massively parallel fashion. It has provided unprecedented scope and resolution for transcriptome profiling and analysis. RNA-seq can not only characterize multiple functional elements of a genome, such as exons, introns and the transcription start sites, but also can identify the splice variants and quantify the expression levels of a gene [34,36].

8 The basic procedure of RNA-seq is described below. First of all, mRNA is converted to short cDNA with adapters for sequencing. Adapters are short nucleotide sequences that only allow the adapter-binding segments to be enriched and sequenced. In addition, adapters function similar to a barcode so that multiple cDNA libraries can mixed together in sequencing [36,37]. With or without amplification, each individual molecule is then sequenced from one end or from both ends using NGS. At present, Illumina, Applied Biosystems SOLiD and Roche 454 Life Science system are three main NGS platforms applied in RNA-seq. Millions of short sequence reads, typically ranging from 30bp to 400bp, are produced from the sequencing [34,36]. Computational approaches are applied to store, process and analyze the hundreds of millions of data generated from sequencing. Raw sequence data are required to go through quality check first in order to guarantee the success of subsequent analysis. Software programs, like FastQC and cutadapt, are developed for the quality control and trimming. Poor quality reads, including contaminated reads, dubious reads (too many Ns called in the sequence) and incomplete reads, are excluded in the quality control [38]. The second step in data analysis is to map reads to the reference genome. Nowadays, there are many statistical algorithms developed to undertake this job. TopHat is one of the advanced read-mapping algorithms that is able to align millions of reads to the reference genome. TopHat can not only map the non-junction reads to the genome by an ultrafast short read mapping program Bowtie, but also can identify the exon-exon splice junctions through analyzing the mapping results. Based on the initial Bowtie’s mapping outputs, TopHat finds splice junctions by building a database of potential splice junctions and then mapping the reads against the predicted junctions for confirmation. Since the reads are too long to map, TopHat first cuts them into short

9 segments and then separately aligns them. After alignment, these small segments are assembled together to form an overall alignment [40,41].Following read mapping, Cufflinks is used to assemble transcripts, estimate their relative abundance and test for differential expressions. Fragments per kilobase of exon per million fragments mapped (FPKM) is applied in Cufflinks to normalize the read counts. Cuffdiff is a program built in Cufflink, responsible for differential expression analysis of genes and transcripts [40,41].

1.6 Introduction to Skeletal Muscle Body movements, from maintaining postures to performing intense exercises, are all facilitated by skeletal muscles. Moreover, skeletal muscle plays a crucial role in the regulation of electrolytes and nutrients. It provides a temporary reserve of protein in the body. Muscle protein turnover is affected by various physiological and pathophysiological conditions, such as fasting, feeding, exercise, aging and diseases [42].

1.6.1 Skeletal Muscle Metabolism

Skeletal muscle is a uniquely plastic tissue that can rapidly increase its rate of energy consumption in response to explosive contraction and can maintain a very low level of energy consumption in the resting state. ATP is the main immediate energy source powering muscle contraction, which is regenerated through several metabolic pathways to avoid ATP depletion. Pathways, mainly including the degradation of phosphocreatine (PCr) and the breakdown of muscle glycogen, dominantly provide energy sources during short high-intensity activities. PCr is a mobilizable reserve of ATP in skeletal muscle, which donates a phosphate group to ADP for the synthesis of

10 ATP in the conditions where ATP is depleted. In addition, muscle glycogen is also energy sources that are quickly converted to glucose during periods of high ATP consumption. The glucose is further metabolized to produce ATP in an anaerobic fashion during intensive exercise [43]. For long mild exercise, oxidative metabolism of carbohydrates and lipids is the dominant way to produce ATP. The aerobic ATP- generating system can produce more ATP than anaerobic glycolysis but require more biochemical steps [43].

1.6.2 Skeletal Muscle Fiber Types Based on the myosin heavy chain (MHC) isoforms, three types of muscle fibers are found in the mammalian skeletal muscles as outlined below [43]. a) Type I muscle fibers are recognized as red slow twitch fibers. Consisting of more mitochondria and myoglobins, this fiber mainly utilizes aerobic metabolism to produce ATP. Type I fibers are fatigue-resistant due to high levels of mitochondrial in them. Postural muscles are largely comprised of this type of fibers. b) Type IIa fibers are red fast twitch fibers. They can generate ATP through oxidative metabolism as these fibers have more mitochondria and a high concentration of myoglobin. In addition, type IIa fibers can also produce ATP by anaerobic glycolysis. Therefore, type IIa fibers are relatively slow to fatigue compared with type

IIb and IIX fibers, whereas contract more quickly than type I fibers. c) Type IIb and IIX fibers are white fast twitch fibers. They can quickly respond to central nervous system (CNS) signals and have stronger contraction because of the fast production of ATP by anaerobic glycolysis and a rapid release and uptake of calcium by the sarcoplasmic reticulum. In the cells of these fibers, there are

11 large amounts of glycogen but a lower concentration of mitochondria and myoglobin, accordingly, type IIb and IIX fibers are easily fatigued [44].

1.7 Adverse Effects of Intense Genetic Selection Genetic selection has made tremendous progress in enhancing traits of economic importance in the poultry industry [45,46]. However, because of a lack of knowledge about the underlying biological processes on which selection may impact, cumulative genetic selection has brought some potentially adverse effects on the fitness of broiler chickens [47]. It has been shown that the overall fitness (i.e., skeletal strength and cardiovascular fitness) of modern broiler chickens has been reduced by the intensive selection for high production efficiency [53][54]. Broiler chickens and turkeys selected for rapid growth rate and high muscle yield are more at risk for physiological and behavioral disorders such as muscle problems, ascites and sudden death syndrome [45,47]. It was reported that the breast and leg muscles from turkeys with high growth rates were loosely packed and degenerated [49]. They hypothesized that the intense selection for growth velocity and muscle mass was what was responsible for the muscle disorder in these turkeys. The muscle disorder was possibly attributed to the insufficient extracellular tissue support of the rapidly growing muscle fibers. They also found that the clinical symptoms of these muscle abnormalities were similar to the focal myopathy. Because focal myopathy is a muscle disease caused by ischemia, they also postulated that the muscle abnormality in fast-growing chickens was probably associated with ischemia, which was further supported by histopathological study [50]. Mainly induced by the narrowing or blockage of arteries, ischemia is a shortage of blood and oxygen supply to tissues. Similar symptoms were found in broiler chickens. Birds with high growth rates had increased incidences of

12 muscle abnormalities, reflected by large fiber diameters, more glycolytic fibers and lower protein catabolism rate [51].

1.8 Project Introduction As feed cost accounts for nearly 70% of the total costs of the modern commercial broiler production and feed ingredient prices continue growing in recent years, improving feed efficiency is becoming essential and urgent for poultry producers. Efficient feed usage can not only significantly enhance profitability for the poultry industry but also minimize animal waste, playing a critical role in sustainable poultry production. Despite great advances that have been achieved in the optimization of broiler growth performance over the past 50 years, much room for improvement in feed efficiency still remains. Additionally, some undesirable side effects have appeared in poultry production with the rapid increase in broiler growth rate and breast muscle yield. A molecular and genetic characterization of feed efficiency in broiler chickens is needed for the further improvement of this trait and anticipating the impacts of continued selection for high feed efficiency. This project aims to characterize the biological and molecular basis of differences between high and low feed efficiency broiler chickens. We expect that there would exist several genetics differences between high and low feed efficiency birds that contribute to the feed efficiency variation. To test our hypothesis, we are utilizing high-throughput RNA-seq approach to study the gene expression differences in breast muscle of 23 commercial broiler chickens with extremely high (n=10) and low (n=13) feed efficiencies. The present study identifies genes and pathways that are differentially regulated between high and low feed efficiency chickens, providing

13 comprehensive knowledge toward understanding the molecular basis of the variation in feed efficiency in broiler chickens.

14 REFERENCES

1. Macdonald JM: of U . S . Broiler Production Cataloging Record :. 2008.

2. Paisner MS, Witcover J: 2020 Global Food Outlook. 2001(August).

3. Donohue M, Cunningham DL: Effects of grain and oilseed prices on the costs of US poultry production. The Journal of Applied Poultry Research 2009:325– 337.

4. Willems OW, Miller SP, Wood BJ: Aspects of selection for feed efficiency in meat producing poultry. Worlds Poult Sci J 2013, 69:77–88.

5. Aggrey SE, Karnuah AB, Sebastian B, Anthony NB: Genetic properties of feed efficiency parameters in meat-type chickens. Genet Sel Evol 2010, 42:25.

6. Luiting P, Urff EM: Residual feed consumption in laying hens. 2. Genetic variation and correlations. Poult Sci 1991, 70:1663–1672.

7. Kennedy BW, van der Werf JH, Meuwissen TH: Genetic and statistical properties of residual feed intake. J Anim Sci 1993, 71:3239–3250.

8. Cai W, Casey DS, Dekkers JCM: Selection response and genetic parameters for residual feed intake in Yorkshire swine. J Anim Sci 2008, 86:287–298.

9. Arthur PF, Archer JA, Johnston DJ, Herd RM, Richardson EC, Parnell PF: Genetic and phenotypic variance and covariance components for feed intake, feed efficiency, and other postweaning traits in Angus cattle. J Anim Sci 2001, 79:2805–2811.

10. Havenstein GB, Ferket PR, Qureshi M a: Growth, livability, and feed conversion of 1957 versus 2001 broilers when fed representative 1957 and 2001 broiler diets. Poult Sci 2003, 82:1500–8.

11. Davis A: Preferences of Growing Fowls for Different Light Intensities in Relation to Age , Strain And Behaviour. Anim Welf 1999, 8:1999.

12. Bregendahl K, Sell JL, Zimmerman DR: Effect of Low-Protein Diets on Growth Performance and Body Composition of Broiler Chicks. Volume 81; 2002:1156–1167.

15 13. Holsheimer JP, Veerkamp CH: Effect of dietary energy, protein, and lysine content on performance and yields of two strains of male broiler chicks. Poult Sci 1992, 71:872–879.

14. Brown WR a, Hubbard SJ, Tickle C, Wilson S a: The chicken as a model for large- scale analysis of vertebrate gene function. Nat Rev Genet 2003, 4:87–98.

15. Blahová J, Dobšíková R, Straková E, Suchý P: Effect of Low Environmental Temperature on Performance and Blood System in Broiler Chickens (Gallus domesticus). Acta Vet Brno 2007, 76:S17–S23.

16. Peter R. Ferket* and Abel G. Gernat: Factors that affect feed intake of meat. International Journal of Poultry Science 2006, 5(10): 905-911.

17. Song Z, Everaert N, Wang Y, Decuypere E, Buyse J: The endocrine control of energy homeostasis in chickens. Gen Comp Endocrinol 2013, 190:112–7.

18. Richards MP: Genetic regulation of feed intake and energy balance in poultry. Poult Sci 2003, 82:907–916.

19. Jensen J: Regulatory peptides and control of food intake in non-mammalian vertebrates. Comp Biochem Physiol A Mol Integr Physiol 2001, 128:471–479.

20. Woods SC, Seeley RJ, Porte Jr. D, Schwartz MW: Signals that regulate food intake and energy homeostasis. Science (80- ) 1998, 280:1378–1383.

21. Bottje WG, Carstens GE: Association of mitochondrial function and feed efficiency in poultry and livestock species. J Anim Sci 2009, 87:E48–E63.

22. Bottje W, Iqbal M, Tang ZX, Cawthon D, Okimoto R, Wing T, Cooper M: Association of mitochondrial function with feed efficiency within a single genetic line of male broilers. Poult Sci 2002, 81:546–555.

23. Iqbal M, Pumford NR, Tang ZX, Lassiter K, Ojano-Dirain C, Wing T, Cooper M, Bottje W: Compromised liver mitochondrial function and complex activity in low feed efficient broilers are associated with higher oxidative stress and differential protein expression. Poult Sci 2005, 84:933–941.

24. Ojano-Dirain CP, Iqbal M, Cawthon D, Swonger S, Wing T, Cooper M, Bottje W: Determination of mitochondrial function and site-specific defects in electron transport in duodenal mitochondria in broilers with low and high feed efficiency. Poult Sci 2004, 83:1394–1403.

16 25. Ojano-Dirain C, Iqbal M, Wing T, Cooper M, Bottje W: Glutathione and respiratory chain complex activity in duodenal mitochondria of broilers with low and high feed efficiency. Poult Sci 2005, 84:782–788.

26. Iqbal M, Pumford NR, Tang ZX, Lassiter K, Wing T, Cooper M, Bottje W: Low feed efficient broilers within a single genetic line exhibit higher oxidative stress and protein expression in breast muscle with lower mitochondrial complex activity. Poult Sci 2004, 83:474–484.

27. Lassiter K, Ojano-Dirain C, Iqbal M, Pumford NR, Tinsley N, Lay J, Liyanage R, Wing T, Cooper M, Bottje W: Differential expression of mitochondrial and extramitochondrial proteins in lymphocytes of male broilers with low and high feed efficiency. Poult Sci 2006, 85:2251–2259.

28. Bottje W, Pumford NR, Ojano-Dirain C, Iqbal M, Lassiter K: Feed efficiency and mitochondrial function. Poult Sci 2006, 85:8–14.

29. Ojano-Dirain C, Toyomizu M, Wing T, Cooper M, Bottje WG: Gene expression in breast muscle and duodenum from low and high feed efficient broilers. Poult Sci 2007, 86:372–381.

30. Mailloux RJ, Harper M-E: Uncoupling proteins and the control of mitochondrial reactive oxygen species production. Free Radic Biol Med 2011, 51:1106–1115.

31. Harper M-E, Mailloux RJ: Mitochondrial proticity and ROS signaling: lessons from the uncoupling proteins. Trends in Endocrinology & Metabolism 2012:451–458.

32. Kong B-W, Song JJ, Lee JY, Hargis BM, Wing T, Lassiter K, Bottje W: Gene expression in breast muscle associated with feed efficiency in a single male broiler line using a chicken 44K oligo microarray. I. Top differentially expressed genes. Poult Sci 2011, 90:2535–47.

33. Bottje WG, Kong B-W, Song JJ, Lee JY, Hargis BM, Lassiter K, Wing T, Hardiman J: Gene expression in breast muscle associated with feed efficiency in a single male broiler line using a chicken 44K microarray. II. Differentially expressed focus genes. Poultry Science 2012:2576–2587.

34. Wang Z, Gerstein M, Snyder M: RNA-Seq: a revolutionary tool for transcriptomics. Nat Rev Genet 2009, 10:57–63.

35. Frith MC, Pheasant M, Mattick JS: The amazing complexity of the human transcriptome. European journal of human genetics : EJHG 2005:894–897.

17 36. Nagalakshmi U, Waern K, Snyder M: RNA-Seq: a method for comprehensive transcriptome analysis. Curr Protoc Mol Biol 2010, Chapter 4:Unit 4.11.1–13.

37. Schiemer J: Illumina TruSeq DNA Adapters De-Mystified. :1–5.

38. Wilhelm BT, Landry J-R: RNA-Seq-quantitative measurement of expression through massively parallel RNA-sequencing. Methods 2009, 48:249–257.

39. Trapnell C, Pachter L, Salzberg SL: TopHat: discovering splice junctions with RNA-Seq. Bioinformatics 2009, 25:1105–1111.

40. Trapnell C, Roberts A, Goff L, Pertea G, Kim D, Kelley DR, Pimentel H, Salzberg SL, Rinn JL, Pachter L: Differential gene and transcript expression analysis of RNA-seq experiments with TopHat and Cufflinks. Nat Protoc 2012, 7:562–78.

41. Trapnell C, Hendrickson DG, Sauvageau M, Goff L, Rinn JL, Pachter L: Differential analysis of gene regulation at transcript resolution with RNA-seq. Nat Biotechnol 2013, 31:46–53.

42. Rennie, MJ, Wackerhage, H, Spangenburg, EE Booth, FW: Control of the size of the human muscle mass. Annu Rev Physiol 2004, 66:799-828.

43. Westerblad H, Bruton JD, Katz A: Skeletal muscle: energy metabolism, fiber types, fatigue and adaptability. Exp Cell Res 2010, 316:3093–3099.

44. Pette D, Staron RS: Myosin isoforms, muscle fiber types, and transitions. Microsc Res Tech 2000, 50:500–509.

45. Deeb N, Lamont SJ: Genetic architecture of growth and body composition in unique chicken populations. J Hered 1998, 93:107–18.

46. Emmerson DA: Commercial approaches to genetic selection for growth and feed conversion in domestic poultry. Poult Sci 1997, 76:1121–1125.

47. Rauw W., Kanis E, Noordhuizen-Stassen E., Grommers F.: Undesirable side effects of selection for high production efficiency in farm animals: a review. Livest Prod Sci 1998, 56:15–33.

48. Julian RJ: Rapid growth problems: ascites and skeletal deformities in broilers. Poult Sci 1998, 77:1773–1780.

49. Wilson BW, Nieberg PS, Buhr RJ, Kelly BJ, Shultz FT: Turkey muscle growth and focal myopathy. Poult Sci 1990, 69:1553–1562.

18 50. Sosnicki AA, Cassens RG, Vimini RJ, Greaser ML: Histopathological and ultrastructural alterations of turkey skeletal muscle. Poult Sci 1991, 70:349– 357.

51. Dransfield E, Sosnicki AA: Relationship between muscle growth and poultry meat quality. Poult Sci 1999, 78:743–746.

19 Chapter 2

MESSENGER RNA SEQUENCING AND PATHWAY ANALYSIS PROVIDE NOVEL INSIGHTS INTO THE BIOLOGICAL BASIS OF CHICKENS’ FEED EFFICIENCY

2.1 Background

Genetic selection has tremendously improved livestock and plant production over the past 50 years [59][60]. Advanced selection technologies have been developed and continually optimized to genetically improve traits of agricultural importance [59][52]. However, these methods have been primarily applied without knowledge of underlying biological changes that may be induced by selection [61][54]. Previous studies reported possible association between selection for improved performance and increased rate of physiological and metabolic disorders in modern breeds [58][56][62]. For example, chickens and turkeys selected for high growth rate have shown increased incidence of muscle disorders, heart failure syndrome and ascites [55][63][64]. A detailed characterization of traits of breeding interest may help to anticipate unfavorable consequences of long-term selection programs and adjust or perhaps redefine breeding objectives accordingly. One of the most important traits in broiler chicken production is feed efficiency (FE), which defines the chicken’s ability to convert feed into body weight. As feed cost represents nearly 70% of the total cost in poultry production, improving FE has been an important goal in broiler chicken breeding programs [4]. Selection for FE in broiler chickens can be accomplished using different measurements and procedures. A widely used measure of FE in broilers is residual feed consumption

20 (RFC), which is defined as the difference between an animal’s actual feed intake and expected feed intake on the basis of body weight and growth [4]. Although moderate heritability, ranging from 0.42 to 0.45, for RFC has been reported in a previous study using more diverse chicken populations [5], to our knowledge this trait exhibits lower heritability (~0.2) in the commercial pure lines, possibly explaining the relatively slow progress in improving FE in commercial breeding programs. Insights into the biological basis of differences in chicken FE are required to develop more efficient and sustainable selection strategies. Previous studies have revealed a link between mitochondrial function and FE in broiler chickens. Lower electron transport chain coupling and greater hydrogen peroxide (H2O2) production were observed in mitochondria of low-FE birds [29]. A microarray gene expression analysis of breast muscle samples from high- and low-FE broiler chickens [39][40] identified 782 differentially expressed genes. Most of the genes up-regulated in high-FE chickens were related to anabolic metabolism, whereas genes up-regulated in low-FE chickens were associated with muscle fiber development, muscle function, cytoskeletal organization and stress response [39]. With the rapid development of next-generation sequencing technologies, RNA sequencing (RNA-seq) has been replacing microarray technology for transcriptome-wide gene expression analysis. Avoiding technical issues inherent to microarray such as cross-hybridization and narrow ranges of signal detection, RNA-seq can provide more accurate and comprehensive information regarding changes in gene expression between different conditions or different phenotypes [65][41][66][67]. Therefore, a global gene expression study using RNA-seq is required for better understanding the molecular basis of FE in broiler chickens.

21 The objective of this study is to characterize the biological basis of differences between high- and low-FE chickens through breast muscle RNA-seq analysis. Using tissue samples from extremely high- and low-FE chickens, the present study identifies genes and pathways differentially regulated in breast muscle between these two groups of chickens, providing important information toward understanding the biological basis of variation in FE in broiler chickens.

2.2 Methods

2.2.1 Animals and Sample Collection

Six groups of 400 male commercial broiler chickens from a cross between three commercial broiler pure lines were sampled at 29 days of age from the field in the Delmarva region of the United States and transferred into individual cages. The birds were individually weighed at the beginning of the test (29 days of age, BW29) and fed ad libitum until 46 days of age, at which time individual body weights (BW46) were recorded. Chickens were euthanized by cervical dislocation at 47 days of age, and then breast muscle samples were obtained from high- and low-FE birds, immediately flash frozen in liquid nitrogen and kept at –80 ºC until further processing.

Body weight after evisceration (BW47) and breast muscle weight (BM) were also recorded and used for estimating the percentage of breast muscle [(BM/BW47)*100]. The total feed consumption of each bird was measured by subtracting the total amount of feed left at the end of the test from the total amount of feed provided to each bird at the beginning of the test. To measure the broiler’s FE, Residual feed consumption (RFC) was calculated using the following equation:

RFC = FC – (a*Row + b1*BW29 + b2*BW46 + c) (3)

22 where FC represents the feed consumption of each bird; Row stands for the effects of cage location on FC; BW29 is the initial (29-day) body weight; BW46 is the ending (46-day) body weight; c is the intercept; and a, b1 and b2 are the partial regression coefficients of FC on Row, BW1 and BW2. After excluding outliers and erroneous data (3.3%) and data from birds with defects (1.2%; leg and wings problem, etc.), samples from clinically healthy broilers exhibiting the highest (n=12) or lowest (n=13) RFC from the 6 groups of 400 birds were selected for cDNA library preparation. Two samples from the high-FE group did not produced enough cDNA libraries, so samples from 23 birds, 10 high- and 13 low-FE, were used for further analysis. The protocols were submitted to, and the use of the collected data and samples for research was approved by, the University of Delaware Agricultural Animal Care and Use Committee.

2.2.2 RNA Isolation The frozen breast muscle samples were smashed into pieces by hammering. Pulverized tissues were stored at –80 °C until RNA extraction. The total RNA was isolated from 70-100mg of fragmented breast muscle tissues using a mirVana™ miRNA Isolation Kit (Ambion®; Austin, TX), according to the manufacturer’s protocol. RNA quantity and quality were assessed using a NanoDrop ND-1000 spectrophotometer (NanoDrop Technologies; Wilmington, DE) and Agilent 2100 Bioanalyzer (Agilent Technologies; Santa Clara, CA). The RNA integrity number was above 8.0 for all RNA samples.

23 2.2.3 RNA-seq Library Preparation and Sequencing In total, 23 cDNA libraries were constructed using an Illumina Truseq stranded RNA sample preparation kit following the manufacturer’s instruction (Illumina Inc.; San Diego, CA). Briefly, polyA containing mRNA molecules were purified by oligo (dT) magnetic beads and subsequently fragmented. The purified RNA fragments were reverse transcribed into first-strand cDNA using SuperScript II Reverse Transcriptase (Invitrogen™; Austin, TX). The second-strand cDNA was synthesized using dUTP instead of dTTP, as a result, the second-strand cDNA was not amplified during PCR because of the polymerase cannot add nucleotide to the dUTP. The double-strand cDNA was adenylated at the 3’ end and ligated to the Illumina indexing adapters. After PCR enrichment, cDNA quantity and quality were assessed using a NanoDrop ND-1000 spectrophotometer and Agilent 2100 Bioanalyzer. The averaged size of synthesized cDNA fragments was approximately 260 bp. cDNA libraries were normalized to 10nM for each sample and then pooled together and sequenced on four lanes of an Illumina Hiseq 2000 sequencer at Delaware Biotechnology Institute, University of Delaware. Approximate 68 million fragments per sample were sequenced by 75-bp from both ends.

2.2.4 Mapping Reads to the Chicken Reference Genome

Before read alignment, the quality of raw sequence reads was checked using the FastQC program, and nucleotide calls with a quality score of 28 or higher were considered very good quality [68]. Sequencing reads from each sample were mapped to the chicken reference genome [Ensembl Galgal4 (GCA_000002315.2)] using the TopHat program [69]. Because only the strand generated during the first-strand synthesis was sequenced, “-library-type fr-firststrand” was applied as one of the

24 parameters in our read alignment using TopHat. Only one alignment for a given read was allowed in our analysis (i.e., -g 1), and both reads from a single sequence fragment were required to be mapped to the reference genome in a concordant manner (i.e., --no-discordant and --no-mixed). To summarize the alignments statistics, the resulting alignment files (SAM files) statistics were analyzed using SAMtools [70].

2.2.5 Differential Expression Analysis Cuffdiff, a companion tool of Cufflinks (v 2.1.1), was used to quantify the gene expression levels and to perform a differential expression test [71]. The fragments counts were normalized via a geometric method, as described previously [47]. Genes with a false discovery rate of less than 5% (i.e., FDR < 0.05) were considered significant.

2.2.6 Nanostring nCounter® Gene Expression Assay

The gene expression data was verified by NanoString nCounter® technology, as described previously [72]. Briefly, 23 RNA samples were submitted to NanoString, Inc. (Seattle, WA USA) for gene expression assay. With 12 housekeeping genes, 192 endogenous transcripts were selected across multiple on-going RNA-seq projects in our laboratory as target sequences to be measured. Designs of specific probes for target sequences were provided by NanoString [72] and were screened to avoid areas of high SNP density. A total of 100 ng of each RNA sample were hybridized to the CodeSet®, which was composed of both capture and reporter probes [72]. After 16 hours incubation, the samples were transferred to the nCounter® Prep Station and Digital Analyzer for transcript quantification. Positive control normalization factors and reference genes were used to normalize the raw data for biological analysis [72].

25 Log2 ratios of gene expression levels between high- and low-FE groups were calculated to compare with the corresponding log ratio values from RNA-seq analysis.

2.2.7 Ingenuity Pathway Analysis Genes differentially expressed (FDR < 0.05) between high- and low-FE birds were included in pathway and function analysis using Ingenuity Pathway Analysis (IPA; Ingenuity® Systems, http://www.ingenuity.com). The functional and canonical pathway analysis was used to identify the significant biological functions and pathways. Functions and pathways with P-value < 0.05 (Fischer’s exact test) were considered to be statistically significant. IPA’s upstream regulator analysis function was used to identify potential transcriptional regulators that could explain the observed changes in gene expression between high- and low-FE chickens. The activation z-score was calculated to predict activation or inhibition of transcriptional regulators based on published findings accessible through the Ingenuity Knowledge Base. Regulators with z-score greater than 2 or less than –2 were considered to be significantly activated or inhibited.

2.3 Results and Discussion

2.3.1 Phenotype Measurements A summary of the phenotype measurements from 23 high-FE (n=10) and low-FE

(n=13) broilers is presented in Table 2.1. Although the initial bird weights (BW29) are not significantly different between these two groups (P = 0.661), the mean body weight of high-FE birds is significantly heavier than that of low-FE birds at the end of the test (P < 0.05), and the high-FE chickens consumed significantly less feed than

26 low-FE birds during the test (P < 0.01). Consequently, the difference in mean RFC values between high- and low-FE broilers is highly significant (P < 0.001). The mean breast muscle weight and breast muscle percentage of the high-FE birds are significantly higher than those of low-FE birds (P < 0.05).

Table 2.1 Statistics of the Measurements from High- and Low-Feed Efficiency (FE) Chickens

Measurements High-FE birds (n=10) Low-FE birds (n=13) Bird weight (Kg), 29-d 1.316±0.140 1.345±0.169 Bird weight (Kg), 46-d 3.093±0.283 2.960±0.176 Weight gain (Kg), 29- to 46-d 1.778±0.188 * 1.615±0.099 * Feed consumption (Kg), 29- to 2.874±0.249 ** 3.325±0.136 ** 46-d Feed conversion ratio a 1.620±0.054 ** 2.063±0.085** Residual feed consumption (Kg)b -0.276±0.040** 0.356±0.048** Breast muscle weight (%BW), 23.2±01.6* 21.6±1.4* 47-d Breast muscle weight (Kg), 47-d 0.721±0.100* 0.648±0.071* The significance between high- and low-FE birds was determined using Fisher’s least significance difference (LSD) test. P ≤ 0.05 is denoted by *; P ≤ 0.01 is denoted by **. a Feed conversion ratio = Weight gained (29- to 46- d) / Feed consumption (29- to 46- d) b Residual feed consumption = FC – (a*Row + b1*BW29 + b2*BW46 + c), where FC represents the feed consumption of each bird; Row stands for the effects of cage location on FE; BW29 is the initial (29-day) body weight; BW46 is the ending (46-day) body weight; c is the intercept; and a, b1 and b2 are the partial regression coefficients of FI on Row, BW1 and BW2.

2.3.2 Transcriptional Profile of Chicken Breast Muscle A total of 23 cDNA libraries were constructed using RNA samples of breast muscle tissues from 10 high- and 13 low-FE broilers and sequenced for 75 cycles from both ends on four lanes. In total, about 1.573 billion of 75-base sequence reads were

27 obtained with an average of 393 million raw reads per lane. No significant difference in the number of reads between these four lanes was observed. The total number of reads for one sample ranges from 50 million to 88 million, with an average of 68 million reads per sample. Based on quality check reports, the averaged quality score of sequence reads is high, approximately 38, with average GC content ranging from 49% to 51%. On average, 80% of the paired-end reads are properly mapped to the chicken reference genome (Ensembl Galgal4). The summary of alignment for all samples is shown in figure 2.1. The relative expression of a gene is normalized as fragments per kilobase of transcript per million mapped reads (FPKM), which is proportional to the number of cDNA fragments originated from the gene transcript. The lowest limit of gene expression value was set to be 0.1 FPKM in at least one of the 23 samples. According to this limit, 14,148 genes are identified as being expressed in the breast muscle tissues. To assess the consistency of the gene expression levels between different samples, the Pearson correlation coefficient was calculated for each pairwise combination of samples [73]. The averaged pairwise correlation coefficient between samples is 0.794, reflecting pretty consistent gene expression profiles.

28 100,000,000" Unmapped"reads"

90,000,000" Improperly"mapped"reads"

80,000,000" Properly"mapped"reads"

70,000,000"

60,000,000"

50,000,000" Read"Number" 40,000,000"

30,000,000"

20,000,000"

10,000,000"

0"

13917" 14000" 22237" 22360" 22370" 22513" 38847" 39348" 39629" 39663" 40012" 40128" 40679" 40707" 40752" 47246" 47337" 47424" 47731" 47762" 48713" 48797" 48883" Sample"ID"

Figure 2.1 The Number of Properly Mapped, Improperly Mapped and Unmapped Reads is Shown for Each Sample.

2.3.3 Gene Differential Expression Analysis

Differentially expressed genes were detected by Cuffdiff, an internal program of Cufflinks. Of 17,107 genes in the Ensemble database (Ensembl Galgal4), 1,059 were identified as significant genes with different expression levels between high- and low-FE broilers (q-value < 0.05) (Figure 2.2). All of this group of 1059 genes have a fold change greater than 1.3, and 642 genes (60.6%) have a fold change above 1.5. Among the 1,059 differentially expressed genes, 327 and 732 genes are down- and up- regulated in high-FE birds, respectively (Figure 2.3). This relative imbalance in the

29 number of down- and up-regulated genes is likely due to the increased breast muscle regeneration and inflammatory response in the high-FE chickens (discussed below). Since muscle development and inflammatory response requires higher levels of activators such as growth factors, hormones and cytokines, the gene expression is likely positively regulated by these activators in the breast muscle of the high-FE birds.

Figure 2.2 Volcano Plot Showing Differentially Expressed Genes between High- and Low-Fe Chickens.

30 ● ● ● ● ● not DE genes ● ● ● ● ● ● Downregulated genes ● ● ● ● ● ● ● ● Upregulalted genes ● ● ● ● ● ● ●●● ● ●●● ● ●●●●● ●● ●●●●● ● ● ●●●●●●●●● ● ●●●●● ● ●● ● 3 ● ● ● ●● ● ● ● ●●● ● ● ●●● ●● ● ●● ●●● ● ● ●●●● ● ●●● ●●●●●●●●● ● ●●●●●●● ● ●●●●● ● ●●●●●●● ● ● ●●●●●●●●●● ●● ●● ●●● ● ● ● ●●●●● ● ●● ● ● ●●●●●● ●● ●●●●● ●● ●● ●● ●● ●● ●●●●● ●● ●●●●●●●●● ●●●●● ●●● ● ●● ● ●●● ●●● ● ● ● ● ● ●●●●●●● ●●● ●● ●●●●●●●● ● ●●● ● ●●●●●●●●●● ● ●●●● ●●●●●●●●●●● ● ●● ●●●●●●●●●●●● ● ● ●●●● ●●●●●●●●●●●●● ● ● ●●●●●●●●●●●●●● ● ● ●● ●●●●●●●●●●●●● ●● ●●●●●●●● ●●● ●●●●●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●●● ●● ●●●●●●●●●●●●●●●●●● ● ● ●●●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●●●● ● 2 ● ●● ●●●●●●●●●●●●●●●●●●● ●● ● ● ●●●●●●●●●●●●●●●●● ● ●●●● ●●●●●●●●●●●●●●●●●● ● ● ●●●●●●●●●●●●●●●●●●● ● ● ●●● ●●●●●●●●●●●●●●●●● ●●● ●●●●●●●●●●●●●●●●●●●●●●●●●● ● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ● ●●●●●●●●●●●●●●●●●●●●●●● ● ●● ●●●●●●●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●●●●● ● ● ●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●●●●●●●● ● ● ● ● ●●●●●●●●●●●●●●●●●●●●●●●●●● ● ● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ● ● ●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ●● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ● ●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● Mean log10(FPKM+1) in high FE ●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ●● 1 ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ●●●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●● ● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ●● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ●●● ●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ● ●●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●● ●●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●●●●●●● ●●●● ●●●●●●●●●●●●●●●●●●●●●●●●●● ●● ●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●●●●●●●●●●●●●●●●● ●●●●●●●●●●●●●●●●●●●●●● ●●●●●●●●●●●●●●●●●●● ●●●●●●●●●●●●●●●● ● ●●●●●●●●●●●● 0 ●●●●●●●●

0 1 2 3 4 Mean log10(FPKM+1) in low FE

Figure 2.3 Comparison of Gene Expression between High- and Low-FE Broiler Chickens.

2.3.4 Confirmation of RNA-seq Data To verify the gene expression data obtained from RNA-seq analysis, we selected 192 target genes (71 significant and 121 non-significant) along with 12 housekeeping genes for the NanoString nCounter® assay. Comparison of the normalized counts from NanoString with FPKM values derived from RNA-seq shows high concordance, with pair-wise Pearson correlation coefficients ranging from 0.70 to

31 0.98. The Pearson correlation coefficients of fold change in gene expression levels (log2 fold change with respect to high- and low-FE groups) between NanoString and RNA-seq results are also high: 0.7532 for all genes and 0.8332 for the 71 significantly differentially expressed genes (Figure 2.4). The correlation of log2 fold change between two analyses is notably affected by lowly expressed genes, and increases by excluding genes with low expression levels (Figure 2.5). This can explain why the significantly differentially expressed genes show greater correlation compared with all the selected genes, because the FPKM values of all the significant target genes are equal or greater than 0.4, whereas 23 genes out of the 121 non-significant target genes have an FPKM value less than 0.4.

● ●

● 2 ● ● Pearson−R = 0.8332 ● ● ●

● ● ● ● ●

● ● ● ● ● ● ● ● ● ● ● ● 1 ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● ● ●

● ● ● ● 0 ● ● ●

●● ●

● ● DE genes' log2 fold_change from NanoString DE genes' log2 fold_change ● ● ● ●

● ● −1 ● ● ● ● ●

−1 0 1 2 DE genes' log2 fold_change from RNA−seq Figure 2.4 Correlation of Log2 Fold-Change between RNA-Seq and Nanostring for Significantly Differentially Expressed Target Genes.

32 0.95$

0.9$

0.85$

0.8$

0.75$ !Correla(on!between!RNA0seq!!and!NanoString!

0.7$ 4$ 6$ 8$ 10$ 12$ 14$ 16$ 18$ 20$ 0.5$ 0.9$ 1.5$ 2.5$ 0.1$ 0.3$ 0.7$ 0.06$ 0.18$ 0.26$ 0.38$ 0.46$ 0.58$ 0.66$ 0.78$ 0.86$ 0.98$ 0.02$ 0.14$ 0.22$ 0.34$ 0.42$ 0.54$ 0.62$ 0.74$ 0.82$ 0.94$ Minimum!FPKM!cutoff!

Figure 2.5 The Correlation of Log2 Fold-Change between RNA-Seq and Nanostring Increased with Gene Expression Levels.

2.3.5 Overview of IPA Analysis To fully interpret the biological implications of the results from the differential expression analysis, all significant genes with their respective log2 fold change ratios were submitted for Ingenuity Knowledge Base analysis. The top 10 up-regulated and top 10 down-regulated genes in high-FE broilers are listed in table 2.2. A summary of the IPA analysis, including top five biological functions and canonical pathways, are presented in Table 2.3. Generally, most of the differentially expressed genes are related to immune response and metabolic processes. Genes up-regulated in high-FE birds are associated with cell movement, growth and death. In

33 contrast, genes up-regulated in low-FE birds are associated with metabolic processes including lipid, nucleic acid and carbohydrate metabolism (Table 2.4). Upstream regulator analysis through IPA predicted the cascade of upstream transcriptional regulators that can potentially explain the differences in gene expression profile between high- and low-FE broilers. A summary of the upstream regulators identified by IPA is presented in Table 2.5. A total of 27 transcriptional regulators are predicted to be activated or inhibited (24 activated and 3 inhibited) in high-FE broilers, of which 24 regulators are considered to be significant with P-value < 0.05 (21 activated and 3 inhibited).

Table 2.2 Top 10 Up-Regulated and Down-Regulated Genes in High-FE Chickens

Top10 up-regulated genes in high-FE chickens Fold-Change from Fold-change from Symbol Description RNA-seq NanoString Matrix metallopeptidase 1 (interstitial MMP1 16.16 1.47 collagenase) NPPA Natriuretic peptide A 6.38 - - - 5.82 -1.01 MYL3 Myosin, light chain 3, alkali 5.37 6.02 TGM4 Transglutaminase 4 5.11 4 C-type lectin domain family 3, member CLEC3A 4.73 - A K60 Interleukin 8-like 1 4.58 - MLANA Melan-A 4.55 2.93 - - 4.5 1.89 PTX3 Pentraxin 3, long 4.44 5.31 Top10 down-regulated genes in high-FE chickens Fold Change from Fold-change from Symbol Description RNA-seq NanoString CCDC81 Coiled-coil domain containing 81 -3.85 -

34 ChaC, cation transport regulator CHAC1 -3.48 - homolog 1 GAL9 Gal 9 -3.28 - MPO Myeloperoxidase -3.19 - Beaded filament structural protein 1, BFSP1 -3.17 -2.13 filensin SDK1 Sidekick cell adhesion molecule 1 -3.01 1.09 SUN3 Sad1 and UNC84 domain containing 3 -2.58 1.01 BG2 Intestinal zipper protein -2.58 - GPR160 G protein-coupled receptor 160 -2.47 - PROKR2 Prokineticin receptor 2 -2.46 -2.08

Table 2.3 Top Biological Functions and Pathways Enriched by Differentially Expressed Genes between High- and Low-FE Chickens.

Top molecular and cellular functions Functions P-value #Molecules Cellular Movement 4.36E-18-2.57E-04 187 Cellular Function and Maintenance 2.38E-16-2.57E-04 241 Cell-To-Cell Signaling and Interaction 1.23E-13-3.09E-04 167 Cellular Growth and Proliferation 2.89E-13-3.02E-04 259 Cell Death and Survival 1.70E-12-2.72E-04 256 Top canonical pathways Pathways P-value Ratio Hepatic Fibrosis / Hepatic Stellate Cell Activation 4.22E-09 23/146 (0.158) Fcg Receptor-mediated Phagocytosis in 1.65E-08 18/102 (0.176) Macrophages and Monocytes Leukocyte Extravasation Signaling 8.56E-07 24/207 (0.116) Role of Tissue Factor in Cancer 1.36E-06 17/116 (0.147) PI3K Signaling in B Lymphocytes 9.94E-06 17/140 (0.121) 1A summary results from Ingenuity® Pathway Analysis (IPA) Software.

Table 2.4 Top Functions Enriched by Genes Up-Regulated in High- or Low-FE Broilers

Function annotation P-value #Molecules Genes up-regulated Lipid metabolism 6.43E-06 -3.51E-02 34 in low-FE chickens Molecular transport 6.43E-06 -3.54E-02 43 Small molecule biochemistry 6.43E-06 -3.54E-02 62

35 Nucleic acid metabolism 3.28E-05 -3.54E-02 21 Carbohydrate metabolism 7.79E-05 -3.54E-02 35 Genes up-regulated Cellular movement 3.78E-26 -7.96E-05 162 in high-FE chickens Cellular function and 1.62E-19 -6.97E-05 194 maintenance Cellular growth and 2.22E-19 -6.75E-05 210 proliferation Cell-to-cell signaling and 6.74E-18 -6.20E-05 145 interaction Cell death and survival 2.44E-15 -7.97E-05 198

Table 2.5 Activated or Inhibited Upstream Transcription Regulators Predicted by Ingenuity® Pathway Analysis (IPA) Software.

Upstream Regulator Activation z-score P-value of overlap Activated CTNNB1 2.687 2.59E-09 JUN 2.923 1.70E-08 FOS 2.277 7.64E-07 HIF1A 2.332 3.85E-06 SP1 3.347 1.33E-05 NFE2L2 2.036 1.94E-05 ETS1 2.91 1.25E-04 SPI1 2.322 2.51E-04 STAT4 2.984 3.25E-04 EGR1 2.95 3.55E-04 ETV1 2.2 5.24E-04 JUNB 2.425 1.20E-03 SIX5 2.2 1.64E-03 RUNX1 2.511 1.69E-03 SREBF2 2.621 4.30E-03 STAT3 3.696 4.77E-03 NKX2-3 2.035 4.85E-03 HMGB1 2.564 6.02E-03

36 XBP1 2.728 1.16E-02 SMAD2 2.211 2.91E-02 MEF2C 2.186 3.11E-02 SNAI1 2 5.55E-02 RELA 2.429 8.34E-02 PAX6 2.191 2.92E-01 Inhibited HOXA10 -2.047 4.02E-04 SPDEF -2.236 3.54E-02 REST -2.201 3.78E-02

2.3.6 Increased Muscle Growth and Remodeling in High-FE Chickens Of all differentially expressed genes, 32 are associated with muscle development, supporting the increased breast muscle weight in the high-FE birds. Among them, both hepatocyte growth factor (HGF) and insulin-like growth factor 2 (IGF2) encode key growth factors that have autocrine or paracrine effects on chicken skeletal muscle development and regeneration [74]. HGF can not only activate the proliferation of quiescent muscle satellite cells, it also can induce the migration of activated satellite cells to the injured sites [75]. IGF2 acts as a crucial regulator in muscle regeneration by stimulating muscle cell differentiation as well as inducing muscle cell hypertrophy [76]. Other muscle growth-related genes that are up-regulated in the high-FE chickens include myogenin (MYOG), cysteine and glycine-rich protein 3 (CSRP3), myoferlin (MYOF), glypican 1 (GPC1), protein tyrosine phosphatase, receptor type, A (PTPRA) and gap junction protein (GJA1). As a member of myogenic regulatory factors (MRFs), MYOG is essential for the fusion of myoblasts into myotubes during muscle growth and regeneration [77]. The CSRP3 gene encodes

37 muscle LIM protein, which is able to increase the activity of MRFs and plays a critical role in enhancing myogenesis [78]. The MYOF-encoded protein is a fundamental modulator for myoblast fusion, highly expressed during muscle repair and regeneration [79]. The stimulatory effects of GPC1 on muscle satellite cell differentiation and myotube formation was reported in turkeys [80]. The protein encoded by the PTPRA gene is a signaling molecule that was found to increase myogenesis of rat muscle L6 cells [81]. The GJA1-encoded protein is a major component in gap junctions and required for myogenesis and regeneration [82]. Collectively, the up-regulation of genes that can positively regulate muscle growth indicates that muscle growth and development is elevated in the high-FE chickens. In addition to genes involved in muscle development, genes associated with muscle hypertrophy, including F-box protein 32 (FBXO32; fold change = –1.879), F-box protein 40 (FBXO40; fold change = –1.879], F-box protein 9 (FBXO9; fold change = –1.347), forkhead box O3 (FOXO3; fold change = –1.540) and myotrophin (MTPN; fold change = 1.426), are found differentially expressed in the breast muscle of chickens with high versus low FE. MTPN, a well-known positive growth factor in promoting muscle growth [83], is up-regulated in high-FE chickens. The increased MTPN expression may indicate that myocyte growth and protein synthesis are augmented in the breast muscle of high-FE birds, accordingly, contributing to breast muscle hypertrophy in these chickens. Furthermore, the down-regulation of FOXO3 and F-box family proteins in high-FE broilers further explains muscle growth differences between high- and low-FE chickens. Protein encoded by FOXO3 is a master regulator of both autophagy-lysosome and ubiquitin-proteasomal pathways, promoting protein degradation and thereby contributing to muscle atrophy [84].

38 Proteins from the F-box family mediate the interaction between substrates and ubiquitin-conjugating enzymes, which facilitate proteolysis in diverse tissues [85]. Of them, the FBXO32-encoded protein, known as atrogin 1, is a well-recognized muscle-specific ubiquitin ligase leading to muscle atrophy in a wide range of diseases [86][87][88]. The FBXO40-encoded protein also has been proposed to play a role in muscle atrophy in mammals [89]. Thus, the decreased expression of atrophy-related genes in breast muscle of high-FE birds suggests that muscle protein loss is reduced in high-FE chickens in contrast to low-FE birds. The transcription of these genes is regulated by the PI3K/AKT signaling pathway, which will be discussed later. Taken together, the up-regulation of MTPN and down-regulation of FOXO3 and FBXO family genes in the high-FE chickens suggest that birds with high FE may have elevated protein synthesis and decreased protein degradation in their breast muscle. Genes associated with (ECM) remodeling are also up-regulated in the high-FE birds. The ECM of skeletal muscle serves as a scaffold for maintaining the structure of muscle and guiding new fiber formation [90]. Muscle regeneration is frequently accompanied by the degradation of ECM because it facilitates satellite cell migration to specific sites for proliferation and fusion into myotubes [77][91]. Therefore, the up-regulation of genes involved in ECM remodeling implies that muscle remodeling is increased in the breast muscle of high-FE chickens. Matrix metalloproteinases (MMPs) are the main endopeptidases responsible for degrading all kinds of ECM, consequently, playing an important role in mediating muscle cell migration and regeneration [92][93]. As presented in Table 2.6, six genes from the MMP family are differentially expressed in our study, all of which are up-regulated in high-FE birds. Of the proteins encoded by these genes,

39 MMP1 and MMP13 belong to MMP collagenases that are capable of cleaving interstitial collagen types I, II and III [94]. Through an in vitro wound-healing assay, MMP1 was able to promote myoblast migration and differentiation by increasing the expression of N-cadherin and β-catenin or pre-MMP-2 and TIMP [95]. MMP13 also plays a role in muscle regeneration, expressing in all muscle cells during muscle regeneration, and its expression level is positively correlated with the extent of muscle damage [96]. MMP9 is a gelatinase that also relates to muscle regeneration [94]. Evidence showed that the expression levels of MMP9 were greatly increased in response to inflammation and the activation of satellite cells in injured muscles [97][98][99]. However, contrary to the positive function of MMP1 on muscle regeneration, a recent study revealed that MMP9 could lead to muscle cell necrosis, inflammation and fibrosis [100]. Collectively, the up-regulation of MMPs in the high-FE birds suggests an augmented muscle remodeling in these birds compared with the low-FE chickens.

Table 2.6 Differentially Expressed Genes from Matrix Metalloproteinases (Mmps) Family

Genes in Fold change from Fold change from dataset RNA-seq Nanostring MMP28 1.397 - MMP9 3.9 3.297 MMP13 2.203 - MMP7 2.026 - MMP27 1.945 - MMP1 16.164 1.371

40 The IPA upstream regulator analysis provides additional support to our conclusions regarding muscle development and remodeling in the high-FE birds. According to the predictions from IPA, several transcriptional factors involved in muscle development are activated in high-FE chickens. As a regulator of postnatal muscle growth, JunB is predicted as being activated in the breast muscle of the high-FE birds (P-value = 1.2E-03, z-score = 2.200). JunB can stimulate myosin expression to elevate protein synthesis, accordingly, maintaining muscle mass and promoting muscle hypertrophy [101]. Additionally, JunB can suppress the transcription of FOXO3 and thereby inhibit protein degradation in muscle [102]; therefore, activation of JunB in the high-FE birds may be one of the causes of the down-regulation of FOXO3 as well as of FBXO32 in breast muscle of these birds. In addition, through the IPA analysis, JunB is predicted to activate the transcription of MMP1, MMP9, MMP13, fibronectin 1 (FN1), heme oxygenase 1 (HMOX1), neutrophil cytosolic factor 2 (NCF2) and interleukin 1 receptor-like 1 (IL1RL1) (Figure 2.6A). As discussed above, FN1, MMP1 and MMP13 are all positively correlated with muscular satellite cell proliferation. Thus, JunB may also have increased myogenesis through activating transcription of these genes in the high-FE birds. Apart from JunB, a main transcriptional factor in the formation of mature sarcomeres, myocyte-specific enhancer factor 2C (MEF2C), is predicted as being activated in breast muscle of high-FE birds [103][104]. Protein encoded by MEF2C is a member of the myocyte enhancer factor 2 (MEF2) family, which directly cooperates with MRFs and enhances skeletal muscle development [105]. In the present study, MEF2C is predicted to be an activated upstream regulator that increases the

41 transcription of GJA1, MMP13, MYOG, myozenin 2 (MYOZ2; fold change = 2.400) and ATPase, Ca++ transporting, , slow twitch 2 (ATP2A2) (fold change = 1.390) (Figure 2.6B). GJA1, MMP13 and MYOG are all involved in myogenesis and exert positive effects on skeletal muscle growth and regeneration, which has been discussed above. Thus, MEF2C’s activation may augment the muscle development in high-FE birds. Moreover, MYOZ2 is also predicted as being up-regulated by MEF2C. The MYOZ2-encoded protein belongs to a family of -interacting proteins that modulates specific skeletal muscle signaling pathways through suppressing calcineurin [106]. It has been shown that MYOZ2 plays a role in regulating myocyte differentiation and promoting slow-oxidative fibers growth [107]. Collectively, MYOZ2 may be more active in breast muscle of high-FE chickens and, consequently, mediates some biological pathways and leads to muscle remodeling in these birds.

Figure 2.6 Upstream Regulators JunB and MEF2C.

42 2.3.7 Growth Hormone (GH) and IGFs/PI3K/AKT Signaling Pathway Over-represented in the Differentially Expressed Genes

Through the IPA canonical pathway analysis, several critical pathways in the regulation of body and muscle growth are over-represented among the differentially expressed genes. One of these pathways is the GH signaling pathway, enriched by 10 genes in our dataset (P-value = 3.25E-04; ratio = 1.32E-01) (Figure 2.7). As a key mediator of body size, GH has an anabolic effect on skeletal muscle development [108]. Through binding to growth hormone receptor (GHR) in muscles, biologically active GH can activate nuclear receptor STAT5 and thereby induce the synthesis and secretion of IGF-1 as well as IGF-2 (fold change = 1.657) [109]. Furthermore, GH can stimulate signaling molecules including PI3K [phosphatidylinositol-4,5-bisphosphate 3-kinase, catalytic subunit beta (PIK3CB; fold change = 1.663); phosphatidylinositol-4,5-bisphosphate 3-kinase, catalytic subunit delta (PIK3CD; fold change = 1.709); phosphoinositide-3-kinase, regulatory subunit 5 (PIK3R5; fold change = 1.564) and PKC (PRKCD; fold change = 1.558)], leading to the activation of the AKT/PKB signaling pathway and STAT5 [110]. Both the PI3K/AKT/PKB pathway and IGFs are crucial contributors to muscle hypertrophy, which will be discussed later. Most of the mapped genes (8 of 10) are up-regulated in the high-FE broilers, indicating that the GH signaling pathway is more activated in the breast muscle of the high-FE birds compared with the low-FE birds. The two down-regulated genes [GHR and insulin-like growth factor binding protein (IGFBP3)] also fit this assumption. Evidence from the literature indicates that the expression of GHR is inversely correlated with the concentration of GH [111][112]. Thus, the down-regulation of GHR gene expression in the high-FE birds may be due to relatively high circulating GH levels in these birds. In spite of the stimulatory effects

43 of GH on IGFBP3 transcription, there may be other modulators inhibiting the expression of IGFBP3 in high-FE chickens, consequently exerting an inhibitory effect on IGF-1 function [113]. Another important finding is that the IGFs/PI3K/Akt signaling pathway is over-represented by the differentially expressed genes. The IGFs/PI3K/Akt signaling pathway plays a key role in the regulation of muscle growth and muscle hypertrophy in a variety of organisms [114][115][116]. Nine differentially expressed genes are mapped to the IGFs/PI3K/Akt pathway (Figure 2.8). Of these, PIK3CD (fold change = 1.709), PIK3CB (fold change = 1.663) and PIK3R5 (fold change = 1.564) are up-regulated in the high-FE chickens, implying that the PI3K complex is more active in the breast muscle of these birds. The up-regulated members of the PI3K complex are predicted to increase PI3K-AKT cascade activity in the high-FE birds by IPA (Figure 2.8). Activated PI3K can induce the phosphorylation of phosphatidylinositol 4,5-bisphosphate (PIP2) to generate phosphatidylinositol-3,4,5-trisphosphate (PIP3). PIP3 acts as a docking site for phosphoinositide-dependent kinase 1 (PDK1) and Akt and subsequently contributes to the activation of Akt [117][118]. On the basis of the IPA prediction, the activated Akt translocate into the nucleus and then inhibit the transcription of the forkhead box O (FOXO) family, which is consistent with the down-regulation of the forkhead box O3 (FOXO3) gene (fold change = –1.540) in our results [114]. As mentioned above, FOXO3 can promote protein degradation and muscle atrophy [84]. Hence, considering the expression profile of the mapped genes, the protein degradation process is predicted to be reduced in the breast muscle of the high-FE chickens as a result of PI3K/Akt pathway activation. In addition, because activation of Akt can up-regulate the transcription of ATP citrate lyase (ACLY)

44 through suppressing the activity of kinase (Gsk3), the increased expression of ACLY in the high-FE birds (fold change = 1.345) lends more support to the assumption that the PI3K/Akt pathway is activated in the high-FE chickens. Apart from repressing protein degradation, the activated PI3K/Akt pathway can promote protein synthesis in muscle via inhibiting Gsk3 [114], which is also inferred in our analysis. Therefore, the up-regulated IGFs/PI3K/Akt pathway suggests increased protein synthesis as well as decreased protein degradation in the breast muscle of the high-FE birds, explaining in part why high-FE broilers have more breast muscle than do low-FE birds.

45

Figure 2.7 Growth Hormone Signaling Pathway Analysis Using Ingenuity Molecule Activity Predictor (MAP).

46

Figure 2.8 IGFs/PI3K/AKT Signaling Pathway Analysis Using Ingenuity Molecule Activity Predictor (MAP).

47 2.3.8 Inflammatory Response in the Breast Muscle of High-FE Broilers In the present study, a large number of the differentially expressed genes (136 genes) are involved in inflammatory response. Most of these genes (124 genes), including genes encoding for interleukin 8 (IL-8) and chemokine (C-X-C motif) ligand 14 (CXCL14), are expressed greater in the high-FE broilers. Although the cellular source of IL-8 and CXCL14 remains unknown in the present study, both not only exert direct effects on immune cell recruitment but also act as paracrine or endocrine factors in skeletal muscles. IL-8 has been recently classified as a myokine that can promote angiogenesis within the muscle [119][120]. CXCL14 is encoded by an obesity-induced gene in mice that inhibits the insulin-induced glucose uptake in cultured myocytes [121]. In addition, the gene encoding for corticotropin-releasing hormone (CRH) is also up-regulated in the high-FE birds (fold change = 2.824). Previous studies have demonstrated that CRH is secreted from nerve terminals and epithelial cells at inflammation sites and has a local proinflammatory effect on resident immune cells [122]. Therefore, it is likely that the elevated transcription of CRH functions to augment an immune response in the breast muscle of high-FE chickens. Apart from its immunomodulatory role, an increase in CRH may have a positive impact on thermogenesis of skeletal muscles in high-FE birds [123]. A series of genes encoding for cytokine receptors are also up-regulated in the high-FE chickens, further indicating an augmented immune response in the breast muscle of the high-FE chickens. These genes include chemokine (C-C motif) receptor 2 (CCR2), chemokine (C-C motif) receptor 5 (CCR5), interleukin 17 receptor A (IL17RA), interleukin 18 receptor 1(IL18R1), interleukin 1 receptor, type I (IL1R1), interleukin 1 receptor, type II (IL1R2), interleukin 1 receptor-like 1 (IL1RL1), interleukin 2 receptor, gamma (IL2RG) and interleukin 5 receptor, alpha (IL5RA).

48 Among them, CCR2 was found to be expressed on infiltrating macrophages and play a crucial role in muscle regeneration [124]. This gene can recruit macrophages to injured muscles, which then produce a high level of IGF-I to promote muscle regeneration [125]. Therefore, the up-regulation of CCR2 suggests that, compared with the low-FE chickens, macrophage infiltration and muscle regeneration are increased in the breast muscle of the high-FE birds. The IPA canonical pathway analysis also supports our hypothesis regarding augmented immune response and active recruitment of immune cells to the breast muscle of the high-FE chickens. Several over-represented pathways involved in inflammatory response are identified in our analysis (Table 2.7). Given that nearly all genes mapped to these pathways are up-regulated in the high-FE chickens, we conclude that these immune-related pathways are activated in the breast muscle of the high-FE chickens. According to the predictions from IPA, a number of transcription factors associated with inflammatory response are also activated in the breast muscle of the high-FE broilers: v-ets erythroblastosis virus E26 oncogene homolog 1 (ETS1), spleen focus forming virus (SFFV), spi-1 proto-oncogene (SPI1), X-box binding protein 1 (XBP1) and runt-related transcription factor 1(RUNX1). Muscle inflammation is a key step in muscle remodeling, which can clean disrupted muscle cells and promote muscle regeneration by releasing growth factors [126]. A variety of circumstances (e.g., muscle injury, exercise and obesity) can activate transcription factors NF-kB and c-Jun/AP-1 in muscle cells, resulting in the expression and secretion of several factors [127]. These factors, including cytokines and other non-protein mediators, can either directly attract circulating immune cells to the muscle or activate resident immune cells, providing chemotactic signals for

49 recruitment [128]. As a consequence, a number of immune cells are recruited to the muscle, phagocytizing the cellular debris and producing cytokines affecting muscle cells [127][129]. The IPA upstream regulator analysis predicts that transcription factors JUN and FOS are activated in the breast muscle of the high-FE birds (Figure 2.9). Protein encoded by JUN and FOS are components of activator protein 1 (AP-1), which is an important transcription factor responding to various physiological and pathological stimuli [130]. Overall, our results suggest that, compared with the low-FE birds, the high-FE birds experienced a more intense muscle restructuring and thereby higher inflammatory responses in the breast muscle.

Table 2.7 Over-represented Pathways Involved in Immune Response

Canonical Pathways involved in inflammatory response P-value Ratio Fcγ Receptor-mediated Phagocytosis in Macrophages and 3.55E-06 1.76E-01 Monocytes Leukocyte Extravasation Signaling 1.23E-04 1.16E-01 PI3K Signaling in B Lymphocytes 7.58E-04 1.21E-01 fMLP Signaling in Neutrophils 3.31E-03 1.08E-01 Natural Killer Cell Signaling 3.39E-03 1.20E-01 Granulocyte Adhesion and Diapedesis 4.17E-03 1.01E-01 FcγRIIB Signaling in B Lymphocytes 4.17E-03 1.36E-01 CD28 Signaling in T Helper Cells 2.13E-02 9.09E-02 Role of JAK1 and JAK3 in γc Cytokine Signaling 2.75E-02 1.19E-01 IL-8 Signaling 2.75E-02 7.69E-02 B Cell Receptor Signaling 3.39E-02 8.19E-02 Integrin Signaling 3.72E-02 7.69E-02 Macropinocytosis Signaling 3.80E-02 1.05E-01 IL-6 Signaling 4.17E-02 8.87E-02 IL-4 Signaling 4.79E-02 1.01E-01

50

Figure 2.9 Upstream Regulator JUN and FOS.

51 2.3.9 Free Radical Scavenging Enriched in the Differentially Expressed Genes between High- and Low-FE Broilers

Several differentially expressed genes in our dataset are involved in the production of reactive oxygen species (ROS). Genes encoding for ROS-generating enzymes, including cytochrome b-245, beta polypeptide (CYBB) [fold-change = 2.08] and NADPH oxidase organizer 1 (NOXO1) [fold-change = 2.38], are all up-regulated in the high-FE birds, suggesting that ROS production is increased in the breast muscle of these birds compared with the low-FE birds. CYBB, also known as NADPH oxidase 2 (NOX2), is a major enzyme responsible for producing superoxide in the sarcoplasmic reticulum [131]. NOXO1, a positive mediator of NOX1 and NOX3, initiates the activity of NOX1 and NOX3 for generating ROS [132]. Moreover, the down-regulation of uncoupling protein 3 (UCP3) [fold-change = -1.67] in the high-FE birds may indicate that mitochondria from the breast muscle of the high-FE chickens have higher electron transport chain coupling compared with that from low-FE chickens. This assumption is consistent with previous findings [29][28]. Because UCP3-mediated uncoupling can attenuate the production of ROS [133], the down-regulation of UCP3 in the high-FE birds may also suggest a higher production of ROS from the mitochondria of the breast muscle of these birds. Collectively, our data suggest that, compared with the low-FE birds, ROS is produced at a higher level in the breast muscle of the high-FE chickens. The IPA downstream effect analysis supports our hypothesis regarding increased ROS production in the breast muscle of high FE-chickens. Processes, including metabolism of reactive oxygen species (P-value = 5.77E-07), synthesis of reactive oxygen species (P-value = 1.75E-06), production of reactive oxygen species (P-value = 4.92E-06) and production of superoxide (P-value = 2.05E-03), are

52 predicted to be increased in the high-FE broilers. In addition, the NRF2-mediated oxidative stress response pathway is over-represented among the differentially expressed genes, with 17 genes (P-value = 6.96E-04; ratio = 0.089) mapped to this pathway (Figure 2.10A). Nuclear factor (erythroid-derived 2)-like 2 (NRF2), also known as NFE2L2, is a key transcription factor in cells that responds to a range of oxidative and xenobiotic stresses [134]. Upon exposure of cells to various stimuli such as ROS and electrophilic compounds, quiescent NRF2 in cytoplasm is activated through phosphatidylinositol 3-kinase (PI3K), RAS and protein kinase C (PKC) signaling pathways [135]. By phosphorylation or binding to actin, activated NRF2 translocates into the nucleus and binds to the antioxidant response elements, initiating the transcription of a number of genes encoding for antioxidants and ROS detoxifying enzymes [136]. A group of NRF2 downstream genes, including v-maf musculoaponeurotic fibrosarcoma oncogene homolog F (avian) (MAFF) [fold-change = 1.61], glutathione S- A3 (GSTA3) [fold-change = 2.03], glutathione S-transferase omega 1 (GSTO1) [fold-change = 1.50], heme oxygenase (decycling) 1 (HMOX1) [fold-change = 1.36], microsomal glutathione S-transferase 1 (MGST1) [fold-change = 1.39], superoxide dismutase 3 (SOD3) [fold-change = 1.85], thioredoxin (TXN) [fold-change = 1.50], peptidylprolyl isomerase B (PPIB) [fold-change = 1.43], aldehyde oxidase 1 (AOX1) [fold-change = -1.63], DnaJ (Hsp40) homolog, subfamily A, member 1 (DNAJA1) [fold-change = -1.39] and DnaJ (Hsp40) homolog, subfamily C, member 1 (DNAJC1) [fold-change = -1.35], are differentially expressed in the current study. Genes encoding for antioxidant proteins, such as SOD3, HMOX1 and TXN, are up-regulated in the high-FE birds. Protein encoded by SOD3 is an extracellular protective enzyme

53 against not only ROS but also inflammation, thus playing a role in tissue recovery [137]. HMOX1 is increased in the condition of oxidative stress and has an effect on protecting cells against ROS and inflammation [138]. The TXN-encoded protein is involved in a range of redox reactions and can decrease the quantity of ROS [139]. The up-regulation of TXN, SOD3 and HMOX1 indicates that an NRF2-mediated antioxidant response is activated in the breast muscle of the high-FE chickens. Additionally, three members from the glutathione s-transferase (GST) group, encoded by GSTO1, GSTA1 and MGST1, are all up-regulated in the high-FE birds. GST is known for its function in detoxification of xenobiotics as well as endogenous metabolites [140]. The increased expression of the GST superfamily also suggests that responses to oxidative stress are elevated in the breast muscle of the high-FE chickens. Although few genes in the NRF2 signaling pathway, including AOX1, DNAJA1 and DNAJC1, are down-regulated in the high-FE chickens, overall there are 14 up-regulated genes mapped to this pathway, indicating that NRF2-mediated oxidative stress response is augmented in the breast muscle of the high-FE birds. Moreover, NRF2 (NFE2L2), a transcription factor, is predicted to be activated in the high-FE broilers (P-value = 1.94E-05; z-score = 2.036) (Figure 2.10B). Taken together, our results suggest a higher level of ROS generated in the breast muscle of high-FE broilers. However, in contrast to our findings, Bottje et al. (2002) reported higher amounts of ROS produced in the breast muscle of their low-FE birds [35]. This inconsistency is likely caused by the difference in broiler chickens between two studies. Male breeders, presumably with relatively low breast muscle yield, were studied in the Bottje et al. [29], whereas we study broilers from a commercial cross

54 with high breast muscle yield. The ancestors of this cross have been intensively selected for the disproportionate growth of breast muscle, and the resulting higher levels of variation in breast muscle in the broiler cross may be responsible for a significant part of the variation in FE in this cross compared to the male breeder strain in the study by Bottje et al. [29]. In regard to broilers in the current study, intensive inflammatory response is possibly a major source of increased ROS in the breast muscle of the high-FE broilers. ROS-generating enzymes, such as NOX in muscle cells, can be stimulated through extracellular inflammatory cytokines including interleukin (IL)-1, IL-6 and IL-8 in a ligand-induced pattern [141][142]. Furthermore, the implied infiltrating immune cells in the breast muscle of high-FE birds may be another cause for increased ROS. It’s well known that immune cells produce a large amount of ROS to support their functions during inflammation [143]. Hence, in our study, strong indications for elevated ROS production in the breast muscle of the high-FE chickens are likely due to augmented inflammatory response, whereas the higher level of ROS observed in the study by Bottje et al. (2002) is possibly from mitochondria of breast muscle cells. Further study of genes associated with free radical scavenging may support our assumption. Indeed, in our study, a large part of these genes are also related to inflammatory response (P-value = 1.03E-23–5.15E-06), indicating that the production of ROS in the high-FE birds is closely associated with an increased immune response in the breast muscle.

Notably, growth factors involving HGF, IGF-1 and fibroblast growth factor (FGF)-2 are also found to be able to induce intracellular generation of ROS in different types of cells [141]. As mentioned above, the breast muscle of high-FE birds have higher expression of HGF and IGF-2, which may play a role in

55 stimulating ROS production in these birds. Moreover, such generated ROS exerts insulin-mimicking effects on the insulin/IGFs signaling pathway, which has been shown to be a second messenger in insulin/IGFs signal transduction under physiological conditions [144]. Therefore, in the breast muscle of the high-FE birds, the insulin/IGFs receptor-signaling pathway may be activated, in part, because of increased ROS production. Higher ROS production may also lead to an increase in intracellular calcium concentration. It has been found that ROS mediates the influx of extracellular Ca2+ and mobilization of intracellular Ca2+ stores [145][146][147]. In this study, genes involved in calcium transport [solute carrier family 8, member B1 (SCL8B1), phospholipase C, beta 2 (PLCB2) and ATPase, Ca++ transporting, cardiac muscle, slow twitch 2 (ATP2A2)] are all up-regulated in the high-FE birds, indicating increased calcium mobilization in the breast muscle of these birds. ATP2A2 encodes sarcoplasmic reticulum Ca2+-ATPase isoform 2 (SERCA2), which is an important pump responsible for muscle relaxation through transporting Ca2+ from the cytosol into the sarcoplasmic reticulum lumen in muscle cells [148]. Because more SERCA2 are needed to maintain calcium homeostasis when high Ca2+ levels are present in cytosols, the up-regulation of ATP2A2 in the high-FE birds may imply a high level of cytosolic Ca2+ in the breast muscle of these chickens compared with the low-FE birds.

56

Figure 2.10 NRF2-mediated Oxidative Stress Response.

57 2.3.10 Transcriptional Regulation of Hypoxia-inducible Factor-1α (HIF1α) Hypoxia-inducible factor-1α (HIF1α) is a key transcription factor that mediates cell adaption to hypoxia through regulation of a variety of gene expression [149]. Although HIF1α mRNA is constantly expressed in cells under both normoxic and hypoxic conditions, the HIF1α protein has a very short half-life in normoxia because of degradation through the ubiquitin-proteasome system [149]. During hypoxia, HIF1α degradation is repressed. As a result, HIF1α translocates into the nucleus and activates downstream genes in response to low O2 tension [149]. In our study, HIF1α mRNA, HIF2β mRNA as well as HIF1α inhibitor hypoxia inducible factor 1, alpha subunit inhibitor (HIF1AN) mRNA are differentially expressed in the breast muscle between high- and low-FE chickens. HIF1α and HIF2β are up-regulated in the high-FE birds (fold change = 1.341 and 1.42, respectively), whereas HIF1α inhibitor HIF1AN is down-regulated in these birds (fold change = –1.343). Although the up-regulated HIF1α and HIF2β mRNA cannot represent increased amounts of stabilized HIF1α protein in the breast muscle of the high-FE chickens, decreased expression of HIF1AN may imply that HIF1α activity is increased in the breast muscle of the high-FE chickens compared with the low-FE birds. This assumption is supported by the expression of HIF1α downstream genes. As a transcription factor, HIF1α is predicted to be activated in the high-FE birds through the IPA upstream regulator analysis (P-value = 3.85E-06; z-score = 2.332; Figure 2.11). Indeed, most of the HIF1α target genes are up-regulated in the high-FE birds, indicating the activation of HIF1α in these birds. Moreover, the HIF1α signaling pathway is over-represented among significantly differentially expressed genes (-log P-value = 3.12; ratio = 1.39E-01). In response to hypoxia or a variety of peptide stimulators under normoxic conditions,

58 AKT/PI3K and MAPK signaling pathways are activated to induce the accumulation of HIF1α in human cells [150][151]. Consequently, the accumulated HIF1α is translocated to the nucleus to modulate the transcription of genes involved in angiogenesis, glucose metabolism, matrix metabolism, erythropoiesis and apoptosis [149]. In our study, with increased expression of PIK3CB, PIK3CD, PIK3R5 and muscle RAS oncogene homolog (MRAS), both the AKT/PI3K and MAPK signaling pathways are predicted to be activated in the high-FE chickens. The activated AKT/PI3K and MAPK signaling pathways may stimulate the induction of HIF1α, as reflected by the up-regulation of HIF1α and its downstream genes [glucose transporter type 3 (GLUT-3), glucose transporter-like protein 5 (GLUT5), matrix metallopeptidase 1 (MMP1), matrix metallopeptidase 7 (MMP7), matrix metallopeptidase 9 (MMP9), matrix metallopeptidase 13 (MMP13), matrix metallopeptidase 27 (MMP27), matrix metallopeptidase 28 (MMP28) and lactate dehydrogenase B (LDHB)] in the high-FE birds. Based on the gene expression profile, we conclude that, compared with the low-FE birds, the activity of HIF1α signaling pathway is increased in the breast muscle of the high-FE birds.

59

Figure 2.11 Upstream Regulator HIF1α and Its Target Genes.

Although it is unclear from our results whether hypoxia and/or mediators such as IGFs induced HIF1α activation in the breast muscle of the high-FE birds, we would like to speculate here about potential mechanisms underlying this activation. It’s widely accepted that inflammation and hypoxia are closely interdependent in a wide array of physiological and pathological conditions [152][153][154][155][156]. Inflammation is frequently accompanied with hypoxia because of the high oxygen consumption of infiltrating immune cells [154]. Assuming an increased inflammatory response in the breast muscle of the high-FE birds, we speculate that the up-regulation of HIF1α is partly caused by an inflammation-induced hypoxia. Alternatively, the up-regulation of HIF1α may be caused by excessive muscle remodeling, which may

60 be the result of selection for breast muscle proportion. Elevated muscle rearrangements may have led to the reconstruction of vasculature, consequently reducing the blood flow and resulting in oxygen deficiency [157] in the breast muscle of the high-FE birds. Furthermore, insulin and IGFs have been shown to be modulators of HIF1α induction during both normoxia and hypoxia [151]. Given that IGF2 is up-regulated in the breast muscle of the high-FE birds, this growth factor may also have contributed to the activation of HIF1α. Finally, the activation of HIF1α may also be partly due to a higher production of ROS in the breast muscle of the high-FE chickens. Studies have found that ROS are essential for the stabilization of HIF1-DNA, thereby triggering HIF1α-induced transcription [158][159]. It was also proposed that cellular ROS-producing proteins could sense changes in cellular oxygen concentration [160]. Evidence indicated that low oxygen tension inhibited mitochondrial electron transport and therefore increased ROS production. The generated ROS then acted as a second messenger that contributed to HIF1α activation [161]. Thus, the ROS production may have been increased in the breast muscle of the high-FE chickens partly because of a relatively low oxygen concentration within this tissue, which in turn may have played a role in HIF1α activation.

2.4 Conclusions The current study provides a global view of gene expression differences in the breast muscle of broiler chickens with extreme high and low FE from a population of a modern commercial high-meat-yield broiler cross. To our knowledge, this study reports for first time the RNA-seq analysis of a trait of selection and breeding importance in chickens. We identified 1,059 genes significantly differentially

61 expressed in the breast muscle between high- and low-FE chickens based on the RNA- seq experiment. Furthermore, we achieved a large-scale validation of our RNA-seq experiment by quantifying the expression of a large number of target genes (192 transcripts + 12 house-keeping genes) using a high-sensitive non-PCR-based method, i.e. NanoString nCounter® Technology[72]. Function and pathway analysis of the differentially expressed genes sheds light on some of the underlying mechanisms that regulate chicken FE. Birds with high FE exhibit higher expression of genes involved in muscle growth, development and remodeling, which may explain why these birds have more breast muscle than do the low-FE chickens. Pathway analysis shows that anabolic pathways, including growth hormone signaling and IGFs/PI3K/AKT signaling pathways, are more activated in the high-FE birds, which may have not only led to increased muscle growth in the high-FE chickens but also contributed to the feed conversion advantages of these birds. Our results also suggest that transcriptional factors JunB and MEF2C play crucial roles in regulating muscle growth and remodeling in high-FE chickens. Furthermore, most of the genes up-regulated in the high-FE birds are associated with inflammatory response and oxidative stress, suggesting augmented inflammation and oxidative stress in the breast muscle of these birds. Our results also show increased activity of HIF1α, which may be caused by a lower oxygen environment in the breast muscle of high-FE chickens. Although no clinical symptoms indicating sickness or muscle damage were observed in the birds used in the current study, some of the molecular changes in the high-FE chickens may be hypothesized to lead to recently reported muscle quality issues in modern broiler chickens such as white striping and wooden breast [162][163][164]. These disorders have been reported

62 to be more frequent in birds with high breast muscle weight and high FE, suggesting that the susceptibility may be primarily induced by breeding for these traits. Further investigation (e.g., histological and protein analysis) would be helpful for examining inflammation and oxidative stress in the breast muscle of high-FE and high-breast-muscle-yield birds.

63 REFERENCES

1. Macdonald JM: of U . S . Broiler Production Cataloging Record :. 2008.

2. Paisner MS, Witcover J: 2020 Global Food Outlook. 2001(August).

3. Donohue M, Cunningham DL: Effects of grain and oilseed prices on the costs of US poultry production. The Journal of Applied Poultry Research 2009:325– 337.

4. Willems OW, Miller SP, Wood BJ: Aspects of selection for feed efficiency in meat producing poultry. Worlds Poult Sci J 2013, 69:77–88.

5. Aggrey SE, Karnuah AB, Sebastian B, Anthony NB: Genetic properties of feed efficiency parameters in meat-type chickens. Genet Sel Evol 2010, 42:25.

6. Luiting P, Urff EM: Residual feed consumption in laying hens. 2. Genetic variation and correlations. Poult Sci 1991, 70:1663–1672.

7. Kennedy BW, van der Werf JH, Meuwissen TH: Genetic and statistical properties of residual feed intake. J Anim Sci 1993, 71:3239–3250.

8. Cai W, Casey DS, Dekkers JCM: Selection response and genetic parameters for residual feed intake in Yorkshire swine. J Anim Sci 2008, 86:287–298.

9. Arthur PF, Archer JA, Johnston DJ, Herd RM, Richardson EC, Parnell PF: Genetic and phenotypic variance and covariance components for feed intake, feed efficiency, and other postweaning traits in Angus cattle. J Anim Sci 2001, 79:2805–2811.

10. Ibarburu M, Bell D: and the Cost of Producing Eggs. 2013.

11. Havenstein GB, Ferket PR, Qureshi M a: Growth, livability, and feed conversion of 1957 versus 2001 broilers when fed representative 1957 and 2001 broiler diets. Poult Sci 2003, 82:1500–8.

12. Davis A: Preferences of Growing Fowls for Different Light Intensities in Relation to Age , Strain And Behaviour. Anim Welf 1999, 8:1999.

64 13. Bregendahl K, Sell JL, Zimmerman DR: Effect of Low-Protein Diets on Growth Performance and Body Composition of Broiler Chicks. Volume 81; 2002:1156–1167.

14. Holsheimer JP, Veerkamp CH: Effect of dietary energy, protein, and lysine content on performance and yields of two strains of male broiler chicks. Poult Sci 1992, 71:872–879.

15. Brown WR a, Hubbard SJ, Tickle C, Wilson S a: The chicken as a model for large- scale analysis of vertebrate gene function. Nat Rev Genet 2003, 4:87–98.

16. Blahová J, Dobšíková R, Straková E, Suchý P: Effect of Low Environmental Temperature on Performance and Blood System in Broiler Chickens (Gallus domesticus). Acta Vet Brno 2007, 76:S17–S23.

17. Factors that affect feed intake of meat .pdf. .

18. Song Z, Everaert N, Wang Y, Decuypere E, Buyse J: The endocrine control of energy homeostasis in chickens. Gen Comp Endocrinol 2013, 190:112–7.

19. Richards MP: Genetic regulation of feed intake and energy balance in poultry. Poult Sci 2003, 82:907–916.

20. Jensen J: Regulatory peptides and control of food intake in non-mammalian vertebrates. Comp Biochem Physiol A Mol Integr Physiol 2001, 128:471–479.

21. Woods SC, Seeley RJ, Porte Jr. D, Schwartz MW: Signals that regulate food intake and energy homeostasis. Science (80- ) 1998, 280:1378–1383.

22. Friedman JM: Leptin and the regulation of body weigh. The Keio journal of medicine 2011:1–9.

23. Friedman-Einat M, Boswell T, Horev G, Girishvarma G, Dunn IC, Talbot RT, Sharp PJ: The chicken leptin gene: has it been cloned?. Gen Comp Endocrinol 1999, 115:354–363.

24. Ohkubo T, Tanaka M, Nakashima K: Structure and tissue distribution of chicken leptin receptor (cOb-R) mRNA. Biochim Biophys Acta 2000, 1491:303–308.

25. Beck B: Neuropeptide Y in normal eating and in genetic and dietary-induced obesity. Philos Trans R Soc Lond B Biol Sci 2006, 361:1159–1185.

65 26. McMinn JE, Wilkinson CW, Havel PJ, Woods SC, Schwartz MW: Effect of intracerebroventricular alpha-MSH on food intake, adiposity, c-Fos induction, and neuropeptide expression. Am J Physiol Regul Integr Comp Physiol 2000, 279:R695–R703.

27. Tachibana T, Sugahara K, Ohgushi A, Ando R, Sashihara K, Yoshimatsu T, Furuse M: Intracerebroventricular injection of exendin (5-39) increases food intake of layer-type chicks but not broiler chicks. Brain Res 2001, 915:234– 237.

28. Bottje WG, Carstens GE: Association of mitochondrial function and feed efficiency in poultry and livestock species. J Anim Sci 2009, 87:E48–E63.

29. Bottje W, Iqbal M, Tang ZX, Cawthon D, Okimoto R, Wing T, Cooper M: Association of mitochondrial function with feed efficiency within a single genetic line of male broilers. Poult Sci 2002, 81:546–555.

30. Iqbal M, Pumford NR, Tang ZX, Lassiter K, Ojano-Dirain C, Wing T, Cooper M, Bottje W: Compromised liver mitochondrial function and complex activity in low feed efficient broilers are associated with higher oxidative stress and differential protein expression. Poult Sci 2005, 84:933–941.

31. Ojano-Dirain CP, Iqbal M, Cawthon D, Swonger S, Wing T, Cooper M, Bottje W: Determination of mitochondrial function and site-specific defects in electron transport in duodenal mitochondria in broilers with low and high feed efficiency. Poult Sci 2004, 83:1394–1403.

32. Ojano-Dirain C, Iqbal M, Wing T, Cooper M, Bottje W: Glutathione and respiratory chain complex activity in duodenal mitochondria of broilers with low and high feed efficiency. Poult Sci 2005, 84:782–788.

33. Iqbal M, Pumford NR, Tang ZX, Lassiter K, Wing T, Cooper M, Bottje W: Low feed efficient broilers within a single genetic line exhibit higher oxidative stress and protein expression in breast muscle with lower mitochondrial complex activity. Poult Sci 2004, 83:474–484.

34. Lassiter K, Ojano-Dirain C, Iqbal M, Pumford NR, Tinsley N, Lay J, Liyanage R, Wing T, Cooper M, Bottje W: Differential expression of mitochondrial and extramitochondrial proteins in lymphocytes of male broilers with low and high feed efficiency. Poult Sci 2006, 85:2251–2259.

35. Bottje W, Pumford NR, Ojano-Dirain C, Iqbal M, Lassiter K: Feed efficiency and mitochondrial function. Poult Sci 2006, 85:8–14.

66 36. Ojano-Dirain C, Toyomizu M, Wing T, Cooper M, Bottje WG: Gene expression in breast muscle and duodenum from low and high feed efficient broilers. Poult Sci 2007, 86:372–381.

37. Mailloux RJ, Harper M-E: Uncoupling proteins and the control of mitochondrial reactive oxygen species production. Free Radic Biol Med 2011, 51:1106–1115.

38. Harper M-E, Mailloux RJ: Mitochondrial proticity and ROS signaling: lessons from the uncoupling proteins. Trends in Endocrinology & Metabolism 2012:451–458.

39. Kong B-W, Song JJ, Lee JY, Hargis BM, Wing T, Lassiter K, Bottje W: Gene expression in breast muscle associated with feed efficiency in a single male broiler line using a chicken 44K oligo microarray. I. Top differentially expressed genes. Poult Sci 2011, 90:2535–47.

40. Bottje WG, Kong B-W, Song JJ, Lee JY, Hargis BM, Lassiter K, Wing T, Hardiman J: Gene expression in breast muscle associated with feed efficiency in a single male broiler line using a chicken 44K microarray. II. Differentially expressed focus genes. Poultry Science 2012:2576–2587.

41. Wang Z, Gerstein M, Snyder M: RNA-Seq: a revolutionary tool for transcriptomics. Nat Rev Genet 2009, 10:57–63.

42. Frith MC, Pheasant M, Mattick JS: The amazing complexity of the human transcriptome. European journal of human genetics : EJHG 2005:894–897.

43. Nagalakshmi U, Waern K, Snyder M: RNA-Seq: a method for comprehensive transcriptome analysis. Curr Protoc Mol Biol 2010, Chapter 4:Unit 4.11.1–13.

44. Schiemer J: Illumina TruSeq DNA Adapters De-Mystified. :1–5.

45. Wilhelm BT, Landry J-R: RNA-Seq-quantitative measurement of expression through massively parallel RNA-sequencing. Methods 2009, 48:249–257.

46. Trapnell C, Pachter L, Salzberg SL: TopHat: discovering splice junctions with RNA-Seq. Bioinformatics 2009, 25:1105–1111.

47. Trapnell C, Roberts A, Goff L, Pertea G, Kim D, Kelley DR, Pimentel H, Salzberg SL, Rinn JL, Pachter L: Differential gene and transcript expression analysis of RNA-seq experiments with TopHat and Cufflinks. Nat Protoc 2012, 7:562–78.

67 48. Trapnell C, Hendrickson DG, Sauvageau M, Goff L, Rinn JL, Pachter L: Differential analysis of gene regulation at transcript resolution with RNA-seq. Nat Biotechnol 2013, 31:46–53.

49. Liu Z, Barrett EJ: Human protein metabolism: its measurement and regulation. Am J Physiol Endocrinol Metab 2002, 283:E1105–E1112.

50. Westerblad H, Bruton JD, Katz A: Skeletal muscle: energy metabolism, fiber types, fatigue and adaptability. Exp Cell Res 2010, 316:3093–3099.

51. Pette D, Staron RS: Myosin isoforms, muscle fiber types, and transitions. Microsc Res Tech 2000, 50:500–509.

52. Deeb N, Lamont SJ: Genetic architecture of growth and body composition in unique chicken populations. J Hered 1998, 93:107–18.

53. Emmerson DA: Commercial approaches to genetic selection for growth and feed conversion in domestic poultry. Poult Sci 1997, 76:1121–1125.

54. Rauw W., Kanis E, Noordhuizen-Stassen E., Grommers F.: Undesirable side effects of selection for high production efficiency in farm animals: a review. Livest Prod Sci 1998, 56:15–33.

55. Julian RJ: Rapid growth problems: ascites and skeletal deformities in broilers. Poult Sci 1998, 77:1773–1780.

56. Wilson BW, Nieberg PS, Buhr RJ, Kelly BJ, Shultz FT: Turkey muscle growth and focal myopathy. Poult Sci 1990, 69:1553–1562.

57. Sosnicki AA, Cassens RG, Vimini RJ, Greaser ML: Histopathological and ultrastructural alterations of turkey skeletal muscle. Poult Sci 1991, 70:349– 357.

58. Dransfield E, Sosnicki AA: Relationship between muscle growth and poultry meat quality. Poult Sci 1999, 78:743–746.

59. Hayes BJ, Lewin H a, Goddard ME: The future of livestock breeding: genomic selection for efficiency, reduced emissions intensity, and adaptation. Trends Genet 2013, 29:206–14.

60. Dekkers JCM, Hospital F: The use of molecular genetics in the improvement of agricultural populations. Nat Rev Genet 2002, 3:22–32.

61. Goddard ME, Hayes BJ: Mapping genes for complex traits in domestic animals and their use in breeding programmes. Nat Rev Genet 2009, 10:381–91.

68 62. Petracci M, Cavani C: Muscle growth and poultry meat quality issues. Nutrients 2012, 4:1–12.

63. Mills LJ, Mitchell M a, Mahon M: Incidence of skeletal muscle damage in selected and unselected strains of turkey. Br Poult Sci 1998, 39 Suppl(January 2014):S27–8.

64. Dunnington EA, Siegel PB: Enzyme activity and organ development in newly hatched chicks selected for high or low eight-week body weight. Poult Sci 1995, 74:761–770.

65. Shendure J: The beginning of the end for microarrays?. Nature methods 2008:585– 587.

66. Marioni JC, Mason CE, Mane SM, Stephens M, Gilad Y: RNA-seq: An assessment of technical reproducibility and comparison with gene expression arrays. Genome Res 2008, 18:1509–1517.

67. Wang C, Gong B, Bushel PR, Thierry-Mieg J, Thierry-Mieg D, Xu J, Fang H, Hong H, Shen J, Su Z, Meehan J, Li X, Yang L, Li H, Łabaj PP, Kreil DP, Megherbi D, Gaj S, Caiment F, van Delft J, Kleinjans J, Scherer A, Devanarayan V, Wang J, Yang Y, Qian H-R, Lancashire LJ, Bessarabova M, Nikolsky Y, Furlanello C, et al.: The concordance between RNA-seq and microarray data depends on chemical treatment and transcript abundance. Nat Biotechnol 2014, advance on.

68. Babraham Bioinformatics [http://www.bioinformatics.babraham.ac.uk/projects/fastqc/]

69. TopHat [http://tophat.cbcb.umd.edu/index.shtml]

70. SAMtools [http://samtools.sourceforge.net]

71. Cufflinks [http://cufflinks.cbcb.umd.edu/]

72. Geiss GK, Bumgarner RE, Birditt B, Dahl T, Dowidar N, Dunaway DL, Fell HP, Ferree S, George RD, Grogan T, James JJ, Maysuria M, Mitton JD, Oliveri P, Osborn JL, Peng T, Ratcliffe AL, Webster PJ, Davidson EH, Hood L, Dimitrov K: Direct multiplexed measurement of gene expression with color- coded probe pairs. Nat Biotechnol 2008, 26:317–25.

73. Inc. SI: JMP® 11 Multivariate Methods. Cary, NC SAS Inst Inc 2013.

69 74. Chargé SBP, Rudnicki MA: Cellular and molecular regulation of muscle regeneration. Physiol Rev 2004, 84:209–238.

75. Gal-Levi R, Leshem Y, Aoki S, Nakamura T, Halevy O: Hepatocyte growth factor plays a dual role in regulating skeletal muscle satellite cell proliferation and differentiation. Biochim Biophys Acta 1998, 1402:39–51.

76. Kaliman P, Canicio J, Testar X, Palacín M, Zorzano A: Insulin-like growth factor- II, phosphatidylinositol 3-kinase, nuclear factor-kappaB and inducible nitric- oxide synthase define a common myogenic signaling pathway. J Biol Chem 1999, 274:17437–17444.

77. Zanou N, Gailly P: Skeletal muscle hypertrophy and regeneration: interplay between the myogenic regulatory factors (MRFs) and insulin-like growth factors (IGFs) pathways. Cell Mol Life Sci 2013, 70:4117–30.

78. Kong Y, Flick MJ, Kudla AJ, Konieczny SF: Muscle LIM protein promotes myogenesis by enhancing the activity of MyoD. Mol Cell Biol 1997, 17:4750– 4760.

79. Demonbreun AR, Lapidos KA, Heretis K, Levin S, Dale R, Pytel P, Svensson EC, McNally EM: Myoferlin regulation by NFAT in muscle injury, regeneration and repair. J Cell Sci 2010, 123:2413–2422.

80. Velleman SG, Song Y, Shin J, McFarland DC: Modulation of turkey myogenic satellite cell differentiation through the shedding of glypican-1. Comp Biochem Physiol - A Mol Integr Physiol 2013, 164:36–43.

81. Lu H, Shah P, Ennis D, Shinder G, Sap J, Le-Tien H, Fantus IG: The differentiation of skeletal muscle cells involves a protein-tyrosine phosphatase- alpha-mediated C-Src signaling pathway. J Biol Chem 2002, 277:46687– 46695.

82. Araya R, Eckardt D, Maxeiner S, Krüger O, Theis M, Willecke K, Sáez JC: Expression of connexins during differentiation and regeneration of skeletal muscle: functional relevance of connexin43. J Cell Sci 2005, 118:27–37.

83. Wang L, Wang Y: Molecular characterization, expression patterns and subcellular localization of Myotrophin (MTPN) gene in porcine skeletal muscle. Molecular Biology Reports 2012:2733–2738.

70 84. Cao DJ, Jiang N, Blagg A, Johnstone JL, Gondalia R, Oh M, Luo X, Yang K-C, Shelton JM, Rothermel B a, Gillette TG, Dorn GW, Hill J a: Mechanical unloading activates FoxO3 to trigger Bnip3-dependent cardiomyocyte atrophy. J Am Heart Assoc 2013, 2:e000016.

85. Kipreos ET, Pagano M: Protein family review The F-box protein family. 2000:1– 7.

86. Sandri M, Sandri C, Gilbert A, Skurk C, Calabria E, Picard A, Walsh K, Schiaffino S, Lecker SH, Goldberg AL: Foxo transcription factors induce the atrophy-related ubiquitin ligase atrogin-1 and cause skeletal muscle atrophy. Cell 2004, 117:399–412.

87. Skurk C, Izumiya Y, Maatz H, Razeghi P, Shiojima I, Sandri M, Sato K, Zeng L, Schiekofer S, Pimentel D, Lecker S, Taegtmeyer H, Goldberg AL, Walsh K: The FOXO3a transcription factor regulates cardiac myocyte size downstream of AKT signaling. J Biol Chem 2005, 280:20814–20823.

88. Gomes MD, Lecker SH, Jagoe RT, Navon A, Goldberg AL: Atrogin-1, a muscle- specific F-box protein highly expressed during muscle atrophy. Proc Natl Acad Sci U S A 2001, 98:14440–14445.

89. Ye J, Zhang Y, Xu J, Zhang Q, Zhu D: FBXO40, a gene encoding a novel muscle- specific F-box protein, is upregulated in denervation-related muscle atrophy. Gene 2007, 404:53–60.

90. Serrano AL, Muñoz-Cánoves P: Regulation and dysregulation of fibrosis in skeletal muscle. Exp Cell Res 2010, 316:3050–8.

91. Chen X, Li Y: Role of matrix metalloproteinases in skeletal muscle. cell Adhes Migr 2009, 3:337–341.

92. Carmeli E, Moas M, Reznick AZ, Coleman R: Matrix metalloproteinases and skeletal muscle: a brief review. Muscle Nerve 2004, 29:191–197.

93. Cornelison DDW: Context matters: in vivo and in vitro influences on muscle satellite cell activity. J Cell Biochem 2008, 105:663–669.

94. Brinckerhoff CE, Matrisian LM: Matrix metalloproteinases: a tail of a frog that became a prince. Nat Rev Mol Cell Biol 2002, 3:207–214.

95. Wang W, Pan H, Murray K, Jefferson BS, Li Y: Matrix metalloproteinase-1 promotes muscle cell migration and differentiation. Am J Pathol 2009, 174:541–549.

71 96. Wu N, Jansen ED, Davidson JM: Comparison of mouse matrix metalloproteinase 13 expression in free-electron laser and scalpel incisions during . J Invest Dermatol 2003, 121:926–932.

97. Kherif S, Lafuma C, Dehaupas M, Lachkar S, Fournier JG, Verdière-Sahuqué M, Fardeau M, Alameddine HS: Expression of matrix metalloproteinases 2 and 9 in regenerating skeletal muscle: a study in experimentally injured and mdx muscles. Dev Biol 1999, 205:158–170.

98. Rullman E, Rundqvist H, Wågsäter D, Fischer H, Eriksson P, Sundberg CJ, Jansson E, Gustafsson T: A single bout of exercise activates matrix metalloproteinase in human skeletal muscle. J Appl Physiol 2007, 102:2346– 2351.

99. Baum O, Ganster M, Baumgartner I, Nieselt K, Djonov V: Basement membrane remodeling in skeletal muscles of patients with limb ischemia involves regulation of matrix metalloproteinases and tissue inhibitor of matrix metalloproteinases. J Vasc Res 2007, 44:202–213.

100. Li H, Mittal A, Makonchuk DY, Bhatnagar S, Kumar A: Matrix metalloproteinase-9 inhibition ameliorates pathogenesis and improves skeletal muscle regeneration in muscular dystrophy. Hum Mol Genet 2009, 18:2584– 2598.

101. Raffaello A, Milan G, Masiero E, Carnio S, Lee D, Lanfranchi G, Goldberg AL, Sandri M: JunB transcription factor maintains skeletal muscle mass and promotes hypertrophy. J Cell Biol 2010, 191:101–113.

102. Braun T, Gautel M: Transcriptional mechanisms regulating skeletal muscle differentiation, growth and homeostasis. Nat Rev Mol Cell Biol 2011, 12:349– 361.

103. Potthoff MJ, Arnold MA, McAnally J, Richardson JA, Bassel-Duby R, Olson EN: Regulation of skeletal muscle sarcomere integrity and postnatal muscle function by Mef2c. Mol Cell Biol 2007, 27:8143–8151.

104. Hinits Y, Hughes SM: Mef2s are required for thick filament formation in nascent muscle fibres. Development 2007, 134:2511–2519.

105. Molkentin JD, Black BL, Martin JF, Olson EN: Cooperative activation of muscle gene expression by MEF2 and myogenic bHLH proteins. Cell 1995, 83:1125– 1136.

72 106. Frey N, Barrientos T, Shelton JM, Frank D, Rütten H, Gehring D, Kuhn C, Lutz M, Rothermel B, Bassel-Duby R, Richardson JA, Katus HA, Hill JA, Olson EN: Mice lacking calsarcin-1 are sensitized to calcineurin signaling and show accelerated cardiomyopathy in response to pathological biomechanical stress. Nat Med 2004, 10:1336–1343.

107. Wan L, Ma J, Wang N, Wang D, Xu G: Molecular cloning and characterization of different expression of MYOZ2 and MYOZ3 in Tianfu goat. PLoS One 2013, 8:e82550.

108. Jørgensen JOL, Jessen N, Pedersen SB, Vestergaard E, Gormsen L, Lund SA, Billestrup N: GH Receptor Signaling in Skeletal Muscle and Adipose Tissue in Human Subjects Following Exposure to an Intravenous GH Bolus. Volume 291; 2006:E899–E905.

109. Goodyer CG, Figueiredo RM, Krackovitch S, De Souza Li L, Manalo JA, Zogopoulos G: Characterization of the growth hormone receptor in human dermal fibroblasts and liver during development. Am J Physiol Endocrinol Metab 2001, 281:E1213–E1220.

110. Carter-Su C, Schwartz J, Smit LS: Molecular mechanism of growth hormone action. Annu Rev Physiol 1996, 58:187–207.

111. Hull KL, Harvey S: Autoregulation of central and peripheral growth hormone receptor mRNA in domestic fowl. J Endocrinol 1998, 156:323–329.

112. Mao JNC, Cogburn LA, Burnside J: Growth hormone down-regulates growth hormone receptor mRNA in chickens but developmental increases in growth hormone receptor mRNA occur independently of growth hormone action. Mol Cell Endocrinol 1997, 129:135–143.

113. Yamada PM, Lee K-W: Perspectives in mammalian IGFBP-3 biology: local vs. systemic action. Am J Physiol Cell Physiol 2009, 296:C954–C976.

114. Schiaffino S, Mammucari C: Regulation of skeletal muscle growth by the IGF1- Akt/PKB pathway: insights from genetic models. Skelet Muscle 2011, 1:4.

115. Duclos MJ: Insulin-like growth factor-I (IGF-1) mRNA levels and chicken muscle growth. J Physiol Pharmacol 2005, 56 Suppl 3:25–35.

73 116. Fuentes EN, Björnsson BT, Valdés JA, Einarsdottir IE, Lorca B, Alvarez M, Molina A: IGF-I/PI3K/Akt and IGF-I/MAPK/ERK pathways in vivo in skeletal muscle are regulated by nutrition and contribute to somatic growth in the fine flounder. Am J Physiol Regul Integr Comp Physiol 2011, 300:R1532– R1542.

117. Ogawa W, Matozaki T, Kasuga M: Role of binding proteins to IRS-1 in insulin signalling. Mol Cell Biochem 1998, 182:13–22.

118. Saltiel A, Kahn C: Insulin signalling and the regulation of glucose and lipid metabolism. Nature 2001, 414:799–806.

119. Pedersen BK, Akerström TCA, Nielsen AR, Fischer CP: Role of myokines in exercise and metabolism. J Appl Physiol 2007, 103:1093–1098.

120. Pedersen BK: Muscles and their myokines. J Exp Biol 2011, 214:337–346.

121. Nara N, Nakayama Y, Okamoto S, Tamura H, Kiyono M, Muraoka M, Tanaka K, Taya C, Shitara H, Ishii R, Yonekawa H, Minokoshi Y, Hara T: Disruption of CXC motif chemokine ligand-14 in mice ameliorates obesity-induced insulin resistance. J Biol Chem 2007, 282:30794–30803.

122. Webster EL, Torpy DJ, Elenkov IJ, Chrousos GP: Corticotropin-releasing hormone and inflammation. Ann N Y Acad Sci 1998, 840:21–32.

123. Solinas G, Summermatter S, Mainieri D, Gubler M, Montani JP, Seydoux J, Smith SR, Dulloo AG: Corticotropin-releasing hormone directly stimulates thermogenesis in skeletal muscle possibly through substrate cycling between de novo lipogenesis and lipid oxidation. Endocrinology 2006, 147:31–38.

124. Lu H, Huang D, Ransohoff RM, Zhou L: Acute skeletal muscle injury: CCL2 expression by both monocytes and injured muscle is required for repair. The FASEB Journal 2011:3344–3355.

125. Lu H, Huang D, Saederup N, Charo IF, Ransohoff RM, Zhou L: Macrophages recruited via CCR2 produce insulin-like growth factor-1 to repair acute skeletal muscle injury. FASEB J 2011, 25:358–369.

126. Tidball JG: Inflammatory processes in muscle injury and repair. Am J Physiol Regul Integr Comp Physiol 2005, 288:R345–R353.

127. Pillon NJ, Bilan PJ, Fink LN, Klip A: Cross-talk between skeletal muscle and immune cells: muscle-derived mediators and metabolic implications. Am J Physiol Endocrinol Metab 2013, 304:E453–65.

74 128. Philippou A, Maridaki M, Theos A, Koutsilieris M: Cytokines in Muscle Damage. Adv Clin Chem 2012, 58:49–87.

129. Glass DJ: Skeletal muscle hypertrophy and atrophy signaling pathways. Int J Biochem Cell Biol 2005, 37:1974–1984.

130. Andreucci JJ, Grant D, Cox DM, Tomc LK, Prywes R, Goldhamer DJ, Rodrigues N, Bédard P-A, McDermott JC: Composition and function of AP-1 transcription complexes during muscle cell differentiation. J Biol Chem 2002, 277:16426–16432.

131. Xia R, Webb JA, Gnall LLM, Cutler K, Abramson JJ: Skeletal muscle sarcoplasmic reticulum contains a NADH-dependent oxidase that generates superoxide. Am J Physiol Cell Physiol 2003, 285:C215–C221.

132. Mofarrahi M, Brandes RP, Gorlach A, Hanze J, Terada LS, Quinn MT, Mayaki D, Petrof B, Hussain SNA: Regulation of proliferation of skeletal muscle precursor cells by NADPH oxidase. Antioxid Redox Signal 2008, 10:559–574.

133. Brand MD, Affourtit C, Esteves TC, Green K, Lambert AJ, Miwa S, Pakay JL, Parker N: Mitochondrial superoxide: Production, biological effects, and activation of uncoupling proteins. Free Radical Biology and Medicine 2004:755–767.

134. Kobayashi A, Kang M-I, Watai Y, Tong KI, Shibata T, Uchida K, Yamamoto M: Oxidative and electrophilic stresses activate Nrf2 through inhibition of ubiquitination activity of Keap1. Mol Cell Biol 2006, 26:221–229.

135. D’Autréaux B, Toledano MB: ROS as signalling molecules: mechanisms that generate specificity in ROS homeostasis. Nat Rev Mol Cell Biol 2007, 8:813– 824.

136. He X, Kan H, Cai L, Ma Q: Nrf2 is critical in defense against high glucose- induced oxidative damage in cardiomyocytes. J Mol Cell Cardiol 2009, 46:47– 58.

137. Laurila JP, Castellone MD, Curcio A, Laatikainen LE, Haaparanta-Solin M, Gronroos TJ, Marjamaki P, Martikainen S, Santoro M, Laukkanen MO: Extracellular superoxide dismutase is a growth regulatory mediator of tissue injury recovery. Mol Ther 2009, 17:448–454.

138. Kim S-J, Lee S-M: NLRP3 inflammasome activation in D-galactosamine and lipopolysaccharide-induced acute liver failure: role of heme oxygenase-1. Free Radic Biol Med 2013, 65:997–1004.

75 139. Wang X, Ling S, Zhao D, Sun Q, Li Q, Wu F, Nie J, Qu L, Wang B, Shen X, Bai Y, Li Y, Li Y: Redox regulation of actin by thioredoxin-1 is mediated by the interaction of the proteins via cysteine 62. Antioxid Redox Signal 2010, 13:565–73.

140. Salinas AE, Wong MG: Glutathione S---a review. Curr Med Chem 1999, 6:279–309.

141. Thannickal VJ, Fanburg BL: Reactive oxygen species in cell signaling. Am J Physiol Lung Cell Mol Physiol 2000, 279:L1005–28.

142. Barbieri E, Sestili P: Reactive Oxygen Species in Skeletal Muscle Signaling. Journal of Signal Transduction 2012:1–17.

143. Victor VM, Rocha M, De La Fuente M: Immune cells: Free radicals and antioxidants in sepsis. International Immunopharmacology 2004:327–347.

144. Bashan N, Kovsan J, Kachko I, Ovadia H, Rudich A: Positive and Negative Regulation of Insulin Signaling by Reactive Oxygen and Nitrogen Species. 2009:27–71.

145. Kumasaka S, Shoji H, Okabe E: Novel mechanisms involved in superoxide anion radical-triggered Ca2+ release from cardiac sarcoplasmic reticulum linked to cyclic ADP- stimulation. Antioxid Redox Signal 1999, 1:55–69.

146. Okabe E, Kato Y, Sasaki H, Saito G, Hess ML, Ito H: Calmodulin participation in oxygen radical-induced cardiac sarcoplasmic reticulum calcium uptake reduction. Arch Biochem Biophys 1987, 255:464–468.

147. Roveri A, Coassin M, Maiorino M, Zamburlini A, van Amsterdam FT, Ratti E, Ursini F: Effect of hydrogen peroxide on calcium homeostasis in smooth muscle cells. Arch Biochem Biophys 1992, 297:265–270.

148. Periasamy M, Kalyanasundaram A: SERCA pump isoforms: their role in calcium transport and disease. Muscle Nerve 2007, 35:430–42.

149. Ke Q, Costa M: Hypoxia-inducible factor-1 (HIF-1). Mol Pharmacol 2006, 70:1469–1480.

150. Stiehl DP, Ã WJ, Wenger RH, Hellwig-bu T, Allee R, Lu D-: Normoxic induction of the hypoxia-inducible factor 1 K by insulin and interleukin-1 L involves the phosphatidylinositol 3-kinase pathway. 2002, 512:157–162.

76 151. Feldser D, Agani F, Iyer N V: Reciprocal Positive Regulation of Hypoxia- inducible Factor 1 α and Insulin-like Growth Factor 2 Growth Factor 2 1. 1999:3915–3918.

152. Eltzschig HK, Carmeliet P: Hypoxia and inflammation. N Engl J Med 2011, 364:656–65.

153. Nizet V, Johnson RS: Interdependence of hypoxic and innate immune responses. Nat Rev Immunol 2009, 9:609–617.

154. Cramer T, Yamanishi Y, Clausen BE, Förster I, Pawlinski R, Mackman N, Haase VH, Jaenisch R, Corr M, Nizet V, Firestein GS, Gerber HP, Ferrara N, Johnson RS: HIF-1α is essential for myeloid cell-mediated inflammation. Cell 2003, 112:645–657.

155. Bartels K, Grenz A, Eltzschig HK: Hypoxia and inflammation are two sides of the same coin. Proc Natl Acad Sci U S A 2013, 110:18351–2.

156. Frede S, Berchner-Pfannschmidt U, Fandrey J: Regulation of Hypoxia-Inducible Factors During Inflammation. Methods in Enzymology 2007.

157. Scheerer N, Dehne N, Stockmann C, Swoboda S, Baba HA, Neugebauer A, Johnson RS, Fandrey J: Myeloid hypoxia-inducible factor-1α is essential for skeletal muscle regeneration in mice. J Immunol 2013, 191:407–14.

158. Chandel NS, Maltepe E, Goldwasser E, Mathieu CE, Simon MC, Schumacker PT: Mitochondrial reactive oxygen species trigger hypoxia-induced transcription. Proc Natl Acad Sci U S A 1998, 95:11715–11720.

159. Semenza GL: Regulation of mammalian O2 homeostasis by hypoxia-inducible factor 1. Annu Rev Cell Dev Biol 1999, 15:551–578.

160. Dröge W: Free radicals in the physiological control of cell function. Physiol Rev 2002, 82:47–95.

161. Chandel NS, McClintock DS, Feliciano CE, Wood TM, Melendez J a, Rodriguez a M, Schumacker PT: Reactive oxygen species generated at mitochondrial complex III stabilize hypoxia-inducible factor-1alpha during hypoxia: a mechanism of O2 sensing. J Biol Chem 2000, 275:25130–8.

162. Petracci M, Mudalal S, Bonfiglio a, Cavani C: Occurrence of white striping under commercial conditions and its impact on breast meat quality in broiler chickens. Poult Sci 2013, 92:1670–5.

77 163. Kuttappan VA, Shivaprasad H, Shaw DP, Valentine BA, Hargis BM, Clark FD, Mckee SR, Owens CM: Pathological changes associated with white striping in broiler breast muscles. Poult Sci 2013:331–338.

164. Sihvo H-K, Immonen K, Puolanne E: Myodegeneration with fibrosis and regeneration in the pectoralis major muscle of broilers. Vet Pathol 2014, 51:619–23.

78 Chapter 3

CHARACTERIZATION OF METABOLIC DIFFERENCES IN THE BREAST MUSCLE OF HIGH AND LOW FEED EFFICIENCY CHICKENS USING RNA-SEQ

3.1 Introduction

As the first domestic animal chosen for complete genome sequencing and assembly [1], the chicken has now become one of the National Institutes of Health- sponsored model organisms for biomedical research (http://www.nih.gov/science/models/gallus/). Divergently selected fat and lean lines of chickens have been widely used for studies of human obesity and have given insights into the biological basis for the development of obesity [2,3]. In addition, the unique physiological features of the chicken make it as a valuable model for understanding the mechanisms of obesity-related metabolic disorders (i.e., insulin resistance and diabetes) [4]. For example, the chicken has proven to be a valid and premier species for the study of atherosclerosis because chickens are naturally hypercholesterolemic and tend to develop spontaneous and diet-induced atherosclerosis in the aorta and coronary arteries in a way similar to humans [5]. Atherosclerosis in chicken is induced by high cholesterol diets, vascular injury or infections, and display similar fatty lesions as humans. Numerous drug intervention studies for atherosclerosis have conducted using chickens throughout the 20th century, contributing greatly to the development of atherosclerosis therapies greatly [6]. Given the contribution of chicken models to

79 biomedical research, transcriptome studies in chickens should provide more comprehensive information on the molecular mechanisms of human diseases. Selected for different purposes, there exists an extensive diversity among domestic chickens, which makes them excellent models for studying the genetic basis of phenotypic traits [7]. Significant phenotypic differences in body composition were observed between commercial broiler chickens with extremely high and low feed efficiency (FE) in our previous study. For example, chickens with high FE also exhibit greater breast muscle yield but less abdominal fat yield than those with low FE (unpublished observations). Since body composition is a complex trait controlled by multiple genetic and environmental factors in humans and other vertebrates [8][9], a global gene expression analysis for these broiler chickens should provide unique insights into the underlying mechanisms that control the divergence of body composition and the development of obesity in humans. Despite no clinical symptoms of infection or muscle damage were observed in the birds used in our previous study, transcriptome analysis of breast muscle of these chickens showed that high-FE chickens (HFE) have significant increases in inflammatory response compared with low-FE birds (LFE). It is well recognized that immune responses and metabolic regulation are highly intertwined and closely associated with chronic inflammation [10]. Many metabolites act as signaling molecules regulating inflammatory responses, while various inflammatory mediators can significantly alter metabolic homeostasis [11][12]. The differences in immune responses between these birds is likely related to variations in metabolic regulation. Characterization of the expression of metabolism-associated genes in these chickens will help elucidate the etiology of the elevated inflammatory response in the breast

80 muscle of HFE broiler chickens and provide more clues to understand the interconnected network of metabolic and immune systems. The aim of this study was to characterize the divergence in metabolic regulation between HFE and LFE chickens through breast muscle RNA-seq analysis and glycogen assay. Utilizing the transcriptome profile that were generated previously by our lab from breast muscle samples of extremely HFE and LFE chickens, we focused on the differentially expressed (DE) genes associated with metabolism and how they contribute to the phenotypic differences between HFE and LFE birds. We also investigated the variation in glycogen metabolism in breast muscle of these HFE and LFE chickens by glycogen assay. The present study revealed a divergence in metabolic regulation between HFE and LFE chickens, providing more information for human studies.

3.2 Methods

3.2.1 Animals and RNA-seq Analysis

Detailed procedures used for animal and sample collection, RNA isolation, RNA-seq library preparation and sequencing, differential expression analysis and validation of RNA-seq data were reported previously. In brief, using the RNA-seq approach, a global gene expression study was performed on breast muscle samples from 23 commercial broiler chickens with extremely high (n=10) and low (n=13) FE. The differentially expressed (DE) genes between the HFE and LFE broiler phenotypes were identified by the Cuffdiff differential expression analysis. A large-scale verification of the RNA-seq data was conducted by quantifying expression levels of 204 target transcripts using the NanoString nCounter® gene expression assay. The

81 phenotypic characteristics of the HFE and LFE chickens were provided in details in our previous study.

3.2.2 Ingenuity Pathway Analysis QIAGEN’s Ingenuity Pathway Analysis Fall 2014 release (IPA; Ingenuity® Systems, http://www.ingenuity.com) was used for canonical pathway, network and function analysis in the current study. All 1,059 significant DE genes (FDR value < 0.05) were uploaded into IPA software program for analysis. The program identified 878 DE genes and conducted core analysis for these genes. Disease & Function analysis of IPA was applied to classify the DE genes based on their functional annotations. Moreover, IPA’s downstream effects analysis is able to identify whether a biological process is significantly activated or inhibited on the basis of gene expression profile. Canonical pathway analysis was utilized to determine the significance of the affected pathways. The additional functionality of IPA Fall 2014 release was used to predict if the significant canonical pathways are activated or inhibited based on the DE genes dataset. IPA upstream regulator analysis was used to predict potential upstream molecules including transcription factor and microRNA that cause changes in the gene expression. The prediction was base on Activation z-score that was calculated on the basis of published findings accessible through the Ingenuity

Knowledge Base. Regulators with z-score > 2 or <-2 were predicted to be significantly activated or inhibited.

3.2.3 Ultimate pH (pHu) Measurement and Analysis

All pHu measurement was performed on the breast muscle at 24h post mortem using a hand-held pH meter (Testo 205 pH measuring instrument, Lenzkirch,

82 Germany, http://www.testo.com). The pH meter was calibrated according to the manufacturer’ instruction once every 100 samples. The resulting pHu data was first adjusted for pH calibration and hatch effects and then analyzed using a two-sample t- test from JMP (http://www.jmp.com/).

3.2.4 Determination of Breast Muscle Glycogen Content Breast muscle tissues from 12 birds were randomly sampled from the 24 RNA- seq samples for glycogen assay. Frozen breast muscle tissues (80-100mg) were hydrolyzed in 300 µl of 30% KOH for 2 hr at 100° C. After fully digested, the samples were cooled on ice for 10 minutes and then 900 µl of 95% ethanol was added for precipitation. The solutions were incubated overnight at -20° C, followed by centrifugation at 3,000 x g for 10 minutes to precipitate the crude glycogen. The resulting glycogen precipitates were dissolved in 100 µl of distilled water, which were acidified to pH=3 with HCL (5N) and subsequently re-centrifuged at 3,000 x g for 10 minutes after adding 100 µl of 95% ethanol. The glycogen pellets were washed once with 100 µl of 95% ethanol. Following centrifugation, the final precipitates were dried and dissolved in 300 µl of distilled water for assay. Glycogen contents were quantified using a commercial glycogen assay kit (Abcam, ab65620). Colorimetric assays were carried out in 96-well plates and read on a microplate reader (SpectraMax

M2, Molecular Devices).

3.2.5 Quantitative Reverse Transcription-PCR (qRT-PCR)

The gene encoding muscle glycogen synthase 1 (GYS1) has not been annotated in Ensembl gene annotation file for Galgal 4 assembly yet; therefore, the expression of GYS1 cannot be detected by RNA-seq analysis. Given that GYS1 plays an important

83 role in glycogen metabolism, we used quantitative qRT-PCR analysis to quantify its expression levels in the RNA samples isolated earlier from 24 breast muscle tissues of the HFE and LFE chickens. To obtain purified RNA, all the 24 RNA samples were treated with DNase I (RNase-free) individually according to the manufacturer’s instruction (Ambion life technologies). Subsequently, each RNA sample was quantified using Nanodrop Spectrophotometer (Nanodrop Technology) and diluted to 10 ng/µl for qRT-PCR assay. We identified exon-intron structure of a partial coding sequence (cds) of chicken GYS1 (AB090806.1) by aligning it with the mouse GYS1. To be able to detect possible contamination of RNA samples by genomic DNA, the primers were designed (using Primer3; http://biotools.umassmed.edu/bioapps/primer3_www.cgi) on two exons spanning one intron. Primer sequences were: CTGTGTGCACCCACAACAT, forward and CGCACGAACTCCTCGTAGTC, reverse. To validate the primer pairs, conventional PCR and electrophoresis were performed for the target sequence using genomic DNA and complementary DNA (cDNA) of two randomly selected birds. Hypoxanthine phosphoribosyltransferase 1 (HPRT1) was tested and applied as a housekeeping gene in this study. Utilizing Power SYBR® Green RNA-to-CT™ 1- Step Kit (Life technologies), the qRT-PCR measurement was performed in triplicate for each RNA sample on an Applied Biosystems 7900HT FAST Real-Time PCR system (Applied Biosystems) using ABI 7900 Fast System SDS Software. Following the manufacturer’s instructions, the 10-µl qRT-PCR reaction mixture contained 200 nM of each primer, 1x RT-PCR Mix, 1x RT Enzyme Mix and 10 ng RNA sample. The qRT-PCR conditions were 48 °C for 30 min for reverse transcription, 95 °C for 10 min, followed by 40 cycles of 95 °C for 15 sec and 60 °C for 1 min. Melt curve

84 analysis (95 °C for 15 sec, 60 °C for 15 seconds and 95 °C for 15 sec) was conducted after all the reactions. The qRT-PCR data were normalized using the housekeeping gene and analyzed by the ΔΔCT method [13]. The fold-change value was calculated as 2- ΔΔCT.

3.3 Results and Discussion

3.3.1 Carbohydrate Metabolism

A number of genes involved in carbohydrate metabolism are divergently expressed in this study, presented in Table 3.1. Genes encoding for glucose and fructose transporter (GLUT), GLUT3, GLUT5 and GLUT6, are all up-regulated in HFE birds, suggesting that glucose uptake is increased in the breast muscle of the HFE chickens. GLUT3 is a member of glucose carriers expressed at high levels in the chicken brain but low in skeletal muscles [14]. It is reported that GLUT3 acts as a beneficial mediator of glucose uptake during glucose deprivation and hypoxia [15][16];[17]. The increased expression of GLUT3 may be, at least in part, induced by a local hypoxic environment implicated in the breast muscle of the HFE chickens. Moreover, amino sugar synthesizing enzymes, including glutamine-fructose-6- phosphate transaminase 2 (GFPT2), phosphoglucomutase 3 (PGM3), UDP-N- acteylglucosamine pyrophosphorylase 1 (UAP1) and UDP-glucose 6-dehydrogenase (UGDH), are likely more active in the HFE chickens due to higher gene expression levels of GFPT2, UAP1 and UGDH, respectively. GFPT2, PGM3 and UAP1 participate in the biosynthesis of UDP-N-acetyl-D-glucosamine (UDP-GlcNAc), which is the precursor of N-acetyl-D-glucosamine for synthesizing glycosylated proteins and cell surface structures (55) (Figure 3.1). UGDH can yield UDP-

85 glucuronate that is involved in biosynthesis (25). Notably, Gene encoding hyaluronan synthases 2 (HAS2) is also up-regulated in the HFE chickens. HAS2 can utilize UDP-GlcNAc and UDP-glucuronate to produce hyaluronan for extracellular matrix (ECM) formation (29). The up-regulation of HAS2 as well as augmented UDP-glucuronate and UDP-GlcNAc synthesis implicate that the HFE chickens may have an increase in hyaluronan production in their breast muscle, which is consistence with the increased ECM remodeling we have discussed before (Nan et al., 2015). Furthermore, numerous evidence have shown that hyaluronan plays a role in inflammation, involved in leukocyte recruitment, inflammatory cell activation as well as cytokine release (97). Therefore, the implied increase in hyaluronan production may be also associated with the elevated inflammatory response in the breast muscle of the HFE birds.

86 Table 3.1 Metabolism-associated DE Genes

Biological Processes Genes Carbohydrate Glucose and fructose transport SLC2A3 é; SLC2A5 é; SLC2A6 é Metabolism Glycolysis/ ALDOC é; ENO2 é; LDHB é; FBP2 ê; FBP1 Gluconeogenesis é; TKT é; BPGM ê; PFKFB3 ê Glycogenesis/ PHKA1 ê; PGM3 é; PGM2 é; GAA ê; Glycogenolysis PPP1R3a ê; PPP1R3C ê; PRKAA2 ê; GYS1 ê Glycerol-3-phosphate Shuttle GPD2 ê; GPD1 ê Amino sugar and nucleotide UGDH é; UAP1 é; PGM3 é; GUSB é; sugar metabolism GFPT2 é; HAS2 é Fatty acid transport FABP3 é; FABP4é Lipid Steroid Biosynthesis DHCR24 é; CYP51A1 é; LSS é; LIPA é Metabolism MSMO1 é; SQLE é Phospholipid metabolism PHKA1ê; CHKA é; LPIN1 ê; AGPAT5 ê; Lipolysis AADACé; LIPGé Fatty acid oxidation ACAA2 é; EHHADH ê; ACOX2 ê Acetyl-CoA synthesis ACLY é Glycerol degradation GKé Reverse cholesterol transport LCAT é; PLTP é Eicosanoid biosynthesis PLA2G7 é; PLA2R1 é; ALOX5 é; ALOX5AP é; LTA4H é Amino Acid Amino Acid Transport SLC1A2 é; SLC1A6 é; SLC25A4 ê; Metabolism SLC25A12 ê; SLC38A1 é; SLC6A6 é; SLC6A9 é SLC7A1é; SLC7A2 ê; SLC7A3 é; SLC7A5ê Valine Degradation I BCKDHB ê; DLD ê; EHHADH ê; HIBADH ê Biosynthesis I ALDH18A1 é; PYCR2 é Biosynthesis PHGDH é; PSAT1 é Arginine Biosynthesis IV ASL é; ASS1 é Asparagine Biosynthesis I ASNS é Glutamine Biosynthesis I GLUL ê; GLS é Immune genes Toll-like receptors TLR4 é; TLR2-1 é; TLR2-2 é; TLR6 é involved in metabolic Cytokine receptors IL1R1 é; TL1R2 é regulation Signaling molecule NFKBIE é

87 édenotes genes up-regulated in the HFE chickens; ê represents genes down-regulated in the HFE birds. AADAC: Arylacetamide Deacetylase; ACAA2: Acetyl-Coenzyme A Acyltransferase 2; ACLY: ATP Citrate Synthase; ACOX2: Acyl-CoA Oxidase 2, Branched Chain; AGPAT5: 1- Acylglycerol-3-Phosphate O-Acyltransferase 5; ALDH18A1: Aldehyde Dehydrogenase 18 Family, Member A1; ALDOC: Aldolase C, Fructose-Bisphosphate; ALOX5: Arachidonate 5- Lipoxygenase; ALOX5AP: Arachidonate 5-Lipoxygenase-Activating Protein; ASL: Argininosuccinate Lyase; ASNS: Asparagine Synthetase (Glutamine-Hydrolyzing); ASS1: Argininosuccinate Synthase 1; BCKDHB: Branched Chain Keto Acid Dehydrogenase E1, Beta Polypeptide; BPGM: 2,3-Bisphosphoglycerate Mutase; CHKA: Choline Kinase Alpha; CYP51A1: Cytochrome P450, Family 51, Subfamily A, Polypeptide 1; DHCR24: 24- Dehydrocholesterol Reductase; DLD: Dihydrolipoamide Dehydrogenase; EHHADH: Enoyl- CoA, Hydratase/3-Hydroxyacyl CoA Dehydrogenase; ENO2: Enolase 2 (Gamma, Neuronal); FABP3: Fatty Acid Binding Protein 3, Muscle And Heart; FABP4: Fatty Acid Binding Protein 4, Adipocyte; FBP1: Fructose-1,6-Bisphosphatase 1; FBP2: Fructose-1,6-Bisphosphatase 2; GAA: Glucosidase, Alpha; Acid; GFPT2: Glutamine-Fructose-6-Phosphate Transaminase 2; GK: Glycerol Kinase; GLUL: Glutamate-Ammonia Ligase; GLS: glutaminase; GPD1: Glycerol-3-Phosphate Dehydrogenase 1; GPD2: Glycerol-3-Phosphate Dehydrogenase 2 (Mitochondrial); GUSB: Glucuronidase, Beta; GYS1: Glycogen Synthase 1 (Muscle); HAS2: 2; HIBADH: 3-Hydroxyisobutyrate Dehydrogenase; IL-1R1: interleukin 1 receptor, type I; IL-1R2: interleukin 1 receptor, type II; LCAT: Lecithin-Cholesterol Acyltransferase; LDHB: Lactate Dehydrogenase B; LIPA: Lipase A, Lysosomal Acid, Cholesterol Esterase; LIPG: Lipase, Endothelial; LPIN1: Lipin 1; LTA4H: Leukotriene A4 Hydrolase; LSS: Lanosterol Synthase; MSMO1: Methylsterol Monooxygenase 1; NFKBIE: nuclear factor of kappa light polypeptide gene enhancer in B-cells inhibitor, epsilon; PFKFB3: 6-Phosphofructo-2-Kinase/Fructose-2,6-Biphosphatase 3; PGM2: Phosphoglucomutase 2; PGM3: Phosphoglucomutase 3; PHKA1: Kinase, Alpha 1 (Muscle); PHGDH: Phosphoglycerate Dehydrogenase; PLA2G7: Phospholipase A2, Group VII; PLA2R1: Phospholipase A2 Receptor 1, 180kDa; PLTP: Phospholipid Transfer Protein; PRKAA2: Protein Kinase, AMP-Activated, Alpha 2 Catalytic Subunit; PPP1R3a: Protein Phosphatase 1, Regulatory Subunit 3A; PPP1R3C: Protein Phosphatase 1, Regulatory Subunit 3C; PSAT1: Phosphoserine Aminotransferase 1; PYCR2: Pyrroline-5-Carboxylate Reductase Family, Member 2; SLC2A3: Solute Carrier Family 2 (Facilitated Glucose Transporter), Member 3; SLC2A5: Solute Carrier Family 2 (Facilitated Glucose/Fructose Transporter), Member 5; SLC2A6: Solute Carrier Family 2 (Facilitated Glucose Transporter), Member 6; SLC1A2: solute carrier family 1 (glial high affinity glutamate transporter), member 2; SLC1A6: solute carrier family 1 (high affinity aspartate/glutamate transporter), member 6; SLC25A4: solute carrier family 25 (mitochondrial carrier; adenine nucleotide translocator), member 4; SLC25A12: solute carrier family 25 (aspartate/glutamate carrier), member 12; SLC38A1: solute carrier family 38, member 1; SLC6A6: solute carrier family 6 (neurotransmitter transporter, taurine), member 6; SLC6A9: solute carrier family 6 (neurotransmitter transporter, glycine), member 9; SLC7A1: solute carrier family 7 (cationic amino acid transporter, y+ system), member 1; SLC7A2: solute carrier family 7 (cationic amino acid transporter, y+ system), member 2; SLC7A3: solute carrier family 7 (cationic amino acid transporter, y+ system), member 3; SLC7A5: solute carrier family 7 (amino acid transporter light chain, L system), member 5; SQLE: Squalene Epoxidase; TLR4: Toll-like receptor 4; TLR2-1: toll-like receptor 2 type1; TLR2-2: toll-like receptor 2 type2; TLR6: Toll-like receptor 6; TKT:

88 transketolase; UAP1: UDP-N-Acteylglucosamine Pyrophosphorylase 1; UGDH: UDP- Glucose 6-Dehydrogenase

3.3.1.1 Glycerol-3-Phosphate Shuttle In order to maintain the balance of NADH and NAD+ in cytosol, glycerol-3- phosphate shuttle is commonly used in muscle to transfer cytosolic NADH produced by glycolysis into mitochondria for oxidative phosphorylation. Moreover, the glycerol-3-phosphate shuttle also functions as an essential link between lipid and carbohydrate metabolism [22]. The shuttle is composed of the combined actions of mitochondrial FAD-dependent glycerophosphate dehydrogenase (mGPDH) and cytoplasmic NAD-linked glycerophosphate dehydrogenase (cGPDH) [22]. cGPDH catalyzes the conversion of carbohydrate-derived dihydroxyacetone phosphate and NADH to NAD+ and glycerol-3-phosphate, which can be then utilized for phospholipids and triglyceride synthesis. In contrast, mGPDH oxidizes the resulting glycerol-3-phosphate to dihydroxyacetone phosphate for gluconeogenesis or oxidative phosphorylation [23] . In the present study, genes encoding mGPDH and cGPDH, GPD2 and GPD1, are both down-regulated in breast muscle of the HFE birds, suggesting that the activity of glycerol-3-phosphate shuttle may be inhibited in breast muscle of the HFE chickens compared with the LFE birds (Figure 3.1). Thus, the decreased function of glycerol-3-phosphate shuttle may implicate that there is an abnormal NADH/NAD+ ratio in the breast muscle of the HFE birds, which is likely associated with the intensified inflammatory response and oxidative stress in these birds. It should be pointed out that skeletal muscle is commonly believed to be unable to carry out gluconeogenesis due to lack of glucose-6-phosphatase complex.

89 Using the same enzymes of gluconeogenesis, carbon skeletons like pyruvate can be converted to glucose-6-phosphate for glycogen and polysaccharides synthesis in skeletal muscles. However, since there is no active glucose-6-phosphatase complex in muscles, the resulting glucose-6-phosphate cannot be further catalyzed to glucose for the maintenance of blood glucose homeostasis [24]. Glucose-6-phosphatase complex is located in the endoplasmic reticulum where its function is coordinated by glucose-6- phosphate catalytic subunits and a glucose-6-phosphate transporter [25]. Although genes encoding glucose-6-phosphatase complex, including glucose-6-phosphatase, catalytic subunit (G6PC), glucose-6-phosphatase, catalytic, 2 (G6PC2) and glucose-6- phosphatase, catalytic, 3 (G6PC3), have been found to be expressed in human skeletal muscles, the role of glucose-6-phosphatase complex in muscle cells is still unclear [25]. Study in mice muscles has implicated that G6PC3 and glucose-6-phosphate transporter could form an active glucose-6-phosphatase which might have unrecognized effects on muscle glucose production [26]. In the present study, G6PC2 and G6PC3 are expressed in breast muscle of both HFE and LFE chickens with mean FPKM values equal to 1.41 and 15.08, respectively. Despite that G6PC2 and G6PC3 are not differentially expressed between the HFE and LFE chicken, the expression of these two genes in chicken breast muscle may provide more information for further study of the activity and function of glucose-6-phosphatase complex in avian skeletal muscles.

90

Figure 3.1 UDP-N-acetyl-D-glucosamine biosynthesis and Glycerol-3-Phosphate Shuttle.

91 3.3.1.2 Transcriptional Control of Carbohydrate Metabolism IPA upstream regulator analysis sheds light into the difference in the modulation of carbohydrate metabolism between the HFE and LFE chickens. Neuron- derived orphan receptor 1 (Nor1), also known as NR4A3, is predicted to be significantly inhibited in the HFE chickens [P-value (NR4A3) = 2.24E-05] based on the expression levels of its target genes (figure 3.2). NR4A3 belongs to nuclear receptor 4A subgroup (NR4As) which is proposed as regulator of glucose metabolism in mammalian skeletal muscles [27]. In vivo and in vitro experiments showed that

NR4As induced the expression of a set of glucose metabolic genes including FBP2, PHKA1, BPGM, GPD1 and PPP1R3C in mice skeletal muscles, modulating glycolysis and gluconeogenesis [28]. Moreover, NR4A3 has been found to regulate the mRNA expression of genes involved in oxidative metabolism such as uncoupling protein 3 (UCP3) (74)(72). In our study, based on Ingenuity Knowledge Base, NR4A3 is postulated as a master mediator down-regulating FBP2, pyruvate dehydrogenase phosphatase 1 (PDP1), myosine, heavy chain 2, skeletal muscle (MYH2), lipin1 (LPIN1) and UCP3 as well as up-regulating Hypoxia-inducible factor 1-alpha (HIF1A) and transforming growth factor, beta 3 (TGFB3) in the HFE chickens. The variation in glucose and oxidative metabolism in HFE and LFE birds is probably associated with the divergent activity of NR4A3 between these broilers.

92

Figure 3.2 Upstream Regulator NR4A3 and Its Target Genes.

3.3.2 Glycogen Depletion in Breast Muscle of the HFE Chickens Glycogen serves as a major supplier of metabolic fuel in muscles supplies in times of fasting or exercise [31]. The amount of glycogen in chicken muscles is a key factor controlling ultimate meat pH (pHu), which influences the poultry meat quality [32]. Moreover, the glycogen content of chicken breast muscle has shown negatively correlated with bird growth rate and breast muscle yield whereas positively correlated with carcass fatness [33] [34][35]. In the present study, despite that no significant difference was observed in pHu between HFE and LFE chickens, the mean pHu value

93 of the HFE group is higher than that of LFE birds (pHu (HFE, n=10) = 5.847±0.039; pHu

(LFE, n=13) = 5.802±0.029). The LFE chickens had 2.7-fold greater breast muscle glycogen content than the HFE birds (P ≤ 0.05). The transcriptome analysis provides more information about the mechanisms underlying glycogen storage differences between the HFE and LFE broiler chickens. The regulation of glycogen metabolism in birds is very similar to that in mammals, which is mainly through reversible phosphorylation control of (GP) and glycogen synthase (GYS) [36]. Protein phosphatase 1 (PP1) is a serine/threonine phosphatase that can dephosphorylate GYS and GP, as a result, activating GYS to promote glycogen synthesis as well as inactivating GP to reduce glycogen breakdown [37]. The glycogenic effects of PP1 require the coordination of PP1c-targeting subunits that target the enzyme to glycogen particles. Proteins like PPP1R3A (Gm) and PPP1R3C (PTG) are members of PP1c-targeting subunits that play a role in the regulation of glycogen metabolism [38]. Evidence show that both PPP1R3A and PPP1R3C can enhance glycogen storage in human muscle cells [39][40]. In the present study, genes encoding PPP1R3A and PPP1R3C are down- regulated in the HFE birds, which may partly explain why there is less glycogen in breast muscle of these chickens. Furthermore, muscle glycogen content is modulated by kinase such as cAMP- dependent protein kinase (PKA). Activation of PKA leads to phosphorylation of GYS and GP, which reduces glycogenesis and stimulate glycogenolysis [41]. Through IPA canonical pathway analysis, PKA signaling pathway is enriched in our dataset with 31 DE genes mapped to it (P-value = 5.83E-04; ratio= 0.081). Based on the gene expression profiles of these mapped molecules, the Molecule Activity Predictor (MAP)

94 tool of IPA predicts that PKA is more active in the HFE chickens. Thus the activated PKA may partly explain why there is less glycogen stored in breast muscle of the HFE birds compared to those of the LFE chickens. The gene encoding muscle glycogen synthase 1 (GYS1) has not been annotated in Ensembl gene annotation file for Galgal 4 assembly yet, therefore the expression of GYS1 cannot be detected by RNA-seq analysis. We used quantitative qRT-PCR analysis to quantify the gene expression levels of GYS1 in the breast muscle samples from HFE and LFE birds. Results show that the LFE chickens have 2.43-fold higher expression level of GYS1 than the HFE chickens, indicating that the HFE chickens have less GYS and therefore a reduced glycogen synthesis capability in their breast muscle compared to the LFE broilers. In an apparent contradiction, the expression of gene encoding phosphorylase kinase, alpha 1 (PHKA1) is 1.66-fold higher in the LFE birds than it is in the HFE chickens. PHKA1 is a skeletal muscle isoform of phosphorylase kinase that can phosphorylate GP and induce glycogenolysis (59). The down-regulated expression of PHKA1 implicates that the activity of glycogen breakdown may be decreased in the HFE birds, which is likely because the glycogen content has already become low in their breast muscle due to decreased synthesis of glycogen. In this sense, the glycogen depletion in breast muscle of the HFE broilers may be largely caused by decreased glycogenesis. Collectively, our data suggests that the divergence of glycogen storage between the HFE and LFE chickens is partly due to the transcriptional and post-transcriptional control of the activity of glycogen synthase and glycogen phosphorylase.

95 3.3.3 Transcriptional Regulation of Lipid Metabolism A large number of differentially expressed genes, approximately 12% of the total DE genes, are associated with lipid metabolism. Based on the Ingenuity Downstream Effects analysis, the transport of lipid [P-value = 9.27E-10; Activation z- score = 2.259] is predicted to be significantly increased in the breast muscle of the HFE chickens compared with the LFE birds. Thirteen DE genes in this study are involved in the fatty acid release. Out of these genes, genes from fatty-acid-binding- proteins (FABPs) family [fatty acid binding protein 4 (FABP4), fatty acid binding protein 3 (FABP3)] are up-regulated in the HFE birds. FABPs are known as intracellular lipid chaperon responsible for lipid transport [43]. Primarily expressed in the skeletal muscle, the protein encoded by FABP3 participates in the uptake and transportation of fatty acids to the mitochondria for beta-oxidation [44][45]. Mice lacking FABP3 exhibited markedly reduced fatty acids uptake in their skeletal muscles [46]. Consequently, the enhanced expression of FABP3 may suggest a rise in the uptake of fatty acids in the breast muscle of the HFE broiler. FABP4 encodes the adipocyte protein 2 (aP2) that has critical impacts on several components of metabolic syndrome through integrating metabolic and inflammatory responses [47][48][49]. aP2 is primarily expressed in mature adipocytes and adipose tissue [43]. aP2- defiecient mice have shown protection against insulin resistance and hyperinsulinemia in the context of dietary or genetic obesity [50][51]. Interestingly, aP2 is also expressed in macrophages, mediating the inflammatory response and cholesterol ester accumulation ([43]. Macrophages lacking aP2 have been found to provide protection against atherosclerosis in the apolipoprotein E-deficient mice [48]. Since aP2 is a regulator coordinating inflammatory and metabolic responses, the up-regulation of

96 FABP4 is likely related to the amplified immune responses in the breast muscle of the HFE chickens and foreboding a disturbance of metabolic homeostasis in these birds.

3.3.3.1 Cholesterol Biosynthesis and Reverse Cholesterol Transport Cholesterol serves as a fundamental structural component of , which mediates membrane integrity and viability, intracellular transport as well as cell signaling [52] [53]. Due to the importance of cholesterol, de novo synthesis of cholesterol occurs in the cytoplasm and endoplasmic reticulum (ER) of most kinds of animal cells [24]. The synthesis starts with converting acetyl-CoA to 3-hydroxy-3- methylglutaryl-coenzyme A (HMG-CoA), which then is converted to mevalonate by HMG-CoA reductase. Synthesized mevalonate is further phosphorylated and decarboxylated to isopentenyl pyrophosphate (IPP) and IPP is condensed to form squalene. Through multiple reactions, squalene is cyclized to lanosterol and thereby converted cholesterol [24]. In the current study, the biosynthesis of cholesterol is implied to be increased in breast muscle of the HFE birds as genes encoding cholesterol metabolic enzymes including 24-dehydrocholesterol reductase (DHCR24), cytochrome P450, family 51, subfamily A, polypeptide 1 (CYP51A1), lanosterol synthase (2,3-oxidosqualene-lanosterol cyclase) (LSS), lipase A, lysosomal acid, cholesterol esterase (LIPA), methylsterol monooxygenase 1 (MSMO1) and squalene epoxidase (SQLE) are all up-regulated in these chickens (Table 3.1). Upstream regulator analysis provides further evidence to support increased cholesterol biosynthesis and explains why the cholesterol synthesis genes are up- regulated in the HFE birds. Through the analysis, sterol regulatory element binding transcription factor 2 (SREBF2) is predicted to be an active regulator in the HFE birds (Z-score = 2.621), increasing the gene expression of CYP51A1, LSS, MSMO1, SQLE

97 and LIPA (figure 3.3). SREBF2 is a key transcription factor that controls cholesterol homeostasis in mammals and birds [54]. In response to intracellular cholesterol depletion, SREBF2 binds to the sterol regulatory element (SRE) and thereby induces the transcription of cholesterol-related genes [55]. Taken together, the activation of SREBF2 in the HFE chickens may in part explain why cholesterol biosynthesis is possibly increased in breast muscle of these birds. As cholesterol cannot be catabolized in peripheral cells such as macrophages and muscle cells, excess cholesterol from extra-hepatic tissues is delivered by high- density lipoprotein (HDL) to the liver for excretion [56]. The whole cholesterol movement process is known as reverse cholesterol transport, which contributes to the prevention of atherosclerosis [57]. In this study, the reverse cholesterol transport pathway is enriched where two genes encoding key components of this pathway show higher expression in the HFE chickens. DE gene, lecithin-cholesterol acyltransferase (LCAT) [fold-change = 1.767], encodes a major enzyme in reverse cholesterol transport pathway converting externalized free cholesterol to cholesteryl ester. Only the esterified cholesterol is able to be absorbed by nascent discoidal HDL, resulting in the formation of mature HDL which then be transported to the liver [58]. The product of phospholipid transfer protein (PLTP) [fold-change = 1.998] is also relevant to the reverse cholesterol transport as it transfers phospholipids from triglyceride rich lipoproteins into HDL [59]. The synthesized HDL act as cholesterol acceptors which carry free cholesterol back into the liver [60] [59]. Collectively, the up-regulation of LCAT and PLTP suggests that reverse cholesterol transport is likely to be more active in breast muscle of the HFE chickens compared with the LFE birds.

98 Given that cholesterol biosynthesis, reverse cholesterol transport and inflammatory response are all strongly implicated in the HFE broilers, the three processes may be interconnected and related to atherosclerosis. Atherosclerosis is a chronic inflammatory disease characterized by cholesterol accumulation and infiltration of macrophage-dominated immune cells in the arterial wall [61][62]. The infiltrating macrophages ingest the cholesterol-rich apolipoprotein B (APOB)- containing lipoproteins deposited in the arterial vasculature, transforming to cholesterol-laden foam cells to promote atherosclerosis progression. Reverse cholesterol transport is required for the efflux of cholesterol from cholesterol-loaded macrophages, which plays a role in alleviation of atherosclerosis [56]. Ingenuity downstream effects analysis provides more insights, predicting that atherosclerosis is increased in the HFE chickens with and overlap p-value =4.04E-09 and z-score = 2.052. The elevated cholesterol synthesis, reverse cholesterol transport and inflammatory response are all probably associated with atherosclerosis in the HFE birds.

3.3.3.2 Lipid Catabolism Several genes involved in the lipolysis and fatty acid oxidation are found to be differentially expressed in breast muscle of the HFE and the LFE birds (Table 3.1).

Lipolytic genes, arylacetamide deacetylase (AADAC) [fold-change = 1.433] and Endothelial Lipase (LIPG) [fold-change = 1.651], express higher in breast muscle of the HFE birds. The enzyme encoded by AADAC has been demonstrated to be a cellular triglyceride lipase in liver, hydrolyzing the newly formed triglyceride stores and increasing fatty acid oxidation [63]. The product of LIPG is an endothelial-derived

99 lipase primarily causing the hydrolysis of phospholipid and HDL, but having little effects on triglyceride hydrolysis [64]. Furthermore, genes involved in fatty acid oxidation show differential expression levels between the HFE and LFE groups. Up-regulated in the HFE chickens, Mitochondrial acetyl-CoA acyltransferase 2 (ACAA2) [fold-change = 1.350] encodes the enzyme catalyzing the cleavage of 3-ketoacyl CoA to yield acetyl-CoA and acyl-CoA for energy production [24]. Interestingly, the transcription levels of peroxisomal enzymes, including enoyl-CoA, hydratase (EHHADH) [fold-change = 1.369] and acyl-CoA oxidase 2 branched chain (ACOX2) [fold-change = 1.420], are lower in the HFE birds. The protein encoded by EHHADH catalyzes the third step of peroxisomal beta-oxidation [24]. ACOX2 encoding enzyme has also shown to be involved in fatty acid oxidation through degrading long branched fatty acids and bile acid intermediates in peroxisome [65]. The down-regulation of EHHADH and ACOX2 may suggest that the peroxisomal fatty acid beta-oxidation is less active in breast muscle of the HFE chickens than the LFE birds. The peroxisomal beta-oxidation is able to degrade a class of compounds with very long aliphatic carbon chain that cannot be oxidized in mitochondria [66]. Eicosanoids including leukotriene and prostaglandins practically undergo primary oxidation in peroxisome, which facilitates the further elimination in the mitochondria [67]. Hence, the reduced peroxisomal beta- oxidation in the HFE birds is possibly leading to a decreased degradation of eicosanoids in breast muscle of the HFE broilers than the LFE birds, which may aggravate the inflammatory response in these birds (Discussed below).

100 3.3.3.3 Lipogenesis A few lipogenic genes are found differentially expressed in the breast muscle between the HFE and the LFE chickens. Gene encoding ATP-citrate lyase (ACLY) is expressed higher in the breast muscle of the HFE chickens compared with the LFE birds. ACLY is a lipogenic enzyme catalyzing the formation of cytosolic acetyl-CoA from citrate that is mainly exported from mitochondria. As acetyl-CoA is the primary building block for lipogenesis and cholesterogenesis, ACLY functions as a link between the energy metabolism of carbohydrate and the production of fatty acids in many tissues. [68]. In comparison with the increased expression of ACLY, genes involved in triglyceride synthesis are expressed at lower levels in the HFE chickens. One member of 1-acylglycerol-3-phosphate O-acyltransferase family, AGPAT5, is down-regulated in the breast muscle of the HFE birds. 1-acylglycerol-3-phosphate O- acyltransferase is an enzyme catalyzing the second step of phospholipid and triglyceride biosynthesis [24]. Another down-regulated gene in the HFE broilers that plays fundamental role in triglyceride synthesis is LPIN1. The enzyme encoded by LPIN1, lipin1, converts phosphatidate to diacylglycerol that is used for triglyceride, phosphatidylcholine and phosphatidylethanolamine biosynthesis [69]. Furthermore, lipin 1 can act as a transcription coactivator that directly interacts with the nuclear receptor peroxisome proliferator-activated receptor α (PPARα) and PPARγ coactivator

1α (PGC-1α), modulating the expression of genes involved in fatty acid metabolism [70]. Recent studies have shown that Lipin 1-deficient mice exhibited enhanced energy expenditure and fatty acid oxidation in the skeletal muscle, whereas muscle- specific lipin 1-overexpressing mice developed obesity-associated insulin resistance [69][71]. In contrast, the lipin 1-expression levels in adipose tissues are positively correlated with insulin sensitivity and negatively correlated with inflammatory

101 cytokine expressions and intramyocellular lipid in human and mice [72] [71]. Despite the lack of knowledge about lipin 1 in chickens, the down-regulation of LPIN1 in the HFE broilers is probably associated with the increased expression of inflammatory cytokine and receptor genes in these birds. It has been demonstrated that the expression of LIPN1 is reduced in NR4A3-silenced mouse C2C12 cells [29]. Hence, based on the Ingenuity upstream regulator analysis, the down-regulation of LPIN1 in the HFE chickens is possibly due to the putative inhibited transcription factor NR4A3. In addition to NR4A3, many other factors, such as sterol accumulation, endoplasmic reticulum stress and high fat diet, have been found to be able to suppress the expression of LPIN1[73][74][75]. In consideration of the augmented cholesterol biosynthesis and inflammatory response implicated in the breast muscle of the HFE birds, the inhibited transcription of LPIN1 is also likely to be, at least in part, associated with these factors.

3.3.4 Amino Acid Metabolism Several DE genes are found to be associated with amino acid metabolism in this study. Of note, 11 out of these DE genes encode membrane transport proteins responsible for the transfer of amino acid (Table 3.1), which indicates a difference in the process of amino acid transport in breast muscle of chickens with high and low FE.

Gene encoding cationic amino acid transporter 1 (SLC7A1) is highly up-regulated in the HFE birds with fold-change = 2.01. Overexpression of SLC7A1 has been found to be able to increase L-arginine uptake in non-hepatic tissues[76]. Thus, the up- regulation of SLC7A1 in the HFE broilers may suggest that the transport of L-arginine is increased in breast muscle of these birds. Since L-arginine has been demonstrated to

102 be beneficial for muscle growth [77], the rise in L-arginine uptake in the HFE chickens is likely contributing to their higher breast muscle yield.

3.3.4.1 The Regulation of Amino Acid Metabolism In addition to DE genes related to amino acid transport, several amino acid metabolic genes are differentially expressed between the HFE and LFE birds (Table 3.1). Genes involved in Valine degradation, including branched chain keto acid dehydrogenase E1, beta polypeptide (BCKDHB), 3-hydroxyisobutyrate dehydrogenase

(HIBADH), enoyl-CoA, hydratase/3-hydroxyacyl CoA dehydrogenase (EHHADH) and dihydrolipoamide dehydrogenase (DLD) are all down-regulated in the HFE chickens, implicating that the degradation of Valine is inhibited in breast muscle of these birds. In contrast, the HFE birds have higher expression levels of genes involved in Proline, Serine, Arginine and Asparagine biosynthesis compared to the LFE chickens, which suggests that the synthesis of these amino acid is increased in breast muscle of the HFE broilers. IPA canonical pathway analysis also shows that pathways including Valine degradation I, Proline biosynthesis I, Serine biosynthesis, Arginine biosynthesis IV and Asparagine biosynthesis I are enriched among the DE genes. In general, the transcriptome analysis indicates that amino acid synthesis is increased whereas degradation is reduced in breast muscle of the HFE chickens compared with the LFE birds. As amino acid anabolism is beneficial to muscle growth[78] [79], a more active amino acid biosynthesis and lowered amino acid degradation are likely to be partial contributors to the greater breast muscle mass in the HFE broilers. Since a wide array of downstream genes of specificity protein 1 (Sp1) are up- regulated, transcription factor Sp1 is predicted to be significantly activated (Z-score = 3.472; P-value = 4.34E-08) in the HFE chickens (Figure 3.3). Sp1 is a ubiquitous

103 transcription factor that has been found to regulate the expression of a vast number of genes in response to insulin and other hormones [80][81]. Studies have shown that mammalian Sp1 protein mediates the actions of amino acid availability and glutamine on the transcription activity of ASNS and ASS1 [82][83]. Therefore, the increased expression of ASNS and ASS1 in this study is probably modulated by Sp1, which leads to an augmented rate of amino acid biosynthesis. Apart from Sp1, activating transcription factor (ATF4) is significantly enriched among the DE genes (P-value = 3.37E-04; Z-score = 1.907), possibly disclosing the underlying mechanisms of the variation in amino acid metabolism between the HFE and LFE chickens (figure 3.3). Binding to the C/EBP-ATF response elements (CARE), ATF4 controls the transcriptions of a number of genes associated with amino acid transport and metabolism in mammals [84]. In the present analysis, ATF4 is predicted to up-regulate the transcription of cationic amino acid transporters SLC7A1 and SLC7A3, consequently, promoting amino acid uptake in breast muscle of the HFE chickens. In addition, genes encoding metabolic enzymes, including ASNS and phosphoglycerate dehydrogenase (PHGDH), are predicated to be up-regulated by ATF4, implicating a role of ATF4 in increasing amino acid anabolism in the HFE birds. Taken together, based on the transcription analysis, the enhanced amino acid uptake and biosynthesis in the HFE chickens are likely in part attributed to the regulation of Sp1 and ATF4 in their breast muscle.

104

Figure 3.3 Upstream Regulator SREBF2, ATF4 and Sp1.

105 3.3.4.2 Glutamine Metabolism Glutamine is one of the main amino acids synthesized and stored in skeletal muscles, exerting a multitude of effects on muscle growth. Apart from the role in ammonia detoxification, glutamine content in skeletal muscle has been proposed to be positively correlated with muscle protein synthesis and accretion in birds [85]. Responding to starvation, birds showed decreased levels of free glutamine content along with inhibited rates of protein synthesis in leg muscles but not in breast muscle. In contrast with mammalian skeletal muscles and bird leg muscles, chicken breast muscle have higher glutaminase (GLS) activity but lower glutamine synthetase (GLUL) activity [86][87] [85]. Since GLUL catalyzes the formation of glutamine using glutamate and ammonia whereas GLS catalyzes the reverse reaction converting glutamine to glutamate, the studies proposed that chicken breast muscle might function as an organ of net glutamine utilization [86][87] [85]. Previous study has revealed a divergence in glutamine metabolism in breast muscle of broiler chickens with high and low residual feed intakes (RFI). Aggrey et al. (2013) found that GLS was down-regulated but GLUL was up-regulated in the breast muscle of low RFI chickens when compared with birds with high RFI at day 42 [88]. In the current study, the mRNA expression of GLS is also lower in the HFE birds than the LFE chickens with fold-change = 1.77. However, opposed to Aggrey’s findings, the transcription of GLUL is significantly down-regulated in the HFE birds with fold- change = 1.55. The incompatibility between our result and Aggrey’s study is likely due to the differences in experimental chickens. The birds used in Aggrey’s research come from an Arkansas Random bred line (4), while we study chickens from a commercial cross with high breast muscle yield. The average body weight of our broilers at 46-day is more than 2.7-fold higher than theirs at 42-day. As the expression

106 levels of GLUL and GLS have shown to be positively correlated with their activities (109), the down-regulation of GLUL and GLS may implicate that compared with the LFE birds, the HFE chickens have reduced glutamine synthesis and breakdown in their breast muscle. Although the glutamine levels cannot be predicted on the basis of our data, the down-regulation of GLUL and GLS suggests that glutamine metabolism is differentially regulated in breast muscles of the HFE and LFE broiler chickens and thereby may have a potential effect on the variation of chicken feed efficiency.

3.3.5 Metabolism and Inflammatory Response

3.3.5.1 Effects of Inflammatory Response on Metabolic Regulation

As discussed in the previous paper, our data suggests an augmented inflammatory response occurred in the breast muscle of the HFE chickens. Genes encoding inflammatory cytokines and receptors, including toll-like receptor 4 (TLR4)[fold-change = 1.52], TLR2-1 [fold-change = 1.76], TLR2-2 [fold-change = 1.77], TLR6 [fold-change = 1.70], interleukin 1 receptor, type 1 (IL1R1) [fold-change = 1.50] and IL1R2 [fold-change = 2.53], are all up-regulated in the HFE birds. TLR4, TLR2 and TLR6 belong to a family of pattern-recognition receptors that are essential for the activation of pro-inflammatory response in the innate immune system [89]. Mounting evidence has indicated that TLR4 and TLR2 not only function in immune response [90][91] but also play a role in the metabolic regulation [92][93][94][95][96]. TLR4-dificient mice displayed an increase in oxidative capacity in skeletal muscles and exhibited lower levels of circulating triglycerides and non-esterified free fatty acids, suggesting that the activation of TLR4 can contribute to augmented release of fatty acids and triglyceride lipolysis in skeletal muscles [92][11]. In the present study,

107 the release of fatty acids [P-value = 3.69E-05; Activation z-score = 3.282] and lipids [P-value = 1.06E-05; Activation z-score = 3.208] are both predicted to be increased in the breast muscle of the HFE chickens compared to the LFE birds. In light of the higher expression of TLRs in the HFE broilers, it’s reasonable to postulate that the enhanced inflammatory response may be one of the causes for the increased fatty acid release in the skeletal muscles of the HFE chickens. In addition to the effects of TLRs on lipid metabolism, TLRs may serve as a modulator of glucose metabolism in the HFE chickens [93][94][95][96]. Recent studies revealed that TLR4 signaling could be stimulated by the elevated saturated fatty acids in skeletal muscles and thereby lead to the activation of NF-KB pathway and subsequently impair the insulin signaling pathway in mammals [97][98]. We found the higher expression of gene encoding for nuclear factor of kappa light polypeptide gene enhancer in B-cells inhibitor, epsilon (NFKBIE) [fold-change = 1.66] in the breast muscle of the HFE chickens. NKKBIE is a protein binding to the NF-KB complex, which inactivates NF-KB through trapping it into the cytoplasm. It has been found that the transcription of NFKBIE mRNA is increased following the activation of NF-KB [99][100]. Therefore, the up-regulation of NFKBIE may suggest that NF-KB is more active in the HFE chickens compared with the LFE birds. Although the mechanism of insulin signaling in chickens is not fully elucidated yet [101], the augmented inflammatory response may exert an effect on insulin actions in the breast muscle of the HFE chickens.

3.3.5.2 Lipid-derived Inflammatory Mediators It is well established that a wide array of lipid-derived mediators play key roles in inflammation [102]. Eicosanoid is one of the well-characterized lipid mediators that orchestrate a variety of immune responses, such as cytokine production, antibody

108 formation, and antigen presentation [103][104]. Consisting of prostaglandins (PGs), leukotrienes (LTs) and lipoxins (LXs), eicosanoids are signaling molecules mostly produced from the oxidation of arachidonic acid that is released from the membrane phospholipids [105]. In the current study, a number of genes that encode the enzymes for eicosanoids biosynthesis are shown to be up-regulated in the breast muscle of the HFE chickens. Two members of phospholipase A2 (PLA2), phospholipase A2, group VII (PLA2G7) and phospholipase A2 receptor 1 (PLA2R1), are expressed at higher levels in the HFE broilers compared to the LFE birds. PLA2 is the rate-limiting enzyme in eicosanoid synthesis, catalyzing the hydrolysis of phospholipid to generate arachidonic acid [106]. The up-regulation of PLA2G7 and PLA2R1 may implicate that there is a rise in the PLA2 activity in the breast muscle of the HFE chickens. Furthermore, arachidonate 5-lipoxygenase (ALOX5) and arachidonate 5- lipoxygenase-activating protein (ALOX5AP) are also up-regulated in the HFE birds. ALOX5 encodes an important enzyme in the synthesis of leukotriene that converts arachidonic acid to leukotriene A4 [107]. The protein encoded by ALOX5AP can increase the activation of ALOX5, contributing to the production of leukotriene in mice [108][109]. Another crucial gene associated with leukotriene biosynthesis, leukotriene A4 hydrolase (LTA4H), is also expressed higher in the HFE chickens. The product of LTA4H is a bifunctional enzyme catalyzing the rate-limiting step of leukotriene B4 biosynthesis and acting as an aminopeptidase [110]. Leukotriene B4 is a chemoattractant that exerts multiple controls in inflammatory process, responsible for leukocytes activation and cytokine production [111]. The elevated expression of ALOX5, ALOX5AP and LTA4H indicates that the biosynthesis of leukotriene B4 is

109 elevated in the breast muscle tissue of the HFE chickens, which results in an amplification of the inflammatory response in these birds.

Figure 3.4 Proposed Network between Immune and Metabolic Systems in Breast Muscles of HFE chickens.

3.4 Conclusions

The present study provides a detailed characterization of the differential expression of metabolic genes in breast muscle of chickens with extreme high and low FE. Our findings show that a large number of membrane transporter genes and enzyme-coding genes were differentially expressed in this study and suggest that metabolic processes, including glucose, lipid and amino acid transport,

110 glycolysis/gluconeogenesis, lipid catabolism, lipogenesis and glutamine metabolism, are divergently regulated in the breast muscle of HFE and LFE chickens. An increased amino acid transfer and biosynthesis and reduced amino acid degradation were implicated in the breast muscle of HFE chickens, probably contributing to the greater breast muscle mass in these birds. Transcription factors, NR4A3, SREBF2, SP1 and ATF4, were predicted to play key roles in modulating the metabolic differences between the HFE and LFE phenotypes. Remarkably, the breast muscle from HFE birds was found to contain less glycogen than that from LFE chickens. qRT-PCR and RNA-seq results indicated that the glycogen deficiency was likely attributed to the transcriptional and post- transcriptional inhibition of glycogen synthesis in the HFE broilers, which might be arisen from metabolic perturbation in the breast muscle of these birds. A reduced activity of glycerol-3-phosphate shuttle was implicated in the breast muscle of the HFE chickens, possibly indicating an abnormal cellular NAD/NADH ratio associated with the increased oxidative stress and inflammatory response found in the HFE birds. Cholesterol biosynthesis and reverse cholesterol transport were indicated to be elevated in the breast muscle of HFE broilers, which might relate to the augmented inflammatory response and link with atherosclerosis. Furthermore, our results revealed a relation between metabolism and inflammatory response in the HFE chickens. The up-regulation of TLR genes in the HFE broilers might have an effect on the fatty acid release and insulin actions in the breast muscle of these birds, while increased biosynthesis of leukotriene B4 and hyaluronan in the HFE chickens likely enhanced the inflammatory response in these broilers.

111 In summary, this study 1) characterized the major discrepancy in metabolic regulation in the breast muscle of HFE and LFE chickens; 2) disclosed a potential disturbance of metabolic homeostasis in the breast muscle of HFE birds that was likely associated with the intensified inflammatory response found in these birds. Given that the chicken is a valuable model organism, our findings could provide more information for studying connection between metabolic disorders and inflammatory response in humans. Further research (e.g., metabolomics and proteomic analysis) would be helpful for understanding the metabolic differences in breast muscle between HFE and LFE birds.

112 REFERENCES

1. Hillier LW, Miller W, Birney E, Warren W, Hardison RC, Ponting CP, Bork P, Burt DW, Groenen MA, Delany ME, Dodgson JB, Chinwalla AT, Cliften PF, Clifton SW, Delehaunty KD, Fronick C, Fulton RS, Graves TA, Kremitzki C, Layman D, Magrini V, McPherson JD, Miner TL, Minx P, Nash WE, Nhan MN, Nelson JO, Oddy LG, Pohl CS, Randall-Maher J, et al.: Sequence and comparative analysis of the chicken genome provide unique perspectives on vertebrate evolution. Nature 2010, 432:695–716.

2. Resnyk CW, Carré W, Wang X, Porter TE, Simon J, Le Bihan-Duval E, Duclos MJ, Aggrey SE, Cogburn L a: Transcriptional analysis of abdominal fat in genetically fat and lean chickens reveals adipokines, lipogenic genes and a link between hemostasis and leanness. BMC Genomics 2013, 14:557.

3. Byerly MS, Simon J, Cogburn LA, Le Bihan-Duval E, Duclos MJ, Aggrey SE, Porter TE: Transcriptional profiling of hypothalamus during development of adiposity in genetically selected fat and lean chickens. Physiol Genomics 2010, 42:157–167.

4. Ji B, Ernest B, Gooding JR, Das S, Saxton AM, Simon J, Dupont J, Métayer- Coustard S, Campagna SR, Voy BH: Transcriptomic and metabolomic profiling of chicken adipose tissue in response to insulin neutralization and fasting. BMC Genomics 2012:441.

5. Clair RWS: The Contribution of Avian Models to Our Understanding of Atherosclerosis and Their Promise for the Future. 1998, 48:565–568.

6. Dome G: Use of the Chicken as an Experimental Animal Model in Atherosclerosis. 2005, 16:151–159.

7. Wong GK-S, Liu B, Wang J, Zhang Y, Yang X, Zhang Z, Meng Q, Zhou J, Li D, Zhang J, Ni P, Li S, Ran L, Li H, Zhang J, Li R, Li S, Zheng H, Lin W, Li G, Wang X, Zhao W, Li J, Ye C, Dai M, Ruan J, Zhou Y, Li Y, He X, Zhang Y, et al.: A genetic variation map for chicken with 2.8 million single- nucleotide polymorphisms. Nature 2004, 432:717–722.

113 8. Hsu F-C, Lenchik L, Nicklas BJ, Lohman K, Register TC, Mychaleckyj J, Langefeld CD, Freedman BI, Bowden DW, Carr JJ: Heritability of body composition measured by DXA in the diabetes heart study. Obes Res 2005, 13:312–319.

9. Gaya LG, Ferraz JBS, Rezende FM, Mourão GB, Mattos EC, Eler JP, Michelan Filho T: Heritability and genetic correlation estimates for performance and carcass and body composition traits in a male broiler line. Poult Sci 2006, 85:837–843.

10. Navab M, Gharavi N, Watson AD: Inflammation and metabolic disorders. Curr Opin Clin Nutr Metab Care 2008, 11:459–464.

11. Glass CK, Olefsky JM: Inflammation and lipid signaling in the etiology of insulin resistance. Cell Metabolism 2012:635–645.

12. Perseghin G, Petersen K, Shulman GI: Cellular mechanism of insulin resistance: potential links with inflammation. Int J Obes Relat Metab Disord 2003, 27 Suppl 3:S6–S11.

13. Schmittgen TD, Livak KJ: Analyzing real-time PCR data by the comparative C(T) method. Nat Protoc 2008, 3:1101–1108.

14. Kono T, Nishida M, Nishiki Y, Seki Y, Sato K, Akiba Y: Characterisation of glucose transporter (GLUT) gene expression in broiler chickens. Br Poult Sci 2005, 46:510–515.

15. O’Rourke JF, Pugh CW, Bartlett SM, Ratcliffe PJ: Identification of hypoxically inducible mRNAs in HeLa cells using differential-display PCR - Role of hypoxia-inducible factor-1. Eur J Biochem 1996, 241:403–410.

16. Bosco MC, Puppo M, Santangelo C, Anfosso L, Pfeffer U, Fardin P, Battaglia F, Varesio L: Hypoxia modifies the transcriptome of primary human monocytes: modulation of novel immune-related genes and identification of CC-chemokine ligand 20 as a new hypoxia-inducible gene. J Immunol 2006, 177:1941–1955.

17. Fattoretti P, Bertoni-Freddari C, Di Stefano G, Casoli T, Gracciotti N, Solazzi M, Pompei P: Quantitative immunohistochemistry of glucose transport protein (Glut3) expression in the rat hippocampus during aging. J Histochem Cytochem 2001, 49:671–672.

18. Konopka JB: N-acetylglucosamine (GlcNAc) functions in cell signaling. Scientifica (Cairo) 2012, 2012:631–632.

114 19. Egger S, Chaikuad A, Kavanagh KL, Oppermann U, Nidetzky B: UDP-glucose dehydrogenase: structure and function of a potential drug target. Biochem Soc Trans 2010, 38:1378–1385.

20. Fraser JR, Laurent TC, Laurent UB: Hyaluronan: its nature, distribution, functions and turnover. J Intern Med 1997, 242:27–33.

21. Taylor KR, Gallo RL: and their proteoglycans: host- associated molecular patterns for initiation and modulation of inflammation. FASEB J 2006, 20:9–22.

22. Mráček T, Drahota Z, Houštěk J: The function and the role of the mitochondrial glycerol-3-phosphate dehydrogenase in mammalian tissues. Biochimica et Biophysica Acta - Bioenergetics 2013:401–410.

23. Mráček T, Pecinová A, Vrbacký M, Drahota Z, Houštěk J: High efficiency of ROS production by glycerophosphate dehydrogenase in mammalian mitochondria. Arch Biochem Biophys 2009, 481:30–36.

24. Nelson DL, Cox MM: Lehninger Principles of Biochemistry 5th ed. Book 2008:1–1294.

25. Hutton JC, O’Brien RM: Glucose-6-phosphatase catalytic subunit gene family. J Biol Chem 2009, 284:29241–29245.

26. Shieh JJ, Pan CJ, Mansfield BC, Chou JY: A potential new role for muscle in blood glucose homeostasis. JBiolChem 2004, 279:26215–26219.

27. Pearen M a, Muscat GEO: Minireview: Nuclear hormone receptor 4A signaling: implications for metabolic disease. Mol Endocrinol 2010, 24:1891–903.

28. Chao LC, Zhang Z, Pei L, Saito T, Tontonoz P, Pilch PF: Nur77 coordinately regulates expression of genes linked to glucose metabolism in skeletal muscle. Mol Endocrinol 2007, 21:2152–63.

29. Pearen MA, Myers SA, Raichur S, Ryall JG, Lynch GS, Muscat GEO: The orphan nuclear receptor, NOR-1, a target of beta-adrenergic signaling, regulates gene expression that controls oxidative metabolism in skeletal muscle. Endocrinology 2008, 149:2853–2865.

115 30. Pearen M a, Eriksson N a, Fitzsimmons RL, Goode JM, Martel N, Andrikopoulos S, Muscat GEO: The nuclear receptor, Nor-1, markedly increases type II oxidative muscle fibers and resistance to fatigue. Mol Endocrinol 2012, 26:372–84.

31. Ivy JL: Muscle glycogen synthesis before and after exercise. Sport Med 1991, 11:6–19.

32. Sibut V, Hennequet-Antier C, Le Bihan-Duval E, Marthey S, Duclos MJ, Berri C: Identification of differentially expressed genes in chickens differing in muscle glycogen content and meat quality. BMC Genomics 2011, 12:112.

33. Berri C, Le Bihan-Duval E, Debut M, Santé-Lhoutellier V, Baéza E, Gigaud V, Jégo Y, Duclos MJ: Consequence of muscle hypertrophy on characteristics of Pectoralis major muscle and breast meat quality of broiler chickens. J Anim Sci 2007, 85:2005–2011.

34. Le Bihan-Duval E, Berri C, Baeza E, Millet N, Beaumont C: Estimation of the genetic parameters of meat characteristics and of their genetic correlations with growth and body composition in an experimental broiler line. Poult Sci 2001, 80:839–843.

35. Alnahhas N, Berri C, Boulay M, Baéza E, Jégo Y, Baumard Y, Chabault M: Selecting broiler chickens for ultimate pH of breast muscle : Analysis of divergent selection experiment and phenotypic consequences on meat quality , growth , and body composition traits 1. 2014:3816–3824.

36. Edwards MR, McMurtry JP, Vasilatos-Younken R: Relative insensitivity of avian skeletal muscle glycogen to nutritive status. Domest Anim Endocrinol 1999, 16:239–247.

37. Montori-Grau M, Guitart M, García-Martínez C, Orozco A, Gómez-Foix A: Differential pattern of glycogen accumulation after protein phosphatase 1 glycogen-targeting subunit PPP1R6 overexpression, compared to PPP1R3C and PPP1R3A, in skeletal muscle cells. BMC Biochemistry 2011:57.

38. Cohen PTW: Protein phosphatase 1--targeted in many directions. J Cell Sci 2002, 115:241–256.

116 39. Lerín C, Montell E, Nolasco T, Clark C, Brady MJ, Newgard CB, Gómez-Foix AM: Regulation and function of the muscle glycogen-targeting subunit of protein phosphatase 1 (GM) in human muscle cells depends on the COOH-terminal region and glycogen content. Diabetes 2003, 52:2221– 2226.

40. Printen JA, Brady MJ, Saltiel AR: PTG, a protein phosphatase 1-binding protein with a role in glycogen metabolism. Science 1997, 275:1475–1478.

41. Nielsen JN, Richter EA: Regulation of glycogen synthase in skeletal muscle during exercise. In Acta Physiologica Scandinavica. Volume 178; 2003:309– 319.

42. Lyon JB, Porter J: The Relation of Phosphosylase in Skeletal Muscle and to Glycogenolysis Heart of Mice. 1963.

43. Furuhashi M, Hotamisligil GS: Fatty acid-binding proteins: role in metabolic diseases and potential as drug targets. Nat Rev Drug Discov 2008, 7:489– 503.

44. Veerkamp JH, van Moerkerk HT: Fatty acid-binding protein and its relation to fatty acid oxidation. Mol Cell Biochem 1993, 123:101–6.

45. Zanotti G: Muscle fatty acid-binding protein. Biochim Biophys Acta 1999, 1441:94–105.

46. Binas B, Han X-X, Erol E, Luiken JJFP, Glatz JFC, Dyck DJ, Motazavi R, Adihetty PJ, Hood DA, Bonen A: A null mutation in H-FABP only partially inhibits skeletal muscle fatty acid metabolism. Am J Physiol Endocrinol Metab 2003, 285:E481–E489.

47. Boord JB, Fazio S, Linton MF: Cytoplasmic fatty acid-binding proteins: emerging roles in metabolism and atherosclerosis. Curr Opin Lipidol 2002, 13:141–147.

48. Makowski L, Boord JB, Maeda K, Babaev VR, Uysal KT, Morgan MA, Parker RA, Suttles J, Fazio S, Hotamisligil GS, Linton MF: Lack of macrophage fatty-acid-binding protein aP2 protects mice deficient in apolipoprotein E against atherosclerosis. Nat Med 2001, 7:699–705.

49. Furuhashi M, Tuncman G, Görgün CZ, Makowski L, Atsumi G, Vaillancourt E, Kono K, Babaev VR, Fazio S, Linton MF, Sulsky R, Robl JA, Parker RA, Hotamisligil GS: Treatment of diabetes and atherosclerosis by inhibiting fatty-acid-binding protein aP2. Nature 2007, 447:959–965.

117 50. Uysal KT, Scheja L, Wiesbrock SM, Bonner-Weir S, Hotamisligil GS: Improved glucose and lipid metabolism in genetically obese mice lacking aP2. Endocrinology 2000, 141:3388–3396.

51. Hotamisligil GS, Johnson RS, Distel RJ, Ellis R, Papaioannou VE, Spiegelman BM: Uncoupling of obesity from insulin resistance through a targeted mutation in aP2, the adipocyte fatty acid binding protein. Science 1996, 274:1377–1379.

52. Incardona JP, Eaton S: Cholesterol in signal transduction. Current Opinion in Cell Biology 2000:193–203.

53. Ohvo-Rekilä H, Ramstedt B, Leppimäki P, Peter Slotte J: Cholesterol interactions with phospholipids in membranes. Progress in Lipid Research 2002:66–97.

54. Alipour F, Hassanabadi A: Effects of Sterol Regulatory Element-binding Protein (SREBP) in chickens. Lipids in Health and Disease 2012:20.

55. Espenshade PJ, Hughes AL: Regulation of sterol synthesis in eukaryotes. Annu Rev Genet 2007, 41:401–427.

56. Ohashi R, Mu H, Wang X, Yao Q, Chen C: Reverse cholesterol transport and cholesterol efflux in atherosclerosis. QJM - Monthly Journal of the Association of Physicians 2005:845–856.

57. Tall AR: An overview of reverse cholesterol transport. Eur Heart J 1998, 19 Suppl A:A31–A35.

58. Daniels TF, Wu XL, Pan Z, Michal JJ, Wright RW, Killinger KM, MacNeil MD, Jiang Z: The Reverse Cholesterol Transport Pathway Improves Understanding of Genetic Networks for Fat Deposition and Muscle Growth in Beef Cattle. PLoS One 2010, 5.

59. Huuskonen J, Olkkonen VM, Jauhiainen M, Ehnholm C: The impact of phospholipid transfer protein (PLTP) on HDL metabolism. Atherosclerosis 2001:269–281.

60. Nishida HI, Nishida T: Phospholipid transfer protein mediates transfer of not only phosphatidylcholine but also cholesterol from phosphatidylcholine- cholesterol vesicles to high density lipoproteins. J Biol Chem 1997, 272:6959–6964.

118 61. Hansson GK, Libby P: The immune response in atherosclerosis: a double- edged sword. Nat Rev Immunol 2006, 6:508–519.

62. Tall AR, Yvan-Charvet L: Cholesterol, inflammation and innate immunity. Nat Rev Immunol 2015, 15:104–116.

63. Lo V, Erickson B, Thomason-Hughes M, Ko KWS, Dolinsky VW, Nelson R, Lehner R: Arylacetamide deacetylase attenuates fatty-acid-induced triacylglycerol accumulation in rat hepatoma cells. J Lipid Res 2010, 51:368–377.

64. Jaye M, Lynch KJ, Krawiec J, Marchadier D, Maugeais C, Doan K, South V, Amin D, Perrone M, Rader DJ: A novel endothelial-derived lipase that modulates HDL metabolism. Nat Genet 1999, 21:424–428.

65. Baumgart E, Vanhooren JC, Fransen M, Marynen P, Puype M, Vandekerckhove J, Leunissen JA, Fahimi HD, Mannaerts GP, van Veldhoven PP: Molecular characterization of the human peroxisomal branched-chain acyl-CoA oxidase: cDNA cloning, chromosomal assignment, tissue distribution, and evidence for the absence of the protein in Zellweger syndrome. Proc Natl Acad Sci U S A 1996, 93:13748–13753.

66. Reddy JK, Hashimoto T: Peroxisomal beta-oxidation and peroxisome proliferator-activated receptor alpha: an adaptive metabolic system. Annu Rev Nutr 2001, 21:193–230.

67. Mannaerts GP, Van Veldhoven PP, Casteels M: Peroxisomal lipid degradation via beta- and alpha-oxidation in mammals. Cell Biochem Biophys 2000, 32 Spring:73–87.

68. Sun T, Hayakawa K, Bateman KS, Fraser ME: Identification of the citrate- binding site of human ATP-citrate lyase using X-ray crystallography. J Biol Chem 2010, 285:27418–27428.

69. Reue K, Dwyer JR: Lipin proteins and metabolic homeostasis. J Lipid Res 2009, 50 Suppl:S109–S114.

70. Finck BN, Gropler MC, Chen Z, Leone TC, Croce M a, Harris TE, Lawrence JC, Kelly DP: Lipin 1 is an inducible amplifier of the hepatic PGC- 1alpha/PPARalpha regulatory pathway. Cell Metab 2006, 4:199–210.

71. Phan J, Reue K: Lipin, a lipodystrophy and obesity gene. Cell Metab 2005, 1:73–83.

119 72. Reue K: The role of lipin 1 in adipogenesis and lipid metabolism. Novartis Found Symp 2007, 286:58–68; discussion 68–71, 162–163, 196–203.

73. Ishimoto K, Nakamura H, Tachibana K, Yamasaki D, Ota A, Hirano KI, Tanaka T, Hamakubo T, Sakai J, Kodama T, Doi T: Sterol-mediated regulation of human lipin 1 gene expression in hepatoblastoma cells. J Biol Chem 2009, 284:22195–22205.

74. Takahashi N, Yoshizaki T, Hiranaka N, Suzuki T, Yui T, Akanuma M, Kanazawa K, Yoshida M, Naito S, Fujiya M, Kohgo Y, Ieko M: Endoplasmic reticulum stress suppresses lipin-1 expression in 3T3-L1 adipocytes. Biochem Biophys Res Commun 2013, 431:25–30.

75. Barroso E, Rodríguez-Calvo R, Serrano-Marco L, Astudillo AM, Balsinde J, Palomer X, Vázquez-Carrera M: The PPARβ/δ activator GW501516 prevents the down-regulation of AMPK caused by a high-fat diet in liver and amplifies the PGC-1α-lipin 1-PPARα pathway leading to increased fatty acid oxidation. Endocrinology 2011, 152:1848–1859.

76. González M, Gallardo V, Rodríguez N, Salomón C, Westermeier F, Enrique GG, Abarzúa F, Leiva A, Casanello P, Sobrevia L: Insulin-stimulated L-arginine transport requires SLC7A1 gene expression and is associated with human umbilical vein relaxation. J Cell Physiol 2011, 226:2916–2924.

77. Wu G, Bazer FW, Davis TA, Kim SW, Li P, Marc Rhoads J, Carey Satterfield M, Smith SB, Spencer TE, Yin Y: Arginine metabolism and nutrition in growth, health and disease. Amino Acids 2009:153–168.

78. Wolfe RR: Regulation of muscle protein by amino acids. J Nutr 2002, 132:3219S–24S.

79. Drummond MJ, Fry CS, Glynn EL, Timmerman KL, Dickinson JM, Walker DK, Gundermann DM, Volpi E, Rasmussen BB: Skeletal muscle amino acid transporter expression is increased in young and older adults following resistance exercise. J Appl Physiol 2011, 111:135–142.

80. Solomon SS, Majumdar G, Martinez-Hernandez A, Raghow R: A critical role of Sp1 transcription factor in regulating gene expression in response to insulin and other hormones. Life Sciences 2008:305–312.

81. Samson SL-A, Wong NCW: Role of Sp1 in insulin regulation of gene expression. J Mol Endocrinol 2002, 29:265–279.

120 82. Van Leung-Pineda, Kilberg MS: Role of Sp1 and Sp3 in the nutrient-regulated expression of the human asparagine synthetase gene. J Biol Chem 2002, 277:16585–16591.

83. Brasse-Lagnel C, Fairand A, Lavoinne A, Husson A: Glutamine Stimulates Argininosuccinate Synthetase Gene Expression through Cytosolic O- Glycosylation of Sp1 in Caco-2 Cells. J Biol Chem 2003, 278:52504–52510.

84. Kilberg MS, Shan J, Su N: ATF4-dependent transcription mediates signaling of amino acid limitation. Trends in Endocrinology and Metabolism 2009:436– 443.

85. Watford M, Wu G: Glutamine metabolism in uricotelic species: Variation in skeletal muscle glutamine synthetase, glutaminase, glutamine levels and rates of protein synthesis. Comp Biochem Physiol - B Biochem Mol Biol 2005, 140:607–614.

86. Wu G, Thompsont JR, Baracos VE: Glutamine metabolism in skeletal muscles from the broiler chick ( Gallus domesticus ) and the laboratory rat ( Rattus norvegicus ). 1991, 274:769–774.

87. Wu G, Chung-Bok M Il, Vincent N, Kowalski TJ, Choi YH, Watford M: Distribution of phosphate-activated glutaminase isozymes in the chicken: absence from liver but presence of high activity in pectoralis muscle. Comp Biochem Physiol - B Biochem Mol Biol 1998, 120:285–290.

88. Aggrey SE, Lee J, Karnuah a B, Rekaya R: Transcriptomic analysis of genes in the nitrogen recycling pathway of meat-type chickens divergently selected for feed efficiency. Anim Genet 2014, 45:215–22.

89. Takeda K, Kaisho T, Akira S: Toll-like receptors. Annu Rev Immunol 2003, 21:335–376.

90. Abasht B, Kaiser MG, Lamont SJ: Toll-like receptor gene expression in cecum and spleen of advanced intercross line chicks infected with Salmonella enterica serovar Enteritidis. Vet Immunol Immunopathol 2008, 123:314–323.

91. Abasht B, Kaiser MG, van der Poel J, Lamont SJ: Genetic lines differ in Toll-like receptor gene expression in spleens of chicks inoculated with Salmonella enterica serovar Enteritidis. Poult Sci 2009, 88:744–749.

121 92. Frisard MI, McMillan RP, Marchand J, Wahlberg KA, Wu Y, Voelker KA, Heilbronn L, Haynie K, Muoio B, Li L, Hulver MW: Toll-like receptor 4 modulates skeletal muscle substrate metabolism. Am J Physiol Endocrinol Metab 2010, 298:E988–E998.

93. Shi H, Kokoeva M V., Inouye K, Tzameli I, Yin H, Flier JS: TLR4 links innate immunity and fatty acid-induced insulin resistance. J Clin Invest 2006, 116:3015–3025.

94. Reyna SM, Ghosh S, Tantiwong P, Meka CSRM, Eagan P, Jenkinson CP, Cersosimo E, Defronzo RA, Coletta DK, Sriwijitkamol A, Musi N: Elevated toll-like receptor 4 expression and signaling in muscle from insulin- resistant subjects. Diabetes 2008, 57:2595–2602.

95. Song MJ, Kim KH, Yoon JM, Kim JB: Activation of Toll-like receptor 4 is associated with insulin resistance in adipocytes. Biochem Biophys Res Commun 2006, 346:739–745.

96. Himes RW, Smith CW: Tlr2 is critical for diet-induced metabolic syndrome in a murine model. FASEB J 2010, 24:731–739.

97. Radin MS, Sinha S, Bhatt BA, Dedousis N, O’Doherty RM: Inhibition or deletion of the lipopolysaccharide receptor Toll-like receptor-4 confers partial protection against lipid-induced insulin resistance in rodent skeletal muscle. Diabetologia 2008, 51:336–346.

98. Sinha S, Perdomo G, Brown NF, O’Doherty RM: Fatty acid-induced insulin resistance in L6 myotubes is prevented by inhibition of activation and nuclear localization of nuclear factor kappa B. J Biol Chem 2004, 279:41294–41301.

99. Tian B, Nowak DE, Jamaluddin M, Wang S, Brasier AR: Identification of direct genomic targets downstream of the nuclear factor-κB transcription factor mediating tumor necrosis factor signaling. J Biol Chem 2005, 280:17435– 17448.

100. Whiteside ST, Epinat JC, Rice NR, Israël A: I kappa B epsilon, a novel member of the I kappa B family, controls RelA and cRel NF-kappa B activity. EMBO J 1997, 16:1413–1426.

101. Dupont J, Tesseraud S, Simon J: Insulin signaling in chicken liver and muscle. Gen Comp Endocrinol 2009, 163:52–57.

122 102. Henderson WR: The role of leukotrienes in inflammation. Annals of Internal Medicine 1994:684–697.

103. Harris SG, Padilla J, Koumas L, Ray D, Phipps RP: Prostaglandins as modulators of immunity. Trends in Immunology 2002:144–150.

104. Harizi H, Corcuff JB, Gualde N: Arachidonic-acid-derived eicosanoids: roles in biology and immunopathology. Trends in Molecular Medicine 2008:461– 469.

105. Greene ER, Huang S, Serhan CN, Panigrahy D: Regulation of inflammation in cancer by eicosanoids. Prostaglandins and Other Lipid Mediators 2011:27– 36.

106. Wymann MP, Schneiter R: Lipid signalling in disease. Nat Rev Mol Cell Biol 2008, 9:162–176.

107. Funk CD, Hoshiko S, Matsumoto T, Rdmark O, Samuelsson B: Characterization of the human 5-lipoxygenase gene. Proc Natl Acad Sci U S A 1989, 86:2587–2591.

108. Abramovitz M, Wong E, Cox ME, Richardson CD, Li C, Vickers PJ: 5- lipoxygenase-activating protein stimulates the utilization of arachidonic acid by 5-lipoxygenase. Eur J Biochem 1993, 215:105–111.

109. Byrum RS, Goulet JL, Griffiths RJ, Koller BH: Role of the 5-lipoxygenase- activating protein (FLAP) in murine acute inflammatory responses. J Exp Med 1997, 185:1065–1075.

110. Thunnissen MM, Nordlund P, Haeggström JZ: Crystal structure of human leukotriene A(4) hydrolase, a bifunctional enzyme in inflammation. Nat Struct Biol 2001, 8:131–135.

123