<<

The Effect of Androstenediol on the APR8/34 in A549 Cells

THESIS

Presented in Partial Fulfillment of the Requirements for the Degree Master of Science in the Graduate School of The Ohio State University

By

Scott Franklin Wray, B.A.

Graduate Program in Pathology

The Ohio State University

2012

Master's Examination Committee:

W. James Waldman, Advisor

John F. Sheridan

Copyright by

Scott Franklin Wray

2012

Abstract

A major problem in fighting influenza virus infections with the current trivalent vaccines

and antivirals is the inherent genetic instability of the RNA-encoded proteins of the virus.

Current vaccines induce neutralizing specific for two major surface

expressed on the virion. These two proteins are encoded by the hemagglutinin and

genes respectively, and in these genes, there is a high rate of point

resulting in antigenic drift of the expressed proteins. A similar problem arises

with antivirals where resistant strains develop or are selected for when antivirals are used

in a widespread fashion. In other words, the virions that are sensitive to the antiviral are

killed, whereas those that are resistant to the antiviral are free to infect susceptible

individuals. Therefore, it would be beneficial to develop anti-influenza strategies that do not rely on treatments directed against the variable genes encoded by the virus. Based on previous work from our laboratory, we hypothesize that androstenediol inhibits influenza replication by modifying the host cell’s interaction with the virus. Our proposal is that treatment with androstenediol (AED), may lead to an approach that is not directed at the variable genes encoded by the . Our results suggest that AED treatment may inhibit , as indicated by the influenza viral matrix protein, M1, gene expression. Additionally, our results exclude certain cellular mechanisms and suggest other possible mechanisms that will need to be examined in more detail.

ii

Dedication

This document is dedicated to my family and friends.

iii

Acknowledgments

I would like to acknowledge all my advisors, fellow laboratory members, and other faculty members for their guidance, encouragement, and support during the course of my training.

iv

Vita

May 1999 ...... Liberty Union-Thurston High

2003...... B.A. Biology, Capital University

2003 to 2004 ...... Graduate Research Associate, Department

of Integrated Biomedical Science, The Ohio

State University

2004 to 2009 ...... Graduate Research Fellow, Department of

Oral Biology, The Ohio State University

2009 to 2011 ...... Graduate Research Associate, Department

of Integrated Biomedical Science, The Ohio

State University

2011 to present ...... Master’s Student, Department of Pathology,

The Ohio State University

Publication(s)

1. Integrin engagement increases histone H3 acetylation and reduces histone H1 association with DNA in murine lung endothelial cells. Molecular Pharmacology. 68(2):

439-46, 2005.

v

Fields of Study

Major Field: Pathology

vi

Table of Contents

Abstract ...... ii

Dedication ...... iii

Acknowledgments...... iv

Vita ...... v

List of Tables ...... iix

List of Figures ...... x

List of Abbreviations ...... xii

Chapter 1: Introduction ...... 1

Chapter 2: Androstenediol Suppresses Influenza Virus-Encoded M1 Gene Expression in

Human Respiratory Epithelial Cells ...... 19

Introduction ...... 19

Methods ...... 23

Results ...... 28

Discussion ...... 31

Chapter 3: Androstenediol’s Effect on Cellular Signaling and a Coalescing of Pattern

Recognition Receptors ...... 42

vii

Introduction ...... 42

Methods ...... 44

Results ...... 48

Discussion ...... 51

Chapter 4: Determining Whether Androstenediol Modifies Influenza Virus-Encoded M1

Gene Expression Through the Androgen Receptor ...... 57

Introduction ...... 57

Methods ...... 59

Results ...... 66

Discussion ...... 69

Chapter 5: General Discussion...... 80

References ...... 91

viii

List of Tables

Table 2.1. Primer and TaqMan Probe Sequences of cytokines M1 and 18S for Real Time

PCR ...... 35

Table 3.1. PCR array analysis of AED activated genes in A549 cell culture with exposure to influenza virus Primer and TaqMan Probe Sequences of cytokines M1 and

18S for Real Time PCR ...... 56

ix

List of Figures

Figure 2.1 Dose Curve of AED and its Effect on Influenza Virus M1 Gene Expression in

A549 Lung Epithelial Cells ...... 36

Figure 2.2 Time Course of AED and its Effect on Influenza Virus M1 Gene Expression in A549 Lung Epithelial Cells ...... 37

Figure 2.3 Data of Influenza Virus M1 Gene Expression in A549 Lung Epithelial Cells

...... 38

Figure 2.4 Multiple Steroid Agonists and Antagonists Effect on Influenza Virus M1

Gene Expression in A549 Lung Epithelial Cells ...... 39

Figure 2.5 The above images were used to determine whether AED had an effect on influenza ...... 40

Figure 2.6 The above images were used to determine whether AED had an effect on influenza viral entry ...... 41

Figure 3.1 Data of Interferon Beta Gene Expression in A549 Lung Epithelial Cells ..... 53

Figure 3.2 Focused microarray analysis of AED activated genes in A549 cell culture without exposure to influenza virus ...... 54

Figure 3.3 Data of Interferon Beta Gene Expression in A549 Lung Epithelial Cells With

Exposure to Influenza Virus ...... 55

Figure 4.1 Data of Influenza Virus M1 Gene Expression in A549 Lung Epithelial Cells

Treated With a Dose Curve of Methyltrienolone (R1881) ...... 74 x

Figure 4.2 Time Course of AED and its Effect on Influenza Virus M1 Gene Expression in A549 Lung Epithelial Cells ...... 75

Figure 4.3 Data of Androgen Receptor Transient Transfection Assay With a Dose Curve

Comparing AED and DHT ...... 76

Figure 4.4 Data of Influenza Virus M1 Gene Expression in A549 Lung Epithelial Cells

Treated With a Dose Curve of Cyproterone Acetate in an Inhibition Assay of Androgen

Receptor ...... 77

Figure 4.5 Both blots used varying concentrations of AR siRNA ...... 78

Fig 4.6 Data of Influenza Virus M1 Gene Expression in A549 Lung Epithelial Cells

Treated With either control siRNA or androgen receptor ...... 79

xi

List of Abbreviations

β-Me β-mercaptoethanol

ΔCT delta (change) cycle threshold

ΔΔCT delta delta cycle threshold

nM nanomolar

ng nanogram

µg microgram

µl microliter

18S 18S ribosomal RNA

A549 human lung adenocarcinoma epithelial cell line

AED 5-androstene-3ß-17ß-diol

AET 5-androstene-3ß-7ß-17ß-triol

AKT serine/threonine protein kinase

ANOVA One way or one factor analysis of variance

APR8 A/Puerto Rico/8

AR androgen receptor

ARE(s) androgen response element(s)

ATII(s) alveolar type II epithelial cell(s)

BC Before Christ

C Celsius

CA cyproterone acetate

xii

CARD caspase activation and recruitment domain

cDNA complimentary DNA

CO2 carbon dioxide

Cort corticosterone

COS7 African green monkey kidney fibroblast-like cell line

CT cycle threshold

DAPI 4',6-diamidino-2-phenylindole

DE dimethyl sulfoxide/ethanol

DEX dexamethasone dH2O distilled water

DHEA dehydroepiandrosterone

DHT dihydrotestosterone

DMSO dimethyl sulfoxide

DNA deoxyribonucleic acid dsRNA double-stranded RNA

Ebp1 ErbB3-binding protein eIF-2α eukaryotic initiation factor 2 alpha

ErbB3 epidermal receptor tyrosine kinase

ERK extracellular signal-related kinase

ETOH ethanol

FAM reporter dye 6-carboxyfluorescein

FBS fetal bovine serum

xiii

Flut flutamide

GAPDH Glyceraldehyde 3-phosphate dehydrogenase

GR glucocorticoid receptor

HA hemagglutinin

HAU(s) hemagglutinin assay unit(s)

HEPES 4-(2-hydoxyethyl)-1-piperazineethanesulfuonic acid

HF hydroxyflutamide

HRP horseradish peroxidase

Hsp90 heat shock protein 90

IFN(s) interferon(s)

IFNAR interferon-α/β receptor

IgA immunoglobulin A

IgG immunoglobulin G

IL-1 interleukin-1

IL-2 interleukin-2

IPS-1 IFN-β promoter stimulator 1

IRAK interleukin receptor-associated kinase

IRF3 interferon regulatory factor 3

IRF7 interferon regulatory factor 7

Kb(s) kilobase(s)

LSD least significant difference

M molar (moles/liter)

xiv

M1 matrix protein

M2 ion channel; matrix protein 2

MDA5 melanoma differentiation-associated gene 5

ml milliliter

mRNA(s) messenger RNA(s)

MyD88 myeloid differentiation primary response gene (88)

NA neuraminidase

NF90 NFAT factor

NFκB nuclear factor κB

NFAT nuclear factor of activated T-cells

nm nanometer

NO nitric oxide

NOS2 nitric oxide synthase 2

NP

NR(s) nuclear receptor(s)

NS1 nonstructural protein 1

NS2 nonstructural protein 2

OAS 2’-5’ oligoadenylate synthetase

PA polymerase acidic protein

PAMPs pathogen-associated molecular patterns

PB1 Polymerase basic protein 1

PB1-F2 polymerase base fragment 2

xv

PB2 Polymerase basic protein 2

PBS Phosphate buffered saline

PCR polymerase chain reaction

p.i. post infection

PI3K phosphatidylinositol-3-kinase

PKR RNA-dependent protein kinase

poly I:C polyinosine-polycytidylic acid

PRRs pattern recognition receptors

pUC18 plasmid cloning vector

R1881 methyltrienolone

RAW 264.7 mouse macrophage cell line

RIG-I retinoic acid inducible gene-I

RLR(s) RIG-I-like receptor(s)

RNA ribonucleic acid

RNAase ribonuclease

RPMI 1640 Roswell Park Memorial Institute 1640 rRNA ribosomal RNA

TAM 6-carboxy-tetramethyl-rhodamine

TANK traf family member-associated NFκB activator

T cell thymus derived cell

TLR(s) toll-like receptor(s)

TLR3 toll-like receptor 3

xvi

TLR7 toll-like receptor 7

TLR8 toll-like receptor 8

TNFα tumor necrosis factor alpha

TRAF6 tumor necrosis factor receptor-associated factor 6

TRAIL TNF-related apoptosis inducing ligand

TRIF TIR-domain-containing adapter-inducing interferon-β

VIC 2′-chloro-7′-phenyl-1,4-dichloro-6-carboxyfluorescein

xvii

Chapter 1: Introduction

1.1 Overview

Today’s influenza viruses are the evolutionary products resulting from the interactions between the host and this genetically unstable virus. In spite of the development of reasonably effective vaccines, influenza infection remains a major health concern globally. For example, in the United States, even during a thirty year period when a major pandemic did not occur, from 1976 to 2006 23,607 deaths occurred each year from influenza-associated deaths (CDC 2010). This should evoke the question as to how the influenza virus can be such a health threat in the era of effective vaccines. Two types of vaccines for influenza virus are currently used, a live, cold-adapted attenuated influenza virus vaccine and a trivalent inactivated vaccine. Both typically contain three influenza viral strains, A(H3N2), A(H1N1), and B, and both induce increased plasma cell production of neutralizing IgA and IgG antibodies specific for the hemagglutinin and neuraminidase of the viral strains used in the vaccines (Cox et al. 2004). However, because the segmented influenza virus RNA (ribonucleic acid) genome is highly amenable to genetic mutation, from and drift, antigenically distinct subtypes emerge each year, requiring adjustment of the vaccine. Even more alarming, antigenic shift can result in the emergence of new viral strains, such as the avian influenza virus and the swine influenza virus, for which humans lack adequate .

1

The emergence of such strains has recently prompted fears of impending pandemics with

catastrophic implications.

In addition to antigenic drift and antigenic shift, which involve modifications of

the virus’ genes that result in protein modification, the virus can further evade innate

immune responses by sequestering and disguising internal structures such as viral dsRNA

(double-stranded RNA) (Short 2009). This, in effect, helps hide the virus from detection by host pattern recognition receptors. Furthermore, the influenza virus has also adapted itself to use the host advantageously to facilitate its own proliferation. For example, by exploiting a common cytokine secondary signaling messenger (phosphatidylinositol-3-

kinase [PI3K]) the virus can actually prevent apoptosis aimed at killing the virus before

its budding from a cell (Ehrhardt et al. 2007). Such evolutionary adaptations regularly

displayed by the influenza virus indicate a need for a treatment that does not focus on the

virus but on the host, making it much more difficult for the virus to simply evade by

mutation.

One treatment that has shown potential to protect animals with an influenza

infection is androstenediol, AED. In fact, studies have been performed in which mice

were infected with 24 HAUs (hemagglutinin assay units) of influenza virus APR8

(A/Puerto Rico/8) after treatment with AED or DMSO (dimethylsulfoxide)/ethanol, a

vehicle control. The results indicated that the AED treatment led to a significant increase

in the survivability when compared to vehicle control. In fact only 12 of the 40 or 30%

of vehicle control-mice survived, while the AED treated mice had 36 of 45 or 80%

survive (Padgett et al. 2000). Thus, this result led to an attempt to discover how this

2 increase in survivability occurred and whether the treatment was having an effect at the cellular level.

Hypothesizing that pharmacological treatment of the host resulted in less severe influenza infection, the focus of this project was to determine if there was a mechanism in which AED, modified the host cell’s interaction with the virus thereby inhibiting viral replication. AED has been shown to enhance the adaptive immune response to influenza virus leading to a higher survival rate of the host. The data also appear to suggest that early events during infection may contribute to the increased survivability of the infected hosts (Padgett et al. 2000). However, the events that occur during the early stage of infection at the primary site of viral replication are only now being discovered.

Like most androgen steroid hormones, AED drives some of its effects through the androgen receptor, and in so doing can have both non-genomic and genomic influences.

In this study, we have attempted to delineate how androstenediol, at concentrations above the physiological concentrations, can influence the replication of the influenza virus. In doing so, knowledge of the mechanism(s) by which AED increases the chance of survival from an influenza virus infection will be generated and may provide important information for the development of new therapeutic strategies for preventing or treating an influenza viral infection.

1.2 Influenza

Influenza, commonly referred to as the flu, is a highly contagious respiratory viral infection. The characteristic signs of an influenza viral infection are high fever, muscle

3 pain, headache, weakness, general discomfort, non-productive cough, sore throat, and runny nose (Monto et al. 2000). According to the Center for Disease Control, in the

Northern hemisphere, a normal influenza season begins in the middle of fall and peaks near the middle of winter. Thus, most transmission occurs during the portion of the year in which temperature decreases and the relative humidity remains low. This observation led to a study that shows the influenza virus remains more stable under cooler conditions

(Lowen et al. 2007). It has also been proposed that lower relative humidity allows ejection of the virus to remain in smaller water droplets, thus allowing it to remain airborne for a longer length of time (Lowen et al. 2007). Thus, one way transmission occurs is from inhaling water droplets containing the influenza virus. Another route of transmission occurs by touching a surface (ex. hand, door knob, etc.), contaminated with water droplets containing influenza virus and then introducing the virus to a respiratory mucosal surface. Although the characteristics of the infection have been noted for thousands of years, understanding how an influenza viral infection could be prevented

(by attempting to inhibit transmission or replication) has only recently been developed as modern molecular techniques have been employed to study the virus.

The influenza virus is believed to have been around since 412 BC (Before Christ), when it was first described by Socrates, and humans have been trying to eliminate it since the first recorded influenza pandemic in 1580 (Ghendon 1994). The first attempt at vaccination was in the mid 1930’s and the results of the experiment were inconclusive, but it was believed that the increase seen in antibodies may have suggested a possible prevention of influenza virus (Francis & Magill 1937). Later, it was shown that vaccines

4

were able to prevent influenza, although it quickly became apparent that the effectiveness

of an influenza virus vaccine generally decreased as the years past. Thus other treatments

such as antivirals were developed and have also shown decreased efficacy after multiple

applications (Boltz et al 2010). To understand why this occurs one may need to

appreciate the interactions between the influenza virus and host’s response.

Influenza A virus is a negative polarity, single-stranded, segmented RNA virus whose 8 RNA segments encode for 11 proteins including the envelope glycoproteins hemagglutinin (HA) and neuraminidase (NA), matrix protein (M1), nucleoprotein (NP), polymerase base fragment 2 (PB1-F2), RNA polymerases PB1, PB2, and PA, ion channel

(M2), and nonstructural proteins (NS1 and NS2) (Clifford et al. 2009). The virus gains entry to the body through an airway passage. Then HA on the binds to sialic acid-containing receptors on the surface of respiratory epithelial cells and induces clatherin-mediated endocytosis. Neuraminidase cleaves the sialic acid and plays an important part in viral entry and release. HA-specific antibodies prevent either attachment of the virus to the host or intra-endosomal fusion, and if HA-specific levels are high enough they may completely prevent the initiation of infection

(Mozdzanowska et al. 1997). Neuraminidase-specific antibodies are also protective, but unlike HA-specific antibodies, the NA specific antibodies are important in suppressing subsequent viral replication and not the initial entry into cells (Kilbourne et al. 1968,

Murphy et al. 1972, Powers et al. 1996). Current vaccines for influenza virus act by inducing neutralizing antibody (primarily IgA and IgG) responses to specific epitopes of

5

HA (i.e., neutralizing epitopes: aa92 – 105, 127 – 133 and 183 – 195) and to NA in a

heterotypic fashion (Li et al. 2003).

The upper respiratory tract is initially infected by the virus, and this causes systemic (i.e., fever, headaches and severe malaise) and local respiratory (i.e., nonproductive cough, sore throat, and rhinitis) symptoms. While the incubation period is approximately 1-5 days, the period of infection that has the greatest chance of communicability is within the first 3 days. Respiratory epithelial cells are the primary targets for infection, and these cells serve as the main factories for the production of large numbers of viral progeny (Schmolke & García-Sastre 2010, Shieh et al. 2010, Yu et al. 2011). As the infection progresses, the virus spreads to lower respiratory tissues and if left unchecked can cause progressive primary viral pneumonia or the more common secondary bacterial pneumonia (Rello & Pop-Vicas 2009). In extreme cases or in susceptible individuals (such as the very old and very young), lethal symptoms can occur such as pulmonary edema and acute respiratory distress syndrome, which is associated with massive infiltration of mononuclear leukocytes (Taubenberger &

Morens 2008). This can be associated with haemophagocytosis in which macrophages, presumably over-activated by high levels of inflammatory cytokines, begin to phagocytose and destroy erythrocytes, leukocytes, and platelets (Taubenberger & Morens

2008). As our study will focus specifically on the early innate response from the lung epithelial cell, we will not be providing much detail on the adaptive response or the transition from the innate to adaptive response, but a thorough review of how innate cells are involved in the regulation of the adaptive immune response during influenza a viral

6

infection can be found in a 2009 review (McGill et al. 2009). Another 2009 article,

details the discovery of other receptors that are essential for the adaptive response to the

influenza virus (Ichinohe et al. 2009) and fits with a short general review of innate

regulation of adaptive immunity (Iwasaki & Medzhitov 2010).

Understanding the human’s response to infection and subsequent recovery should

lead to an important health question. Why doesn’t an infection or vaccination create an

adaptive immune response in which anti-viral antibodies and life-long immunological memory prevent future infections? The reason is based on the fact that the antibodies that are generated in response to the antigens of the virus (or vaccine) only negatively select a very specific subtype. In addition, that negative selection of the antibody- neutralized strains actually positively selects alternative viral subtypes that are not recognized by circulating antibody. In other words, the presence of a specific antibody puts immunological pressure on the mixed influenza virus population and helps facilitate

productive antigenic shift and antigenic drift leading to multiple viral subtypes for which

there is no immunological memory. Antigenic shift is the result of two different strains

mixing their gene segments to form a new infectious subtype, involving animals, such as

poultry or swine (Nelson & Holmes 2007, Treanor 2004). High variability can also be

caused by antigenic drift or the random accumulation of in viral genes coding

for proteins (Nelson & Holmes 2007, Treanor 2004). This high mutation rate is

attributed to a less than efficient viral RNA polymerase combined with the lack of

proofreading during nucleic acid replication. The result of frequent random accumulation

of errors in the RNA leads to amino acid changes, which exposes different antigenic

7

surfaces. Thus the ability of the virus to change so readily is the major reason why

researchers in the influenza virus field have decided to examine whether modifying the

cell that the virus infects is a better way to combat the influenza virus.

1.3 Cellular Recognition and Reaction to Influenza Viral- Transcriptional control of the innate immune response to influenza virus

At the time a cell becomes infected with influenza virus, the respiratory epithelial cell sends out a call for help; the virus triggers that alarm signal. This occurs because innate immune recognition detects conserved microbial products that are unique to microorganisms. The microbial products that are recognized are typically essential to microbial function and are conserved among closely related species. For example, hemagglutinin and neuraminidase are viral products that are not manufactured by eukaryotic cells. Normally the structures that could recognize these conserved pathogen- associated products would not recognize host components; their presence would thus signal an infection. These conserved targets of innate immune recognition are called pathogen-associated molecular patterns (PAMPs) (Pichlmar & Reis, 2007) and the receptors on or within host cells are called pattern-recognition receptors (PRRs) (Meylan et al. 2006).

PRRs function to induce phagocytosis, activate proinflammatory signaling pathways, activate complement cascades, and induce apoptosis The PRRs include Toll- like receptors (TLRs) and cytoplasmic retinoid acid inducible gene-I (RIG-I)-like receptors (RLRs). In mammals, PRRs that are capable of detecting an infecting pathogen

8

drive the transcription of genes that activate cells of the innate and

subsequently stimulate the . Although PRRs were initially

thought to be transmembrane proteins expressed on the surface of the cell, we now know

that several of the PRRs, such as TLR3 (toll like receptor 3) and TLR8 (toll like receptor

8), are localized on the internal surfaces of endosomal/lysosomal compartments (Gorden

et al. 2006). In this configuration, they are more likely to interact with molecular patterns

of viral replication. For example, poly I:C (polyinosine-polycytidylic acid), which mimics viral double-stranded RNA, stimulates TLR3 (Kimura-Takeuchi et al. 1992), whereas imiquimod or resiquimod, which mimic viral single-stranded RNA activate

TLR7 (toll like receptor 7) or TLR8 (Hattermann et al. 2007).

Although stimulation of any of the PRRs results in the activation of a shared group of inflammatory cytokine genes, each PRR is responsible for its own ‘fingerprint’ that drives the immune response towards an anti-viral, anti-bacterial, anti-parasite, or anti-fungal response (Akira et al. 2006, Kawai et al. 2010). Briefly, after interaction with their respective ligands, all TLRs except TLR3 utilize the adapter protein, MyD88

(myeloid differentiation primary response gene (88)), to activate TRAF6 (tumor necrosis factor receptor-associated factor 6) through the IRAK (interleukin receptor-associated kinase) pathway. In contrast, TLR3 uses the adapter protein, TRIF (TIR-domain- containing adapter-inducing interferon-β), in place of MyD88, to activate TRAF6 (Hoebe et al. 2003, Oshiumi et al. 2003, Yamamoto et al. 2003). Regardless, both MyD88 and

TRIF activation lead to NFκB (nuclear factor κB) transcriptional activity. The activation of NFκB results in the transcription of the cytokine genes shared by all members of the

9

TLR family. Somewhat unique to TLR3 ligation is the induction of IRF3 (interferon

regulatory factor 3), which activates transcription of the anti-viral type I IFNs

(interferons) (Yamamoto et al. 2003). Likewise, TLR8 activation drives type I IFN

expression albeit by utilizing IRF7 (interferon regulatory factor 7) instead of IRF3

(Matsumoto et al. 2004). Thus, although influenza-mediated activation of the TLRs,

drives the expression of the common ‘inflammatory’ genes through NFκB, their specific

activation of TLR3 and TLR8 would selectively drive the expression of a subset of genes

responsible for activation of the appropriate anti-viral immune response (i.e., the type I

interferons). Those TLR3 and TLR8-induced genes would, in part, be dependent on the

transcriptional activity of IRF3 and IRF7. Recognition of the influenza virus by TLRs in some epithelial cells are limited at best, because TLRs are undetectable or only detectable after upregulation by cytokines such as TNFα (tumor necrosis factor alpha) or IFN β

(Tissari et al. 2005). Because of this, RLRs are thought to be the detection methods for the ‘infected’ respiratory epithelial cells (Siren et al. 2006).

There are several RLRs including RNA helicase RIG-I, melanoma differentiation- associated gene-5 (MDA-5), and RNA-dependent protein kinase (PKR) that are known to recognize some form of dsRNA. Although PKR is able to recognize dsRNA, this regulation only occurs after IFN β has activated it (Clemens & Elia 1997, Samuel 2001).

Currently the most likely PRR involved in the early recognition of influenza virus by epithelial cells is RIG-I. At first it was not clear whether MDA5 and RIG-I were redundant, because both were thought to have showed the ability to bind to poly I:C (a synthetic polyinosinic-polycytidylic acid double-stranded RNA or a substitute for

10

influenza virus activation). Until recently it was not known that RIG-I was unaffected by

this interaction and currently it has been shown that RIG-I, not MDA5, is the essential

RLR in early cellular recognition of the influenza virus (Pichlmar & Reis 2007).

After recognizing the influenza virus, RIG-I initiates antiviral and pro- inflammatory signaling (Le Goffic et al. 2007). The signaling begins with RIG-I via the recruitment domain interacting with the IFN β promoter stimulator 1 (IPS-1), which then leads to the activation of IRFs 3 and 7 through Traf family member-associated NFκB activator (TANK)-binding kinase 1- and Inhibitor of kappa B kinase i-dependent phosphorylation. IPS-1 also activates NFκB via Fas-associated death domain and regulated intramembrane proteolysis-1-dependent pathways. IRFs 3 and 7 coordinately activate IFN β promoter, while NFκB activates pro-inflammatory cytokines. Thus IFN β is one of the major coalescing points of all the currently known PRRs in the epithelial cell.

IFN β, an antiviral cytokine, is one of only a limited number of cytokines and chemokines that have previously been shown to produce a response to a viral infection in epithelilal cells (Gern et al. 2003, Guillot et al. 2005, Julkunen et al. 2001). The secretion of IFN β is one of the initial antiviral events in viral-infected cells and it has been shown to act on specific receptors of neighboring cells to prevent infection by inducing a number of antiviral genes including PKR and ribonuclease (RNAase) L/2’-5’ oligoadenylate synthetase (OAS) (Goodburn et al. 2000, Haller et al. 2007). The importance of this pathway during an influenza viral infection has been confirmed in vivo by IFN β receptor knockout mice (Bergmann et al. 2000, Durbin et al. 1996). PKR, an

11

IFN β induced, dsRNA-activated cytokine is capable of blocking protein synthesis through its ability to phosphorylate a subunit of eukaryotic initiation factor 2 alpha (eIF-

2α), thus inhibiting viral replication (Saelens et al. 2001). Thus the IFN β gene is a suitable target to study the early signaling that may potentially affect the replication of influenza viruses.

1.4 Influenza Virus Inhibition and Usage of the Cellular Defense

When influenza viruses enter a mammalian cell, it begins an intracellular battle.

As mentioned previously, the cell activates a number of antiviral defenses to combat the virus; the virus, on the other hand, attempts to circumvent these cellular defenses. To counteract the antiviral effects of IFN induction and PKR activation, many eukaryotic viruses, including influenza, have developed strategies to block PKR activity and inhibit the post-transcriptional processing of cellular precursor mRNAs (messenger RNAs)

(Haller & Weber 2007, Krug et al. 2003). The influenza virus also prevents the cell from recognizing viral dsRNA by binding the dsRNA with a viral nonstructural protein, NS1

(Cheng et al. 2009). NS1 proteins sequester and disguise the dsRNA from host dsRNA receptors, like RIG-I and TLR3, therefore hindering the cellular immune defense mechanisms (Min et al. 2006, Newby et al. 2007).

As the influenza infection continues, a switch from host mRNA transcription to genomic viral mRNA occurs, resulting in synthesis of viral proteins (Engelhardt & Fodor

2006). Another function of NS1 is the inhibition of the posttranscriptional processing of cellular antiviral pre-mRNAs by binding cellular proteins that are required for this

12

process (Nemeroff et al. 1998). At the same time, this does not affect viral mRNAs,

because viral mRNAs are produced by the viral polymerase and not the cellular

polymerase. Thus, many of the host genes upregulated in the presence of the influenza

virus are unable to be translated into proteins, while the virus is still capable of producing

viral mRNA and protein. Beyond the attempts to counteract the defense that the cell has

evolved, influenza uses other cytokines of the cellular defense to facilitate its own

proliferation.

Just as with other cellular defensive signaling, apoptosis and/or the apoptotic process have been shown to be detrimental (Kurokawa et al. 1999) and essential to the replication of the influenza virus (Wurzer et al. 2003, Wurzer et al. 2004). There are at least three ways that apoptosis is regulated in an influenza infected cell: caspase activation and recruitment domain (CARD, type of death domain), several viral proteins, and transcription of apoptotic factors (Ludwig et al. 2006, Mazur et al. 2008, Morris et al.

2005). The apoptotic pathway most relevant to this study is the transcription of apoptotic factors. An example of the transcription of apoptotic factors would be Fas, Fas Ligand, and TRAIL (TNF-related apoptosis inducing ligand), which are expressed in an NFkB- dependent manner and activate the initiator caspases (caspase 8 and 9) (Ludwig et al.

2006). This led us to understand that a successful balance of the apoptotic process is essential for influenza viral propagation. NS1 sequestering viral RNA is an example of the virus attempting to prevent too early of an apoptotic response from the cell, which would result in little to no replication before cell death. An interruption in this delicate balance may result in decreased replication of the influenza virus. Thus, if such a

13

treatment as AED were able to disrupt the molecular balancing act that the virus uses, it

may result in the limitation of viral replication.

1.5 Androgen Receptor

AED is a naturally occurring hormone in the human body, as it is a metabolite of

dehydroepiandrosterone (DHEA) a steroid produced by the human adrenal cortex. To

understand how AED has an effect on a cell, one must look at the target molecule that

AED interacts with at the cellular level, which is a nuclear receptor. Nuclear receptors

are proteins that are differentiated by the ligands binding them, and the steroid response

element recognition that interacts with the nuclear receptor (Bain et al. 2007). Most

nuclear receptors contain a conserved set of regions consisting of an amino-terminal

domain, a central DNA-binding domain, and a carboxy-terminal ligand-binding domain.

To become more acquainted with how a nuclear receptor functions, we will focus on one

nuclear receptor, the androgen receptor.

In the cytoplasm androgen receptors (ARs) are inactive steroid receptors. The androgen receptors exist in complexes with chaperones, including heat-shock proteins 70

and 90 that hold their ligand binding domain in constitutively open states (Black et al.

2004, Picard et al. 1987, Hager et al. 2000, Picard et al. 1987, Pratt et al. 2004). Once

bound by an androgen, such as AED, androgen receptors are known to form homodimers

thus activating a nongenomic pathway in the cytosol or traveling to the nucleus where the

androgen receptor binds to androgen response elements (AREs) (Claessens et al. 1996,

Rennie et al. 1993, Shaffer et al. 2004, Verrijdt et al. 1999, Verrijdt et al. 2000). The

14 androgen complex (androgen bound to androgen receptor) is known to activate the

PI3K/AKT (serine/threonine protein kinase) pathway in the cytosol without needing to translocate into the nucleus, although most of the known affects appear to result from transcriptional regulation (Baron et al. 2004, Evans 1988, Giguere et al. 1988,

Mangelsdorf et al. 1995).

Androgen binding an androgen receptor induces a conformational change that reveals the nuclear-localization signal overlapping the DNA-binding domain (Jenster et al. 1993, Simental et al. 1991, Zhou et al. 1994). The translocation of the AR into the nucleus occurs once the NLS can be recognized by an import receptor such as importin-α

(Fahrenkrog et al. 2004, Paschal 2002). Inside the nucleus, more than 500 genes are regulated by androgen/androgen receptor complex in epithelial cells, and these genes are also regulated by many other differing functions (Nelson et al. 2002). Occupancy of the

AREs (androgen response elements) by ARs is known to be cyclic, but the precise timing and co-regulator recruitment have not been described in detail (Kang et al. 2002). One problem that arises is that androgen responsive genes, those bound by an AR at the ARE, have shown that most AREs are further than 10 kilobases (kbs) upstream and/or downstream from the target gene’s transcription start site (Bolton et al. 2007).

Furthermore, a significant portion of androgen responsive genes have been shown to be further than 54 kb away from the ARE (Bolton et al. 2007). Another problem, as with many genes, is that the androgen responsive genes may result in secondary genes being affected. Thus, there are many challenges in determining which genes are direct primary targets of the androgen receptor. Something else to consider is that the androgen

15 receptor, like other nuclear receptors, not only acts as a transcription factor, but also has been found to directly or indirectly interact with other proteins found in the cell (Jasavala et al. 2007, Hedman et al. 2006, Honda 2007, Honda et al. 2008, Zhang et al. 2002).

1.6 Interaction of Influenza Viral Subunits and Androgen Receptors

The influenza virus has been shown to interact with many different host cellular proteins, including several cellular proteins that have also been found to interact with subunits of the influenza viral polymerase complex (Deng et al. 2006, Honda et al 2007,

Honda et al. 2008, Momose et al. 2002). The phase in viral replication where these cellular proteins have been found to interact with the virus is the period after the virus escapes the endosome. After exiting the endosome, viral polymerase complex subunits

(PB1, PB2, and PA (polymerase acidic protein)) and nucleoprotein (NP) enters the host cell nucleus, where synthesis of vRNA, positive-strand RNA, or cRNA occurs. During the transport of the influenza viral polymerase complex subunits, one or more of the subunits are known to interact with at least 9 to 11 human cellular proteins (Deng et al.

2006, Honda 2008, Momose et al. 2002). Three of these cellular proteins have been identified: Ebp1 (ErbB3[epidermal receptor tyrosine kinase]-binding protein), Hsp90

(Heat Shock Protein 90), and Ran binding protein 5 (Honda 2007, Honda et al. 2008,

Naito et al. 2007). Even more notable is that two of these identified cellular proteins

(Ebp-1 and Hsp90) have also been shown to interact with the AR (Honda 2007, Honda et al. 2008, Zhang et al. 2002).

16

Ebp1’s carboxy-terminal proximal end has been shown to interact with the

catalytic region, for RNA polymerization, of PB1 (Honda et al. 2007). Thus, Ebp1 has

been shown to inhibit RNA synthesis of the influenza virus by binding to PB1 (Honda

2007, Honda et al. 2008, Zhang et al. 2002). This carboxy-terminal region of Ebp1, which interacts with PB1, has also been shown to contain a LXXLL motif that is important for mediating interactions with nuclear receptors, including the amino-terminal domain of AR (Zhang et al. 2002). Ebp1 is known to have many other cellular roles including, but not limited to involvement in apoptosis, protein translation, growth regulation, and the ErbB3 signal transduction pathway (Honda et al. 2007). AR is also known to interact with Hsp90, although there is a lack of specifics as to which domains interact. It is known which domains Hsp90 interacts with PB2; Hsp90s N-terminal domain and its middle region (Momose et al. 2002). Recently many more proteins have been discovered that interact with proteins of the influenza virus, in fact many have been shown to interact with subunits of the viral polymerase complex (Konig et al. 2010).

Together these studies show the potential for an interaction between AR and subunits of the influenza virion. Understanding the complexes that both the androgen receptor and influenza virus form may lead to an effective target to inhibit viral replication.

1.7 Hypothesis

Prior results have shown that treatment with AED increases the survival of the host after infection with influenza (Padgett et al. 1997, Padgett et al. 1998, Padgett et al.

17

1999, Padgett et al. 2000, Padgett et al. 2000b). In addition, it has been shown in our laboratory that AED appears to act through the androgen receptor in RAW 264.7 cells, a mouse macrophage cell line. This along with the knowledge that the AR and proteins of the influenza virus are known to interact with the same cellular proteins led us to our hypothesis that AED treatment would result in decreased replication of the influenza virus in lung epithelial cells. After determining that inhibition of replication may occur

(Fig. 2.1), we developed a central hypothesis that androstenediol inhibits influenza virus replication by modifying the lung epithelial cell’s interaction with the virus. There are three corollary hypotheses that must be evaluated in order to suggest how this may occur.

First, we hypothesize that AED, through the AR, directly influences the virus encoded polymerase and in so doing inhibits viral replication. To determine whether AED is having a direct effect on the influenza viral polymerase complex through its binding to

AR, we will perform binding assays. Our second corollary hypothesizes that AED can enhance the innate immune response in respiratory epithelial cells. In order to determine whether AED can alter the innate immune response we will determine the gene expression of IFN β. Thirdly, we hypothesized that AED modifies other cellular signaling pathways during infection thus affecting influenza viral replication. To investigate this we will use pathway finder arrays to look at multiple pathways to determine whether AED is able to modify these pathways during an influenza infection.

Taken together these studies are aimed at determining if and how AED inhibits the in vitro replication of influenza virus in human lung epithelial cells.

18

Chapter 2: Androstenediol Suppresses Influenza Virus-Encoded M1 Gene Expression in Human Respiratory Epithelial Cells

2.1 Introduction

The influenza virus apparently was first described by Socrates in 412 BC and

since the first recorded influenza pandemic in 1580, humans have been trying to

eliminate it (Ghendon 1994). The first attempt at influenza virus vaccination was in the mid 1930’s on a small group of human subjects but the tests were unable to determine if the vaccine would prevent influenza; however there was an increase seen in the number of antibodies against the influenza virus suggesting that prevention of influenza-related

disease was possible (Francis & Magill 1937). Shortly thereafter, it was shown that

vaccines were, in fact, able to prevent influenza infection; the was first

introduced as a licensed product in the United States in 1944. Soon after the

development of these first anti-influenza vaccines it became apparent that the

effectiveness of an influenza virus vaccine decreased continually year after year as the

influenza virus mutated and escaped the protection afforded by the vaccine-induced

antibodies that were highly specific to the original viral . In other words, the

mutated progeny virions were not recognized by the specific vaccine-induced antibodies.

As it became clear that development of long-lived protection against influenza virus with a vaccine was going to be elusive, efforts were redirected toward prevention of viral replication. The first antiviral for the influenza virus was developed in 1966 and

19

was intended to inhibit the release of influenza viral progeny from infected cells.

Currently, antivirals act in two distinct manners: first there are inhibitors of

neuraminidase (NA) and second there are inhibitors of matrix protein 2 (M2). The

current neuraminidase inhibitors include zanamivir and oseltamivir, while the inhibitors

of M2 include amantadine and rimantadine. The neuraminidase inhibitors act as

competitive inhibitors for sialic acid, thus preventing new viral particles from being

released from infected cells. Sialic acid is known to play an important role in human

influenza viral infections by allowing the hemagglutinin to bind to the cell surface before

both the entrance and release of the virion from the cell. The other antivirals act as matrix protein 2 (M2) inhibitors by affecting the influenza virus M2 ion channel. The

M2 also plays a crucial step in the life cycle of the influenza virus as it allows hydrogen ions to pass through the membrane, enter the virion, and result in a dissociation of the viral matrix protein (M1) from the ribonucleoprotein. This results in the uncoating of the virus and exposes the contents of the virion to the cytoplasm of the cell. However, similar to the vaccination strategy against influenza virus, the current anti-virals target the function of highly variable virus-encoded genes and thus can become ineffective. An example of resistance to these antivirals was shown during the 2005 and 2006 influenza seasons when the rates of resistance to M2 inhibitors reach levels greater than 90% in many countries (Bright et al. 2006). Given the poor ability to prevent or treat influenza viral infection a new approach to combating the virus that does not target highly variable virus-encoded genes is required. One possible approach could be through the use of, the androgen hormone androstenediol (AED).

20

Previous studies had shown that AED was an effective treatment for influenza in

a murine model (Padgett et al. 1997, Padgett et al. 1998, Padgett et al. 1999, Padgett et al.

2000, Padgett et al. 2000b). The hypothesis that AED might be an effective treatment for

influenza viral infection first came from the work of Loria et al., who showed that the

steroid hormone dehydroepiandrosterone (DHEA) regulated host immunity (Loria et al.

1996). Not only was DHEA able to regulate the host immunity, but it also significantly

increased the protective effect against both viral and bacterial infections. This research

was later followed up with the determination that DHEA’s effects occurred through its

metabolite, AED. Early testing of an enteroviral challenge showed that AED was able to

provide better protection than DHEA (Loria & Padgett 1991). Based on this successful

treatment of an enteroviral infection with AED, experiments were designed to investigate

whether AED would have a similar impact during an influenza viral infection.

This led to three individual studies in mice, which when collapsed into a single

data set, revealed that AED treatment significantly increased survival following a lethal challenge with influenza APR8 virus (Padgett et al. 1997). In fact, 80% of the AED treated mice survived compared to only 30% in the control group (Padgett et al. 1997).

In addition, these studies showed AED significantly increased inflammation at day 3 post infection (p.i.) compared to non-AED treated controls (Padgett et al. 1997). This increased inflammation on day 3 p.i., was associated with accumulation of cells in the draining mediastinal lymph node of infected mice (Padgett et al. 1997). This inflammatory infiltrate resolved by day 7 p.i. in AED-treated mice while continuing in the control mice (Padgett et al. 1997). Thus, these data suggest that AED treatment

21

resulted in an earlier response to the virus and a quicker clearing of the virus from the

lungs when compared to the control mice.

Because AED appeared to act on the early response to the virus, it was decided

that these experiments would target the earliest site of the antiviral response to the influenza virus, and the virus’s target for replication, the epithelial cell. Thus, we proposed that by modifying the epithelial cell, early viral proliferation would be limited thus preventing viral spread. This strategy seemed plausible since the influenza virion is limited by its small coding capacity and thus relies upon the host’s cellular machinery for many stages of its replication cycle.

To gain a better understanding of the cellular pathogenesis associated with the influenza virus, and to determine the immune mechanisms importance for combating the infection at a cellular level, we chose to look at the primary site of infection, the respiratory tract. The respiratory tract contains multiple cell types, but the site where viral proliferation occurs is the alveolar type II epithelial cell (ATIIs). The alveolar type

II epithelial cells have been shown to be a primary target for influenza viral infection

(Wang J et al. 2009). Thus, the cell line A549, that has been characterized as a type II pulmonary epithelial cell line (Foster et al. 1998, Zhang et al. 2005) and has also been shown to contain both α2-3 and α2-6 sialic acid, with high levels of α2-6 sialic acid, was chosen for study (Kumari et al. 2007, Gulati rt al. 2005). Having a cell displaying α2-6

sialic acid is important because human influenza viral strains have been shown to

predominantly attach to this sialic acid in humans (Ibricevic et al. 2006, Thompson et al.

2006). Taken together, these findings guided us to consider the effects of AED on

22

influenza viral replication in A549 cells. We hypothesized that treatment with AED

would significantly limit viral replication, as shown by M1 gene expression, in the A549

lung epithelial cells.

2.2 Methods

Cell Culture: The A549 (CCL-185™) human epithelial cell line was cultured in Roswell

Park Memorial Institute 1640 (RPMI 1640) media (Invitrogen, Carlsbad, CA) adjusted to

0.01 Molar (M) 4-(2-hydoxyethyl)-1-piperazineethanesulfuonic acid (HEPES), 10-5 M β-

mercaptoethanol (β-Me), 0.075% sodium bicarbonate, 1.5mM L-glutamine, 50 units/ml

(U/ml) penicillin, and 50 µg/ml streptomycin. The RPMI 1640 culture media was supplemented with 10% fetal bovine serum (FBS) (Biocell Laboratories, Inc., Rancho

Dominguez, CA) and cultures were grown in a 37°C incubator in the presence of 5%

CO2 (carbon dioxide).

Experimental Stimulation and Infection: Approximately one day prior to experimental

stimulation, culture supernatants were removed and refreshed with RPMI 1640

supplemented with 10% FBS. The A549 cellular monolayer was then trypsinized with

GIBCO® Trypsin (Invitrogen), collected in conical tubes, and centrifuged. The media

was then removed and freshly made RPMI 1640 was added to resuspend the cell pellet.

A sample of this suspension was diluted with trypan blue stain and the cell number and

viability were calculated using a counting chamber. Cells were then plated in single

culture dishes at approximately 6 x105 cells/ml in RPMI 1640 supplemented with 10%

23

FBS and allowed to incubate overnight at 37°C incubator in the presence of 5% CO2.

AED (Sigma Aldrich, St. Louis, MO), androstenetriol (AET) (a gift from the Loria

laboratory), dehydroepiandrosterone (DHEA) (Sigma Aldrich), dihydrotestosterone

(DHT) (Sigma Aldrich), flutamide (Flut), hydroxyflutamide (HF), corticosterone (Cort),

and dexamethasone (DEX) (Sigma Aldrich) were prepared fresh before each experiment

at a concentration of 5 x 10-3 M – 1 x 10-4 M solution in a solution of dimethylsulfoxide

(DMSO) and ethanol (ETOH) referred to from herein as DMSO/ETOH (which also served as the vehicle control). The final concentration of DMSO/ETOH in culture media was 0.1% by volume. In all experiments, media was removed and replaced with fresh media containing 10% FBS. Treatment groups were added to stock media preparations at a concentration of 1 µl/ml of 5 x 10-3 M – 1 x 10-4 M solution in a solution of

DMSO/ETOH. Samples that did not receive AED were treated with 1 µl/ml of

DMSO/ETOH or culture media lacking any other solution, as a media control.

Concurrently, treatment groups received media containing 1 hemagglutinatin assay unit

(HAU) of influenza A/PR8/34 (H1N1) virus per ml of culture media.

RNA Isolation: Total RNA isolation was performed by using TRIzol ReagentTM

(Invitrogen) according to the manufacturer’s protocol. The TRIzol method accomplishes this by using a phenol/chloroform based phase separation protocol and is followed by an isopropanol-induced RNA precipitation. RNA precipitates were then washed by adding 1 ml of 75% ETOH. After the RNA wash, the 75% ETOH was decanted and the RNA pellet was resuspended in a range of 20-50 µl of distilled water (dH2O), depending on

24

pellet size. Total RNA concentrations and purity were determined by a

spectrophotometer. The RNA method was selected on the spectrophotometer to analyze

at 260 nm and 280 nm wavelengths. Concentration of RNA were determined by

measuring the absorbance at 260 nm in a spectrophotometer using quartz cuvettes. Pure

preparations of RNA have an absorbance reading 260/280 value of 1.8 to 2.0. The

values were recorded and used to determine the volume required to equal 1µg of total

RNA. Then the 1µg of total RNA was used as a template to synthesize complimentary

DNA (cDNA) by , Promega A3500 reverse transcriptase kit

(Promega, Madison, WI). The manufacturer’s protocol was followed using the random

primer mixtures and then upon completion the final volume was adjusted from 20 µl to

50 µl.

Real Time Polymerase Chain Recation (PCR) analysis of gene expression was accomplished using the TaqMan multiplex method of gene amplification. The final concentration for the PCR reaction was 900 nM for the primers and 100 nM for the probe. The M1 probe was labeled at the 5’ end with the reporter dye 6- carboxyfluorescein (FAM) and at the 3’ end with the quencher dye 6-carboxy-

tetramethyl-rhodamine (TAM). 18S was included in each well and used to standardize

the relative concentrations of cDNA. The 18S probe was labeled at the 5’ end with 2′-

chloro-7′-phenyl-1,4-dichloro-6-carboxyfluorescein (VIC) probe and at the 3’ end with

TAM. Samples were prepared by adding 12.5 μl of TaqMan Universal PCR Master Mix

(2X) (Applied Biosystems, Branchburg, NJ) with 2.5 μl of cDNA, 2.5 μl cytokine

25 primer/probe mix, 2.5 μl 18S, ribosomal RNA (rRNA), primer/probe mix, and 5 μl sterile for a total volume of 25 μl. Table 2.1 contains a list of the sequences for primers and probes that were synthesized by Applied Biosystems, Inc. Amplification was performed on the Applied Biosystems ABI Prism 7700 Gene Amplification System (Applied

Biosystems, Foster City, CA) using a two-step process with 40 cycles of 15 second denaturing phase (95°C) and a 1 minute anneal/extension phase (60°C). Sequences for the primer and probe sets for M1 was obtained from van Elden et al. and the sets for 18S rRNA was obtained from Dr. Michael Caligiuri’s (Ohio State University Comprehensive

Cancer Center, Columbus, Ohio).

Quantitation of Gene Expression: Measurement of Real Time PCR data depends on the amount of fluorescent dye detected and the initial amount of amplicon, cDNA template, for amplification. There are excessive quantities of both the amplicon and PCR reagents at the start of the amplification. The detection of the reporter dye early in the Real Time

PCR should follow an exponential growth pattern as a result of each new PCR cycle.

Once the number of amplicons rise and the PCR reagents are exhausted, the rate of increase detection of reporter dye slows into a linear growth pattern. In the final stage the growth pattern reaches a plateau, a point where so much of the reagents have been used that no new reporter dye is detected with the next PCR cycle. Because all samples begin with the same amount of PCR reagents, the point at which they enter into the linear growth phase indicates a similar amount of consumption of the reagents, but will vary based do to the difference in amount of starting amplicon. Then we manually set a

26

threshold line to intersect the linear growth phase of each graph line and use this to

determine the cycle at which all reactions are at the same relative level of amplification.

The number of cycles that have occurred at this intersection is given the name cycle

threshold or CT. All CT for a particular dye and amplicons were determined from the

same threshold. Once all the CTs were established, there were several methods that could

be used to determine the relative gene expression. We chose to use the comparative CT

method. All the methods have their advantages and disadvantages. The CT method

begins by subtracting the internal control (18S rRNA) from the CT of the experimental

gene (M1 in this case), which results in a ΔCT value for each reaction. ΔCT values

standardize for individual differences in the starting amount of total RNA for each

reaction. The ΔCT values for each of the treatment groups are then averaged and the

averaged ΔCT for the control group is subtracted from the average ΔCT for each test

group. This calibrates datum relative to each experiment’s control group. The resulting

ΔΔCT is then power transformed using the formula 2(-ΔΔCT). This will result in a value of

one for the control group and another number for each treatment group.

Immunocytochemistry: A549 cells were grown on 15-mm coverslips and allowed to

grow for 24 hours before being washed twice with PBS. We treated and infected the cells

simultaneously and allowed the virus to attach to the cells by placing the cells at 4°C for

30 minutes. Then the cells were placed in a 37°C incubator for 30 minutes or 1 hour.

Cells were washed with either acidic PBS (PBS/HCl, 4°C) or PBS (4°C), followed by

two subsequent wash steps with PBS (4°C). Cells were then fixed for 10 minutes with

27

3.7% formaldehyde (in PBS) at room temperature. After washing, cells were

permeabilised with acetone at -20 C for 5 minutes. Coverslips were then washed with

PBS and blocked with 10% goat serum in PBS for 20 minutes at 37C. All steps beyond this point were performed at room temperature in a humidity chamber. After blocking, cells were incubated with mouse anti-influenza a (Serotec) in PBS for 30 minutes. After

washes with Tween PBS and a single wash with PBS, cells were incubated with an Alexa

fluor goat anti-mouse 488 for 1 hour. Cells were washed three times with Tween PBS and

a single wash with PBS. To visualize cell nuclei and mount cells to slide a DAPI (4',6- diamidino-2-phenylindole) fluorescent mounting medium was used courteously provided by Dr. Yang (Glaser Lab). Fluorescence was visualized using a fluorescence microscope

(Glaser Lab). Analysis of images was performed with ImagePro 6.2 (Fischer Lab).

Statistical Analysis: Statistical analyses were carried out using StatView®. One way or one factor analysis of variance (ANOVA) with M1 ΔCT as a between subjects factor was used to compare differences within studies. Follow up testing; using least significant difference (LSD) post-hoc test was used to make comparisons between treatment groups.

2.3 Results

Influenza Virus M1 gene expression in A549 cells

A549 human lung epithelial cells were treated with AED, DE (vehicle control), or

RPMI/10% FBS (media control) in combination with infection by influenza A/PR9 virus

28 to determine whether or not the steroid had any effect on replication of the virus. Gene expression of the late stage influenza M1 was used to determine the effect that AED had on influenza viral replication. First a dose curve was performed in order to determine the dosage was optimal for this study (Fig. 2.1). Analysis of these data showed that a concentration of 1 x 10-6 M AED resulted in no significant effect on M1 gene expression. However, a decrease of M1 gene expression was observed at a concentration of 5 x 10-6 M AED and was the amount used in all further studies.

M1 expression was not detected at time points earlier than 6 hours following infection (data not shown). The results at 6 hours p.i. (Fig 2.2), were variable and M1 gene expression was not always detected at this time. Thus, from the time course it was determined that the most appropriate time to further examine M1 expression was 24 hours. The results of this study showed that introducing the virus and treating the epithelial cells with AED at the same time for a period of 24 hours resulted in a significantly decrease in M1 gene expression (greater than 50%), while the vehicular control had no significant affect (Fig.2.3). It was then decided to test whether AED was able to affect viral replication after a cell had become infected. The results showed that treatment of a cell with AED that was already infected with the influenza virus did not result in a decrease in replication as shown by M1 gene expression (data not shown).

After determining that AED was able to decrease influenza viral replication if treatment was given simultaneously with infection, we decided to determine if any other nuclear receptor agonists/antagonists had a similar effect on the viral replication in the A549 cell line.

29

Comparing different agonists and antagonists of nuclear receptors

The knowledge that the AR is a ligand-dependent transcription factor led us to performed experiments to examine the effect treatment with other steroid hormone nuclear receptor agonists and antagonists. The treatments were comprised of what are mainly known as androgen agonists (AED, AET, DHEA, and DHT), androgen antagonists (CA, Flut, and HF), and glucocorticoid agonists (Cort and DEX). The androgen agonists we chose to use in this experiment were DHEA, of which AED is a metabolite; AET, another DHEA metabolite; and DHT, a well-known potent androgen.

Multiple androgen antagonists, CA, Flut, and HF, were also used in these experiments because all known antagonists of the androgen receptor are recognized to have some agonistic abilities and/or are shown to be metabolized in cells. The glucocorticoids, Cort and DEX, were used because it was previously shown that AED may act to inhibit the effects of glucocorticoids.

The only treatment that indicated a decrease in M1 gene expression was AED, while several of the other treatments appear to indicate an increased level of M1 gene expression (Fig. 2.4). In fact the glucocorticoid agonists, Cort and DEX, showed about a nine-fold increase in M1 gene expression. This large increase in M1 gene expression was not shown in all glucocorticoid agonists as cortisol was tested later and did not show an effect (data not shown). Although AED has been shown in other studies to have an estrogenic effect; estrogen receptor agonists and antagonists were not examined in this study as A549 cells have been shown to lack expression of that specific nuclear receptor.

30

Thus, these findings indicated that a clear difference between AED and the other tested agonists/antagonists of nuclear receptors. Only AED inhibited M1 gene expression.

Ability of Androstenediol to Inhibit Influenza Viral Entry

To determine whether AED prevented the influenza virus from entering the A549 cells we performed an immunocytochemistry assay. We performed many experiments

(not shown) to determine several different important details: the amount of primary and secondary antibody, the time to allow the influenza virus to enter the cell, type of wash, and blocking agent. In order to perform this experiment we had to allow the influenza virus to attach to the cells, before allowing them a certain time period (30 minutes or 1 hour) to enter. To determine the amount of influenza virus in the cells an acid wash stripped any excess virus from the surface of the cells. Then the virus remaining in the

A549s were cross linked in the cell with formaldehyde and the cells were later permeabilized with acetone to allow the antibodies specific for the influenza virus to enter the cell and bind to the virus. AED treatment did not result in a significant difference when compared to vehicular control (Fig. 2.5 and 2.6).

2.4 Discussion

The study began by determining the best time (Fig. 2.1) and best dosage (Fig. 2.2) of AED to use at the time of infection in the treatment of A549 respiratory epithelial cells. These studies showed that AED decreased gene expression of the late stage influenza viral protein M1 (Fig 2.3). The reason M1 was chosen was because its

31

expression was shown to be important for the development of a mature influenza virion

and decreased expression of this gene resulted in reduced viral budding of the influenza

virus (Chiang et al. 2008, Gomez et al. 2000, Hui et al. 2004). Bourmakina and Garcia-

Sastre have shown that a threshold amount of is essential for the assembly of

viral components into a virion and that release of the virion does not take place until a

sufficient level of M1 protein accumulates at the plasma membrane of a cell (Bourmakina

& Garcia-Sastre 2005). This is unlike other influenza viral proteins such as NA in which

levels can fluctuate with minimal effects on mature infectious virus levels. However,

when M1 protein levels are decreased there is a decrease shown in mature influenza

virions (Bourmakina & Garcia-Sastre 2005). Thus M1 gene expression was chosen as a surrogate for viral replication to study the effects of AED on A549 cells.

Our laboratory had studied AED as an androgen receptor agonist and its abilities to alter immune function both in vivo and in vitro. As discussed earlier in this chapter, other studies have shown AED and other beta-androgens have an ability to inhibit other viral and bacterial infections (Loria & Padgett 1991, Loria et al. 1996). While these studies determined that all the beta-androgens had varying abilities to modify viral or bacterial replication in vitro, we showed that the beta-androgens (DHEA and AET) did not affect influenza viral replication (Fig 2.4). Like the beta-androgens, the other treatment group that was used as an androgen agonist, DHT, did not show a significant modification of the replication of the influenza virus. Androgen antagonists (Flut and

HF) did appear to show a marginal, but insignificant increase in M1, while CA resulted in a nearly three-fold increase in M1 gene expression (Fig. 2.4). Thus adding this to the

32

AED results suggests that AR may be playing a role in influenza viral replication as

shown by M1 gene expression levels.

The only other treatments to show any modification of M1 levels was the

glucocorticoid agonists (Cort and DEX) with a significant, nearly nine-fold increase in

M1 gene expression (Fig. 2.4). Glucocorticoids have been used in treatment for influenza

viral infection for over fifty years, but more recently there have been warnings issued by

World Health Organization that prolonged use or high dosages of glucocorticoids can result in increased levels of influenza viral replication (Writing Committee of WHO,

2008). This suggests that glucocorticoids, such as dexamethasone, when given to patients who have a lower respiratory infection, may result in an increased viral replication in the respiratory tract. A question that remained was whether the virus was able to enter the cell when treated with AED.

The reason we must look into whether the virus was entering the cell is to determine whether which future experiments should be performed to determine how

AED is modifying the M1 gene replication. There are a couple of different ways found in the literature to answer this question: western blotting of a specific gene such as M1

(Eierhoff 2009) or immunofluorescence imaging with a microscope (Eierhoff 2009,

Karlas 2010). Our group decided to use immunofluorescence imaging with a microscope, which showed (Fig. 2.5 & Fig. 2.6) that AED does not appear to modify the amount of virus entering the A549s.

The experiments described above demonstrate that AED decreased M1 gene expression when AED treatment was given at the time of an influenza viral infection.

33

Remarkably, AED does not decrease M1 gene expression when it is given post-infection.

AED, also does not appear to prevent the influenza virus from entering A549 cells. This would seem to suggest that the modified cellular or influenza virion has to be treated early on before the influenza virus is able to begin to modify the cellular environment.

Even more surprising was the fact that although androgen antagonist, CA, results in a significant increase, the other androgen hormones do not appear to affect a decrease in

M1 gene expression.

34

Cytokine Sequence Size Source

M1 Forward 5’-GGACTGCAGCGTAGACGCTT-3’ 189 van Elden

Reverse 5’-CATCCTGTTGTATATGAGGCCCAT-3’

Probe: 5’-CTCAGTTATTCTGCTGGTGCACTTGCCA-3’

18S Forward 5’-CGGTACCACATCCAAGGAA-3’ 187 Caliguiri

Reverse 5’-GCTGGAATTACCGCGGCT-3’

Probe: 5’-TGCTGGCACCAGACTTGCCCTC-3

Table 2.1: Primer and TaqMan Probe Sequences of cytokines M1 and 18S for Real Time

PCR

35

Dose Curve of AED 1.2

1

Fold Change) Fold 0.8 -

0.6 * 0.4

0.2

0

M1 Gene Expression (N M1 Expression Gene CTL DE AED 5x10-6 AED 1x10-6 AED 1x10-7 AED 1X10-8 Treatments (Concentrations in Moles/Liter or M)

Fig. 2.1: Dose Curve of AED and its Effect on Influenza Virus M1 Gene Expression in A549 Lung Epithelial Cells. Treatment regimens are represented along the x-axis and the results are reported as fold increase over control cultures. A549 cells (6 x 105 cells/ml) were untreated control (RPMI/FBS), treated with DE (DMSO/ETOH), or varying concentrations (5x10-6 M, 1x10-6 M, 1x10-7 M, 1x10-8 M of AED (in DMSO/ETOH); and infected with influenza virus APR8/34 for 24 hours. M1, a viral gene upregulated late during influenza viral replication, was used to determine results. Treatments were in triplicate. Analysis of gene expression was performed by Real Time PCR. The symbol indicates a significant difference (p<0.05, LSD) between groups.

36

Influenza Time Course 4.5 4.0 3.5 3.0 2.5 2.0

(M1 Gene Expression) Gene (M1 1.5 1.0 0.5 0.0 6 Hr 6 Hr 6 Hr 12 Hr 12 Hr 12 Hr 24 Hr 24 Hr 24 Hr Infect Ctl Infect DE Infect Infect Ctl Infect DE Infect Infect Ctl Infect DE Infect Fold ChangeFold - AED AED AED N Treatments

Fig. 2.2: Time Course of AED and its Effect on Influenza Virus M1 Gene Expression in A549 Lung Epithelial Cells. Treatment regimens are represented along the x-axis and the results are reported as fold increase over control culture at the 6 hour time point. A549 cells (6 x 105 cells/ml) were untreated control (RPMI/FBS), treated with DE (DMSO/ETOH), or AED (5 x 10-6 M in DMSO/ETOH); and infected with influenza APR8/34 virus for varying periods. M1, a viral gene upregulated late during influenza viral replication, was used to assess viral replication. Treatment groups were performed in duplicate. Analysis of gene expression was performed by Real Time PCR.

37

M1 Gene Expression 1.2

1

0.8

0.6 * 0.4

0.2 Fofld Change Fofld

- 0

N CTL DE AED Treatment

Figure 2.3: Data of Influenza Virus M1 Gene Expression in A549 Lung Epithelial Cells. Treatment regimens are represented along the x-axis and the results are reported as fold-increase over control cultures. A549 cells (6 x 105 cells/ml) were untreated control (RPMI/FBS), treated with DE (DMSO/ETOH), or AED (5 x 10-6 M in DMSO/ETOH); and infected with influenza APR8/34 virus for a period of 24 hours. M1, a viral gene upregulated late during influenza viral replication, levels were decreased by nearly 60%. Analysis of gene expression was performed by Real Time PCR. There were 5 samples per treatment group. The symbol indicates a significant difference (p<0.05, LSD) between groups.

38

M1 Gene Expression 10 * * 9 8 7 6 5 4 * 3 2 Fold Difference Fold - 1 N 0 DE AED AET DHEA DHT Flut HF CA Cort DEX

Treatment

Figure 2.4: Multiple Steroid Agonists and Antagonists Effect on Influenza Virus M1 Gene Expression in A549 Lung Epithelial Cells. Treatment regimens are represented along the x-axis and the results are reported as fold increase over DMSO/Ethanol (vehicular control) cultures. A549 cells (6 x 105 cells/ml) were treated with DE (DMSO/ETOH), AED (5 x 10-6 M in DMSO/ETOH) , androstenetriol (AET, 5 x 10-6 M in DMSO/ETOH), dehydroepiandrosterone (DHEA, 5 x 10-6 M in DMSO/ETOH), dihydrotestosterone (DHT, 1 x 10-7 M in DMSO/ETOH), flutamide (Flut, 1 x 10-7 M in DMSO/ETOH), hydroxyflutamide (HF, 1 x 10-7 M in DMSO/ETOH), cyproterone acetate (CA, 1 x 10-7 M in DMSO/ETOH), corticosterone (Cort, 1 x 10-7 M in DMSO/ETOH), and dexamethasone (DEX, 1 x 10-7 M in DMSO/ETOH); and infected with influenza APR8/34 virus for a period of 24 hours. M1, a viral gene upregulated late during influenza viral replication, levels were decreased only with AED treatment. M1 levels were appeared to be increased by the three androgen antagonists (approximately two to three fold) and the two glucocorticoids (Cort and DEX) show a nearly nine fold increase in expression. Analysis of gene expression was performed by Real Time PCR. There were 6 samples per treatment group. The symbol indicates a significant difference (p<0.05, LSD) between groups.

39

Figure 2.5: The above images were used to determine whether AED had an effect on influenza viral entry. A549s were grown for 24 hrs on cover slips and infected with 10 HAUs of influenza APR/8. Treatment occurred at the same time as infection. Primary antibody (serotec; mouse anti influenza a) used in the experiment was at a concentration of 1:100 and secondary (Alexa Fluor 488 goat anti-mouse) at a concentration of 1:100. In the images above the cells were visualized with a Leica fluorescent microscope using a L5 filter at 40X. Analyses of images were performed using ImagePro 6.2.

40

Figure 2.6: The above images were used to determine whether AED had an effect on influenza viral entry. A549s were grown for 24 hrs on cover slips and infected with 40 HAUs of influenza APR/8. Treatment occurred at the same time as infection. Primary antibody (serotec; mouse anti influenza a) used in the experiment was at a concentration of 1:100 and secondary (Alexa Fluor 488 goat anti-mouse) at a concentration of 1:100. In the images above the cells were visualized with a Leica fluorescent microscope using a L5 filter at 40X. Analyses of images were performed using ImagePro 6.2.

41

Chapter 3: Androstenediol’s Effect on Cellular Signaling and a Coalescing of Pattern Recognition Receptors

3.1 Introduction

The goal in the previous chapter was to determine whether androstenediol (AED)

was able to inhibit viral replication at a cellular level: results showed that the influenza

virus matrix protein 1 gene level decreased upon AED treatment. In addition, previous

studies by our lab show that AED contributed to an earlier upregulation of the immune

response to the influenza virus and was also likely clearing the virus from the lungs much

quicker than the control mice. In addition to the general immunological finding,

published studies showed that AED treated mice were able to augment expression of

cytokines of both the innate (i.e., interleukin-1, [IL-1] and tumor necrosis factor alpha,

[TNFα]) and adaptive (i.e., interleukin-2, [IL-2] and interferon gamma, [IFN-γ]) immune responses (Padgett et al. 2000b). In a previous study, AED augmented the expression of

IL-1, IL-2, and TNFα by preventing the glucocorticoid-mediated suppression of these cytokines (Padgett et al. 2000). Thus, the results from the previous chapter showed that

AED treatment limited replication of the virus, next we will determine the mechanism by which replication is limited at the cellular level. The previous studies along with the knowledge that other androgenic hormones tested on epithelial cells have shown regulation of at least 500 genes (DePrimo et al. 2002), led us to examine whether any major pathways or specific genes were modified by AED in epithelial cells. To test

42

these pathways we decided to again look into respiratory epithelial cells, since they are

the primary target for influenza virus and typically show elevated production of cytokines

in response to influenza viral infection.

Although often shown to produce only a limited number of cytokines, respiratory

epithelial cells are one of the major components in the innate immune response to

influenza viral infection. This response all begins with the recognition of the influenza

virus by the epithelial cell, which then leads to an attempt to prevent replication of the

virus and expression of warning signals to other cells. Respiratory epithelial cells have

intracellular sensors for viral products that once activated begin a signaling cascade that

collectively culminates in the expression of the interferon beta (IFN β) gene (Le Goffic et

al. 2007). It has been shown that retinoic acid-inducible gene-I (RIG-I) is the essential

RIG-I-like receptors (RLR) in early cellular recognition of the influenza virus although

other pattern recognition receptors (PRRs) have been shown to play a role as well

(Rehwinkel & Reis 2010). After recognizing the influenza virus, RIG-I signaling

initiates the antiviral response through the interferon beta promoter and pro-inflammatory signaling through nuclear factor kappa B (NFκB) (Rehwinkel & Reis 2010). All of the known PRRs that have been shown to recognize viruses, including the influenza virus, initiate signaling pathways that coalesce at the activation of interferon regulatory factor-3

(IRF3), interferon regulatory factor-7 (IRF7), and/or NFκB which leads to transcriptional induction of the IFN β gene (Kawai & Akira 2007, Kawai & Akira 2008, Lee & Kim

2007, Takeuchi & Akira 2010).

43

The effects of IFN β occur through the binding of IFN β to a cellular surface

receptor, interferon-α/β receptor (IFNAR) on both the infected cell and neighboring cells

(Biron 2001, Darnell et al. 1994, Uze et al. 2007). Binding and activation of IFNAR leads to expression of numerous IFN-stimulated genes which act to prevent replication of the influenza virus. These pathways are of such importance that the influenza virus has a

specific gene, nonstructural protein (NS) 1, whose primary function is to bind to and

directly inhibit RIG-I, thus preventing the recognition of the influenza virus and

transcription of IFN β (Gack et al. 2009, Guo et al. 2007, Haye et al. 2009 ). As has been

shown, activation of IFN β is associated with the prevention of influenza viral replication.

Therefore, we decided to test the hypothesis, that AED modified virus-induced signaling

mechanisms and augmented the immediate early antiviral immune response in the

infected cell, by examining IFN β gene expression. Furthermore, we hypothesized that

AED would likely alter viral modifications in cellular cytokines or pathways similar to

results that were augmented in the animal experiments (Padgett et al. 2000b).

3.2 Methods

Cell Culture: The A549 (CCL-185™) human epithelial cell line was cultured in

Rosswell Park Memorial Institute 1640 (RPMI 1640) media (Invitrogen, Carlsbad, CA)

adjusted to 0.01 Molar (M) 4-(2-hydoxyethyl)-1-piperazineethanesulfuonic acid

(HEPES), 10-5 M β-mercaptoethanol (β-Me), 0.075% sodium bicarbonate, 1.5mM L-

44 glutamine, 50 Units/ml (U/ml) penicillin, and 50 µg/ml streptomycin. The RPMI culture media was supplemented with 10% fetal bovine serum (FBS) (Biocell Laboratories, Inc.,

Rancho Dominguez, CA) and cultures were grown in a 37°C incubator in the presence of

5% CO2.

Experimental Stimulation and Infection: Approximately one day prior to experimental stimulation, culture supernatants were removed and refreshed with RPMI 1640 supplemented with 10% FBS. The A549 cellular monolayer was then trypsinized with

GIBCO® Trypsin (Invitrogen), collected in conical tubes, and centrifuged. The media was then removed and freshly made RPMI 1640 was added to resuspend the cell pellet.

A sample of this suspension was diluted with trypan blue stain and the cell number was calculated using a counting chamber. Cells were then plated in single culture dishes at approximately 6 x105 cells/ml in RPMI 1640 supplemented with 10% FBS and allowed to incubate overnight at 37°C incubator in the presence of 5% CO2. AED (Sigma

Aldrich, St. Louis, MO) was prepared fresh before each experiment at a concentration of

5 x 10-3 M solution in a solution of dimethylsulfoxide (DMSO) and ethanol (ETOH) referred to from herein as DMSO/ETOH, also used as the vehicular control. Final concentration of DMSO/ETOH in culture media was 0.1% by volume. In all experiments, media was removed and replaced with fresh media containing 10% FBS.

Samples that did not receive AED were treated with 1 µl/ml of DMSO/ETOH or culture media lacking any other solution, as a media control. Concurrently, treatment groups’

45 media had added 1 hemagglutinating unit (HAU) of Influenza A/PR8/34 (H1N1) virus per ml of culture media.

Focused Microarrays: A GEArrayTM Q Series Kit (Frederick, MD) was used to obtain gene expression in up to 96 cDNA fragments from genes associated with a specific biological pathway. The kit is commercially available and includes RNA preparation; probe synthesis, hybridization, and chemiluminescent detection. Imaging was taken with

GENEGnomeTM a product commercially available through Syngene (Frederick, MD), while GEArray AnalyzerTM performed acquisition and analysis. Membranes were spotted with negative controls (pUC18, etc.) and housekeeping genes (β-actin, GAPDH, etc.). All these products are commercially available through SABiosciences a Qiagen

Company (Frederick, MD).

RNA Isolation: Total RNA isolation was performed after cells were incubated for 4 hr or

24hr by using TRIzol ReagentTM (Invitrogen) according to the manufacturer’s protocol as described in chapter two or Qiagen (Valencia, CA) RNeasy MiniKit. Total RNA concentrations were determined by spectrophotometric readings at 260 and 280 nm wavelengths.

Real Time Polymerase Chain Reaction (PCR) analysis of gene expression was accomplished using the TaqMan multiplex method of gene amplification. The final concentration for the PCR reaction was 900 nM for the primers and 100 nM for the probe. The IFN-beta probe was labeled at the 5’ end with the reporter dye 6- 46 carboxyfluorescein (FAM) and at the 3’ end with the quencher dye 6-carboxy- tetramethyl-rhodamine (TAM). 18S was included in each well and used to standardize the relative concentrations of cDNA. The 18S probe was labeled as shown before.

Samples were prepared by adding 12.5 μl of TaqMan Universal PCR Master Mix (2X)

(Applied Biosystems, Branchburg, NJ) with 2.5 μl of cDNA, 2.5 μl cytokine primer/probe mix, 2.5 μl 18S, ribosomal RNA (rRNA), primer/probe mix, and 5 μl sterile for a total volume of 25 μl. Amplification was performed on the Applied Biosystems ABI

Prism 7700 Gene Amplification System (Applied Biosystems, Foster City, CA) as described previously. Sequences for the primer and probe sets for human IFN-beta was designed using Primer Express software (Applied Biosystems) and the sets for 18S rRNA was obtained from Dr. Michael Caligiuri’s laboratory.

Focused PCR Arrays: A Human Signal Transduction PathwayFinder PCR Array

(Frederick, MD) was used to obtain gene expression in up to 84 key genes representative of 18 different signal transduction pathways. The RT² SYBR® Green qPCR Master

Mixes contained all of the reagents and buffers required for SYBR® Green based real- time polymerase chain reactions in real-time PCR. Each master mix included real-time

PCR buffer, a high-performance HotStart DNA Taq polymerase, nucleotides, and

ROX®. Added the master mix to PCR tubes along with the template and primers were in the microarray plates. Microarray plates have negative controls (pUC18, etc.) and housekeeping genes (β-actin, GAPDH, etc.). All these products are commercially available through SABiosciences a Qiagen Company (Frederick, MD).

47

Quantitation of Gene Expression: Quantitation was described before with the only two

modifications being the experimental gene being tested and sample size. The CT method

begins by subtracting the internal control (18S rRNA) from the CT of the experimental

gene (IFN-beta in this case), which results in a ΔCT value for each reaction.

Statistical Analysis: Statistical analyses were carried out using StatView®. One way or

one factor analysis of variance (ANOVA) with M1 ΔCT as a between subjects factor was

used to compare differences with studies. Follow up testing, using least significant

difference (LSD) post-hoc test to indicate any significance. Data analysis was performed

on the PCR Array by an integrated web-based software package for the PCR Array

System automatically performs all ΔΔCT based fold-change calculations from the uploaded raw threshold cycle data.

3.3 Results

Interferon Beta Gene Expression in A549 Cells Without Influenza Viral Infection

In the previous experiment our data showed that AED decreased viral replication as shown by M1 gene expression. Thus we now wanted to determine if this decreased expression occurred through the innate immune response of the epithelial cell, thus PRRs.

The gene expression of IFNβ was used as a surrogate to determine the effect that AED had on PRRs, as all known PRRs signaling influenza virus infection are shown to

48 coalesce at the induction of IFN β gene expression. A549 human lung epithelial cells were treated with AED or DE (vehicle control) to determine whether or not the steroid had any effect on IFN β gene expression. This study which allowed the uninfected cells to be treated with AED for 24 hours showed that AED did not significantly alter the gene expression of IFN β as compared to the vehicular control, thus showing no effect (Fig.

3.1).

Focused Microarray in A549 Cells Without Influenza Viral Infection

Knowing that the AED did not affect IFN β on its own without infection we decided to see if it had any effect on the A549 cells in the absence of influenza viral infection. Thus, we again used A549 cells that were treated with AED or DE to determine whether or not the steroid alone had any effect on multiple cellular pathways.

This study allowed the cells to be treated with AED for 4 hours and showed that AED did not significantly alter the expression of any of the 84 genes that were tested, thus again showing no effect of AED alone (Fig. 3.2).

Interferon Beta Gene Expression in A549 Cells With Influenza Viral Infection

Having determined that AED treatment did not have any effect on multiple genes without stimuli, we decided to study IFN beta gene expression with influenza virus infection. A549 human lung epithelial cells were treated with AED or DE to determine

49

whether or not the steroid had any effect on IFN beta gene expression during an infection

with the influenza virus. The study which allowed the cells to be treated with AED and infected with influenza virus for 24 hours showed that AED did not significantly increase gene expression of IFN β. In fact the infected cells that were treated with AED had a significant decrease in the amount of IFN β gene expression (Fig. 3.3).

PCR Array of A549 Cells With Influenza Viral Infection

Since AED treatment did not affect IFN β gene expression when the cells were infected by influenza virus, we decided to examine whether there was any noticeable effect of AED treatment of influenza viral infection on the multiple cellular pathways.

The study which allowed the epithelial cells to be treated with AED and infected with the influenza virus for 24 hours, showed that AED did significantly increase expression of

22 genes involved in 14 pathways by at least 2.5-fold (Table 3.1). Two major pathways,

NFκB and NFAT (Nuclear Factor of Activated T-cells), showed multiple genes being increased along with a gene or two with expression levels increased by 9-fold or greater

(Fig. 3.4). Also two genes that had previously been shown to have been augmented in the mouse model increased significantly in this array were TNF and IL-2 (Fig. 3.4). Thus these findings indicate that the AED-mediated effect may result from an ability to prevent the suppression of gene induction by the influenza virus.

50

3.4 Discussion

These studies showed that AED was unable to significantly modify gene

expression of those genes on the PCR array without influenza viral infection, including

that of IFN β gene expression. The effect of AED did not appear to be occurring through

pattern recognition receptors as IFN β gene expression did not show an increase during

influenza viral infection. In fact, it appears to show a significant decrease in IFN β gene

expression. This led us to examine whether AED could be having an effect on other

pathways of the epithelial cell that may result in decreased viral replication.

Thus the AED treated samples that were concurrently infected with influenza

virus indicated several pathways that may be involved in the inhibition of the viral replication and also showed potential for increased adaptive immune signaling. IL-2,

which has been shown to be weakly expressed in A549s (Hurteau et al. 2007), was shown

to have nearly a ten-fold increase in expression when treated with AED in the presence of

influenza virus infection. Thus this is similar to the AED augmented expression of IL-2

shown in the previous animal model (Padgett et al 2000b). The binding of IL-2, a factor

in both the NFκB and NFAT pathways, to the IL-2 receptor results in the differentiation,

growth, and survival in specific T cells (Beadling and Smith 2002, Chow et al. 1999,

Iwashima et. al 2002, Lindemann et al. 2003). Thus this is unlikely to be contributing to

the decreased influenza viral replication at the cellular level for respiratory epithelial cells

in this in vitro model. But there were other genes that were significantly increased in the

array when treated with AED in the presence of influenza viral infection that were also

surrogates of the NFκB and NFAT pathway.

51

Some of the NFκB pathway surrogates shown to be increased in this array such as nitric oxide synthase 2 (NOS2) and tumor necrosis factor (TNF), have been determined to lead to an increase in nitric oxide (NO) (de Vera et al. 1996, Vila-del Sol et al. 2007).

Although NOS2 and TNF expression are detrimental to the lungs, NO has been shown to be an effective inhibitor of the influenza viral replication (Rimmelzwaan et al. 1999). An increase in NO has been previously shown to correlate with the inhibition of viral RNA synthesis and has been suggested to affect an early step in the replication cycle of the influenza virus (Rimmelzwaan et al. 1999). The other significantly increased pathway,

NFAT pathway, had surrogates increased in this array such as fas ligand and cluster of differentiation 5. These surrogates have been shown to require an accumulation of NFAT for increased expression of their genes, thus suggesting that NFAT has increased expression as well. NFAT factor, NF90, has recently been shown to have a direct effect on influenza viral replication (Wang P et al. 2009). It has been suggested that NF90 decreases early viral replication by targeting the influenza virus nucleoprotein function in the nucleus (Wang P et al. 2009). Thus the increased expression in surrogates of NFAT pathway may lead to a mechanism by which AED contributes to the decreased viral gene expression. Our data, along with data from previous studies (Rimmelzwaan et al. 1999 and Wang P et al. 2009) would suggest that a further examination of NO and NF90 would be beneficial to confirm whether one or both of these products play an essential role in the decreased influenza viral expression we have observed with AED treatment.

Thus, this leads us to question how AED treatment results in the modification of these signaling pathways, and what role the androgen receptor has in these effects.

52

Figure 3.1: Data of Interferon Beta Gene Expression in A549 Lung Epithelial Cells. Treatment regimens are represented along the x-axis and the results are reported as fold increase over control cultures. A549 cells (6 x 105 cells/ml) were treated with DE (DMSO/ETOH) and AED (5 x 10-6 M in DMSO/ETOH) for a period of 24 hours. IFN- beta is a cellular gene that coalesces all the pattern recognition receptors at a single point. IFN-beta levels showed no significant change in cells that were uninfected. Treatment groups were performed in triplicate. Analysis of gene expression was performed by Real Time PCR.

53

Figure 3.2: Focused microarray analysis of AED activated genes in A549 cell culture without exposure to influenza virus. A549 cells (6x105cells/ml) were treated with DE (DMSO/ETOH) and AED (5 x 10-6 M in DMSO/ETOH). The microarray has 84 key genes representative of 18 different signal transduction pathways. Up-regulated genes are those with a greater than 2-fold change and down regulated genes are those that are at a 50% decrease. There were also genes on the microarray such as GAPDH to be used as controls. There was no significant change seen between the two treatments.

54

IFN ß Gene Expression 1.2

1

0.8

0.6 *

0.4

Fold Difference Fold 0.2 -

N 0 DE AED Treatment

Figure 3.3: Data of Interferon Beta Gene Expression in A549 Lung Epithelial Cells With Exposure to Influenza Virus. Treatment regimens are represented along the x- axis and the results are reported as fold increase over control cultures. A549 cells (6 x 105 cells/ml) were treated with DE (DMSO/ETOH) and AED (5 x 10-6 M in DMSO/ETOH) and infected with influenza virus for a period of 24 hours. IFN-beta is a cellular gene that coalesce all the pattern recognition receptors at a single point. IFN-beta levels have appeared to show a significant decrease in cells that were infected. Analysis of gene expression was performed by Real Time PCR. Treatment groups were performed in triplicate. The symbol indicates a significant difference (p<0.05, LSD) between groups.

55

Comparing AED gene expression to vehicle control(DMSO:EtOH) Fold Change Genes Pathways Genes Involved In: 11.4716 p450 enzyme aromatase CREB Pathway 10.8528 cluster of differentiation NFAT Pathway 9.6465 interleukin 2 NFkB Pathway NFAT Pathway Calcium and Protein Kinase C Pathway 8.2249 lymphotoxin alpha Insulin Pathway 7.5685 homeobox protein engrailed 1 Hedgehog Pathway and Retinoic Acid Pathway 6.8211 colony stimulating factor 2 Calcium and Protein Kinase C Pathway and LDL Pathway 6.774 kallikrein-related peptidase 2 Androgen Pathway 6.4531 E-selectin LDL Pathway 6.2333 Proto-oncogene protein Wnt-1 Hedgehog Pathway 6.0629 nitric oxide synthase 2 NFkB Pathway Jak-Stat Pathway Phospholipase C Pathway 6.021 matrix , stromelysin 2 Jak-Stat Pathway 5.9794 fas ligand NFAT Pathway 5.9381 leptin Insulin Pathway 4.5631 Proto-oncogene protein Wnt-2 Hedgehog Pathway 4.2281 platelet/endothelial cell adhesion molecule-1 NFkB Pathway 4.0278 Wnt-1 induced secreted protein 1 Wnt Pathway 3.5064 hexokinase 2 Insulin Pathway 3.1167 selectin P ligand LDL Pathway 3.0525 vascular cell adhesion molecule-1 NFkB Pathway Phospholipase C Pathway LDL Pathway 2.9897 interleukin 4 Jak-Stat Pathway 2.9282 cyclin-dependent kinase inhibitor 2B TGF-ß Pathway 2.9079 tumor necrosis factor NFkB Pathway

Table 3.1: PCR array analysis of AED activated genes in A549 cell culture with exposure to influenza virus. A549 cells (6x105cells/ml) were treated with DE (DMSO/ETOH) and AED (5 x 10-6 M in DMSO/ETOH) with infection of influenza virus for a period of 24 hours. The PCR array has 84 key genes representative of 18 different signal transduction pathways. Up-regulated genes are those with a greater than 2.5-fold change and down regulated genes are those that are at a 67% decrease. There were also genes in the PCR array such as GAPDH to be used as controls. There were 22 genes involved in 14 pathways that were upregulated in the AED treated PCR array as compared to the DMSO:EtOH.

56

Chapter 4: Determining Whether Androstenediol Modifies Influenza Virus-Encoded M1 Gene Expression Through the Androgen Receptor

4.1 Introduction

Nuclear receptors (NRs) vary in their ability to bind hormones and in their functions once hormones are bound to the intracellular receptor. The primary function of a nuclear receptor is to activate genes by acting as a ligand-activated transcription factor

(Robinson-Rechavi et al. 2003), although another postulated function of nuclear receptors is to act through non-genomic signaling (Foradori et al. 2008). Characterized by their conserved DNA-binding domains, NRs are divided into at least four classes by dimerization patterns and response elements (Denayer et al. 2010, Germain et al. 2006,

Glass 1994). Two of these classes contain known receptors for the ligand, androstenediol

(AED): androgen receptor (AR) and estrogen receptor. Interestingly, the lung epithelial cell line (A549) used in our studies has been shown to lack expression of the estrogen receptor. However, the AR has not only been shown to be expressed in the A549 cell line (Provost et al. 2000), but the AR has previously been demonstrated in our lab to bind to a promoter complex in another cell line upon treatment of AED (Farrow 2007). Thus, our studies focused on the androgen receptor as our lab has previously investigated and has seen effects of AED through the activation of the androgen receptor.

Thus if AED is acting through the AR in A549s it could be effecting the virus by modifying cellular genes or physically interacting with other proteins that may have an effect on influenza viral replication. The modifying of cellular genes or transcriptional 57

effect of androgens have been shown to be the result of a direct binding of the androgen

receptor dimer to an androgen response elements on genes (Wong et al. 1993). In fact a recent study suggested that nearly two hundred genes have been shown to be modified in

A549s upon treatment with testosterone (Mikkonen et al. 2010). Another way AED may be resulting in the modification of influenza viral replication would be through AED bound AR physically interacting with proteins that have been shown to interact with subunits of the influenza virus. In fact two cellular proteins (Ebp-1 and Hsp90) that have

been shown to interact with subunits of the influenza virus polymerase have also been

shown to interact with the AR (Honda 2008, Honda et al. 2007, Zhang et al. 2002).

Ebp1, which is known to physically interact with the AR, has been shown to inhibit RNA

synthesis of the influenza virus by binding to PB1 (Honda 2008, Honda et al. 2007,

Zhang et al. 2002). HSP90, is the other protein known to interact with the AR and bind

to the influenza viral protein PB2 (Momose et al. 2002). Other interactions with the AR

have been shown as well, including the effects of AR in cells treated with AED.

In fact previous studies from our laboratory have shown that AED-induced activation of the AR is able to alter the effect of the activated glucocorticoid receptor

(GR). This affect has been shown to modify the effects of activated glucocorticoid transcription activation, gene expression patterns during wound healing, and the ability to ameliorate the suppressive effects on the gene expression of TNF-α (Farow 2007, Head et al. 2006). The activated-GR has been shown to antagonize pro-inflammatory gene expression. The mechanism by which activated-GR suppresses pro-inflammatory signaling pathways is through the interaction/inhibition of the function of both AP1 and

58

NFκB (De Bosscher et al. 2000, De Bosscher et al. 2003). This information combined

with our previous observations that indicated an increase in expression of multiple genes

of the NFκB pathway upon AED-treatment of influenza virus infected cells (Table 3.1) and an increase in M1 gene expression (Fig. 2.3) upon treatment with glucocorticoids led us to our hypothesis. We hypothesized that the activated GR effects on influenza viral replication would be abrogated with treatment of AED. Specifically, we hypothesized that the AED-treatment observations of decreased influenza viral replication would be the result of the AED-bound androgen receptor.

4.2 Methods

Cell Culture: The A549 (CCL-185™) human epithelial cell line was cultured in Roswell

Park Memorial Institute 1640 (RPMI 1640) media (Invitrogen, Carlsbad, CA) adjusted to

0.01 Molar (M) 4-(2-hydoxyethyl)-1-piperazineethanesulfuonic acid (HEPES), 10-5 M β- mercaptoethanol (β-Me), 0.075% sodium bicarbonate, 1.5mM L-glutamine, 50 Units/ml

(U/ml) penicillin, and 50 µg/ml streptomycin. The RPMI culture media was supplemented with 10% fetal bovine serum (FBS) (Biocell Laboratories, Inc., Rancho

Dominguez, CA) and cultures were grown in a 37°C incubator in the presence of 5%

CO2.

Experimental Stimulation and Infection: Approximately one day prior to experimental stimulation, culture supernatants were removed and refreshed with RPMI 1640 supplemented with 10% FBS. The A549 cellular monolayer was then trypsinized with

59

GIBCO® Trypsin (Invitrogen), collected in conical tubes, and centrifuged. The media

was then removed and freshly made RPMI 1640 was added to resuspend the cell pellet.

A sample of this suspension was diluted with trypan blue stain and the cell number was

calculated using a counting chamber. Cells were then plated in single culture dishes at

approximately 6 x105 cells/ml in RPMI 1640 supplemented with 10% FBS and allowed

to incubate overnight at 37°C incubator in the presence of 5% CO2. AED (Sigma

Aldrich, St. Louis, MO), dihydrotestosterone (DHT) (Sigma Aldrich), cyproterone

acetate (CA) (Sigma Aldrich), dexamethasone (DEX) (Sigma Aldrich), and R1881

(Methyltrienolone) (Sima Aldrich) was prepared fresh before each experiment at a

concentration of 5 x 10-3 M – 1 x 10-4 M solution in a solution of dimethylsulfoxide

(DMSO) and ethanol (ETOH) referred to from herein as DMSO/ETOH, also used as the

vehicular control. Final concentration of DMSO/ETOH in culture media was 0.1% by

volume. In all experiments, media was removed and replaced with fresh media containing 10% FBS. Treatment groups were added to stock media preparations at a concentration of 1 µl/ml of 5 x 10-3 M – 1 x 10-4 M solution in a solution of

DMSO/ETOH. Samples that did not receive AED were treated with 1 µl/ml of

DMSO/ETOH or culture media lacking any other solution, as a media control.

Concurrently, treatment groups’ media had 1 hemagglutinating unit (HAU) of Influenza

Virus A/PR8/34 added per ml of culture media.

RNA Isolation: Total RNA was isolation was performed by using TRIzol ReagentTM

(Invitrogen) according to the manufacturer’s protocol. The TRIzol method accomplishes

60

this by using a phenol/chloroform based phase separation protocol and is followed by an

isopropanol-induced RNA precipitation. RNA precipitates were then washed by adding 1 ml of 75% ETOH. After the RNA wash, the 75% ETOH was decanted and the RNA pellet was resuspended in a range of 20-50 µl of distilled water (dH2O), depending on pellet size. Total RNA concentrations and purity were determined by a spectrophotometer. The RNA method was selected on the spectrophotometer to analyze at 260 nm and 280 nm wavelengths. Concentration of RNA should be determined by measuring the absorbance at 260 nm in a spectrophotometer using quartz cuvettes. Pure

preparations of RNA have an absorbance reading 260/280 value of 1.8 to 2.0. To make

sure the readings are significant the readings of the 260 nm should be between 0.15 and

1.0. The numbers were recorded and placed in Microsoft® Excel® file to determine the

volume required to equal 1µg of total RNA. Then the 1µg of total was used as a template

to synthesize complimentary DNA (cDNA) by reverse transcriptase, Promega A3500

reverse transcriptase kit (Promega, Madison, WI). The manufacturer’s protocol was

followed using the random primer mixtures and then upon completion the final volume

was adjusted from 20 µl to 50 µl.

Real Time Polymerase Chain Reaction (PCR) analysis of gene expression was

accomplished using the TaqMan multiplex method of gene amplification. The final

concentration for the PCR reaction was 900 nM for the primers and 100 nM for the

probe. The M1 probe was labeled at the 5’ end with the reporter dye 6-

carboxyfluorescein (FAM) and at the 3’ end with the quencher dye 6-carboxy-

tetramethyl-rhodamine (TAM). 18S was included in each well and used to standardize

61 the relative concentrations of cDNA. The 18S probe was labeled at the 5’ end with 2′- chloro-7′-phenyl-1,4-dichloro-6-carboxyfluorescein (VIC) probe and at the 3’ end with

TAM. Samples were prepared by adding 12.5 μl of TaqMan Universal PCR Master Mix

(2X) (Applied Biosystems, Branchburg, NJ) with 2.5 μl of cDNA, 2.5 μl cytokine primer/probe mix, 2.5 μl 18S, ribosomal RNA (rRNA), primer/probe mix, and 5 μl sterile for a total volume of 25 μl. Table 2.1 contains a list of the sequences for primers and probes that were synthesized by Applied Biosystems, Inc. Amplification was performed on the Applied Biosystems ABI Prism 7700 Gene Amplification System (Applied

Biosystems, Foster City, CA) using a two-step process with 40 cycles of 15 second denaturing phase (95°C) and a 1 minute anneal/extension phase (60°C). Sequences for the primer and probe sets for M1 was obtained from van Elden et al. and the sets for 18S rRNA was obtained from Dr. Michael Caligiuri’s laboratory.

Quantitation of Gene Expression: Measurement of Real Time PCR data depends on the amount of fluorescent dye detection and the initial amount of amplicon, cDNA template, for amplification. There are excessive quantities of both the amplicon and PCR reagents at the start of the amplification. The detection of the reporter dye early in the Real Time

PCR should follow an exponential growth pattern as a result of each new PCR cycle.

Once the number of amplicons rise and the PCR reagents are exhausted, the rate of increase detection of reporter dye slows into a linear growth pattern. In the final stage the growth pattern reaches a plateau, a point where so much of the reagents have been used that no new reporter dye is detected with the next PCR cycle. Because all samples begin

62

with the same amount of PCR reagents, the point at which they enter into the linear

growth phase indicates a similar amount of consumption of the reagents, but will vary

based do to the difference in amount of starting amplicon. Then we manually set a

threshold line to intersect the linear growth phase of each graph line and use this to

determine the cycle at which all reactions are at the same relative level of amplification.

The number of cycles that have occurred at this intersection is given the name cycle

threshold or CT. All CT for a particular dye and amplicons are determined from the same

threshold. Once all the CTs are established, there are several methods that could be used

to determine the relative gene expression, but our lab has chosen to use the comparative

CT method. All the methods have their advantages and disadvantages. The CT method

begins by subtracting the internal control (18S rRNA) from the CT of the experimental

gene (M1 in this case), which results in a ΔCT value for each reaction. ΔCT values

standardize for individual differences in the starting amount of total RNA for each

reaction. The ΔCT values for each of the treatment groups are then averaged and the

averaged ΔCT for the control group is subtracted from the average ΔCT for each test

group. This calibrates datum relative to each experiment’s control group. The resulting

ΔΔCT is then power transformed using the formula 2(-ΔΔCT). This will result in a value of

one for the control group and another number for each treatment group.

Transient transfection assay:

COS7 African green monkey kidney fibroblast-like cell line suitable for transfection was obtained from Dr. Jeanette Marketon’s laboratory. The COS7 cells were seeded at 4x105

63

cells/well (24-well plate) in DMEM supplemented with 10% FBS charcoal-stripped.

Cells were incubated at 37°C in 5% CO2 for 24 hr prior to transfection. Effectene

Transfection Reagent (Qiagen) was used to transfect cells with a firefly luciferase reporter plasmid, 160 ng of pGL3-mouse mammary tumor virus (MMTV) or pGL3 control and 20 ng of the expression vectors for AR ( pARo ) vectors, using 20 ng of pRL-

TK (renilla luciferase reporter plasmid control) for normalization. The AR expression

plasmid was provided by Dr. Tony Hollenberg’s laboratory. Fresh medium with multiple

concentrations of AED (Sigma), DHT (Sigma), or vehicle (0.1% DMSO:EtOH final

concentration) in fresh media were added to the monolayers of cells for an additional 24

h. Firefly and Renilla luciferase activities were assayed with a Dual-Luciferase™

Reporter Assay (Promega). Firefly luciferase activities in were read in triplicate samples

and were normalized for Renilla activities. Luminescent signals were read with a Packard

microplate luminometer.

siRNA:

AR siRNA or negative control siRNA was transfected into A549 cells using

DharmaFECT 2 (Thermo Fisher Scientific) according to the manufacturer's instructions;

48 h after transfection treatment and infection of cells were performed as listed above.

Expression of M1 gene expression was determined by real-time PCR. The efficiency of

siRNA silencing was determined by Western blotting.

64

Total cellular protein was isolated using protein extraction reagent (Thermo Fisher

Scientific). Varying concentrations of protein was subjected to SDS-PAGE on NuPAGE

Novex 4–12% bis-Tris precast gels (Invitrogen). Proteins were then transferred to

nitrocellulose membranes by semidry blotting and blocked in 5% nonfat dry milk in Tris-

buffered saline with 0.05% Tween 20 (TBST). All blocking and antibody steps were

performed while rocking. Membranes were then incubated with AR antibody (sc-815), specific for the N20 region of the AR, in 5% milk in TBST overnight at 4 C. Membranes were then washed three times with TBST and then incubated with goat antirabbit IgG-

horseradish peroxidase (HRP) (sc-2004; Santa Cruz) in TBST for 1 hour at room temperature. Membranes were again washed with TBST and chemiluminescence detected using SuperSignal West Pico Chemiluminscence Substrate (Thermo Fisher Scientific) according to the manufacturer's instructions and exposed to autoradiographic film.

Statistical Analysis: Statistical analyses were carried out using StatView®. One way or one factor analysis of variance (ANOVA) with M1 ΔCT as a between subjects factor was used to compare differences with studies. Follow up testing, using least significant difference (LSD) post-hoc test to indicate any significance. Data analysis was performed

on the PCR Array by an integrated web-based software package for the PCR Array

System automatically performs all ΔΔCT based fold-change calculations from the uploaded raw threshold cycle data.

65

4.3 Results

Effect of the Methyltrienolone (R1881) on Influenza Viral Replication

Knowing that AED was able to both activate genes such as those involved in the

NFκB pathway and suppress replication of the influenza virus as suggested by M1 gene expression, we wanted to test whether other androgens acted in a similar manner. Thus we wanted to more thoroughly test a dose curve of a reference androgen, R1881, to determine if it had an effect similar to AED. The gene expression of M1 was used to determine the effect that R1881 had on the influenza viral replication. This study allowed cells to be treated with multiple concentrations of R1881 for 24 hours. The dose curve showed even at the highest concentration of 5 x 10-6 M that R1881 did not significantly increase or decrease M1 gene expression (Fig. 4.1). This further demonstrated what had been seen in other experiments in our lab that AED, like other androgens, has unique treatment effects.

M1 Expression in A549 Cells with AED and the Glucocorticoid DEX

In the prior experiments, the datum indicated that DEX had the ability to significantly increase M1 gene expression nearly nine fold (Fig. 4.2). We also showed that AED increased expression of multiple cytokines in influenza virus infected cells.

Thus we now wanted to determine if this increased expression of these cytokines may abrogate the effect of the activated-glucocorticoid receptor (GR) in the epithelial cells, as

66 has been a noted ability of AED previously in our lab (Farrow 2007). The gene expression of M1 was used to determine the effect of AED on the activation of GR by

DEX. A549 human lung epithelial cells were treated with AED, DE (vehicle control),

DEX, or AED and DEX combined to determine whether or not AED had any effect on the increased M1 expression seen with the treatment of DEX. This study which allowed the cells to be treated with DEX and AED simultaneously for 24 hours showed that DEX was able to overwhelm the decrease normally seen in gene expression of M1 as compared to AED treatment alone (Fig. 4.2). Although DEX was able to show both a dramatic increase in influenza viral replication and an ability to overwhelm AED’s inhibition of replication another glucocorticoid, cortisol did not have either of these effects (data not shown).

Androstenediols Ability to Activate the Androgen Receptor

To test whether AED caused the androgen receptor to act directly on androgen responsive genes we performed an androgen receptor transactivation assay. In order to perform this experiment we added androgen receptors to cells lacking the expression of the androgen receptor (COS7) via a transient transfection assay to examine multiple levels of AED and DHT. For the androgen receptor transactivation assay, an androgen- dependent reporter gene was added to test a dose response curve for both AED and DHT to determine if the androgen receptor appeared to be acting on the androgen reporter gene. We saw that from a range of 1x10-8 M to 1x10-5 M that DHT did appear to bind

67 and activate the expression of the androgen receptor (Fig. 4.3), but treatment with AED did not appear to have much if only a limited effect at a1x10-5 M concentration.

Determination of the Effect of an Antagonist of the Androgen Receptor

Agonists of the androgen receptor did not affect influenza viral replication as treatment with AED did; we choose to test whether a known antagonist of the androgen receptor would be able to abrogate the suggested decrease in influenza viral replication that had been shown by AED. Cyproterone acetate (CA) is an generally shown as an AR antagonist and was shown to block the effects of AED, although our lab has also shown that AED, has acted as an antagonist of CA in certain situations such as the ability to annul the decreased expression of a specific gene, I kappa B alpha (IκBα) that had been demonstrated by CA. IκBα is an inhibitor of the transcriptional factor NFκB and is used as a surrogate to show increased activation of NFκB. Thus we tested CA to determine if it was acting as an antagonist as had been shown to function in some of our labs experiments with AED. If this did occur it should have led to an abrogation of the decrease of M1 gene expression seen in A549 cells treated with AED. The results we saw were that CA at 1x10-6 M and 5x10-6 M did not result in any abolition of the AED decrease on M1 gene expression (Fig. 4.4). CA at a concentration of 1x10-7 M did indicate a marginal, yet insignificant abolition of the AED effect (Fig. 4.4). This concentration of CA was also the only one that indicated an insignificant, but greater than two fold increase in M1 gene expression when the antagonist alone was added with influenza virus (Fig. 4.4).

68

Androstenediols Ability to Affect Influenza Viral Replication Upon Treatment with

Androgen Receptor siRNA

To determine whether AED was acting through the androgen receptor to affect influenza viral replication we used AR siRNA to knock down the androgen receptor. In order to perform this experiment we added AR siRNA at varying concentrations. The androgen receptor appeared to be knocked down at all concentrations of AR siRNA used

(Fig. 4.5) and thus we chose the lowest concentration of 50 ngs to go forward. There was no significant difference between AR siRNA and control siRNA M1 real time results

(Fig. 4.6). The significant difference on M1 gene expression between DE-treated and

AED-treated cells was not abrogated (Fig. 4.6). Thus treatment with AR siRNA did not appear to have any effect on the significant decrease of M1 gene expression caused by

AED-treatment.

4.4 Discussion

The data in this chapter focuses on the androgen receptor as a mechanism for

AED-mediated suppression of influenza virus replication. We began by using a reference androgen, R1881, to determine whether it showed similar abilities as AED. Figure 4.1 suggests that R1881 is unable to result in any decrease in influenza viral replication and is not acting the under the same mechanisms as AED. Thus this result further indicated that AED appeared to be acting uniquely compared to other agonists of the androgen receptor (Fig. 4.1 and Fig. 2.4), as this had been shown to be the case in our lab before.

69

We then decided to further investigate whether the effect of AED could be similar to that

which was seen in another cell line (Farrow 2007).

We next choose to use a GR agonist, dexamethasone (DEX), as it is known to be

immunosuppressive, anti-inflammatory, and interfering with NFκB activation in A549s

(Jang et al. 2007). It was interesting to see that DEX was able to overwhelm the decrease

in influenza viral replication normally seen with AED treatment (Fig 4.2). This could be

suggesting that an interaction between AR, GR, and NFκB is occurring at a site of

transcription, as many papers dealing with interaction between GR and NFκB

demonstrate that activated GR inhibits transcription at NFκB sites (Barnes 2006, Kleinert et al. 1996, Mukaida et al. 1994, Ohtsuka et al., 1996). This is not always the case as some have shown that the activated GR has the ability to interact directly with NFκB

(Widen et al. 2003) or the NFκB pathway in the cytosol (De Bosscher et al. 2003). This

suggests it is likely that some point prior to the activation of cellular genes these

interactions are taking place. Although another surprising result was that cortisol did not

have any effect on M1 gene expression levels, on its own or with AED treatment. This

not only leads to questions remaining about how DEX is causing its effects, but also

brings us back to the question surrounding how AED is causing its effects on influenza

viral replication. Thus we then decided to investigate whether the effect of AED could be

having its effects directly through transcription activation of the AR. Another reason to

investigate this possible mechanism is the focused pathway array indicated that treatment

with AED during influenza viral infection was resulting in increased transcription of multiple genes.

70

We then designed an experiment to examine whether AED is able to induce transcriptional activation of the AR. Thus the AR was transfected into a cell line, COS7, that lacks the AR and a reporter gene assay was performed. The androgen receptor was transfected transiently and the datum indicated that AED suggested a marginal, but insignificant 50% increase in the transcriptional activator of this androgen receptor at the highest concentration (Fig. 4.3). Comparing this to the increase of nearly 500% in transcriptional activity shown by DHT suggested that AED was at best a potentially weak activator of this androgen receptor. Thus this data suggests that AED did not appear to be acting significantly through the AR in this specific reporter gene assay, but this reporter gene assay was optimized to test specific androgens such as DHT. These results were not inconsistent with a similar experiment that had been performed with a stable transfection of an androgen receptor and treatment with another beta androstene, DHEA, which indicated that DHEA was shown to weakly activate transcriptional activity as compared to DHT (Roy et al. 2006). Thus these experiments did not lead to any conclusive results on how AED may be inhibiting influenza viral replication, but should lead to further investigation on whether AED may act through the androgen receptor transcriptionally. So we next decided to use an antagonist to inhibit the effects of AED through the AR.

To examine whether we could restore the level of influenza viral replication, we attempted to inhibit AED activation of the androgen receptor by treating the cells with varying concentrations of an androgen receptor antagonist, cyproterone acetate (CA).

The dose curve of CA again indicated that the antagonist of the androgen receptor had a

71 marginal increase in influenza viral replication as determined by the expression of the influenza viral gene M1 at a concentration of 1x10-7 M (Fig. 4.4). The higher concentrations of CA did not appear to indicate any significant modification of influenza viral replication (Fig. 4.4). The only concentration of CA that may suggest that it is acting marginally yet insignificantly as an androgen receptor antagonist to AED was at

1x10-7 M (Fig. 4.4). Thus as has been shown in our lab before, AED may be the one that is suppressing the effect of cyproterone acetate (Farrow 2007). Further investigation was needed to determine if AED is able to act on the androgen receptor to facilitate this decrease seen in influenza viral gene M1.

Thus we decided to try to knock down the androgen receptor with siRNA and attempted to inhibit AED activation of the androgen receptor by transfecting the cells with varying concentrations of an androgen receptor siRNA. The AR siRNA indicated that AED was not acting through the AR, unlike what was previously noted in shown in other cells in our laboratory. Thus AED must be acting through another mechanism or receptor. We then decided to determine whether AED was preventing the influenza virus from entering the cell. To determine this we performed immunocytochemistry, as has been performed in the literature. Our findings indicated that this likely was not the mechanism by which AED was decreasing viral replication.

Thus the only conclusions we were able to demonstrate is AEDs effect on influenza viral replication is not occurring through the androgen receptor and DEX is able to overwhelm the effects of AED. Later experiments also concluded that AED is not causing a decrease in the amount of influenza virus entering A549 cells. Thus further

72 investigation will be required to elucidate the interaction of AED in relation to the decrease replication of the influenza A virus in A549 cells.

73

R1881 Affect on M1 Expression

1.6

1.4

1.2

1

0.8

0.6

0.4

0.2

0

Fold Difference RFUs (M1 RNA expression; ddCt) CTL DE AED 5 R1881 6 R1881 7 R1881 - N Treatment

Figure 4.1: Data of Influenza Virus M1 Gene Expression in A549 Lung Epithelial Cells Treated With a Dose Curve of Methyltrienolone (R1881). Treatment regimens are represented along the x-axis and the results are reported as fold increase over control cultures. A549 cells (6 x 105 cells/ml) were untreated control (RPMI/FBS), treated with DE (DMSO/ETOH), AED (5 x 10-6 M in DMSO/ETOH), or varying levels of methyltrienolone (R1881); and infected with influenza virus APR8/34 for a period of 24 hours. AED treatment appeared to show a decrease, while none of the R1881 doses appeared to result in any decrease. Treatments were performed in triplicate. The symbol indicates a significant difference (p<0.05, LSD) between groups and vehicle control.

74

M1 Gene Expression 12

10

8

6 Difference Fold -

N 4

2

0 CTL DE AED DE/DEX AED/DEX Treatments

Fig. 4.2: AED, DEX, and AED/DEX Effect on Influenza Virus M1 Gene Expression in A549 Lung Epithelial Cells. Treatment regimens are represented along the x-axis and the results are reported as fold increase over control cultures. A549 cells (6 x 105 cells/ml) were treated with DE (DMSO/ETOH), AED (5 x 10-6 M in DMSO/ETOH) dexamethasone (DEX, 1 x 10-7 M in DMSO/ETOH), and AED co-treated with DEX; and infected with influenza virus APR8/34 for a period of 24 hours. M1, a viral gene upregulate late during influenza viral replication, levels were decreased with AED treatment, while the DEX showed a nine fold increase in expression. The AED along with DEX treatment showed that DEX was able to inhibit and overwhelm the effects of AED on M1 gene expression. Treatments were in triplicate. Analysis of gene expression was performed by Real Time PCR. The symbol indicates a significant difference (p<0.05, LSD) between group(s) and vehicle control.

75

Androgen Receptor Transcriptional Activation Assay

0.007 0.006 0.005 0.004 AED 0.003 DHT

RFU Expression 0.002 0.001 0 0 0.001 0.01 0.1 1 10

Treatment in micromolar

Figure 4.3: Data of Androgen Receptor Transient Transfection Assay With a Dose Curve Comparing AED and DHT. Treatment concentrations are represented along the x-axis and the results are reported as fold increase over control cultures. COS7 cells (5 x 104 cells/ml) were transfected for ~20 hours and then treated with DHT (dihydrotestosterone) and AED (androstenediol) for a period of 24 hours. Cells were then lysed and expression of the androgen receptor was determined by a dual luciferase assay. As is shown in the graph above DHT showed a nearly six fold increases at high levels, while AED only was able to produce a slightly less than two fold increase. Analysis of luminescence was performed by a microplate luminometer. The symbol indicates a significant difference (p<0.05, LSD) between group(s) and vehicle control.

76

Dose Curve of Cyproterone Acetate Effect on M1

Expression with AED Treatment 4 3.5 3 2.5 2 1.5 1 0.5 RFUs (M1 RNA expression;ddCT) RFUs (M1 RNA 0 CTL DE AED 5x10- 5x10- 1x10- 1x10- 1x10- 1x10- 6M CA 6M 6M CA 6M 7M CA 7M CA/AED CA/AED CA/AED

Treatments

Figure 4.4: Data of Influenza Virus M1 Gene Expression in A549 Lung Epithelial Cells Treated With a Dose Curve of Cyproterone Acetate in an Inhibition Assay of Androgen Receptor. Treatment regimens are represented along the x-axis and the results are reported as fold increase over control cultures. A549 cells (6 x 105 cells/ml) were untreated control (RPMI/FBS), treated with DE (DMSO/ETOH), treated with AED (5 x 10-6 M in DMSO/ETOH), treated with varying concentrations of cyproterone acetate (CA), or cotreated with varying concentrations of cyproterone acetate and a constant concentration of 5 x 10-6 M AED (CA/AED); and infected with influenza virus APR8/34 for a period of 24 hours. Treatment of CA was given two hours prior to treatment with AED and infection of influenza virus. M1 levels were decreased with AED treatment, while cyproterone acetate did not appear to decrease M1 levels on its own and did not restore M1 expression when treatment was given prior to AED treatment and influenza viral infection. Treatments were performed in quadruplicate. Analysis of gene expression was performed by Real Time PCR. The symbol indicates a significant difference (p<0.05, LSD) between groups and vehicle control.

77

Figure 4.5: Both blots used varying concentrations of AR siRNA. The image on the left of the figure is using a primary antibody (sc-816) specific for the C19 region of the AR, while the blot on the right used primary antibody (sc-815) specific for the N20 region of the AR. The primary antibody using the C19 region appears degraded as it does not detect AR at the 110 kda weight, while the primary antibody using the N20 region does indicate in the untreated cells that the AR is present at 110 kda, while the siRNA concentrations suggests almost a complete knockdown of the AR in the A549 cell line.

78

siRNA (control and androgen receptor) Effect on M1 Expression with AED Treatment

4

3.5

3

2.5

2

1.5 *

RFUs (M1 RNARFUs ddC (M1 expression; 1

0.5 *

0 CTL DE AED con/CTL con/DE con/AED AR/CTL AR/DE AR/AED

Treatments

Fig 4.6: Data of Influenza Virus M1 Gene Expression in A549 Lung Epithelial Cells Treated With either control siRNA or androgen receptor. Treatment regimens are represented along the x-axis and the results are reported as fold increase over control cultures. AR (androgen receptor) sirna or control sirna was placed on the A549 cells for 48 hours at a concentration of 50 nM. A549 cells were then left untreated control (RPMI), treated with vehicle DE (DMSO/ETOH), or AED (androstenediol at 5 x 10-6 M in DMSO/ETOH); and infected with influenza virus APR8/34 for a period of 24 hours. M1, a viral gene upregulated late during influenza viral replication was used in the analysis of gene expression performed by Real Time PCR. There were triplicate samples per treatment group. Analysis of gene expression was performed by Real Time PCR. The * symbol indicates a significant difference (p<0.05, LSD) between group(s) and vehicle control.

79

Chapter 5: General Discussion

The purpose of the study was to determine if and how androstenediol (AED) had

an effect on influenza viral replication in human lung epithelial cells. These studies were

designed to test the hypothesis that AED did not act directly on production of viral

proteins in the cell, as do most current antivirals and vaccines, but would modify the

cellular environment in which the virus replicates. Multiple cellular proteins have been

shown to be vital to both the influenza virus’s ability to replicate and the cell’s ability to hinder viral replication. Following influenza viral entry into the lung epithelial cell the virus has been shown to interact with thousands of proteins and multiple pathways.

Studies have shown that modification of a specific cellular protein may lead to an impairment of replication of the influenza virus, because the virus requires cellular proteins to facilitate its replication (Konig et al. 2010). As such, regulating a cellular pathway that interacts with the influenza virus is a reasonable approach to inhibiting viral replication. An example of this approach was seen when the Raf/MEK/ERK mitogenic kinase pathway was inhibited, viral production was impaired (Ludwig et al. 2004,

Olshlager et al. 2004, Pleschka et. al 2001). The reason this approach works to limit viral replication is that the virus is unable to replace the lost function of the cellular protein or pathway that is being modified.

80

Our initial studies were performed to determine whether AED would have a

similar reduction in influenza viral infection at the cellular level, as had been shown in

the animal models. To investigate influenza viral replication we used a type II lung

epithelial cell line, A549s, which is known as the primary cell type for influenza’s viral

infection. Then for analysis we chose to use a surrogate for influenza viral replication,

M1 gene expression, to determine if there was any effect. M1 gene expression was

chosen as it has been shown to be critical for the development of a mature influenza virus

and decreased expression of this gene results in reduced viral budding of the influenza

virus (Bourmakina & Garcia-Sastre 2005, Chiang et al. 2008, Gomez et al. 2000, Hui et

al. 2004). Other influenza virion proteins such as NA can increase or decrease and not

affect the amount of mature virion, but when M1 protein levels decrease there is also a

decrease in mature influenza virions (Bourmakina & Garcia-Sastre 2005). Our results suggested that AED was able to significantly decrease influenza viral replication as indicated by reduced M1 gene expression (Fig. 2.1, 2.2, 2.3). However, the same effect did not occur with several other hormones able to bind the androgen receptor or glucocorticoid receptor (Fig. 2.4). In fact some of the hormones even appeared to increase influenza viral replication. Although M1 gene expression was shown to decrease, this affect was only seen with AED treatment prior to influenza viral infection or with simultaneous treatment and infection. Thus, cells that were infected and subsequently treated with AED did not result in a decrease of M1 gene expression.

Further investigations were used to determine whether this effect was the result of AED

81 modifying the cellular environment thus resulting in the impairment of influenza viral replication.

A primary objective of these studies was to determine if AED treatment was able to modify cellular pathways that the influenza virus uses in viral replication or combat the effects of the influenza virus on innate viral defense. One potential mechanism that could modify the cellular environment is the increased activation of the pattern recognition receptors (PRRs) in the epithelial cell. The main purpose for increased activation of

PRRs in epithelial cells in vitro is to signal neighboring cells and allow them to develop resistance to infection. The two most well-known groups of PRRs for viral gene products are the Toll-like receptors (TLRs) and the RIG-I-like receptors (RLRs). At the time that the research began, RLRs were just being discovered and most research described cells as recognizing the influenza virus through TLRs. However, more recently it has been shown that RLRs and more specifically RIG-I and TLR3 are the PRRs that recognize the existence of the influenza virus in epithelial cells (Le Goffic et al. 2007, Tissari et al.

2005). As the literature was beginning to shift we decided the best way to study the recognition of the pathogen by PRRs would be to study the point in the pathways where the data from the literature coalesced, the induction of the interferon beta (IFN β) gene.

The IFN β gene is shown to be activated by Nuclear Factor-KappaB (NFκB) or interferon regulatory factor 3 (IRF3) in the PRR pathways (Seth et al. 2006). To investigate the potential of AED to regulate IFN β gene expression, we used real time

PCR examining the expression of the IFN β gene (Fig. 3.1). The results of these studies suggested that IFN β gene expression did not appear to be upregulated in A549 cells with

82

AED treatment alone. Even when infected with the influenza virus and treated with AED we did not see an increase in IFN β gene expression (Fig. 3.3). The datum indicated that there was a significant decrease observed in IFN β gene expression that was likely the result of decreased influenza viral replication. We therefore did not think that this pathway was the likely mechanism of the AED-regulated decrease in M1 gene expression. Thus, our next experiments tested whether AED may be modifying other major pathways in the cell that have been shown to play a role in the ability of the influenza virus to replicate itself.

The pathways that the influenza virus interacts with comprise “a significantly greater number of human proteins than expected … even when compared to other viruses” (Shapira et al. 2009). In fact, there are as many as 1745 human genes that have been shown to be important for a portion of the influenza virus life cycle, if not for direct replication of the influenza virus (Shapira et al. 2009). This means that approximately six percent of all human protein coding genes have the potential to play a role in the replicative life cycle of the influenza virus. Two prior studies investigating cellular interactions with the influenza virus used the A549 cell line and showed that approximately 300 genes were important for the influenza virus life cycle (Karlas et al.

2010, Konig et al. 2010, Stertz & Shaw 2011). Many of proteins that interact with the influenza virus, specifically influenza virus A/PR8, fall into six major pathways that include the protein 53 pathway, promyelocytic leukemia pathway, tumor necrosis factor receptor/CD95 Fas-mediated (TNFR/Fas-mediated apoptotic pathway), NFκB, and

WNT/ß-catenin pathway (Shapira et al. 2009). Thus, this led us to our hypothesis that the 83

effect we were seeing in AED-treatment of an influenza virus infected cells was likely the

result of a modification of the cellular environment.

Our experimental design examined the gene expression of 84 genes from multiple

pathways using a commercial pathway array. In our first experiment, we used uninfected

cells treated with AED and found no effect on gene expression (Fig. 3.2). This was not

surprising as AED has previously not shown the ability to have independent effects on transcription, but had been shown to act as a co-factor that affected transcription. We then used influenza virus infected cells treated with AED and found 22 genes involved in 14 different pathways that were upregulated as compared to influenza virus infected cells that were treated with the vehicle control (Table 3.1). In this experiment, two of the six major pathways, that was upregulated when treated with AED and infected as compared to control, was found to interact with proteins of the influenza virus, NFκB and

TNFR/Fas-mediated apoptotic pathway. The NFκB and TNFR/Fas-mediated apoptotic

pathways are two of the pathways that have also been shown to physically interact with

two polymerase viral subunits PB1 and PB2, and with NP of the influenza virus (Shapira

et al. 2009). This is of great significance as it has been shown that the lack of a

functional PB1 or PB2 might interfere with the viral polymerase (Wasilenko et al. 2008).

In fact a study has shown that epidermal receptor tyrosine kinase binding protein, a

protein that also interacts with the androgen receptor, binds PB1 and interferes with the

viral replication (Honda et al. 2007, Zhang et al. 2002). Other evidence suggests that the

influenza virus attempts to indirectly regulate the expression of genes in the NFκB and

TNFR/Fas-mediated apoptotic pathways (Shapira et al. 2009). Thus, our primary 84

conclusion was that the decreased influenza viral replication seen in AED-treated influenza infected cells was likely the result of the modification of cellular pathways.

Other genes and pathways we found upregulated, also led to greater understanding of how AED is functioning in influenza virus infected cells.

Activation of the NFAT pathway was unexpected. Little of the current literature describes the interactions among the NFAT pathway, epithelial cells, and influenza viral infection. Actually when a PUBMED search was performed with NFAT and one of the other two terms (epithelial cells or influenza viral infection) a total of just three papers were found. Although a recent study looking for interactions of cellular pathways with influenza viral proteins indicated activation of the NFAT pathway during infection, it suggested that further studies were necessary (Shapira et al. 2009). Two of the other genes shown to be upregulated in the current study were the p450 enzyme aromatase and kallikrein-related peptidase 2 (Table 3.1), and both genes are known to be regulated by androgens. Thus AED’s effects on these genes, along with the previous studies in our lab show that AED could have been acting through the androgen receptor in the A549 cell.

To understand the mechanism of the AED treatment effect on viral replication, we

turned to study of the androgen receptor, even though AED had also been shown to

interact with the estrogen receptor (ER) (Kuiper et al. 1997). Our lab ruled out the

estrogen receptor as having a possible role in the AED-mediated changes seen in the

A549 cells, as the ER was not shown to be expressed in the A549 cells (Croxtall et al.

1994). Previous literature also pointed out that the AR has been shown to directly

interact with proteins that play a direct role in the early events of influenza virus

85

replication (Honda 2008, Honda et al. 2007, Zhang et al. 2002). We wanted to further

investigate whether AED was acting differently compared to other androgens. The first

experiment was to determine if a known reference androgen, R1881, had an effect similar to AED on influenza viral replication in A549 cells. The results indicated that R1881 did not have an effect on the influenza virus as determined by M1 gene expression (Fig. 4.1).

Thus, R1881 acted differently from AED and did not result in any modification of M1 gene expression. This was not unexpected as the side chains of different receptor ligands alter the domains of the steroid receptor, thus resulting in differing interactions with various cellular proteins. Even previous studies had shown that AED acted differently than other androgens in the RAW264.7 cell line (Farrow 2007).

Previously, data in RAW264.7 cell line indicated that AED was able to inhibit

some of the effects of dexamethasone (DEX), a known glucocorticoid receptor agonist

(Farrow 2007). Thus we decided to determine whether AED may also be able to inhibit

the increased M1 gene expression seen with DEX treatment in A549 cells (Fig. 2.4).

When we co-treated A549 cells with AED and DEX, datum indicated that AED did not

significantly affect influenza viral replication when compared to treatment with DEX

alone (Fig. 4.2). This indicated that DEX treatment was able to overwhelm the decrease

in influenza viral replication shown with AED treatment. Normally glucocorticoids, such

as DEX, are given to patients to suppress inflammation. However, in cell lines such as

A549 DEX was shown by Greenberg and colleagues to suppress cellular proliferation

through at least two signaling pathways: ERK/MAPK pathway and the cell cycle regulators (Greenberg et al. 2002). Another study using the A549 cell line suggested that

86

DEX treatment was able to upregulate nearly 1500 genes and downregulate nearly 1200 genes (Mikkonen et al. 2010). Thus, further experiments should be performed to understand how DEX is overwhelming the effects of AED on influenza viral replication.

But looking at how AED acts transcriptionally could lead to further understanding of its effects.

We then decided to assess the ability of AED to act directly on gene transcription through the androgen receptor. To test the transcriptional activity of AED, COS7 cells transfected with an AR expression vector, and an Effectene Transfection Reagent

(Qiagen) was used to transfect the cells with a firefly luciferase reporter plasmid.

Activity was measured over a range of AED concentrations and was compared with a known agonist of the AR, DHT. DHT showed a significant increase in activity of nearly

500% of the reporter plasmid at concentrations as low as 0.1 µM (Fig. 4.3). AED treatment led to a moderate, but not significant increase, of approximately 50% at the highest concentration (Fig. 4.3). Thus AED did not appear to be acting through the AR in this specific reporter gene assay. This would not rule out AED acting transcriptionally through an androgen receptor as this reporter gene assay was optimized to test specific androgens such as DHT. The abilities of AED to drive transcription needed further examination, as we have shown other androgens such as DHT also differ from AED in their ability to affect influenza viral replication.

We then chose to use an AR antagonist to attempt to inhibit AEDs effects on influenza viral replication. The reason we chose to use cyproterone acetate (CA) as an antagonist of androgens was because our prior datum (Fig. 2.4) indicated that influenza

87

virus infected A549 cells had a significant increase in M1 gene expression upon

treatment with CA. Thus it appeared to suggest that this AR antagonist resulted in

increased replication of the influenza virus. Treatment with CA alone at 1x10-7 M again

appeared to have a moderate, but not significant, three-fold increase in M1 gene expression, but this effect was not seen at other concentrations (Fig. 4.4). When treated with AED and varying concentrations of CA, the only concentration of CA that appeared to have an ability to counteract the inhibition of M1 gene expression was 1x10-7 M (Fig.

4.4). This led to the final AR experiment to determine if AED’s effects were through this nuclear receptor in A549s. To determine if the AR in A549s played a role in the decreased M1 gene expression the AR was knocked down with AR siRNA. The AR siRNA, at a concentration as low as 50 nM, in this experiment was able to knockdown the AR, as verified by western blot (Fig. 4.5). Knockdown of AR with AR siRNA prior to treatment with AED and infection with the influenza virus did not modify the significant decrease in viral replication between AED treatment and vehicle control (Fig.

4.6). Thus the culmination of all the AR experiments indicates that AED does not act through the AR in A549s.

Unlike what was previously shown in Raw264.7 cells AED is not acting through the androgen receptor in A549s, thus another experiment was needed to investigate whether AED was modifying the entrance of the influenza virus into A549s. The experiment chosen to investigate viral entrance was an immunocytochemical assay, at one hour post-infection. The immunocytochemical assay used is a method found in the literature to determine influenza viral entrance, but if a further experiment would be

88 needed for validation a western blot assay has also been shown in the literature, as well.

The data collected from the immunocytochemical assay suggested that AED did not significantly decrease viral entrance into the A549 cells (Fig. 2.5 and Fig. 2.6). Thus the culmination of research both through literature and experimentally led us to determine that AED does not act does not act through the androgen receptor or on the cellular entrance of influenza virus.

In conclusion, our data has shown that AED treatment has the ability to reduce replication of an influenza virus in an epithelial cell line. These data also suggest that

AED inhibits viral replication by modifying the cellular environment. This finding was supported by the activation of multiple cellular pathways in influenza virus infected cells treated with AED. The data has ruled out several potential cellular mechanisms that

AED could have been acting through, as suggested in the literature and previous experiments in our lab. The reporter gene assay and siRNA experiment indicated that

AED does not act through the AR, thus ruling out the possibility of AED-bound AR acting directly with proteins of the influenza virus. AED also does not appear to act on, and did not enhance the PRRs, involved in innate antiviral defense of the lung epithelial cells. Most importantly though, AED treatment did have the ability to suppress influenza viral replication in A549 cells and appears to achieve this effect by modifying the cellular environment. The data suggests a potential mechanism of how AED is inhibiting viral replication of the influenza virus. As our data indicated, the gene that was most upregulated of any on the gene array during influenza viral infection was a p450 gene, an aromatase. This aromatase is activated to help metabolize drugs, such as AED and is

89 known to be transcriptionally regulated by the constitutive androstane receptor (CAR)

(Wada, T et al 2009). CAR is usually found in the cytosol or nucleus, but in influenza virus infected cells it relocated to the cellular surface during the replication phase of the viral process (Takahashi et al 2008). This nuclear receptor plays a role during influenza viral infection and binds to steroid hormones, such as AED. This would be a viable option for future experiments of determining the mechanism of influenza viral replication suppression found to be the result of AED treatment.

These findings are significant because AED is inhibiting viral replication and doing so without directly interacting on the highly variable proteins of the influenza virus, such as hemagglutinin and neuraminidase. By modifying the cellular pathways such as metabolism and others as indicated in the gene array, this allows AED treatment to circumvent one of the major strengths that the influenza virus uses to combat all current treatments, which is its ability to mutate itself. This is not only significant because of the AED findings showing the decrease in influenza viral replication, but also shows that AED does not act through the AR in all types of cells. The most significant finding of these results is that this could lead to treatment(s) that circumvent one of the major problems with current therapy, the rapidly evolving proteins of the influenza virus.

This treatment is novel and could lead to future treatments that also act on the initial site of infection of the influenza virus – the human lung epithelial cell.

90

References

Akira S, Uematsu S, Takeuchi, O. Pathogen recognition and innate immunity. Cell 124: 783–801, 2006.

Bain DL, Heneghan AF, Connaghan-Jones KD, Miura MT. Nuclear receptor structure: implications for function. Annu Rev Physiol. 69: 201-220, 2007.

Barnes PJ. Corticosteroid effects on cell signalling. Eur Respir J. 27(2): 413-426, 2006.

Baron S, Manin M, Beaudoin C, Leotoing L, Communal Y, Veyssiere G, Morel L. Androgen Receptor Mediates Non-genomic Activation of Phosphatidylinositol 3-OH Kinase in Androgen-sensitive Epithelial Cells. J Biol Chem. 279(15): 14579-14586, 2004.

Beadling C, Smith KA. DNA array analysis of interleukin-2-regulated immediate/early genes. Med Immunol. 1(1): 2, 2002.

Bergmann M, Garcia-Sastre A, Carnero E, Pehamberger H, Wolff K, Palese P, Muster T. Influenza Virus NS1 Protein Counteracts PKR-Mediated Inhibition of Replication. Journal of 74(13): 6203-6206, 2000.

Biron CA. Interferons [alpha] and [beta] as immune regulators—a new look. Immunity 14: 661–664, 2001.

Black B, Paschal B, Intracellular organization and function of the androgen receptor. TRENDS in Endocrinology and Metabolism. 15 (9): 411-417, 2004.

Bolton EC, So AY, Chaivorapol C, Haqq CM, Li H, Yamamoto KR. Cell- and gene- specific regulation of primary target genes by the androgen receptor. Genes Dev. 21(16): 2005-2017, 2007.

Boltz DA, Aldridge JR Jr, Webster RG, Govorkova EA. Drugs in development for influenza. Drugs 70(11): 1349-62, 2010.

Bourmakina SV, Garcia-Sastre A. The morphology and composition of influenza A virus particles are not affected by low levels of M1 and M2 proteins in infected cells. J Virol. 79(12): 7926-7932, 2005.

91

Bright RA, Shay DK, Shu B, Cox NJ, Klimov AI. Adamantane resistance among influenza A viruses isolated early during the 2005-2006 influenza season in the United States. JAMA. 295(8): 891-894, 2006.

CDC. Estimates of Deaths Associated with Seasonal Influenza --- United

States, 1976-2007. MMWR 59(33);1057-1062, 2010.

Cheng A, Wong SM, Yuan YA. Structural basis for dsRNA recognition by NS1 protein of influenza A virus. Cell Research 19(2): 187-195, 2009.

Chiang C, Chen GW, Shih SR. Mutations at alternative 5' splice sites of M1 mRNA negatively affect influenza A virus viability and growth rate. J Virol. 82(21): 10873- 10886, 2008.

Chow CW, Rincón M, Davis RJ. Requirement for transcription factor NFAT in interleukin-2 expression. Mol Cell Biol. 19(3): 2300-2307, 1999.

Claessens F, Alen P, Devos A, Peeters B, Verhoeven G, Rombauts W. The Androgen- specific Probasin Response Element 2 Interacts Differentially with Androgen and Glucocorticoid Receptors. J Biol Chem. 271: 19013-19016, 1996.

Clemens MJ, Elia A. The double-stranded RNA-dependent protein kinase PKR: structure and function. J Interferon Cytokine Res. 17(9): 503-524, 1997.

Clifford M, Twigg J, Upton C. Evidence for a novel gene associated with human influenza A viruses. Virology 6: 198, 2009.

Coombs KM, Berard A, Xu W, Krokhin O, Meng X, Cortens JP, Kobasa D, Wilkins J, Brown EG. Quantitative Proteomic Analyses of Influenza virus-Infected Cultured Human Lung Cells. J Virol. 84(20): 10888-10906, 2010.

Cox, RJ, Brokstad, KA, Ogra, P. Influenza Virus: Immunity and Vaccination Strategies. Comparison of the Immune Response to Inactivated and Live, Attenuated Influenza Vaccines. Scandinavian Journal of Immunology 59: 1–15, 2004.

Croxtall JD, Emmas C, White JO, Choudhary Q, Flower RJ. Tamoxifen inhibits growth of oestrogen receptor-negative A549 cells. Biochem Pharmacol. 47(2): 197-202, 1994.

Darnell JE Jr., Kerr IM, Stark GR. Jak-STAT pathways and transcriptional activation in response to IFNs and other extracellular signaling proteins. Science 264: 1415–1421, 1994.

92

Denayer S, Helsen C, Thorrez L, Haelens A, Claessens F. The rules of DNA recognition by the androgen receptor. Mol Endocrinol. 24(5): 898-913, 2010.

Deng T, Engelhardt OG, Thomas B, Akoulitchev AV, Brownlee GG, Fodor E. Role of Ran Binding Protein 5 in Nuclear Import and Assembly of the Influenza Virus RNA Polymerase Complex. J Virol 80(24): 11911-11919, 2006.

DePrimo SE, Diehn M, Nelson JB, Reiter RE, Matese J, Fero M, Tibshirani R, Brown PO, Brooks JD. Transcriptional programs activated by exposure of human prostate cancer cells to androgen. Genome Biol. 3(7): RESEARCH0032, 2002.

De Bosscher K, Vanden Berghe W, Haegeman G. Mechanisms of anti-inflammatory action and of immunosuppression by glucocorticoids: negative interference of activated glucocorticoid receptor with transcription factors. J Neuroimmunol. 1;109(1): 16-22, 2000.

De Bosscher K, Vanden Berghe W, Haegeman G. The interplay between the glucocorticoid receptor and nuclear factor-kappaB or activator protein-1: molecular mechanisms for gene repression. Endocr Rev. 24(4): 488-522, 2003. de Vera ME, Shapiro RA, Nussler AK, Mudgett JS, Simmons RL, Morris SM Jr, Billiar TR, Geller DA. Transcriptional regulation of human inducible nitric oxide synthase (NOS2) gene by cytokines: initial analysis of the human NOS2 promoter. Proc Natl Acad Sci U S A. 6;93(3): 1054-1059, 1996.

Durbin JE, Hackenmiller R, Simon MC, Levy DE. Targeted disruption of the mouse Stat1 gene results in compromised innate immunity to . Cell 84(3): 443-450, 1996.

Eierhoff T, Ludwig S, Ehrhardt C. The influenza A virus matrix protein as a marker to monitor initial virus internalisation. Biol Chem. 390(5-6): 509-15, 2009.

Ehrhardt C, Wolff T, Pleschka S, Planz O, Beermann W, Bode JG, Schmolke M, Ludwig S. Influenza A virus NS1 protein activates the PI3K/Akt pathway to mediate antiapoptotic signaling responses. J Virol. 81(7): 3058-3067, 2007.

Engelhardt OG, Fodor E. Functional association between viral and cellular transcription during influenza virus infection. Review Medical Virology 16(5): 329-345, 2006.

Evans RM. The steroid and thyroid hormone receptor superfamily. Science 240(4854): 889-895, 1988.

Fahrenkrog B, Koser J, Aebi U. The nuclear pore complex: a jack of all trades? Trends Biochem Sci 29(4): 175-182, 2004. 93

Farrow MJ. The Effect of Androstenediol on gene expression and NF-κB activation in vitro (Doctoral dissertation). Ohio State University, Columbus, OH. 2007.

Foradori CD, Weiser MJ, Handa RJ. Non-genomic actions of androgens. Front Neuroendocrinol. 29(2): 169-181, 2008.

Foster KA, Oster CG, Mayer MM, Avery ML, Audus KL. Characterization of the A549 cell line as a type II pulmonary epithelial cell model for drug metabolism. Exp Cell Res. 243(2): 359-366, 1998.

Francis T Jr, Magill TP. The antibody response of human subjects vaccinated with the virus of human influenza. J Exp Med. 65: 251-259, 1937.

Gack MU, Albrecht RA, Urano T, Inn KS, Huang IC, Carnero E, Farzan M, Inoue S, Jung JU, García-Sastre A. Influenza A virus NS1 targets the ubiquitin ligase TRIM25 to evade recognition by the host viral RNA sensor RIG-I. Cell Host Microbe. 5(5): 439-449, 2009.

Germain P, Staels B, Dacquet C, Spedding M, Laudet V. Overview of nomenclature of nuclear receptors. Pharmacol Rev. 58(4): 685-704, 2006.

Gern JE, French DA, Grindle KA, Brockman-Schneider RA, Konno S, Busse WW. Double-stranded RNA induces the synthesis of specific chemokines by bronchial epithelial cells. Am J Respir Cell Mol Biol. 28(6): 731-737, 2003.

Ghendon Y. Introduction to pandemic influenza through history. European Journal of Epidemiology 10: 451-453, 1994.

Giguere V, Yang N, Segui P, Evans RM. Identification of a new class of steroid hormone receptors. Nature 7;331(6151): 91-94, 1988.

Glass CK. Differential recognition of target genes by nuclear receptor monomers, dimers, and heterodimers. Endocr Rev. 15(3): 391-407, 1994.

Gómez-Puertas P, Albo C, Pérez-Pastrana E, Vivo A, Portela A. Influenza virus matrix protein is the major driving force in virus budding. J Virol. 74(24): 11538-11547, 2000.

Goodbourn S, Didcock L, Randall RE. Interferons: cell signalling, immune modulation, antiviral response and virus countermeasures. J Gen Virol. 81: 2341-2364, 2000.

Gorden KK, Qiu X, Battiste JJ, Wightman PP, Vasilakos JP, Alkan SS. Oligodeoxynucleotides differentially modulate activation of TLR7 and TLR8 by imidazoquinolines. J Immunol. 177(11): 8164-8170, 2006. 94

Greenberg AK, Hu J, Basu S, Hay J, Reibman J, Yie TA, Tchou-Wong KM, Rom WN, Lee TC. Glucocorticoids inhibit lung cancer cell growth through both the extracellular signal-related kinase pathway and cell cycle regulators. Am J Respir Cell Mol Biol. 27(3): 320-328, 2002.

Guillot L, Goffic R, Bloch S, Escriou N, Akira S, Chignard M, and Si-Tahar M. Involvement of Toll-like Receptor 3 in the Immune Response of Lung Epithelial Cells to Double-stranded RNA and Influenza A Virus. J Biol Chem. 280(7): 5571-5580, 2005.

Gulati U, Wu W, Gulati S, Kumari K, Waner JL, Air GM. Mismatched hemagglutinin and neuraminidase specificities in recent human H3N2 influenza viruses. Virology. 339(1): 12-20, 2005.

Guo Z, Chen LM, Zeng H, Gomez JA, Plowden J, Fujita T, Katz JM, Donis RO, Sambhara S. NS1 protein of influenza A virus inhibits the function of intracytoplasmic pathogen sensor, RIG-I. Am J Respir Cell Mol Biol. 36(3): 263-269, 2007.

Hager GL, Lim CS, Elbi C, Baumann CT. Trafficking of nuclear receptors in living cells. J Steroid Biochem Mol Biol. 74(5): 249-254, 2000.

Haller O, Kochs G, Weber F. Interferon, Mx, and viral countermeasures. Cytokine & Growth Factor Reviews 18: 425-433, 2007.

Haller O, Weber F. Pathogenic viruses: smart manipulators of the interferon system. Curr Top Microbiol Immunol. 316: 315-334, 2007.

Hattermann K, Picard S, Borgeat M, Leclerc P, Pouliot M, Borgeat P. The Toll-like receptor 7/8-ligand resiquimod (R-848) primes human neutrophils for leukotriene B4, prostaglandin E2 and platelet-activating factor biosynthesis. FASEB J. (7): 1575-1585, 2007.

Haye K, Burmakina S, Moran T, García-Sastre A, Fernandez-Sesma A. The NS1 protein of a human influenza virus inhibits type I interferon production and the induction of antiviral responses in primary human dendritic and respiratory epithelial cells. J Virol. 83(13): 6849-6862, 2009.

Head C.C., Farrow M.J., Sheridan J.F., and Padgett D.A. Androstenediol reduces the anti-inflammatory effects of restraint stress during wound healing. Brain, Behavior and Immunity. 20: 590-596, 2006.

Hedman E, Widén C, Asadi A, Dinnetz I, Schröder WP, Gustafsson JA, Wikström AC. Proteomic identification of glucocorticoid receptor interacting proteins. Proteomics 6(10): 3114-26, 2006. 95

Hoebe K., Du X., Georgel E., Janssen, K. Tabeta, S. O. Kim, J. Goode, P. Lin, N. Mann, S. Mudd. Identification of Lps2 as a key transducer of MyD88-independent TIR signalling. Nature 424: 743-748, 2003.

Honda A, Okamoto T, Ishihama A. Host factor Ebp1: Selective inhibitor of influenza virus transcriptase. Genes to Cells 12: 133-142, 2007.

Honda A. Role of host protein Ebp1 in influenza virus growth: Intracellular localization of Ebp1 in virus-infected and uninfected cells. Journal of Biotechnology 133: 208-212, 2008.

Hui EK, Yap EM, An DS, Chen IS, Nayak DP. Inhibition of influenza virus matrix (M1) protein expression and virus replication by U6 promoter-driven and lentivirus-mediated delivery of siRNA. J Gen Virol. 85(Pt 7): 1877-1884, 2004.

Hurteau GJ, Carlson JA, Spivack SD, Brock GJ. Overexpression of the microRNA hsa- miR-200c leads to reduced expression of transcription factor 8 and increased expression of E-cadherin. Cancer Res. 67(17): 7972-7976, 2007.

Ibricevic A, Pekosz A, Walter MJ, Newby C, Battaile JT, Brown EG, Holtzman MJ, Brody SL. Influenza virus receptor specificity and cell tropism in mouse and human airway epithelial cells. J Virol. 80(15): 7469-7480, 2006.

Ichinohe T, Lee HK, Ogura Y, Flavell R, Iwasaki A. Inflammasome recognition of influenza virus is essential for adaptive immune responses. J Exp Med. 206(1): 79-87, 2009.

Iwasaki A, Medzhitov R. Regulation of adaptive immunity by the innate immune system. Science 327(5963): 291-5, 2010.

Iwashima M, Takamatsu M, Yamagishi H, Hatanaka Y, Huang YY, McGinty C, Yamasaki S, Koike T. Genetic evidence for Shc requirement in TCR-induced c-Rel nuclear translocation and IL-2 expression. Proc Natl Acad Sci U S A 99(7): 4544-4549, 2002.

Jang BC, Lim KJ, Suh MH, Park JG, Suh SI. Dexamethasone suppresses interleukin- 1beta-induced human beta-defensin 2 mRNA expression: involvement of p38 MAPK, JNK, MKP-1, and NF-kappaB transcriptional factor in A549 cells. FEMS Immunol Med Microbiol. 51(1): 171-84, 2007.

Jasavala R, Martinez H, Thumar J, Andaya A, Gingras AC, Eng JK, Aebersold R, Han DK, Wright ME. Identification of putative androgen receptor interaction protein modules: cytoskeleton and endosomes modulate androgen receptor signaling in prostate cancer cells. Mol Cell Proteomics. 6(2):252-71, 2007. 96

Jenster G, Trapman J, Brinkmann AO. Nuclear import of the human androgen receptor. Biochem J. 293 (Pt 3): 761-768, 1993.

Julkunen I, Sareneva T, Pirhonen J, Ronni T, Melen K, Matikainen S. Molecular pathogenesis of influenza A virus infection and virus-induced regulation of cytokine gene expression. Cytokine Growth Factor Rev. 12(2-3): 171-180, 2001.

Kang Z, Pirskanen A, Janne OA, Palvimo JJ. Involvement of proteasome in the dynamic assembly of the androgen receptor transcription complex. J Biol Chem. 277: 48366– 48371, 2002.

Karlas A, Machuy N, Shin Y, Pleissner KP, Artarini A, Heuer D, Becker D, Khalil H, Ogilvie LA, Hess S, Mäurer AP, Müller E, Wolff T, Rudel T, Meyer TF. Genome-wide RNAi screen identifies human host factors crucial for influenza virus replication. Nature. 463(7282): 818-822, 2010.

Kawai T, Akira S. Antiviral signaling through pattern recognition receptors. J Biochem. 141(2): 137-145, 2007.

Kawai T, Akira S. Toll-like receptor and RIG-I-like receptor signaling. Ann N Y Acad Sci. 1143: 1-20, 2008.

Kawai T, Akira S. The role of pattern-recognition receptors in innate immunity: update on Toll-like receptors. Nature Immunology 11: 373–384, 2010.

Kilbourne ED, Laver WG, Schulman JL, Webster RG. Antiviral activity of antiserum specific for an influenza virus neuraminidase. J Virol 2: 281–288, 1968.

Kimura-Takeuchi M, Majde JA, Toth LA, Krueger JM. The role of double-stranded RNA in induction of the acute-phase response in an abortive influenza virus infection model. J Infect Dis. 166(6): 1266-1275, 1992.

Kleinert H, Euchenhofer C, Ihrig-Biedert I, Förstermann U. Glucocorticoids inhibit the induction of nitric oxide synthase II by down-regulating cytokine-induced activity of transcription factor nuclear factor-kappa B. Mol Pharmacol. 49(1): 15-21, 1996.

König R, Stertz S, Zhou Y, Inoue A, Hoffmann HH, Bhattacharyya S, Alamares JG, Tscherne DM, Ortigoza MB, Liang Y, Gao Q, Andrews SE, Bandyopadhyay S, De Jesus P, Tu BP, Pache L, Shih C, Orth A, Bonamy G, Miraglia L, Ideker T, García-Sastre A, Young JA, Palese P, Shaw ML, Chanda SK. Human host factors required for influenza virus replication. Nature. 463(7282): 813-817, 2010.

97

Krug RM, Yuan W, Noah DL, Latham AG. Intracellular warfare between human influenza viruses and human cells: the roles of the viral NS1 protein. Virology. 309(2): 181-189, 2003.

Kuiper GG, Carlsson B, Grandien K, Enmark E, Häggblad J, Nilsson S, Gustafsson JA. Comparison of the ligand binding specificity and transcript tissue distribution of estrogen receptors alpha and beta. Endocrinology. 138(3): 863-870, 1997.

Kumari K, Gulati S, Smith DF, Gulati U, Cummings RD, Air GM. Receptor binding specificity of recent human H3N2 influenza viruses. Virology. 9; 4: 42, 2007.

Kurokawa M, Koyama AH, Yasuoka S, Adachi A. Influenza virus overcomes apoptosis by rapid multiplication. International Journal of Molecular Medicine 3, 527-530, 1999.

Le Goffic R, Pothlichet J, Vitour D, Fujita T, Meurs E, Chignard M, and Si-Tahar M. Cutting Edge: Influenza A Virus Activates TLR3-Dependent Inflammatory and RIG-I- Dependent Antiviral Responses in Human Lung Epithelial Epithelial Cells. J Immunol. 178: 3368-3372, 2007.

Lee MS, Kim YJ. Signaling pathways downstream of pattern-recognition receptors and their cross talk. Annu Rev Biochem. 76: 447-480, 2007.

Li H, Ding J, Chen YH. Recombinant protein comprising multi-neutralizing epitopes induced high titer of antibodies against influenza A virus. Immunobiology. 207(5): 305- 313, 2003.

Lindemann MJ, Benczik M, Gaffen SL.Anti-apoptotic signaling by the interleukin-2 receptor reveals a function for cytoplasmic tyrosine residues within the common gamma (gamma c) receptor subunit. J Biol Chem. 278(12): 10239-10249, 2003.

Liu X, Sun L, Yu M, Wang Z, Xu C, Xue Q, Zhang K, Ye X, Kitamura Y, Liu W. Cyclophilin A interacts with influenza A virus M1 protein and impairs the early stage of the viral replication. Cell Microbiol. 11(5): 730-741, 2009. Loria RM, Padgett DA. Androstenediol regulates systemic resistanceagainst lethal infections in mice. Arch Virol. 127: 103–115, 1991.

Loria RM, Padgett DA, Huynh PN. Regulation of the immune response by dehydroepiandrosterone and its metabolites. J Endocrinol. 150 Suppl: S209-220, 1996.

Lowen AC, Mubareka S, Steel J, Palese P. Influenza virus transmission is dependent on relative humidity and temperature. PLoS Pathogens 3: 1470-1476, 2007.

Ludwig S, Wolff T, Ehrhardt C, et al. MEK inhibition impairs propagation without emergence of resistant variants. FEBS Lett 561: 37–43, 2004. 98

Ludwig S, Pleschka S, Planz O, Wolff T. Ringing the alarm bells: signalling and apoptosis in influenza virus infected cells. Cell Microbiol. 8(3): 375-386, 2006. Mangelsdorf DJ, Thummel C, Beato M, Herrlich P, Schutz G, Umesono K, Blumberg B, Kastner P, Mark M, Chambon P, Evans RM. The nuclear receptor superfamily: the second decade. Cell 83: 835–839, 1995. Matsumoto M, Funami K, Oshiumi H, Seya T. Toll-like receptor 3: a link between toll like receptor, interferon and viruses. Microbiol Immunol. 48(3): 147-154, 2004.

Mazur I, Anhlan D, Mitzner D, Wixler L, Schubert U, Ludwig S. The proapoptotic influenza A virus protein PB1-F2 regulates viral polymerase activity by interaction with the PB1 protein. Cell Microbiol. 10(5): 1140-1152, 2008.

McGill J, Heusel JW, Legge KL. Innate immune control and regulation of influenza virus infections. J Leukoc Biol. 86(4):803-12, 2009.

Meylan E, Tschopp J, Karin M. Intracellular pattern recognition receptors in the host response. Nature. 442(7098): 39-44, 2006.

Mikkonen L, Pihlajamaa P, Sahu B, Zhang FP, Jänne OA. Androgen receptor and androgen-dependent gene expression in lung. Mol Cell Endocrinol. 12; 317(1-2): 14-24, 2010.

Min JY, Krug RM. The primary function of RNA binding by the influenza A virus NS1 protein in infected cells: inhibiting the 2'–5' oligo (A) synthetase/RNase L pathway. Proc Natl Acad Sci USA 103: 7100–7105, 2006.

Momose F, Naito T, Yano K, Sugimoto S, Morikawa Y, Nagata K. Identification of Hsp90 as a Stimulatory Host Factor Involved in Influenza Virus RNA Synthesis. J Biol Chem. 277(47): 45306-45314, 2002.

Monto AS, Gravenstein S, Elliott M, Colopy M, Schweinle J. Clinical Signs and Symptoms Predicting Influenza Infection. Arch Intern Med. 160(21): 3243-3247, 2000.

Morris SJ, Nightingale K, Smith H, Sweet C. Influenza A virus-induced apoptosis is a multifactorial process: exploiting reverse genetics to elucidate the role of influenza A virus proteins in virus-induced apoptosis. Virology. 335(2): 198-211, 2005.

Mozdzanowska K, Furchner M, Washko G, Mozdzanowski J, Gerhard W. A pulmonary influenza virus infection in SCID mice can be cured by treatment with hemagglutinin- specific antibodies that display very low virus-neutralizing activity in vitro. J Virol. 71(6): 4347-4355, 1997.

99

Mukaida N, Morita M, Ishikawa Y, Rice N, Okamoto S, Kasahara T, Matsushima K. Novel mechanism of glucocorticoid-mediated gene repression. Nuclear factor-kappa B is target for glucocorticoid-mediated interleukin 8 gene repression. J Biol Chem. 269(18): 13289-13295, 1994.

Murphy BR, Kasel JA, Chanock RM. Association of serum anti-neuraminidase antibody with resistance to influenza in man. New Engl J Med. 286: 1329–1332, 1972.

Naito T, Momose F, Kawaguchi A, Nagata K. Involvement of Hsp90 in Assembly and Nuclear Import of Influenza Virus RNA Polymerase Subunits. J Virol 81(3): 1339-1349, 2007.

Nelson MI, Holmes EC. The evolution of epidemic influenza. Nat Rev Genet. 8(3): 196- 205, 2007.

Nelson PS, Clegg N, Arnold H, Ferguson C, Bonham M, White J, Hood L, Lin B. The program of androgen-responsive genes in neoplastic prostate epithelium. Proc Natl Acad Sci U S A 99(18): 11890-11895, 2002.

Nemeroff ME, Barabino SM, Li Y, Keller W, Krug RM. Influenza virus NS1 protein interacts with the cellular 30 kDa subunit of CPSF and inhibits 3'end formation of cellular pre-mRNAs. Molecular Cell 1(7): 991-1000, 1998.

Newby CM, Sabin L, Pekosz A. The RNA binding domain of influenza A virus NS1 protein affects secretion of tumor necrosis factor alpha, interleukin-6, and interferon in primary murine tracheal epithelial cells. J Virol. 81: 9469–9480, 2007.

Ohtsuka T, Kubota A, Hirano T, Watanabe K, Yoshida H, Tsurufuji M, Iizuka Y, Konishi K, Tsurufuji S. Glucocorticoid-mediated gene suppression of rat cytokine- induced neutrophil chemoattractant CINC/gro, a member of the interleukin-8 family, through impairment of NF-kappa B activation. J Biol Chem. 271(3): 1651-1659, 1996.

Olschlager V, Pleschka S, Fischer T, et al. Lung-specific expression of active Raf kinase results in increased mortality of influenza A virus-infected mice. Oncogene 23: 6639– 6646, 2004.

Oshiumi, H., M. Matsumoto, K. Funami, T. Akazawa, and T. Seya. TICAM-1, an adaptor molecule that participates in Toll-like receptor 3-mediated interferon-induction. Nat. Immunol. 4: 161-167, 2003.

Padgett DA, Loria RM, Sheridan JF. Endocrine regulation of the immune response to influenza virus infection with a metabolite of DHEA-androstenediol. J Neuroimmunol 78(1-2): 203-211, 1997.

100

Padgett DA, Loria RM, Sheridan JF. Steroid hormone regulation of antiviral immunity. Ann N Y Acad Sci 917: 935-943, 2000.

Padgett DA, MacCallum RC, Loria RM, Sheridan JF. Androstenediol-induced restoration of responsiveness to influenza vaccination in mice. J Gerontol A Biol Sci Med Sci. 55(9): B418-424, 2000.

Padgett DA, MacCallum RC, Sheridan JF. Stress exacerbates age-related decrements in the immune response to an experimental influenza viral infection. J Gerontol A Biol Sci Med Sci. 53(5): B347-353, 1998.

Padgett DA, Sheridan JF. Androstenediol (AED) prevents neuroendocrine-mediated suppression of the immune response to an influenza viral infection. J Neuroimmunol. 98(2): 121-129, 1999.

Paschal BM. Translocation through the nuclear pore complex. Trends Biochem Sci 27(12): 593-596, 2002.

Picard D, Yamamoto K. Two signals mediate hormone-dependent nuclear localization of glucocortoid receptor. EMBO J. 6, 3333-3340, 1987.

Pichlmar A and Reis e Sousa C. Innate Recognition of Viruses. Immunity 27: 370-383, 2007.

Pleschka S, Wolff T, Ehrhardt C, et al. Influenza virus propagation is impaired by inhibition of the Raf/MEK/ERK signalling cascade. Nat Cell Biol. 3: 301-305, 2001.

Powers DC, Kilbourne ED, Johansson BE. Neuraminidase-specific antibody responses to inactivated influenza virus vaccine in young and elderly adults. Clin Diagn Lab Immunol. 3: 511–516, 1996.

Pratt, WB, Galigniana MD, Morishima Y, Murphy PJ. Role of molecular chaperones in steroid receptor action. Essays Biochem. 40: 41-58, 2004.

Provost PR, Blomquist CH, Godin C, Huang XF, Flamand N, Luu-The V, Nadeau D, Tremblay Y. Androgen formation and metabolism in the pulmonary epithelial cell line A549: expression of 17beta-hydroxysteroid dehydrogenase type 5 and 3alpha- hydroxysteroid dehydrogenase type 3. Endocrinology. 141(8): 2786-2794, 2000.

Rehwinkel J, Reis e Sousa C. RIGorous detection: exposing virus through RNA sensing. Science. 327(5963): 284-286, 2010.

Rello J, Pop-Vicas A. Clinical review: primary influenza viral pneumonia. Crit Care. 13(6): 235, 2009. 101

Rennie PS, Bruchovsky N, Leco KJ, Sheppard PC, McQueen SA, Cheng H, Snoek R, Hamel A, Bock ME, MacDonald BS, Nickel BE, Chang C, Liao S, Cattini PA and Matusik RJ. Characterization of two cis-acting DNA elements involved in the androgen regulation of the probasin gene. Mol Endo. 7: 23-36, 1993.

Rimmelzwaan GF, Baars MM, de Lijster P, Fouchier RA, Osterhaus AD. Inhibition of influenza virus replication by nitric oxide. J Virol. 73(10): 8880-8883, 1999.

Robinson-Rechavi M, Escriva Garcia H, Laudet V. The nuclear receptor superfamily. J Cell Sci. 116(4): 585-586, 2003.

Roy P, Franks S, Read M, Huhtaniemi IT. Determination of androgen bioactivity in human serum samples using a recombinant cell based in vitro bioassay. J Steroid Biochem Mol Biol. 101(1): 68-77, 2006.

Saelens X, Kalai M, Vandenabeele P. Translation inhibition in apoptosis: caspase- dependent PKR activation and eIF2-alpha phosphorylation. J Biol Chem. 276(45): 41620-41628, 2001.

Samuel CE. Antiviral actions of interferons. Clin Microbiol Rev. 14(4): 778-809, 2001.

Schmolke M, García-Sastre A. Evasion of innate and adaptive immune responses by influenza A virus. Cell Microbiol. 12(7): 873-80, 2010.

Seth RB, Sun L, Chen ZJ. Antiviral innate immunity pathways. Cell Res. 16(2): 141- 147, 2006.

Shaffer PL, Jivan A, Dollins DE, Claessens F, Gewrith D. Structural basis of androgen receptor binding to selective androgen response elements. Proc Natl Acad Sci U S A. 101(14): 4758-4763, 2004.

Shapira SD, Gat-Viks I, Shum BO, Dricot A, de Grace MM, Wu L, Gupta PB, Hao T, Silver SJ, Root DE, Hill DE, Regev A, Hacohen N. A physical and regulatory map of host-influenza interactions reveals pathways in H1N1 infection. Cell 139(7): 1255-1267, 2009.

Shieh WJ, Blau DM, Denison AM, Deleon-Carnes M, Adem P, Bhatnagar J, Sumner J, Liu L, Patel M, Batten B, Greer P, Jones T, Smith C, Bartlett J, Montague J, White E, Rollin D, Gao R, Seales C, Jost H, Metcalfe M, Goldsmith CS, Humphrey C, Schmitz A, Drew C, Paddock C, Uyeki TM, Zaki SR. 2009 pandemic influenza A (H1N1): pathology and pathogenesis of 100 fatal cases in the United States. Am J Pathol. 177(1): 166-75, 2010.

102

Short, JAL. Viral evasion of interferon stimulated genes. Bioscience Horizons 2(2): 212- 224, 2009.

Simental JA, Sar M, Lane MV, French FS, Wilson EM. Transcriptional activation and nuclear targeting signals of the human androgen receptor. J Biol Chem. 266: 510-518, 1991.

Sirén J, Imaizumi T, Sarkar D, Pietilä T, Noah DL, Lin R, Hiscott J, Krug RM, Fisher PB, Julkunen I, Matikainen S. Retinoic acid inducible gene-I and mda-5 are involved in influenza A virus-induced expression of antiviral cytokines. Microbes Infect. 8(8):2013- 20, 2006.

Stertz S, Shaw ML. Uncovering the global host cell requirements for influenza virus replication via RNAi screening. Microbes Infect. 13(5):516-25, 2011 .

Takahashi T, Moriyama Y, Ikari A, Sugatani J, Suzuki T, Miwa M. Surface localization of the nuclear receptor CAR in influenza A virus-infected cells. Biochem Biophys Res Commun. 368(3): 550-555, 2008.

Takeuchi O, Akira S. Pattern recognition receptors and inflammation. Cell 140(6): 805- 820, 2010.

Taubenberger JK, Morens DM. The pathology of influenza virus infections. Annu Rev Pathol. 3: 499-522, 2008.

Thompson CI, Barclay WS, Zambon MC, Pickles RJ. Infection of human airway epithelium by human and avian strains of influenza a virus. J Virol. 80(16): 8060-8068, 2006.

Tissari J, Sirén J, Meri S, Julkunen I, Matikainen S. IFN-alpha enhances TLR3-mediated antiviral cytokine expression in human endothelial and epithelial cells by up-regulating TLR3 expression. J Immunol. 174: 4289-4294, 2005.

Treanor J. Influenza vaccine--outmaneuvering antigenic shift and drift. N Engl J Med. 350(3): 218-220, 2004.

Uzé G, Schreiber G, Piehler J, Pellegrini S. The receptor of the type I interferon family. Curr Top Microbiol Immunol. 316: 71-95, 2007. van Elden, L. J., Nijhuis, M., Schipper, P., Schuurman, R. & van Loon, A. M. Simultaneous detection of influenza viruses A and B using real-time quantitative PCR. J Clin Microbiol. 39, 196–200, 2001.

103

Vester, D., E. Rapp, D. Gade, Y. Genzel, and U. Reichl. Quantitative analysis of cellular proteome alterations in human influenza A virus-infected mammalian cell lines. Proteomics 9: 3316-3327, 2009.

Verrijdt G, Schoenmakers E, Alen P, Haelens A, Peeters B, Rombauts W, Claessens F. Androgen specificity of a response unit upstream of the human secretory component gene is mediated by differential receptor binding to an essential androgen response element. Mol Endocrinol. 13(9): 1558-1570, 1999.

Verrijdt G, Schoenmakers E, Haelens A, Peeters B, Verhoeven G, Rombauts W, Claessens F. Change of specificity mutations in androgen-selective enhancers. Evidence for a role of differential DNA binding by the androgen receptor. J Biol Chem. 275(16): 12298-12305, 2000.

Vila-del Sol V, Díaz-Muñoz MD, Fresno M. Requirement of tumor necrosis factor alpha and nuclear factor-kappaB in the induction by IFN-gamma of inducible nitric oxide synthase in macrophages.J Leukoc Biol. 81(1): 272-283, 2007.

Wada T, Gao J, Xie W. PXR and CAR in energy metabolism. Trends Endocrinol Metab. 20(6): 273-279, 2009.

Wang J, Oberley-Deegan R, Wang S, Nikrad M, Funk CJ, Hartshorn KL, Mason RJ. Differentiated human alveolar type II cells secrete antiviral IL-29 (IFN-lambda 1) in response to influenza A infection. J Immunol. 182(3): 1296-1304, 2009.

Wang P, Song W, Mok BW, Zhao P, Qin K, Lai A, Smith GJ, Zhang J, Lin T, Guan Y, Chen H. Nuclear factor 90 negatively regulates influenza virus replication by interacting with viral nucleoprotein. J Virol. 83(16): 7850-7861, 2009.

Wasilenko JL, Lee CW, Sarmento L, Spackman E, Kapczynski DR, Suarez DL, Pantin- Jackwood MJ. NP, PB1, and PB2 viral genes contribute to altered replication of H5N1 avian influenza viruses in chickens. J Virol. 82(9): 4544-4553, 2008.

Widén C, Gustafsson JA, Wikström AC. Cytosolic glucocorticoid receptor interaction with nuclear factor-kappa B proteins in rat liver cells. Biochem J. 373(Pt 1): 211-220, 2003.

Wong CI, Zhou ZX, Sar M, Wilson EM. Steroid requirement for androgen receptor dimerization and DNA binding. Modulation by intramolecular interactions between the NH2-terminal and steroid-binding domains. J Biol Chem. 5;268(25): 19004-12, 1993.

Writing Committee of the Second World Health Organization Consultation on Clinical Aspects of Human Infection with Avian Influenza A (H5N1) Virus. Update on Avian Influenza A (H5N1) Virus Infection in Humans. N Engl J Med. 358: 261-273, 2008. 104

Wurzer WJ, Planz O, Ehrhardt C, Giner M, Silberzohn T, Pleschka S, Ludwig S. Caspase 3 activation is essential for efficient influenza virus propagation. EMBO J. 22: 2717- 2728, 2003.

Wurzer WJ, Ehrhardt C, Pleschka S, Berberich-Siebelt F, Wolff T, Walczak H, Planz O, Ludwig S. NF-κB dependent induction of TRAIL and Fas/FasL is crucial for efficient influenza virus propogation. J Biol Chem. 279: 30931-30937, 2004.

Yamamoto M, Sato S, Hemmi H, Hoshino K, Kaisho T, Sanjo H, Takeuchi O, Sugiyama M, Okabe M, Takeda K, Akira S. Role of adaptor TRIF in the MyD88-independent toll- like receptor signaling pathway. Science 301(5633): 640-643, 2003.

Yu WC, Chan RW, Wang J, Travanty EA, Nicholls JM, Peiris JS, Mason RJ, Chan MC. Viral replication and innate host responses in primary human alveolar epithelial cells and alveolar macrophages infected with influenza H5N1 and H1N1 viruses. J Virol. Epub May 4, 2011.

Zhang L, Bukreyev A, Thompson CI, Watson B, Peeples ME, Collins PL, Pickles RJ. Infection of ciliated cells by human parainfluenza virus type 3 in an in vitro model of human airway epithelium. J Virol. 79(2): 1113-1124, 2005.

Zhang Y, Fondell JD, Wang Q, Xia X, Cheng A, Lu ML, Hamburger AW. Repression of androgen receptor mediated transcription by the ErbB-3 binding protein, Ebp1. Oncogene 21: 56090-56108, 2002.

Zhou ZX, Sar M, Simental JA, Lane MV, Wilson EM. A ligand-dependent bipartite nuclear targeting signal in the human androgen receptor. Requirement for the DNA- binding domain and modulation by NH2-terminal and carboxyl-terminal sequences. J. Biol. Chem. 269: 13115-13123, 1994.

105